Vous êtes sur la page 1sur 341

Dr. Andrew A.

Snelling
Education
PhD, Geology, University of Sydney, Sydney, Australia, 1982
BSc, Applied Geology, The University of New South Wales, Sydney,
Australia, First Class Honours, 1975
Professional Experience

Field, mine, and research geologist, various mining companies,


Australia

, Australian Nuclear Science and Technology Organisation


(ANSTO), Consultant researcher and writer , Australia, 19831992

Geological consultant, Koongarra uranium project, Denison Australia PL, 19831992

Collaborative researcher and writer, Commonwealth Scientific and Industrial Organisation


(CSIRO), Australia, 19811987

Professor of geology, Institute for Creation Research, San Diego, CA, 19982007

Staff member, Creation Science Foundation (later Answers in GenesisAustralia), Australia, 1983
1998

Founding editor, Creation Ex Nihilo Technical Journal (now Journal of Creation), 19841998

Researcher and editor, Radioisotopes and the Age of The Earth (RATE), 19972005

Editor-in-chief, Proceedings of the Sixth International Conference on Creationism, 2008

Director of Research, Answers in Genesis, Petersburg, KY, 2007present

Professional Affiliations
Geological Society of Australia /Geological Society of America /Geological Association of Canada/
Mineralogical Society of America /Society of Economic Geologists /Society for Geology Applied to Mineral
Deposits / Creation Research Society /Creation Geology Society
Dr. Andrew A. Snelling is perhaps one of the world's leading researchers in flood geology.He worked for a
number of years in the mining industry throughout Australia undertaking mineral exploration surveys and field
research. He has also been a consultant research geologist for more than a decade to the Australian Nuclear
Science and Technology Organization and the US Nuclear Regulatory Commission for internationally
funded research on the geology and geochemistry of uranium ore deposits as analogues of nuclear waste
disposal sites..His primary research interests include radioisotopic methods for the dating of rocks, formation of
igneous and metamorphic rocks, and ore deposits. He is one of a controlled number permitted to take rock
samples from the Grand Canyon.He was also a founding member of the RATE group (Radioisotopes and the
Age of The Earth).
Andrew completed a Bachelor of Science degree in Applied Geology with First Class Honours at The University
of New South Wales in Sydney, and graduated a Doctor of Philosophy (in geology) at The University of Sydney,
for his thesis entitled A geochemical study of the Koongarra uranium deposit, Northern Territory,
Australia. Between studies and since, Andrew worked for six years in the exploration and mining
industries in Tasmania, New South Wales, Victoria, Western Australia and the Northern Territory variously as a
field, mine and research geologist. Full-time with the Australian creation ministry from 1983 to 1998, he was
during this time also called upon as a geological consultant to the Koongarra uranium project (19831992).
Consequently, he was involved in research projects with various CSIRO (Commonwealth Scientific and
Industrial Research Organisation), ANSTO (Australian Nuclear Science and Technology Organisation)
and University scientists across Australia, and with scientists from the USA, Britain, Japan, Sweden and the
International Atomic Energy Agency. As a result of this research, Andrew was involved in writing scientific
papers that were published in international scientific journals.Andrew has been involved in extensive
creationist research in Australia and overseas, including the formation of all types of mineral deposits,
radioactivity in rocks and radioisotopic dating, and the formation of metamorphic and igneous rocks,
sedimentary strata and landscape features (e.g. Grand Canyon, USA, and Ayers Rock, Australia) within the
creation framework for earth history. As well as writing regularly and extensively in international creationist
publications, Andrew has travelled around Australia and widely overseas (USA, UK, New Zealand, South Africa,
Korea, Indonesia, Hong Kong, China) speaking in schools, churches, colleges and universities, particularly on
the overwhelming scientific evidence consistent with the Global Flood and the Creation.

THE RADIOMETRIC DATING

Radioisotopes and the Age of the Earth 4


Radiometric Dating: Back to Basics 9
Radiometric Dating: Problems with the Assumptions ..11
Radiometric Dating: Making Sense of the Patterns .13
Radioactive Dating Failure Recent New Zealand Lava Flows Yield Ages of Millions of Years ...14
Radioactive Dating Method Under Fire 16
RATE Radioisotopes and the Age of the Earth .19
U-Th-Pb Dating: An Example of False Isochrons 20
The Failure of U-Th-Pb Dating at Koongarra, Australia .25
Determination of the Radioisotope Decay Constants and Half-Lives: Rubidium-87 (87Rb) ..43
***

CARBON DATING

Carbon-14 Dating Understanding the Basics 50


Carbon-14 in Fossils and Diamonds An Evolution Dilemma ...51
A Creationist Puzzle 50,000-Year-Old-Fossils ..53
Measurable 14C in Fossilized Organic Materials: Conrming the Young Earth Creation-Flood Model 54
Radiocarbon Ages for Fossil Ammonites and Wood in Cretaceous Strata near Redding, California ..63
Radiocarbon in Diamonds Confirmed ..71
Geological Conflict Young Radiocarbon Date for Ancient Fossil Wood Challenges Fossil Dating.72
Radiocarbon in an ancient fossil tree stump casts doubt on traditional rock/fossil dating ....75

Determination of the Radioisotope Decay Constants and Half-Lives: Lutetium-176 (176Lu).76

***
Determination of the Radioisotope Decay Constants and Half-Lives: Rhenium-187 (187Re)85
Determination of the Radioisotope Decay Constants and Half-Lives: Samarium-147 (147Sm).95

DATING RESULTS

Significance of Highly Discordant Radioisotope Dates for Precambrian Amphibolites in Grand Canyon, USA
..105
Whole-Rock K-Ar Model and Isochron, and Rb-Sr, Sm-Nd, and Pb-Pb Isochron, Dating of the Somerset
Dam Layered Mafic Intrusion, Australia
..117
Radioisotopes in the Diabase Sill (Upper Precambrian) at Bass Rapids, Grand Canyon, Arizona An
Application and Test of the Isochron Dating Method
.130
Discordant Potassium-Argon Model and Isochron Ages for Cardenas Basalt (Middle Proterozoic) and
Associated Diabase of Eastern Grand Canyon, Arizona
..140
The Relevance of Rb-Sr, Sm-Nd, and Pb-Pb Isotope Systematics to Elucidation of the Genesis and History of
Recent Andesite Flows at Mt. Ngauruhoe, New Zealand, and the Implications for Radioisotopic Da
.150
The Cause of Anomalous Potassium-Argon Ages for Recent Andesite Flows at Mt. Ngauruhoe, New
Zealand, and the Implications for Potassium-Argon Dating
.161
The Fallacies of Radioactive Dating of Rocks Basalt Lava Flows in Grand Canyon .172
Radioisotope Dating of Rocks in the Grand Canyon 174
Radioisotope Dating of Rocks in the Grand Canyon 175
Radioactive Dating in Conflict! Fossil Wood in Ancient Lava Flow Yields Radiocarbon 178
The age of Australian Uranium A case study of the Koongarra uranium deposit .179
Helium Diffusion Rates Support Accelerated Nuclear Decay ..183

METEORS

Radioisotope Dating of Meteorites: I 195


Radioisotope Dating of Meteorites: II ..239
Radioisotope Dating of Meteorites: III ..277

Radioisotopes and the Age of the Earth


by Dr. Russell Humphreys, Dr. Steve Austin, Dr. Don DeYoung, Eugene Chafn, Dr. Andrew A. Snelling, and Dr. John
Baumgardneron March 9, 2011
Abstract
RATE is an acronym applied to a research project investigating
radioisotope dating sponsored by the Institute for Creation Research
and the Creation Research Society. It stands for Radioisotopes and
the Age of The Earth. This article summarizes the purpose, history,
and intermediate findings of the RATE project five years into an eightyear effort. It reports on the latest status of the research on helium
diffusion through minerals in granitic rock, accelerated nuclear decay theory, radiohalos, isochron discordance studies, case
studies in rock dating, and carbon-14 in deep geologic strata. Each of the RATE scientists will present separate technical
papers at the Fifth International Conference on Creationism on the details of this research.
Keywords: radioisotopes, isotopes, age, dating, nuclear decay, accelerated nuclear decay
This paper was originally published in the Proceedings of the Fifth International Conference on Creationism, pp.337348
(2003) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh (www.csfpittsburgh.
org).
Introduction
The conventional scientific view typically expressed today is that the earth is about 4.6 billion years old and the universe
between 10 and 20 billion years old. These estimates are based primarily on the abundances of parent and daughter
radioisotopes and the implications of stellar and cosmological models. Yet, a literal interpretation of Scripture and much
scientific evidence has been gathered to indicate that the creation of the earth, the solar system, and the universe occurred
only a few thousand years ago.One of the principal forces which has traditionally driven estimates of an old age for the earth
is the necessity for long periods of time for evolution. Even before radioactivity was discovered in the 1890s, estimates of
the age of the earth were growing longer and longer as the complex nature of life became more evident. However, it has
never been demonstrated that the evolution of life from inorganic chemicals has occurred or that life has evolved from
simple life forms to the complex ones we see today. Living systems, even the simplest ones, are based upon symbolic
language structures of extreme complexity. There is no hint in the laws of chemistry and physics that matter on its own can
ever generate symbolic language regardless of the time allowed. Because it has no solution to this fundamental difficulty,
evolutionary dogma is now facing a major crisis that long periods of time simply cannot mend.Young-earth creationists on
the other hand are not convinced that long periods of time have transpired since the origin of the earthand some include
the origin of the entire universe. In defending a young-earth position, they typically point to important assumptions
underlying these dating schemes. For example, when a parent isotope decays into a daughter isotope, the initial
concentration of the daughter isotope may affect the estimate of time since the process started. Creationists in some cases
question the conventional assumption that the initial amount of daughter product is small or at least can be tightly
constrained. Isochron methods attempt to remove this uncertainty, but the results are not wholly satisfactory. Also often
questioned by creationists are the assumptions that the quantities of the parent and daughter isotopes have not been
altered by non-radioactive processes such as migration and transport, and that the rate of decay from parent to daughter
has been constant during the period under consideration. Most researchers attempt to justify each of these three
assumptions, but ultimately no one can be certain if the conditions have been met, particularly over long periods of time.
It is hypothesized by the RATE group that at some time in the past much higher rates of radioisotope decay have occurred,
leading to the production of large quantities of daughter products in a short period of time. It has been suggested that these
increased decay rates may have been associated with the rock-forming processes on the early earth.The RATE group
suspects that large amounts of radioactive decay may have occurred during the first two and a half days of creation as part
of the supernatural creation process. The jury is still out and, until we complete our research phase, this thesis remains
tentative. The presence of supernatural process during creation is essential to our approach, however. Scripture talks of at
least two major events which occurred after creation, the Judgment in the Garden of Eden and the Flood. It would seem
appropriate to consider at least that an original distribution of elements could have been mixed, and radioactive processes
speeded up during one or both of these events.
History of the RATE Project
On July 5, 1997 a group of young-earth creationist researchers met in San Diego, California to address the issue of
reconciling radioisotopes and the age of the earth as reported by Vardiman.1 It was recognized by the group that this was a
significant problem which must be addressed if young-earth creationism was to continue to have a significant impact on the
issue of origins both within and outside the creation community. The group, which has since become known as RATE,
decided that the primary approach would be to explore accelerated rates of decay of radioisotopes during one or more of
the Creation, Fall, and Flood events. A second approach would be to address the mixing of mantle and crustal reservoirs
since the origin of the earth. Additional processes and issues have been suggested and explored as part of the research.
The focus of the RATE research would be primarily on long-age isotopes and their use as chronometers.By February, 2003
six annual meetings of RATE had been held by the principal investigators. These meetings included reports, discussion,
allocation of funds, and research decisions. During the third meeting thirteen research experiments were identified as shown
in Tables 1 and 2. A brief description of each experiment, expected results, the estimated cost, and estimated time were
developed. At the annual meeting in 2001 the importance to this project of 14C in deeply buried fossil material was identified,
and a research thrust on this topic was added. The cost of the eight-year project was estimated to be about $500,000.
Before 2002 about 80% of these funds had been raised through private donations. Two major reports were planned for
RATE. The first report, a 675-page book was published in December 2000 entitled, Radioisotopes and the Age of the Earth:
A Young-Earth Creationist Research Initiative. It contains an introduction to the project, a report on the literature searches by
the principal investigators on most of the topics of concern, a glossary, and a set of research proposals. The purpose of the
initial report was to stake a claim. It was also published to lend structure and direction to the effort and to inform contributors
about what they could expect from their donations. The second and final book is planned to be published in 2005 and is

expected to be titled Radioisotopes and the Age of the Earth: A Young-Earth Creationist Research Report. It will report on
the findings of the five-year research phase.
Intermediate Results of RATE Research
Helium DiffusionDr. D. Russell Humphreys, PI
Two decades ago, it was reported by Gentry, Glish, and McBay2 that up to 58% of the helium (a daughter product of
uranium and thorium decay) generated during the alleged 1.5 billion year age of the Precambrian granodiorite beneath the
Jemez Mountains near Los Alamos, New Mexico, was still in the zircons embedded in the biotite crystals contained within
the granodiorite. Yet, the zircons were so small (see Fig. 1 for a picture of typical zircons) that they should not have retained
the helium for even a small fraction of that time. The high retentions suggest to us and many other creationists that the
helium has not had time to diffuse out of the zirconsthat accelerated nuclear decay produced over a billion years worth of
helium only thousands of years ago. Such accelerated decay could reduce the radiometric timescale from gigayears down
to months.A theoretical creationist model, based on observed helium retention, of diffusion rates of helium over a period of
6,000 years was reported by Humphreys3 and Humphreys et al.4). It compares well with laboratory measurements in Jemez
zircons, as shown in Fig. 2. The solid dots show the diffusion coefficient as a function of inverse temperature for the
measurements with the Jemez zircons and the solid lines through empty squares show the theoretical predictions from the
theoretical model. There is a five-order-of-magnitude difference (100,000) between the predictions of diffusion for the
evolutionist and creationist models. The measured diffusion rates of He predict that helium would leak out of a zircon/biotite
matrix in a period of time on the order of thousands of years, not hundreds of millions of years. This is consistent with the
high concentrations of helium still found in the Jemez granodiorite.
Table 1. High priority RATE experiments.
Experiment

Description

Expected Results

Time

He Diffusion

Acquisition of data on which to base a claim that the


amount of He in rocks today should not be so high if it was
produced by nuclear decay over millions of years. If the He
Determine He diffusion rates throughwas produced within the most recent thousands of years, it
minerals under various conditions
would be expected to remain still in the rocks as observed. 2 years

Isochron
Discordance

Increased evidence for discordance among isotopic dating


methods using isochrons for mineral components of FloodConstruct 5-point mineral and whole-rockrelated rocks. Based on the consistency of the discordance
isochrons on selected basaltic rocksfrom these specimens and others, infer the processes
formed during the Flood
which led to the distribution of isotopes.
2 years

Conduct a literature search for evidence


and models of accelerated nuclear decayIncrease evidence that nuclear decay can vary radically in
and adapt to a creationist worldview, if response to changes in cosmological constants and
Nuclear
appropriate. Complete studies on a and benvironmental effects. Associate other likely effects with
Decay Theory decay.
biblical statements and observational data.
2 years

Radiohalos

Determine the geological distribution of Resolve the question if Po halos are special evidence for
Po
halos,
their
proximity
tocreated rocks only , or could they also occur in Flood rocks.
concentrations of U and the relationshipThis effort may also allow inferences about the process of
to different halo types.
radioisotope decay and halo formation.
5 years

Fission
Tracks

Estimate nuclear decay rates during theFission track estimates of nuclear decay rates are thought
Flood using the fission track method.to be absolute following rock formation and do not inherit
Select an initial sample from a tuff bed in prior evidence of decay. It is important to know if decay
the Muav Formation of Grand Canyon. rates were accelerated during the Flood.
2 years

Additional laboratory measurements and modeling studies of helium diffusion in zircon are expected to lead to a further
refinement of the creationist model. The data of Fig. 2 indicate an age between 4,000 and 14,000 years since the helium
began to diffuse from the zircons. This is far short of the 1.5 billion year evolutionist age! We believe that the final results will
resoundingly support our hypothesis concerning diffusion and radiogenic helium.
Table 2. Lower priority RATE experiments.
Experiment
Uranium
Halos

Time (years)
(U)/Thorium

(Th)
4

Case Studies in Rock Dating

Biblical Word Studies

Pu in OKLO Reactor

Allende Meteorite Origin

Diffusion of Ar in Biotite

Origin of Chemical Elements

Cosmology and Nuclear Decay

Search for Carbon-14 *

* Added in 2001
Nuclear Decay TheoryDr. Eugene F. Chaffin, PI

Fig. 1. Zircons from the Muav Tuff, Grand Canyon, Arizona


(Courtesy of Geotrack International Laboratory).

The quantum theory of alpha and beta decay are being reviewed by Chaffin,5,6 with extensions of the standard models
being explored to see if they could lead to accelerated decay during episodic variations of the coupling constants. Variations
in the radii of compactified extra dimensions and consequent variation in coupling constants over the history of the universe
could cause accelerated decay. If, during early Creation Week, say the first 2+ days before the creation of plants, such
variations were to occur, they could lead to accelerated nuclear decay, thus adjusting isotopic abundances, without giving
unacceptable doses of radiation to life.Grasses, herbs, and fruit

trees which could have been damaged by high radiation.


Fig. 2. Plot of diffusion coefficient of He in zircon vs. inverse temperature.
Fig. 3. Plot of nuclear potential energy vs. radial distance from the center of a nucleus.
These variations may help explain the abundances of radioisotopes, including radioactive equilibrium found in decay chains
such as the uranium series, within the young-earth time frame. We are also exploring the tunneling theory of alpha decay to
see how much change in half-life is possible without drastically affecting other measurable properties of nuclei. For only
slight changes in the depth of the nuclear potential well (see Fig. 3), abrupt changes in the number of nodes of the alpha
particle wave function occur which can lead to drastic changes in half-life. Also, the half-life depends exponentially on the
shape of the potential well, so that even slight changes are effective in accelerating alpha-decay. PIThe significance of
radiohalos is due to the fact they represent a physical, integral historical record of the decay of radioisotopes in the
radiocenters over a period of time as discussed by Snelling7 and Snelling and Armitage.8 The darkening of the minerals
surrounding the radiocenters is caused by damage to their crystal structure by alpha particles produced by nuclear decay.
As part of a systematic effort to investigate radiohalo occurrences in granitic rocks globally and throughout the geologic
record, suitable samples have been collected from the La Posta (southern California), Stone Mountain (near Atlanta,
Georgia) and Cooma (southern New South Wales, Australia) plutons.
Fig. 4. Number of radiohalos vs. type and location.
The biotite crystals in all these granites contain
abundant 238U, 210Po, and 214Po radiohalos. The occurrence
ratio is approximately five 210Po radiohalos for every 214Po
and 238U radiohalo, which occur roughly in equal numbers
except in the Cooma pluton (see Fig. 4). While these
radiohalos are homogeneously distributed throughout the
mineralogically uniform Stone Mountain pluton, they are
almost exclusively concentrated in the muscovite-biotite
granodiorite core of the La Posta pluton. Furthermore,
there are four to five times more of all these radiohalos in
the associated late-stage, Indian Hills granite (southern
California).Hydrothermal fluids are invariably concentrated
in the last liquid phases during the rapid convective
cooling of granite plutons as discussed by Snelling and
Woodmorappe,9 so this pattern of radiohalo occurrence in
the La Posta pluton and Indian Hills granite strongly
suggests that the Po radiohalos have formed as a result of
late hydrothermal fluid transport of Po radioisotopes locally
within the biotite flakes separating them from their
parent 238U in the zircons.The Cooma granite was
produced by partial melting at the center of a regional
metamorphic complex. Thus, this research has the
potential to demonstrate that both the cooling of granite
plutons and regional metamorphism occurred within
weeks to months, not over millions of years, because of
the short half-life of 218Po. Radiohalo occurrences in other
granitic plutons at many levels in the geologic record are
also under continuing investigation.
Isochron DiscordanceDr. Steven A. Austin, PI
Field observations, petrographic study, and geochemical
analysis by Snelling, Austin, and Hoesch10 indicate that a
95-meter-thick sill in sharp contact with the intruded
Hakatai shale near Bass Rapids in Grand Canyon was
well mixed isotopically when emplaced. However, after

intrusion, it segregated mineralogically and chemically by crystal settling. Such a condition of thorough isotopic mixing
followed by rapid chemical segregation is ideally suited to test the assumptions that underlie whole-rock and mineral
isochron dating. Both creationists and evolutionists should accept the well-mixed initial isotopic condition of the original
magma body.
Fig. 5. Isochron age vs. half-life and mode of decay.
New K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotope data from eleven whole-rock samples (eight diabase, three granophyre)
and six mineral phases separated from one of the whole-rock diabase samples yield discordant whole-rock and mineral
isochron ages. These isochron ages range from 842164Ma (whole-rock K-Ar) to 1375170Ma (mineral Sm-Nd). (See
Fig. 5 for a graph of the isochron age versus half-life and type of decay for each of the four radioisotope systems
investigated.) Although significant discordance exists between the K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotope methods,
each method appears to yield concordant ages internally between whole rocks and minerals. Internal concordance is best
illustrated by the Rb-Sr whole-rock and mineral isochron ages of 105546Ma and 105948Ma, respectively. It is,
therefore, argued that only changing radioisotope decay rates in the past could account for these discordant isochron ages
for the same geologic event. Furthermore, these data are consistent with alpha decay having been accelerated more than
beta decay, and with a greater acceleration factor for a greater present half-life.
Case Studies in Rock DatingDr. Andrew A. Snelling, PI
Snelling11,12 earlier reported having obtained K-Ar model ages for recent andesites collected from Mt. Ngauruhoe in New
Zealand. Dates of less than 0.27 to 3.5Ma could not be reproduced, even from splits of the same samples from the same
flow, the explanation being variations in the 40Ar* (radiogenic 40Ar) content in excess of the zero-age amount. It was
concluded that this excess 40Ar* had been inherited by these magmas during their genesis in the upper mantle.
Two samples from each of the lava flows and deposits have now been analyzed for Rb-Sr, Sm-Nd, and Pb-Pb isotopes.
Together with the trace and rare earth element analyses, they further elucidate the petrogenetic history of these andesites,
including crustal components which may have contaminated originally pure basalt magmas. Whereas valid isochron ages
cannot be obtained from this isotopic data except by subjective manipulation, depleted mantle Nd model ages of 8011594
Ma and positive Nd(to) values suggest the original basalt magmas were generated from partial melting of the residual solids
in old depleted upper mantle, while the large positive Sr(to) values and the 87Sr/86Sr ratios suggest contamination during their
ascent with basement greywackes to produce the andesite magmas. Consequently, evidence continues to accumulate that
systematic mixing of mantle and crustal sources makes it nearly impossible to obtain unambiguous radioisotopic results in
these environments.

Fig. 6. Petrogenetic model of melt formation near a subducting slab, based on Tatsumi17 and Davies and
Stevenson.18 Mixing and inheritance of radioisotopes invalidate conventional age dating.
The petrogenetic model therefore favored by Gamble et al.,13 which is consistent with all the isotopic data discussed in
Snelling,14 and shown in Fig. 6, is based on Tatsumi15 and Davies and Stevenson.16 This model envisages a zone of melt
formation approximately coincident to the volcanic front, which includes Ruapehu and Ngauruhoe, and a melt generation
region delimited by the interface of the subducting slab, the base of the arc lithosphere (of continental New Zealand) and
two vertical columns, one delineating the volcanic front, the other, the coupled back-arc basin. Fluids liberated from the
descending slab ascend into and enrich the overlying periodite down to higher pressures, where the amphibole breaks down
giving rise to amphibole dehydration, while progressive dehydration reactions in the slab itself lead to fluid transfer from the
slab into the mantle wedge, both processes producing partial melting as amphibole breaks down over the depth range 112
19km as discussed by Tatsumi19 and Davies and Stevenson.20 The lower density melt then rises and pools in the
upwelling melt column, eventually penetrating upwards into the overlying arc lithosphere to fill magma chambers that then

erupt when full.The Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic ratios in the samples of this study of recent (19491975)
andesite lava flows at Mt. Ngauruhoe, New Zealand, as anticipated, do not yield any meaningful age information, even by
selective manipulation of the data. Instead, these data provide evidence of the mantle source, of magma genesis, and of the
crustal contamination of the parental basalt magmas. By implication, the radioisotopic ratios in ancient lavas found
throughout the geologic record must similarly express the fundamental characteristics of their geochemistry. They therefore
must also strongly reflect the magmatic origin of the lavas from mantle and crustal sources and any history of mixing or
contamination in their petrogenesis which can dramatically distort any inferred isotopic age. Even though radioisotopic
decay has undoubtedly occurred during the earths history, conventional radioisotopic dating of these rocks therefore cannot
provide valid absolute ages for them. This is especially so if accelerated nuclear decay accompanied the catastrophic
geologic and tectonic processes responsible for the mixing of the radioisotopic decay products during magma genesis.
Fifteen rock samples have also been collected from the Somerset Dam gabbro intrusion near Brisbane, Australia
(Snelling21), probably a well-preserved, unmetamorphosed subvolcanic magma chamber. The samples were processed
and submitted to various laboratories for whole-rock major and trace element analyses and for K-Ar, Rb-Sr, Sm-Nd, and PbPb radioisotopic analyses. Additionally, one of the gabbro samples from one of the cyclic units was separated into its mineral
constituents using heavy liquids, and the resultant plagioclase, augite, olivine, and magnetite-ilmenite concentrates, along
with a duplicate piece of the whole-rock, submitted for K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic analyses.
The objective of this study was not only to compare the different dating methods, but also to compare whole-rock and
mineral isochron ages and to test whether there are variations in the radioisotopes between the cyclic units, and between
the macrolayers within them. From these studies it may be possible to infer how mixing occurs in a magma chamber and
demonstrate that radioisotopic compositions of crustal rock may reflect characteristics of the magma sources in the mantle
rather than the ages of the intrusion.
Significant Amounts of 14C in Deep StrataDr. John R. Baumgardner, PI
Fig. 7. Distribution of 14C values for biogenic samples
from the radiocarbon literature. Given their position
in the geological record, all these samples should
contain no detectable 14C according to the
conventional geological timescale.
Fig. 8. Histogram representation of AMS14C analysis
of ten coal samples undertaken by the RATE 14C
research project.
According to the conventional geologic time-scale,
organic materials older than about 250,000 years
should be utterly 14C dead. This is because the
half-life of 14C, only 5,730 years, is so short. 250,000
years of decay (corresponding to 43.6 half-lives)
reduces the number of initial14C atoms by a factor of
710-14. A gram of modern carbon contains about 6
1010 14C atoms, so not a single 14C atom should
remain after 250,000 years. The astonishing result,
however, is that, almost without exception, when
tested by accelerator mass spectrometer (AMS)
methods, organic samples from every portion of the
Phanerozoic record show detectable and
reproducible amounts of 14C! This reality has been
established as dozens of AMS laboratories around
the world over the last 20 years have sought
desperately to understand why organic samples
from deep within the geological record, thought to
be tens to hundreds of millions of years old, should
consistently contain 0.10.5% of the modern level
of 14C. Believing this 14C had to be contamination,
they have mounted an intense quest to identify and
eliminate sources of contamination in their AMS
procedures. But despite improvements in
techniques, this level of 14C, on the order of 0.10.5
percent modern carbon (pmc), continues to be
reported for samples that, given their location in the
geological record, should be entirely 14C dead.
Many scores of such measurements are readily
available
in
the
standard
peer-reviewed
radiocarbon
literature
as
documented
by
Giem22 and Baumgardner et al.23 and displayed in Fig. 7. Measurable 14C at roughly uniform values in pre-Flood organic
materials fossilized in Flood strata, of course, represents powerful support for the young earth Creation-Flood model.Aware
of this, Snelling24,25,26,27,28 analyzed the 14C content of fossilized wood conventionally regarded as 14C dead because it
was derived from Tertiary, Mesozoic, and upper Paleozoic strata having conventional ages of 40 to 250 million years. All
samples were analyzed using AMS technology by a reputable commercial laboratory, with some duplicate samples also
tested by a specialist laboratory in a major research institute. Measurable 14C well above background was obtained in all
cases.More recently, as a check on the AMS results in the peer-reviewed literature, the RATE team acquired a suite of ten
coal samples from the U.S. Department of Energy Coal Repository. These samples represent important U.S. coal deposits
and span the geological record from Carboniferous to Eocene. The 14C measurements by one of the best AMS laboratories
in the world for these ten samples are displayed in graphical form in Fig. 8 and discussed in Baumgardner et al.29 The 14C
levels for these samples fall nicely within the range of values shown in Fig. 7. We conclude that the well-documented
evidence of 14C in fossil organic material provides compelling support for the young earth Creation-Flood model and
represents a severe challenge for the uniformitarian assumptions underlying the long half-life radioisotope methods.
Tentative Conclusions

At this point in the RATE research several tentative conclusions are beginning to emerge, based on the literature searches,
theoretical studies, and laboratory findings. Although some are firmer than others, the following conclusions are likely to be
in the final report. There will likely also be additional conclusions which are too early to include at this time. The tentative
conclusions will only be reported here in outline form. More detail and justifications for most of these conclusions are
discussed in the referenced papers in these Proceedings.
Conventional radioisotope dating methods are unreliable.
Discordance among different dating methods is common.
Key assumptions underlying radioisotope dating methods are untenable.
Mixing of mantle and crustal sources also mixes their isotopic signatures.
Residual 14C appears to be present in all fossil biogenic material.
Massive nuclear decay has occurred in rocks.
Large quantities of daughter elements like Pb, He, and Ar are present.
Many of the daughter elements are in proximity to the parent elements.
Fission tracks and radiohalos are numerous.
Isotopic mixing between the earths mantle and crust has occurred.
Lava flows exhibit isotopic characteristics of the mantle.
Isotopic data suggest basalts were generated from melting of old mantle.
Isotopic data also suggest basalt magmas were contaminated during their ascent.
Residual He and radiohalos suggest recent nuclear decay.
Large quantities of He are still present in many granites today.
If He was formed millions of years ago, it should have already escaped.
Experimentally-determined diffusion rates of He agree with recent production of He.
Po halos appear to have formed during rapid cooling of granite plutons during the Flood (eliminating millions of years).
If the cooling of the plutons was rapid, then metamorphism was also rapid during the Flood (eliminating millions of years).
Massive nuclear decay, radiohalos, helium diffusion, and deep 14C all imply accelerated decay.
Massive nuclear decay requires higher decay rates before the present.
Radiohalos formed during the Flood require decay rates higher than observed today.
Helium diffusion data imply the decay occurred within thousands of years ago.
Deep 14C implies the decay occurred within thousands of years ago.
Studies in theoretical physics suggest accelerated nuclear decay can occur.
Variation in compactified dimensions could affect coupling constants.
Consequent variation in coupling constants could cause accelerated decay.
Changes in potential well depth change the -particle wave function.
Changes in the -particle wave function change decay half-lives.
Summary
The basic conclusion of this research is that conventional radioisotopic dating methods are unreliable. The chief reason is
that uniformitarianism is not a legitimate model of earth history. Observational evidence supports the recent occurrence of a
global catastrophic Flood. Because the earth has suffered a major tectonic catastrophe corresponding to the Flood, the
uniformitarian assumptions that are applied to obtain age estimates from radioisotopic data are simply not true. Intermediate
results from RATE support a young-earth, catastrophic, creationist model.Two remaining years in the research phase will be
needed to complete the analysis of samples yet being processed and theoretical studies still being made. By the end of the
research phase the final report should be based on a larger data set than was available for this paper. A few research
projects within RATE such as Fission Tracks and Biblical Word Studies that have not been discussed in this paper are also
expected to contribute to the final report. It is apparent that significant progress has been made in explaining the presence of
large quantities of nuclear decay products in a young-earth time frame. The evidence should be stronger and more
convincing by the time the research project is completed in 2005. We also hope that by then a more detailed young-earth
creationist model of the history of radioactive decay will also have been developed.
Radiometric Dating: Back to Basics
by Dr. Andrew A. Snelling on June 17, 2009; last featured February 19, 2014
Radiometric dating is often used to prove rocks
are millions of years old. Once you understand the
basic science, however, you can see how wrong
assumptions lead to incorrect dates.
Shop Now
Radiometric Dating 101
PART 1: Back to Basics
PART 2: Problems with the Assumptions
PART 3: Making Sense of the Patterns
This three-part series will help you properly understand radiometric dating, the assumptions that lead to inaccurate dates,
and the clues about what really happened in the past.
Most people think that radioactive dating has proven the earth is billions of years old. After all, textbooks, media, and
museums glibly present ages of millions of years as fact.
Yet few people know how radiometric dating works or bother to ask what assumptions drive the conclusions. So lets take a
closer look and see how reliable this dating method really is.
AtomsBasics We Observe Today
Each chemical element, such as carbon and oxygen, consists of atoms. Each atom is thought to be made up of three basic
parts.The nucleus contains protons (tiny particles each with a single positive electric charge) and neutrons (particles without
any electric charge). Orbiting around the nucleus are electrons (tiny particles each with a single negative electric charge).

The atoms of each element may vary slightly in the numbers of neutrons within their nuclei. These variations are called
isotopes of that element. While the number of neutrons varies, every atom of any element always has the same number of
protons and electrons.So, for example, every carbon atom contains six protons and six electrons, but the number of
neutrons in each nucleus can be six, seven, or even eight. Therefore, carbon has three isotopes (variations), which are
specified carbon-12, carbon-13, and carbon-14 (Figure 1).
Radioactive Decay
Some isotopes are radioactive; that is, they are unstable because their nuclei are too large. To achieve stability, the atom
must make adjustments, particularly in its nucleus. In some cases, the isotopes eject particles, primarily neutrons and
protons. (These are the moving particles measured by Geiger counters and the like.) The end result is a stable atom, but of
a different chemical element (not carbon) because the atom now has a different number of protons and electrons.This
process of changing one element (designated as the parent isotope) into another element (referred to as the daughter
isotope) is called radioactive decay. The parent isotopes that decay are called radioisotopes.Actually, it isnt really a decay
process in the normal sense of the word, like the decay of fruit. The daughter atoms are not lesser in quality than the parent
atoms from which they were produced. Both are complete atoms in every sense of the word.Geologists regularly use five
parent isotopes to date rocks: uranium-238, uranium-235, potassium-40, rubidium-87, and samarium-147. These parent
radioisotopes change into daughter lead-206, lead-207, argon-40, strontium-87, and neodymium-143 isotopes, respectively.
Thus geologists refer to uranium-lead (two versions), potassium-argon, rubidium-strontium, or samarium-neodymium dates
for rocks. Note that the carbon-14 (or radiocarbon) method is not used to date rocks because most rocks do not contain
carbon.
Chemical Analysis of Rocks Today
Geologists cant use just any old rock for dating. They must find rocks that have the isotopes listed above, even if these
isotopes are present only in minute amounts. Most often, this is a rock body, or unit, that has formed from the cooling of
molten rock material (called magma). Examples are granites (formed by cooling under the ground) and basalts (formed by
cooling of lava at the earths surface).The next step is to measure the amount of the parent and daughter isotopes in a
sample of the rock unit. Specially equipped laboratories can do this with accuracy and precision. So, in general, few people
quarrel with the resulting chemical
analyses.It is the interpretation of
these chemical analyses that
raises potential problems. To
understand how geologists read
the age of a rock from these
chemical analyses, lets use the
analogy of an hourglass clock
(Figure 2).
In an hourglass, grains of fine
sand fall at a steady rate from the
top bowl to the bottom. After one
hour, all the sand has fallen into
the bottom bowl. So, after only half
an hour, half the sand should be in
the top bowl, and the other half
should be in the bottom bowl.
Suppose that a person did not
observe when the hourglass was
turned over. He walks into the
room when half the sand is in the
top bowl, and half the sand is in
the bottom bowl. Most people
would assume that the clock
started half an hour earlier.
By way of analogy, the sand
grains in the top bowl represent
atoms of the parent radioisotope
(uranium-238, potassium-40, etc.)
(Figure 2). The falling sand
represents radioactive decay, and
the sand at the bottom represents
the daughter isotope (lead-206,
argon-40, etc).
When a geologist tests a rock
sample, he assumes all the
daughter atoms were produced by
the decay of the parent since the
rock formed. So if he knows the
rate at which the parent decays,
he can calculate how long it took
for the daughter (measured in the
rock today) to form.
But what if the assumptions are
wrong? For example, what if
radioactive material was added to
the top bowl or if the decay rate
has changed? Future articles will
explore the assumptions that can

lead to incorrect dates and how the creation model helps us make better sense of the patterns of radioactive dates we find
in the rocks today.

Radiometric Dating: Problems with the Assumptions


by Dr. Andrew A. Snelling on October 1, 2009; last featured August 4, 2010
Radiometric dating is often used to prove rocks are
millions of years old. Once you understand the basic
science, however, you can see how wrong
assumptions lead to incorrect dates.
Radiometric Dating 101
PART 1: Back to Basics
PART 2: Problems with the Assumptions
PART 3: Making Sense of the Patterns
This three-part series will help you properly
understand radiometric dating, the assumptions that lead to inaccurate dates, and the clues about what really happened in
the past.Most people think that radioactive dating has proven the earth is billions of years old. Yet this view is based on a
misunderstanding of how radiometric dating works. Part 1 (in the previous issue) explained how scientists observe unstable
atoms changing into stable atoms in the present. Part 2 explains how scientists run into problems when they make
assumptions about what happened in the unobserved past.
The Hourglass ClockAn Analogy for Dating Rocks
An hourglass is a helpful analogy to explain how geologists calculate the ages of rocks. When we look at sand in an
hourglass, we can estimate how much time has passed based on the amount of sand that has fallen to the bottom.
Radioactive rocks offer a similar clock. Radioactive atoms, such as uranium (the parent isotopes), decay into stable atoms,
such as lead (the daughter isotopes), at a measurable rate. To date a radioactive rock, geologists first measure the sand
grains in the top glass bowl (the parent radioisotope, such as uranium-238 or potassium-40).They also measure the sand
grains in the bottom bowl (the daughter isotope, such as lead-206 or argon-40, respectively). Based on these observations
and the known rate of radioactive decay, they estimate the time it has taken for the daughter isotope to accumulate in the
rock.However, unlike the hourglass whose accuracy can be tested by turning it upside down and comparing it to trustworthy
clocks, the reliability of the radioactive clock is subject to three unprovable assumptions. No geologist was present when
the rocks were formed to see their contents, and no geologist was present to measure how fast the radioactive clock has
been running through the millions of years that supposedly passed after the rock was formed.
Assumption 1: Conditions at Time Zero
No geologists were present when most rocks formed, so they cannot test whether the original rocks already contained
daughter isotopes alongside their parent radioisotopes. For example, with regard to the volcanic lavas that erupted, flowed,
and cooled to form rocks in the unobserved past, evolutionary geologists simply assume that none of the daughter argon-40
atoms was in the lava rocks.For the other radioactive clocks, it is assumed that by analyzing multiple samples of a rock
body, or unit, today it is possible to determine how much of the daughter isotopes (lead, strontium, or neodymium) were
present when the rock formed (via the so-called isochron technique, which is still based on unproven assumptions 2 and 3).
Yet lava flows that have occurred in the present have been tested soon after they erupted, and they invariably contained
much more argon-40 than expected.1 For example, when a sample of the lava in the Mt. St. Helens crater (that had been
observed to form and cool in 1986) (Figure 1) was analyzed in 1996, it contained so much argon-40 that it had a calculated
age of 350,000 years!2 Similarly, lava flows on the sides of Mt. Ngauruhoe, New Zealand (Figure 2), known to be less than
50 years old, yielded ages of up to 3.5 million years.3

Click here to view larger picture (PDF format).


So it is logical to conclude that if recent lava flows of known age yield incorrect old potassium-argon ages due to the extra
argon-40 that they inherited from the erupting volcanoes, then ancient lava flows of unknown ages could likewise have
inherited extra argon-40 and yield excessively old ages.
There are similar problems with the other radioactive clocks. For example, consider the dating of Grand Canyons basalts
(rocks formed by lava cooling at the earths surface). We find places on the North Rim where volcanoes erupted after the
Canyon was formed, sending lavas cascading over the walls and down into the Canyon.Obviously, these eruptions took
place very recently, after the Canyons layers were deposited (Figure 3). These basalts yield ages of up to 1 million years
based on the amounts of potassium and argon isotopes in the rocks. But when we date the rocks using the rubidium and
strontium isotopes, we get an age of 1.143 billion years. This is the same age that we get for the basalt layers deep below
the walls of the eastern Grand Canyon.4How could both lavasone at the top and one at the bottom of the Canyonbe the
same age based on these parent and daughter isotopes? One solution is that both the recent and early lava flows inherited
the same rubidium-strontium chemistrynot agefrom the same source, deep in the earths upper mantle. This source
already had both rubidium and strontium.To make matters even worse for the claimed reliability of these radiometric dating
methods, these same basalts that flowed from the top of the Canyon yield a samarium-neodymium age of about 916 million
years,5 and a uranium-lead age of about 2.6 billion years!6
Assumption 2: No Contamination
The problems with contamination, as with inheritance, are already well-documented in the textbooks on radioactive dating of
rocks.7 Unlike the hourglass, where its two bowls are sealed, the radioactive clock in rocks is open to contamination by
gain or loss of parent or daughter isotopes because of waters flowing in the ground from rainfall and from the molten rocks
beneath volcanoes. Similarly, as molten lava rises through a conduit from deep inside the earth to be erupted through a
volcano, pieces of the conduit wallrocks and their isotopes can mix into the lava and contaminate it.Because of such
contamination, the less than 50-year-old lava flows at Mt. Ngauruhoe, New Zealand (Figure 4), yield a rubidium-strontium
age of 133 million years, a samarium-neodymium age of 197 million years, and a uranium-lead age of 3.908 billion
years!8
Assumption 3: Constant Decay Rate
Physicists have carefully measured the radioactive decay rates of parent radioisotopes in laboratories over the last 100 or
so years and have found them to be essentially constant (within the measurement error margins). Furthermore, they have
not been able to significantly change these decay rates by heat, pressure, or electrical and magnetic fields. So geologists
have assumed these radioactive decay rates have been constant for billions of years.
However, this is an enormous extrapolation of seven orders of magnitude back through immense spans of unobserved time
without any concrete proof that such an extrapolation is credible. Nevertheless, geologists insist the radioactive decay rates
have always been constant, because it makes these radioactive clocks work!New evidence, however, has recently been
discovered that can only be explained by the radioactive decay rates not having been constant in the past. 9 For example,
the radioactive decay of uranium in tiny crystals in a New Mexico granite (Figure 5) yields a uranium-lead age of 1.5 billion
years. Yet the same uranium decay also produced abundant helium, but only 6,000 years worth of that helium was found to
have leaked out of the tiny crystals.This means that the uranium must have decayed very rapidly over the same 6,000 years
that the helium was leaking. The rate of uranium decay must have been at least 250,000 times faster than todays measured
rate! For more details see Don DeYoungs Thousands . . . Not Billions (Master Books, Green Forest, Arkansas, 2005), pages
6578.The assumptions on which the radioactive dating is based are not only unprovable but plagued with problems. As this
article has illustrated, rocks may have inherited parent and daughter isotopes from their sources, or they may have been
contaminated when they moved through other rocks to their current locations. Or inflowing water may have mixed isotopes
into the rocks. In addition, the radioactive decay rates have not been constant.So if these clocks are based on faulty
assumptions and yield unreliable results, then scientists should not trust or promote the claimed radioactive ages of
countless millions of years, especially since they contradict the true history of the universe.
Radiometric Dating: Making Sense of the Patterns

by Dr. Andrew A. Snelling on January 1, 2010; last featured March 16, 2011
Radiometric dating methods sometimes yield
conflicting results, but the technique itself is
scientific and reliable, and once the results are
interpreted in a creation framework, they yield
clear patterns that help us better understand
the earths history since creation six thousand
years ago.
Shop Now
Radiometric Dating 101
PART 1: Radiometric Dating: Back to Basics
PART 2: Radiometric Dating: Problems with the Assumptions
PART 3: Making Sense of the Patterns
This three-part series will help you properly understand radiometric dating, the assumptions that lead to inaccurate dates,
and the clues about what really happened in the past.Part Two of this series showed that the same rocks can yield very
different ages, depending on which radiometric dating technique you use. These inconsistent results are due to the
problems of inheritance and contamination, which cause the rocks chemistry to differ from the assumptions of standard
radioactive clocks.Furthermore, new evidence indicates that radioactive elements in the rocks, which are used to date the
rocks, decayed at much faster rates during some past event (or events) in the last 6,000 years. So the claimed ages of
many millions of years, which are based on todays slow decay rates, are totally unreliable.Does this mean we should throw
out the radioactive clocks? Surprisingly, they are useful!The general principles of using radioisotopes to date rocks are
sound; its just that the assumptions have been wrong and led to exaggerated dates. While the clocks cannot yield absolute
dates for rocks, they can provide relative ages that allow us to compare any two rock units and know which one formed first.
They also allow us to compare rock units in different areas of the world to find which ones formed at the same time.
Furthermore, if physicists examine why the same rocks yield different dates, they may discover new clues about the unusual
behavior of radioactive elements during the past.With the help of this growing body of information, creation geologists hope
to piece together a better understanding of the precise sequence of events in earths history, from Creation Week to the
Flood and beyond.
Different Dates for the Same Rocks
Usually geologists do not use all four main radioactive clocks to date a rock unit. This is considered an unnecessary waste
of time and money. After all, if these clocks really do work, then they should all yield the same age for a given rock unit.
Sometimes though, using different parent radioisotopes to date different samples (or minerals) from the same rock unit does
yield different ages, hinting that something is amiss.1Recently, creationist researchers have utilized all four common
radioactive clocks to date the same samples from the same rock units.2 Among these were four rock units far down in the
Grand Canyon rock sequence (Figure 1), chosen because they are well known and characterized. These were as follows:
Radiometric
Ages
of
Rock
Samples: Figures 1 through 5, and
Table 1. Click the picture to view a
larger, pdf version.
Cardenas Basalt (lava flows deep in
the east Canyon sequence) (Figure
2).
Bass Rapids diabase sill (where
basalt magma squeezed between
layers and cooled) (Figure 3).
Brahma amphibolites (basalt lava
flows deep in the Canyon sequence
that later metamorphosed) (Figure 4).
Elves Chasm Granodiorite (a granite
regarded as the oldest Canyon rock
unit) (Figure 5).
Table 1 lists the dates obtained from
each rock unit. Figure 6 (see below)
graphically illustrates the range in the
supposed ages of these rock units, obtained by utilizing all four radioactive clocks.
It is immediately apparent that the ages for each rock unit do not agree. Indeed, in the Cardenas Basalt, for example, the
samarium-neodymium age is three times the potassium-argon age.
Nevertheless, the ages follow three obvious patterns. Two techniques (potassium-argon age and rubidiumstrontium)always yield younger ages than two other techniques (uranium-lead and samarium-neodymium). Furthermore, the
potassium-argon ages are always younger than the rubidium-strontium ages. And often the samarium-neodymium ages are
younger than the uranium-lead ages.What then do these patterns mean? All the radioactive clocks in each rock unit should
have started ticking at the same time, the instant that each rock unit was formed. So how do we explain that they have
each recorded different ages?The answer is simple but profound. Each of the radioactive elements must have decayed at
different, faster rates in the past!In the case of the Cardenas Basalt, while the potassium-argon clock ticked through 516
million years, two other clocks ticked through 1,111 million years and 1,588 million years. So if these clocks ticked at such
different rates in the past, not only are they inaccurate, but these rocks may not be millions of years old!

Patterns in the Radiometric Ages: Figures 6


and 7. Click the picture to view a larger, pdf
version.
But how could radioactive decay rates have
been different in the past? Creationist
researchers dont fully understand yet.
However, the observed age patterns provide
clues. Potassium and rubidium decay
radioactively by the process known as beta ()
decay, whereas uranium and neodymium
decay via alpha () decay (Figure 6). The
former always gives younger ages. We see
another pattern within beta decay. Potassium
today decays faster than rubidium and always
gives younger ages.Both of these patterns
suggest something happened in the past
inside the nuclei of these parent atoms to
accelerate their decay. The decay rate varied
based on the stability or instability of the
parent atoms. Research is continuing.
Relative Ages
Look again at Figure 1, which is a geologic
diagram depicting the rock layers in the walls
of the Grand Canyon, along with the rock units
deep in the inner gorge along the Colorado
River. This diagram shows that the radiometric
dating methods accurately show the top rock
layer is younger than the layers beneath
it.Thats logical because the sediment making
up that layer was deposited on top of, and
therefore after, the layers below. So reading this diagram tells us basic information about the time that rock layers and rock
units were formed relative to other layers.Based on the radioactive clocks, we can conclude that these four rock units deep
in the gorge (Table 1) are all older in a relative sense than the horizontal sedimentary layers in the Canyon walls.
Conventionally the lowermost or oldest of these horizontal sedimentary layers is labeled early to middle Cambrian3 and thus
regarded as about 510520 million years old.4 All the rocks below it are then labeled Precambrian and regarded as older
than 542 million years.So accordingly all four dated rock units (Table 1) are also Precambrian. And apart from the
potassium-argon age for the Cardenas Basalt, all the radioactive clocks have correctly shown that these four rock units were
formed earlier than Cambrian, so they are pre-Cambrian. (But the passage of time between these Precambrian rock units
and the horizontal sedimentary layers above them was a maximum of about 1,700 yearsthe time between creation and
the Floodnot millions of years.)Similarly, in the relative sense the Brahma amphibolites and Elves Chasm Granodiorite are
older (by hours or days) than the Cardenas Basalt and Bass Rapids diabase sill (Figure 1). Once again, the radioactive
clocks have correctly shown that those two rock units are older than the rock units above them.Why then should we expect
the radioactive clocks to yield relative ages that follow a logical pattern? (Actually, younger sedimentary layers yield a similar
general pattern,5 Figure 7.) The answer is again simple but profound! The radioactive clocks in the rock units at the bottom
of the Grand Canyon, formed during Creation Week, have been ticking for longer than the radioactive clocks in the younger
sedimentary layers higher up in the sequence that were formed later during the Flood.
Conclusion
Although it is a mistake to accept radioactive dates of millions of years, the clocks can still be useful to us, in principle, to
date the relative sequence of rock formation during earth history.The different clocks have ticked at different, faster rates in
the past, so the standard old ages are certainly not accurate, correct, or absolute. However, because the radioactive clocks
in rocks that formed early in earth history have been ticking longer, they should generally yield older radioactive ages than
rock layers formed later.So it is possible that relative radioactive ages of rocks, in addition to mineral contents and other rock
features, could be used to compare and correlate similar rocks in other areas to find which ones formed at the same time
during the earth history.
Radioactive Dating Failure
Recent New Zealand Lava Flows Yield Ages of Millions of Years
by Dr. Andrew A. Snelling on December 1, 1999
Originally published in Creation 22, no 1 (December 1999):
18-21.
Recent New Zealand lava flows yield ages of millions of
years.
Standing roughly in the centre of New Zealands North
Island, Mt Ngauruhoe is New Zealands newest volcano
and one of the most active (Figures 1 and 2). It is not as
well publicized as its larger close neighbour MT Ruapehu,
which has erupted briefly several times in the last five years.However, Mt Ngauruhoe is an
imposing, almost perfect cone that rises more than 1,000 metres (3,300 feet) above the
surrounding landscape to an elevation of 2,291 m (7,500 feet) above sea level1 (Figure 3).
Eruptions from a central 400 m (1,300 foot) wide crater have constructed the cones steep
(33) outer slopes.
Figure 1. The location of Mt Ngauruhoe, central North Island, New Zealand. (click image for
larger view)
Mt Ngauruhoe is thought to have been active for at least 2,500 years, with more than 70
eruptive periods since 1839, when European settlers first recorded a steam eruption.2 Of

course, before that, the Maoris witnessed many eruptions from the mountain. The first lava eruption seen by Europeans
occurred in 1870.3 Then there were ash eruptions every few years until a major explosive eruption in AprilMay 1948,
followed by lava flowing down the northwestern slopes in February 1949.4,5 The estimated lava volume was about 575,000
cubic metres (20 million cubic feet).
Figure 2. Aerial view, looking south at sunrise, of volcanoes Mt Ngauruhoe (foreground) and MT Ruapehu (background).
The eruption lasting from 13 May 1954 to 10 March 1955 began with an explosive ejection of ash
and blocks.6,7 Then almost 8 million cubic metres (280 million cubic feet) of lava flowed from the
crater in a series of 17 distinct flows on the following 1954 dates:
June 4, 30
July 8, 9, 10, 11, 13, 14, 23, 28, 29, 30
August 15(?), 18
September 16, 18, 26
These flows are still distinguishable today on the northwestern and western slopes of Ngauruhoe
(Figure 4). The 18 August flow was more than 18 m (55 feet) thick and still warm almost a year
after congealing. Explosions of ash completed this long eruptive period.
Figure 3. Mt Ngauruhoe as seen looking north from near MT Ruapehu.
Afterwards, Ngauruhoe steamed almost continuously, with many small ash eruptions8 (Figure 5).
Cannon-like, highly explosive eruptions in January and March 1974 threw out large quantities of
ash as a column into the atmosphere, and as avalanches flowing down the cones sides. Blocks
weighing up to 1,000 tonnes were hurled 100 m (330 feet). However, the most violent explosions
occurred on 19 February 1975, accompanied by what eye-witnesses described as atmospheric
shock waves.9 Blocks up to 30 m (100 ft) across were catapulted up to 3 km (almost 2 miles).
The eruption plume was 1113 km (78 miles) high.Turbulent avalanches of ash and blocks
swept down Ngauruhoes sides at about 60 km (35 miles) per hour.10 It is estimated that at least
3.4 million cubic metres (120 million cubic feet) of ash and blocks were ejected in 7 hours.11up>
No further eruptions have occurred since.
Dating the rocks
Figure 4. View from the Mangateopopo Valley at the base of Mt Ngauruhoe, showing the darkercoloured recent lava flows on its northwestern slopes.
Radioactive dating in general depends on three major assumptions:
When the rock forms (hardens) there should only be parent radioactive atoms in the rock and no
daughter radiogenic (derived by radioactive decay of another element) atoms;512After hardening,
the rock must remain a closed system, that is, no parent or daughter atoms should be added to or removed from the rock by
external influences such as percolating groundwaters; and
The radioactive decay rate must remain constant.
If any of these assumptions are violated, then the technique fails and any dates are false.
The potassium-argon (KAr) dating method is often used to date volcanic rocks (and by extension, nearby fossils). In using
this method, it is assumed that there was no daughter radiogenic argon ( 40Ar*) in rocks when they formed.13 For volcanic
rocks which cool from molten lavas, this would seem to be a reasonable assumption. Because argon is a gas, it should
escape to the atmosphere due to the intense heat of the lavas. Of course, no geologist was present to test this assumption
by observing ancient lavas when they cooled, but we can study modern lava flows.
Potassium-argon dates
Figure 5. Small ash eruption, Mt Ngauruhoe.
Figure 6. Inset: Andesite of the June 30, 1954 flow, Mt Ngauruhoe, seen at
60 times magnification under a geological microscope. Different minerals
have different colours. All are embedded in a fine-grained matrix.Eleven
samples were collected from five recent lava flows during field work in
January 1996two each from the 11 February 1949, 4 June 1954, and 14
July 1954 flows and from the 19 February 1975 avalanche deposits, and
three from the 30 June 1954 flow14 (Figure 6). The darker recent lavas were
clearly visible and each one easily identified (with the aid of maps) on the
northwestern slopes against the lighter-coloured older portions of the cone
(Figures 4 and 7). All flows were typically made up of jumbled blocks of
congealed lava, resulting in rough, jagged, clinkery surfaces (Figure 8).
The samples were sent progressively in batches to Geochron Laboratories in
Cambridge, Boston (USA), for whole-rock potassium-argon (KAr) datingfirst a piece of one sample from each flow, then
a piece of the second sample from each flow after the first set of results was received, and finally, a piece of the third sample
from the 30 June 1954 flow.15 To also test the consistency of results within samples, second pieces of two of the 30 June
1954 lava samples were also sent for analysis.Geochron is a respected commercial laboratory, the KAr lab manager
having a Ph.D. in KAr dating. No specific location or expected age information was supplied to the laboratory. However, the
samples were described as probably young with very little argon in them so as to ensure extra care was taken during the
analytical work.
Figure 7. Map of the northwestern slopes of Mt Ngauruhoe showing the lava flows of
1949 and 1954, and the 1975 avalanche deposits.3,4 (Click image for larger view)
The dates obtained from the KAr analyses are listed in Table 1.16 The ages
range from <0.27 to 3.5 ( 0.2) million years for rocks which wereobserved to have
cooled from lavas 2550 years ago. One sample from each flow yielded ages of
<0.27 or <0.29 million years while all the other samples gave ages of millions of
years. The low age samples were all processed by the laboratory in the same
batch, suggesting a systematic lab problem. So the lab manager kindly re-checked
his equipment and re-ran several of the samples, producing similar results. This
ruled out a systematic lab error and confirmed that the low results were real.
Furthermore, repeat measurements on samples already analyzed (A#2 and B#2
in Table 1) did not reproduce the same results, but this was not surprising given the
analytical uncertainties at such low levels of argon. Clearly, the argon content varies
greatly within these rocks. Some geochronologists would say <0.27 million years is

actually the correct date, but how would they know that 3.5 million years was not in fact the correct age if they did not
already know the lava flows were recent?!Because these rocks are known to be less than 50 years old, it is apparent from
the analytical data that these KAr ages are due to excess argon inherited from the magma source area deep in the
earth.17 Thus, when the lavas cooled, they contained appreciable (non-zero) concentrations of normal 40Ar, which is
indistinguishable from daughter radiogenic 40Ar* derived by radioactive decay of parent 40K. This violates assumption (1) of
radioactive dating, and so the KAr method fails the test. This same failure is also known to occur in many other rocks,
including both recent volcanics18and ancient crustal rocks.19
Conclusions
Figure 8. The June 30, 1954 lava flow, showing the jumbled blocks of congealed lava which give it
a rough, jagged, clinkery surface.The radioactive potassium-argon dating method has been
demonstrated to fail on 1949, 1954, and 1975 lava flows at Mt Ngauruhoe, New Zealand, in spite
of the quality of the laboratorys KAr analytical work. Argon gas, brought up from deep inside the
earth within the molten rock, was already present in the lavas when they cooled. We know the
true ages of the rocks because they were observed to form less than 50 years ago. Yet they yield
ages up to 3.5 million years which are thus false. How can we trust the use of this same dating
on rocks whose ages we dont know? If the method fails on rocks when we have an independent
eye-witness account, then why should we trust it on other rocks where there are no independent
historical cross-checks?
The KAr (potassium-argon) dating method
Fossils are almost never dated by radiometric methods, since they rarely contain suitable
radioactive elements. A common way of dating fossils (and rocks which do not contain radioactive
elements) is by dating an associated volcanic rock. This is commonly done using the KAr
method. It depends on the rate at which radioactive potassium decays into the gas argon.The K
Ar method works on the assumption that the clock begins to tick the moment that the rock hardens. That is, it assumes
that no argon derived by radioactive decay was present initially, but after the lava cooled and solidified, the argon from
radioactive decay was unable to escape and started to accumulate. However, it is well-known that if a radiometric date
contradicts a fossil-derived (evolutionary) age, the date is discarded as erroneous. See Lubenow, M.,The Pigs Took It
All, Creation 17(3):3638, 1995.
FLOW DATE

SAMPLE

LAB CODE

KAr AGE (million years)

11 February 1949

R-11714

<0.27

R-11511

1.0 0.2

R-11715

<0.27

R-11512

1.5 0.1

A #1

R-11718

<0.27

A #2

R-12106

1.3 0.3

B #1

R-12003

3.5 0.2

B #2

R-12107

0.8 0.2

R-11513

1.2 0.2

R-11509

1.0 0.2

R-11716

<0.29

R-11510

1.0 0.2

R-11717

<0.27

4 June 1954

30 June 30, 1954

14 July 1954

19 February 1975

Table 1. Potassium-argon dates of recent Mt Ngauruhoe (New Zealand) lava flows.20


Ed. note, this Creation magazine article by Dr Snelling is based on his technical paper21, which has far more detail about
research methods and answers to possible criticisms than was possible in Creation magazine.
Radioactive Dating Method Under Fire
by Dr. Andrew A. Snelling on March 1, 1992
Originally published in Creation 14, no 2 (March 1992): 43-47.
In most peoples minds today, the radioactive dating of the earths rocks by geologists has supposedly proved that the earth
is billions of years old.
Yet most people really dont know much about these radioactive dating methods. So slick and convincing are the
presentations of results, particularly in glossy media and museum propaganda, that no one even bothers to question how
these dating methods work, what assumptions are involved, and how reliable they are.Such questions, however, are highly
relevant. The answers are not only instructive, but demolish the evolutionary geologists case for a 4.5-billion-year old earth.
This in turn allows the evidence for a young earth and universe1 to speak more loudly in support of the scriptural
chronology of a 6,000-7,000 year age, which of course leaves no time for any big bang and molecules-to-man evolutionary
scenarios.Recently, the radioactive dating method which geologists (and physicists) have considered to be perhaps the
most reliable has come under heavy fire. The big surprise is that the attack has come from an evolutionary geologist and
has been published in a secular scientific journal! But more of that in a moment. First, let s find out how radioactive dating
methods are supposed to work.
Radioactive dating explained
Some types (technically known as isotopes) of parent elements such as uranium, thorium, potassium and rubidium are
said to be radioactive because the nuclei of the atoms are unstable, resulting in readjustments between the particles
(primarily neutrons and protons) in the nuclei with time. To achieve stability, some particles are ejected from the atoms, and

these moving particles constitute the radioactivity measured by Geiger counters and the like. The end result is stable atoms
of the daughter elements lead, argon, and strontium respectively.Thus the first step in the radioactive dating technique is to
measure the amounts of the parent and daughter elements (isotopes) in a rock sample via chemical analyses. This is done
in specially equipped laboratories with sophisticated instruments capable of very good precision and accuracy, so in general
there is no quarrel with the resulting chemical analyses.However, it is with the interpretation of the chemical analyses of the
radioactive parents and resultant daughters that the problems with radioactive dating of rocks begin. In order to interpret
these chemical analyses, geochronologists must make three vital assumptions, otherwise the radioactive clock cannot be
made to read the age of the rocks. These assumptions are:
the initial conditions are known;
the system has been closed; and
the radioactive decay rate has remained constant.
So that these assumptions are easily understood, they are best explained in the context of the hourglass analogy (see
Figure 1). Grains of fine sand fall at a steady rate from the top glass bowl to the bottom. At time t = 0, the hourglass is turned
upside-down so that all the sand starts in the top bowl. By time t = 1 hour, all the sand is supposed to have fallen into the
bottom glass bowl.Now this clock works because the initial conditions are knownthat is, all the sand grains are in the top
glass bowl and none are in the bottom one. If there is already some sand in the bottom glass bowl, then unless this amount
is known the hourglass clock cannot tell the time. Similarly, if the system has not remained closed (for example, if sand
were somehow added or subtracted), then the calculation of the elapsed time, based on comparing the amounts of sand in
the two glass bowls, will again lead to an incorrect conclusion. And finally, if the rate at which the sand grains fall from the
top glass bowl to the bottom one varies (for example, moisture causes some clogging of the sand in the constriction
between the two glass bowls), then again the hourglass clock will be inaccurate.
Unproven assumptions
The radioactive decay of parent isotopes of uranium, thorium, potassium, and rubidium to daughter isotopes of lead, argon
and strontium respectively is analogous to our hourglass clock, including these three assumptions. However, in the case of
these radioactive clocks these three assumptions can be shown to be not only unprovable, but invalid, rendering these
clocks virtually useless.In the case of the initial conditions, no scientist can ever be sure as to what they were, because no
scientist was present here on the earth at its origin. Thus the amount of daughter isotope that has actually been derived
from the parent isotope by radioactive decay is unknown, since some of the daughter isotope might have been present with
its respective parent isotope at the time of the earths origin.So geochronologists have assumed that the uranium, thorium
and lead isotopic composition of particular meteorites is equivalent to the initial composition of these isotopes when the
earth came into existence. This is assumed because it is supposed that these meteorites represent fragments from another
planet in the solar system similar to our earth that disintegrated very early in the history of the solar system. However, not all
meteorites have the same uraniumthorium- lead isotopic composition, so why should the isotopic composition of these
particular meteorites be considered to be the correct composition for the earth at its origin rather than some other
composition found in other meteorites?
An hourglass clock tells us the elapsed time by comparing the amount of sand in the
top bowl (parent) with the amount in the bottom bowl (daughter).Furthermore, even if
todays scientists believe they have the methods, for example graphical and
mathematical, for determining how much of the daughter isotope might have been
present either at the origin of the earth or the origin of the rock being dated, no one can
ever be sure that these answers are correct, because there was no scientist present
at the beginning to observe those initial conditions, even though the scientists
calculations may be extremely logical.Similarly, there is no way that it can be proved
that these radioactive systems have been closed through all the supposed millions of
years of decay of parent isotopes into daughter isotopes. Again, the main reason for
this is because no scientist has been present to observe everywhere these radioactive
systems and so report that they have been closed through all their history. Indeed, the
evidence indicates the very opposite, that is, that these systems have been open to all
sorts of external influences.For example, it is known that uranium is generally a mobile
element in the natural environment, particularly in groundwaters near the earths
surface. Thus, if a rock sample is analysed at or near the earths surface for its uranium
and lead isotopes, it would be incorrect to assume that all the uranium and lead in the
sample were there only because of the amounts placed in the rock at its origin and
because of undisturbed radioactive decay from uranium into lead. Some of the uranium
might have been leached out of the rock sample, hence making the rock appear older than it really is according to this
radioactive clock. Or, some uranium might have been deposited by groundwaters into the sample, thus making it appear
younger than what it really is.Indeed, geochronologists often plot the chemical analyses of the isotopes, expressed as
isotope ratios, on graphs, and these often show that the parent-daughter systems have not been closed, but open.
Furthermore, by interpretation of these graphs they often claim to be able to quantify the loss or gain and thus overcome this
difficulty to still read the radioactive clock. However, once again this interpretation to overcome this problem of the
invalidated closed-system assumption cannot be proved, but is merely assumed to be correct because it makes the
radioactive clock work.The final assumption is, of course, that the radioactive decay rates have remained constant.
However, once again, this assumption can in no way be proved, because there were no human observers present right
throughout the earths history to be constantly measuring the radioactive decay rates and to have recorded them.It is special
pleading on the part of geochronologists and physicists to say that the radioactive decay rates have been carefully
measured in laboratories for the past 80 or 90 years and that no significant variation of these rates has been measured. The
bottom line is really that 80 or 90 years of measurements are being extrapolated backwards in time to the origin of the
earth, believed by evolutionists to be 4.5 billion years ago. That is an enormous extrapolation. In any other field of scientific
research, if scientists or mathematicians were to extrapolate results over that many orders of magnitude, thereby assuming
continuity of results over such enormous spans of unobserved time, they would be literally laughed out of court by fellow
scientists and mathematicians. Yet geochronologists are allowed to do this with impunity, primarily because it gives the
desired millions and billions of years that evolutionists require, and because it makes these radioactive clocks work!So we
have seen that none of these three basic assumptions which are foundational to all the radioactive dating techniques can be
proved. Indeed, we have also seen that each of these three assumptions is invalid, not only because no scientist has been
present from the origin of the earth to see what it was like then and to report as an eyewitness all that has happened
everywhere since, but because we know of observations contrary to these assumptions.
The isochron dating method

Apart from the initial conditions, the major problem facing geochronologists is that geological systems are invariably open to
external influences. Thus, analyses of radioisotopes often produce results that reflect loss, or sometimes gain, of either
parent or daughter isotope, rendering single radioactive age determinations suspect. Thus geochronologists tackle the
problem by performing a number of radioactive age determinations on a group of samples from the rock under investigation,
hoping to pin-point a pattern that will enable the calculation of the desired true age.If these multiple isotopic analyses of
various rock samples, and minerals within those rock samples, are from the same geological unit, then geochronologists
can also use what is known as the isochron age determination method. This method is supposed to allow some of the more
uncertain assumptions of the normal age calculating method to be circumvented and so permit a higher degree of
confidence in the resulting age estimate. Consequently, geochronologists favour this isochron method and so it has
become very popular, particularly with rubidium-strontium, samarium-neodymium and uranium-lead isotopic systems.
The isochron method works as follows. If a
number of rock samples from a single
geological unit are carefully collected, then
it is claimed that it is reasonable to
assume that each rock sample from that
geological unit formed at the same time,
and therefore ultimately has the same age.
However, from experience it is known that
each rock sample differs in the amounts of
both daughter and parent isotopes
contained.A graph is then constructed so
as to plot the amount of daughter isotope
against the amount of parent isotope, so
the isotopic analysis of each rock sample
will then be represented as a single point
on this graph. Often these data points,
plotted on the graph of daughter isotope
composition against parent isotope
composition, form a linear array through
which a sloping line can usually be drawn
with a high degree of fit of the data points
to the line, as shown in Figure 2. This is
because those samples with larger amounts of parent isotope have correspondingly larger amounts of daughter isotope,
and those samples with smaller amounts of parent isotope have correspondingly smaller amounts of daughter isotope,
assuming of course that all the daughter isotope has been produced by radioactive decay from the parent isotope.
This line is then interpreted as an effect produced by radioactive decay to give an age interpretation. Since all of these rock
samples are supposed to have been formed at the same time because they come from the same geological unit, this line is
called an isochron (from the Greek isos equal, and chronos time) or line of equal age. Furthermore, it can be shown
mathematically that the slope of the line can then be used to calculate the isochron age of the geological unit from which
the rock samples came.This method has become popular because no knowledge or assumptions about the initial conditions
of parent and daughter isotopes need be made. Furthermore, since the analytical equipment determines isotopic ratios, not
absolute abundances of isotopes, parent and daughter isotopes are usually expressed as ratios relative to a reference
isotope whose abundance is not affected by radioactive decay, thus providing easy application of the method and more
confidence in its results.While the assumptions of constant decay rate and a closed system are again necessary, the
isochron method also has two other critical assumptionsthe rock samples must represent the one unit that formed at the
same time geologically, and the daughter isotope was uniformly distributed through all the samples when the rock unit
formed. Because of the apparent success of this isochron method, it has become in recent years the cornerstone of
radioactive dating in geology.
Isochron dating questioned
However, it is this isochron dating method that has recently come under fire. Writing in the international journal Chemical
Geology,2 Y.F. Zheng of the Geochemical Institute at the University of Gottingen in Germany says:
The Rb-Sr isochron method has been one of the most important approaches in isotopic geochronology. But some of the
basic assumptions of the method are being questioned at the present time. As first developed the method assumed a
system to have: (1) the same age; (2) the same initial 87Sr/86Sr ratio; and (3) acted as a closed system. Meanwhile, the
goodness of fit of experimental data points in a plot of 87Sr/86Sr vs. 87Rb/86Sr served as a check of these assumptions.
However, as the method was gradually applied to a large range of geological problems, it soon became apparent that a
linear relationship between 87Sr/86Sr and 87Rb/86Sr ratios could sometimes yield an anomalous isochron which had no distinct
geological meaning. A number of anomalous isochrons have been reported in the literature and various terms have been
invented, such as apparent isochron (Baadsgaard et al., 1976), mantle isochron and pseudoisochron (Brooks et al., 1976a,
b), secondary isochron (Field and Ra- Heim, 1980). inherited isochron (Roddick and Compston, 1977), source isochron
(Compston and Chappell, 1979), erupted isochron (Betton, 1979; Munksgaard, 1984), mixing line (Bell and Powell, 1969;
Faure, 1977; Christoph, 1986) and mixing isochron (Zheng, 1986; Qin, 1988). Even a suite of samples which do not have
identical ages and initial 87Sr/86Sr ratios can be fitted to isochrons, such as aerial isochrons (Kohler and Muller-Sohnius,
1980; Haack et al., 1982).3
He went on to say:
Evidently, the theoretical basis of the classical Rb-Sr isochron is being challenged and some limitations of its basic
assumptions are being revealed. Some of what this paper contains is not new to isotopic geochronologists, but it is drawn
together here for the first time and is placed in a context within unifying general models for Rb-Sr dating.4However, Zhengs
paper really isnt the first time that these problems with the isochron dating method have been comprehensively highlighted
and treated mathematically. It was in fact creation scientists who first comprehensively pointed to the problems with the
isochron dating method. In a series of short articles published in the Bible-Science Newsletter in 1981, Dr. Russell Arndts,
Professor of Chemistry at St Cloud State University in Minnesota, and Dr. William Overn, a former engineer and physicist
with the National Aeronautics and Space Administration (NASA), showed how isochrons were in fact often a result of the
mixing of the radioisotopes from different sources.5 They also illustrated this with various examples from the geological
literature. They concluded:
It is clear that mixing of pre-existent materials will yield a linear array of isotopic ratios. We need not assume that the
isotopes, assumed to be daughter isotopes, were in fact produced in the rock by radioactive decay. Thus the assumption of

immense ages has not been proven. The straight lines, which seem to make radiometric data meaningful, are easily
assumed to be the result of simple mixing.(their emphasis)6They go on to suggest that the concept of mixing a material
from wide ranges seems to suggest that the earth has undergone widespread stirring. Such processes do not of course
always involve the actual physical movement of rock, rock-forming components such as mineral grains, or molten materials,
but more often involve the mixing of chemical components via fluxes of fluids, principally water, through the rocks. Zheng
concurs with this in his paper when he speaks of geological processes such as hydrothermal (hot water) alteration,
metasomatism, and metamorphism, the latter two involving changes in rocks due to fluids, temperature, and pressure.
Zheng admits:
In some cases, gain or loss of Rb and Sr from the rocks is so regular that a linear array can be produced on the
conventional isochron diagram and a biased isochron results from the altered rocks to give spurious age and initial 87Sr/86Sr
ratio estimates.7
At the end of his paper, Zheng wrote:
In conclusion, some of the basic assumptions of the conventional Rb-Sr isochron method have to be modified and an
observed isochron does not certainly define a valid age information for a geological system, even if a goodness of fit of the
experimental data points is obtained in plotting 87Sr/86Sr vs. 87Rb/86Sr. This problem cannot be overlooked, especially in
evaluating the numerical time scale. Similar questions can also arise in applying Sm-Nd and U-Pb isochron methods.8
And as if to make the point even more succinctly and clearly, Zheng also wrote in the abstract (or summary) of his paper:
As it is impossible to distinguish a valid isochron from an apparent isochron in the light of Rb-Sr isotopic data alone, caution
must be taken in explaining the Rb-Sr isochron age of any geological system.One could hardly expect a more emphatic and
complete demolition job on the isochron dating method than that! Notice also that Zheng extends his criticism to the
traditional uranium-lead (UPb) and currently-in-vogue samarium-neodymium (Sm-Nd) isochron methods.
Conclusions
Given now these criticisms from an evolutionist geochemist/geochronologist in the open scientific literature, one wonders
how quickly geochronologists world-wide will rigorously re-examine the isochron method and the results it has produced
over the past few decades. Of course, abandoning the method could hardly be countenanced, as it would mean abandoning
what has become one of the foundational cornerstones to the whole evolutionary view of the geological development of the
earth with its millions of years time-scale.Nevertheless, this attack on radioactive dating by an evolutionist in the open
scientific literature is a timely reminder that there are problems with these methods. creationists need not compromise with
the evolutionists time-scale because it is being propped up by these faulty dating methods. Rather, we should place our
confidence in the 6,000-7,000 year chronology.
Radioisotopes and the Age of the Earth
by Dr. Andrew A. Snelling on October 31, 2007
Abstract
The RATE research project demonstrated
that creationists could support a larger-scale
collaborative research effort, particularly if it
delivered significant breakthroughs on a key
challenging issues.
Keywords: radioisotopes, age of the earth,
RATE, Institute for Creation Research,
Creation Research Society, fission tracks,
radiohalos, radioisotope, dating methods,
radiocarbon, carbon-14, decay
The 19972005 RATE (Radioisotopes and the Age of the Earth) research project at the Institute for Creation Research (cosponsored by the Creation Research Society) demonstrated that creationists could support a larger-scale collaborative
research effort, particularly if it delivered significant breakthroughs on a key challenging issue. The primary focus of this
research effort was the radioactive methods for dating rocks that supposedly yield age estimates of millions and billions of
years and thus provide support for the claimed multi-billion year age for the earth. The research team assembled for this
project included:
Larry Vardiman,
D. Russell Humphreys,
Eugene F. Chaffin,
Donald DeYoung,
John R. Baumgardner,
Steven A. Austin,
Andrew A. Snelling,

Ph.D. Atmospheric Science (project co-ordinator)


Ph.D. Physics (helium diffusion)
Ph.D. Physics (theoretical models)
Ph.D. Physics
Ph.D. Geophysics (radiocarbon)
Ph.D. Geology (rock dating)
Ph.D. Geology (rock dating, fission tracks, radiohalos)

Steven W. Boyd, Ph.D. Hebraic and Cognate Studies There were numerous significant outcomes from this project:
There is visible physical evidence in rocks, namely, fission tracks and radiohalos, that a lot of nuclear decay has occurred
through earth history. Uranium atoms decay in two ways. Some uranium atoms spontaneously break apart (split or fission)
into two smaller atoms. The energy of this fission process causes the two smaller atoms to fly apart, leaving observable
linear scars called fission tracks in the host minerals that can be seen under a microscope. In most other uranium atoms, the
size of their nuclei makes them unstable, and so, radiation particles called alpha-particles are ejected from them. These
alpha-particles are like little bullets that damage the host minerals leaving physical scars. Because the alpha-particles are
ejected in all directions from around where the uranium atoms are concentrated, the result seen in cross section is a halo of
visible physical damage which can be seen under a microscope. These are called radioactive halos, abbreviated to
radiohalos. Some minerals from many levels in the geologic record were found to have high concentrations of fission
tracks and numerous radiohalos consistent with much nuclear decay having occurred, equivalent to hundreds of millions of
years worth of decay at todays slow rates. This would suggest that, the large quantity of nuclear decay must have occurred
at much faster rates than those measured today.There are often systematic differences in the radioisotope age estimates
provided by the four main radioactive dating methods from the same samples of rock units. Unstable parent atoms decay
into daughter atoms of different elements, so measuring quantities of parent and daughter atoms in rocks and minerals, and
knowing the rates at which this decay occurs, enables the calculation of when the decay process began in that rock or

mineral, which is then deemed its age estimate. An example of the results obtained is provided by the rock layer of volcanic
origin at Bass Rapids in the Grand Canyon, which yielded the following age estimates:
841.5 million years (potassium-argon)
1,060 million years (rubidium-strontium)
1,250 million years (uranium-lead)
1,379 million years (samarium-neodymium)
These four methods should have yielded the same age estimate for this volcanic rock layer because the decay of each of
the four parent atoms all began at the same time when this volcanic rock layer formed. One way these different age
estimates can be reconciled is if the different parent atoms decayed at different faster rates in the past. The parent atoms
which give the older ages decayed much more, and thus much faster, relative to the other parent atoms.There is evidence
that nuclear decay rates were grossly accelerated during a recent catastrophic episode or episodes. They are the
systematic differences in radioisotope age estimates for the same rock units, as explained in item two above, which can only
be reconciled by grossly accelerated decay rates in the past. There are co-existing uranium and polonium radiohalos in the
same mineral grains in granites from around the world. Because polonium has a fleeting existence, the polonium radiohalos
had to have formed within hours and days. However, the source of the polonium had to be the uranium which was also at
the same time producing the uranium radiohalos. So, the uranium had to be decaying extremely rapidly to supply sufficient
polonium quickly enough to form the adjacent polonium radiohalos. And finally, helium gas is a by-product of the radioactive
decay of uranium within minerals. However, this helium gas easily leaks out of the host minerals. Thus two age estimates
can be calculated for these mineral grainsone based on radioactive decay of uranium to lead, and the other based on the
rate at which the helium leaks out of the mineral grains. For certain mineral crystals it was found the uranium-lead
radioactive age estimate was 1.5 billion years, yet the helium leak age was only about 6,000 years. Because the latter is
based on experimentally verified physical laws, it can be concluded that a tremendous amount of radioactive decay (which
would take 1.5 billion years at today's decay rates) must have occurred catastrophically during some event in the last 6,000
years!There are significant detectable levels of radiocarbon (carbon-14) intrinsic within ancient coal and diamonds. Samples
from coal layers conventionally dated at 40320 million years old all yielded radiocarbon age estimates of around 50,000
years, implying that they were all deposited recently, at the same time and in the same event. Interestingly, diamonds
conventionally dated at 12 billion years old gave only slightly older radiocarbon age estimates. When it is considered that
radiocarbon levels and production rates were different in the past, these radiocarbon age estimates for these coal layers and
diamonds are direct evidence of a young earth.The mechanisms associated with how radioactive decay occurs within the
nuclei of the parent atomswhen theoretically adjustedchange decay rates. Very tiny adjustments to the nuclear forces
could produce very large changes in decay rates. It is realized that changes in fundamental constants, and also greatly
accelerated nuclear decay, are radical suggestions.Because of the RATE research results, the long-age radioactive
methods for dating rocks can now be more easily demonstrated to often be faulty, since there are problems with the three
crucial assumptions on which they are based:There are uncertainties as to the absence or presence of daughter atoms
when the rocks formed, because there is much evidence of the rocks having inherited daughter atoms that were not formed
by radioactive decay in those rocks.There is abundant evidence of widespread open-system behavior of parent and
daughter atoms. Rocks are often contaminated with extra parent and daughter atoms produced apart from radioactive
decay. Parent and daughter atoms are also removed by various geologic processes (for example, leaching by fluids)
subsequent to the rocks forming.
Nuclear decay rates may well have changed in the past.
Much research, even reported in the conventional scientific literature, has found that rocks of known age often yield
erroneously old radioactive age estimates because either one of the first two assumptions, or both, can be demonstrated to
be false. And if the radioactive clocks have not always ticked at the currently measured slow rates but were grossly
accelerated in the past, then these radioactive dating methods cannot be used to provide reliable age estimates for rocks.
After all, if these clocks dont work on rocks of known ages, how can they be trusted on rocks of unknown ages? To be
sure, there is a systematic trend of radioactive age estimates for rocks according to their positions in the geologic record, but
this would be expected if nuclear decay was grossly accelerated systematically when the rock layers were forming. For
example, rocks laid down early in the Flood would yield older ages than rocks laid down later during the Flood because the
earlier rocks would have experienced more accelerated radioactive decay.
U-Th-Pb Dating: An Example of False Isochrons
by Dr. Andrew A. Snelling on December 9, 2009
Abstract
As with other isochron methods, the U-Pb isochron
method has been questioned in the open literature,
because often an excellent line of best fit between ratios
obtained from a set of good cogenetic samples gives a
resultant isochron and yields a derived age that has no
distinct geological meaning. At Koongarra, Australia, UTh-Pb isotopic studies of uranium ore, host rocks, and
soils have produced an array of false isochrons that yield ages that are geologically meaningless. Even a claimed nearconcordant U-Pb age of 862 Ma on one uraninite grain is identical to a false Pb-Pb isochron age, but neither can be
connected to any geological event. Open system behavior of the U-Th-Pb system is clearly the norm, as is the resultant
mixing of radiogenic Pb with common or background Pb, even in soils in the surrounding region. Because no geologically
meaningful results can be interpreted from the U-Th-Pb data at Koongarra (three uraninite grains even yield a 232Th/208Pb
age of 0 Ma), serious questions must be asked about the validity of the fundamental/foundational basis of the U-Th-Pb
dating method. This makes the task of creationists building their model for the geological record much easier, since claims
of U-Th-Pb radiometric dating having proven the claimed great antiquity of the earth, its strata and fossils can be safely
side-stepped.
Keywords: geochronology, U-Th-Pb isotopes, isochrons, uranium ore, soils
This paper was originally published in the Proceedings of the Third International Conference on Creationism, pp. 497504
(1994) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh.
Introduction
Radiometric dating has now been used for almost 50 years to establish beyond doubt the earths multibillion year
geological column. Although this column and its age was firmly settled well before the advent of radiometric dating, the
latter has been successfully used to help quantify the ages of the strata and the fossils in the column, so that in many

peoples minds today radiometric dating has proved the presumed antiquity of the earth. Of the various methods, uraniumthorium-lead (U-Th-Pb) was the first used and it is still widely employed today, particularly when zircons are present in the
rocks to be dated. But the method does not always give the expected results, leading to fundamental questions about its
validity.In his conclusion in a recent paper exposing shortcomings and criticizing the validity of the popular rubidiumstrontium (Rb-Sr) isochron method, Zheng wrote:
. . . some of the basic assumptions of the conventional Rb-Sr isochron method have to be modified and an observed
isochron does not certainly define a valid age information for a geological system, even if a goodness of fit of the
experimental data points is obtained in plotting 87Sr/86Sr vs. 87Rb/86Sr. This problem cannot be overlooked, especially in
evaluating the numerical timescale. Similar questions can also arise in applying Sm-Nd and U-Pb isochron methods.1
Amongst the concerns voiced by Zheng were the problems being found with anomalous isochrons, that is, where there is an
apparent linear relationship between 87Sr/86Sr and 87Rb/86Sr ratios, even an excellent line of best fit between ratios obtained
from good cogenetic samples, and yet the resultant isochron and derived age have no distinct geological meaning. Zheng
documented the copious reporting of this problem in the literature where various names had been given to these anomalous
isochrons, such as apparent isochron, mantle isochron and pseudoisochron, secondary isochron, source isochron, erupted
isochron, mixing line, and mixing isochron.
Similar anomalous or false isochrons are commonly obtained from U-Th-Pb data, which is hardly surprising given the
common open system behavior of the U-Th-Pb system. Yet in the literature these problems are commonly glossed over or
pushed aside, but their increasing occurrence from a variety of geological settings does seriously raise the question as to
whether U-Th-Pb data ever yields any valid age information. One such geological setting that yields these false U-Th-Pb
isochrons is the Koongarra uranium deposit and the surrounding area (Northern Territory, Australia).
The Koongarra Area
The Koongarra area is 250 km east of Darwin (Northern Territory, Australia) at latitude 1252'S and longitude 13250'E. The
regional geology has been described in detail by Needham and Stuart-Smith2 and by Needham,3, 4 while
Snelling5 describes the Koongarra uranium deposit and the areas local geology.The Koongarra uranium deposit occurs in a
metamorphic terrain that has an Archean basement consisting of domes of granitoids and granitic gneisses (the Nanambu
Complex), the nearest outcrop being 5 km to the north. Some of the lowermost overlying Lower Proterozoic metasediments
were accreted to these domes during amphibolite grade regional metamorphism (estimated to represent conditions of 58
kb and 550630C) at 18001870 Ma. Multiple isoclinal recumbent folding accompanied metamorphism. The Lower
Proterozoic Cahill Formation flanking the Nanambu Complex has been divided into two members. The lower member is
dominated by a thick basal dolomite and passes transitionally upwards into the psammitic upper member, which is largely
feldspathic schist and quartzite. The uranium mineralization at Koongarra is associated with graphitic horizons within
chloritized quartz-mica (feldspar garnet) schists overlying the basal dolomite in the lower member. A 150 Ma period of
weathering and erosion followed metamorphism. A thick sequence of essentially flat-lying sandstones (the Middle
Proterozoic Kombolgie Formation) was then deposited unconformably on the Archean- Lower Proterozoic basement and
metasediments. At Koongarra subsequent reverse faulting has juxtaposed the lower Cahill Formation schists and Kombolgie
Formation sandstone.Owing to the isoclinal recumbent folding of metasedimentary units of the Cahill Formation, the typical
rock sequence encountered at Koongarra is probably a tectono-stratigraphy (from youngest to oldest.)
muscovite-biotite-quartz-feldspar schist (at least 180 m thick)
garnet-muscovite-biotite-quartz schist (90100 m thick)
sulphide-rich graphite-mica-quartz schist (garnet) (about 25 m thick)
distinctive graphite-quartz-chlorite schist marker unit (58 m thick)
quartz-chlorite schist (illite, garnet, sillimanite, muscovite) (50 m thick)the mineralized zone
reverse fault breccia (57 m thick)
sandstone of the Kombolgie Formation
Polyphase deformation accompanied metamorphism of the original sediments, that were probably dolomite, shales and
siltstones. Johnston6 identified a D2 event as responsible for the dominant S 2 foliation of the schist sequence, which at
Koongarra dips at 55 to the south-east. The dominant structural feature, however, is the reverse fault system that dips at
about 60 to the south-east, sub-parallel to the dominant S2 foliation and lithological boundaries, just below the mineralized
zone.
The Uranium Deposit
There are two discrete uranium orebodies at Koongarra, separated by a 100 m wide barren zone. The main (No. 1) orebody
has a strike length of 450 m and persists to 100 m depth. Secondary uranium mineralization is present in the weathered
schists, from below the surficial sand cover to the base of weathering at depths varying between 25 and 30 m. This
secondary mineralization has been derived from decomposition and leaching of the primary mineralized zone, and forms a
tongue-like fan of ore-grade material dispersed down-slope for about 80 m to the south-east. The primary uranium
mineralized zone in cross-section is a series of partially coalescing lenses, which together form an elongated wedge dipping
at 55 to the south-east within the host quartz-chlorite schist unit, subparallel to the reverse fault. True widths average 30 m
at the top of the primary mineralized zone but taper out at about 100 m below surface and along strike.
Superimposed on the primary prograde metamorphic mineral assemblages of the host schist units is a distinct and
extensive primary alteration halo associated, and cogenetic, with the uranium mineralization. This alteration extends for up
to 1.5 km from the ore in a direction perpendicular to the host quartz-chlorite schist unit, because the mineralization is
essentially stratabound. The outer zone of the alteration halo is most extensively developed in the semi-pelitic schists, and is
manifested by the pseudomorphous replacement of biotite by chlorite, rutile and quartz, and feldspar by sericite. Silicification
has also occurred in fault planes and within the Kombolgie Formation sandstone beneath the mineralization, particularly
adjacent to the reverse fault.Association of this outer halo alteration with the mineralization is demonstrated by the apparent
symmetrical distribution of this alteration about the orebody. In the inner alteration zone, less than 50 m from ore, the
metamorphic rock fabric is disrupted, and quartz is replaced by pervasive chlorite and phengitic mica, and garnet by chlorite.
Uranium mineralization is only present where this alteration has taken place.The primary ore consists of uraninite veins and
veinlets (110 mm thick) that cross-cut the S 2 foliation of the brecciated and hydrothermally altered quartz-chlorite schist
host. Groups of uraninite veinlets are intimately intergrown with chlorite, which forms the matrix to the host breccias. Small
(10100 m) euhedral and subhedral uraninite grains are finely disseminated in the chloritic alteration adjacent to veins, but
these grains may coalesce to form clusters, strings, and massive uraninite. Coarse colloform and botryoidal uraninite
masses and uraninite spherules with internal lacework textures have also been noted, but the bulk of the ore appears to be
of the disseminated type, with thin (<0.5 mm) discontinuous wisps and streaks of uraninite, and continuous strings both
parallel and discordant to the foliation (S2), and parallel to phyllosilicate (001) cleavage planes.Associated with the ore are
minor volumes (up to 5%) of sulphides, which include galena and lesser chalcopyrite, bornite, and pyrite, with rare grains of
native gold, clausthalite (PbSe), gersdorffite-cobaltite (NiAsS-CoAsS) and mackinawite (Fe, Ni), S. Galena is the most

abundant, commonly occurring as cubes (510 m wide) disseminated in uraninite or gangue, and as stringers and veinlets
particularly filling thin fractures within uraninite. Galena may also overgrow clausthalite, and replace pyrite and chalcopyrite.
Chlorite, predominantly magnesium chlorite, is the principal gangue, and its intimate association with the uraninite indicates
that the two minerals formed together.Oxidation and alteration of uraninite within the primary ore zone has produced a
variety of secondary uranium minerals, principally uranyl silicates.7 Uraninite veins, even veins over 1 cm wide, have been
completely altered in situ. Within the primary ore zone this in situ replacement of uraninite is most pronounced immediately
above the reverse fault breccia, and this alteration and oxidation diminish upwards stratigraphically. It is accompanied by
hematite staining of the schists, the more intense hematite alteration in and near the reverse fault breccia being due to
hematite replacement of chlorite. The secondary mineralization of the dispersion fan in the weathered schist above the No. 1
orebody is characterized by uranyl phosphates found exclusively in the tail of the fan. Away from the tail uranium is
dispersed in the weathered schists and adsorbed onto clays and iron oxides.The age of the uranium mineralization is
problematic. The mineralization, however, must post-date both the Kombolgie Formation sandstone and the Koongarra
reverse fault, since it occupies the breccia zones generated by the post-Kombolgie reverse faulting. The pattern of alteration
which is intimately associated with the ore also crosses the reverse fault into the Kombolgie sandstone beneath the ore
zone, so this again implies that the ore was formed after the reverse fault and therefore is younger than both the Kombolgie
sandstone and the reverse fault. Because of these geological constraints, Page, Compston, and Needham8 suggested the
mineralization was younger than 16001688 Ma because of their determination of the timing of the Kombolgie Formation
deposition to that period. Sm-Nd isotopic data obtained on Koongarra uraninites9, 10 appears to narrow down the timing of
mineralization to 15501650 Ma. It is unclear as to when deep ground-water circulation began to cause oxidation and
alteration of the primary uraninite ore at depth, but Airey, Golian, and Lever11 suggest that the weathering of the primary ore
to produce the secondary dispersion fan in the weathered schists above the No. 1 orebody seems to have begun only in the
last 13 Ma.
U-Th-Pb Data
Dating of the primary ore
Hills and Richards12 isotopically analyzed individual grains of uraninite and galena that had been handpicked from drill core.
Only one of the five uraninite samples gave a near-concordant age of 862 Ma, that is, the sample plotted almost on the
standard concordia curve, and Hills and Richards13 interpreted this as recording fresh formation of Pb-free uraninite at 870
Ma. The other four uraninite samples all lie well below concordia and do not conform to any regular linear array. Hills and
Richards were left with two possible interpretations. On the one hand, preferential loss of the intermediate daughter products
of 238U (that is, escape of radon, a gas) would cause vertical displacement of points below an episodic-loss line, but this
would only produce a significant Pb isotopic effect if the loss had persisted for a very long proportion of the life of the
uraninite (which is incidentally not only feasible but likely). Alternatively, they suggested that contamination by small amounts
of an older (pre-900 Ma) Pb could cause such a pattern as on their concordia plot, to which they added mixing lines that
they postulated arose from the restoration to each uraninite sample of the galena which separated from it.This of course
assumes that the Pb in the galenas was also derived predominantly from uranium decay. They plotted their Pb ratios in all
their uraninite samples on a standard 207Pb/206Pb diagram, and contended that the pattern of data points did not conform to a
simple age interpretation. Instead, they contended that the scatter of points could be contained between two lines radiating
from the diagrams origin, lines that essentially represented isochrons for uraninites and galenas from the Ranger and
Nabarlek uranium deposits in the same geological region. From the positions of the Koongarra uraninites and galenas on
these diagrams they claimed that the galenas contained left-over radiogenic Pb from earlier uraninites as old as 17001800
Ma (the age of the Ranger uranium mineralization), these earlier uraninites being obliterated by the uranium having
remobilized at 870 Ma, the age of the lone Pb-free uraninite sample.In a separate study Carr and Dean14 isotopically
analyzed whole-rock samples from the Koongarra primary ore zone. These were samples of drill core that had been
crushed. Their isotopic data on four samples were plotted on a U-Pb isochron diagram and indicated a non-systematic
relationship between the 238U parent and the 206Pb daughter. In other words, the quantities of 206Pb could not simply be
accounted for by radioactive decay of 238U, implying open system behavior. They also plotted their four results on a
standard 207Pb/206Pb isochron diagram and found that these samples fell on a very poorly defined linear array whose
apparent age they did not quantify.
Dating of weathered rocks and soils
Carr and Dean15 also isotopically analyzed a further nine whole-rock samples from the weathered schist zone at
Koongarra. Some of these samples were again crushed drill core, but the majority were crushed percussion drill chips.
When their isotopic data were plotted on a U-Pb isochron diagram six of the nine samples plotted close to the reference
1000 Ma isochron, while the other three were widely scattered. However, on the 207Pb/206Pb diagram all nine weathered rock
samples plotted on a linear array which gave an apparent isochron age of 127050 Ma.In an unrelated investigation,
Dickson, Gulson, and Snelling16, 17 collected soil samples from above the mineralization at Koongarra and from
surrounding areas, and these were analyzed for Pb isotopes to see if there was any Pb-isotopic dispersion halo around the
mineralization sufficiently large enough to warrant the use of Pb-isotopic analyses of soils as an exploration technique to find
new uranium orebodies. The technique did in fact work, Pb-isotopic traces of the deeply buried No. 2 orebody mineralization
being found in the soils above. This mineralization, 40 m below the surface, is blind to other detection techniques.Dickson,
Gulson, and Snelling18 found that all 113 soil samples from their two studies were highly correlated (r = 0.99986) on a
standard 207Pb/206Pb diagram, yielding an apparent (false) isochron representing an age of 144520 Ma for the samples.
However, most of the soil samples consisted of detritus eroded from the Middle Proterozoic Kombolgie sandstone, so
because the samples from near the mineralization gave a radiogenic Pb signature Dickson et al. interpreted the false
isochron as being due to mixing of radiogenic Pb from the uranium mineralization with the common Pb from the
sandstone.
Discussion
Snelling19 has already highlighted a telling omission by Hills and Richards.20 Having included all the Pb isotopic ratios they
had obtained on their five uraninite samples, they tabulated also the derived ages, except for those obtainable from 208Pb.
These Th-derived dates should normally be regarded as the most reliable, since Th is less mobile in geochemical
environments and therefore open system behavior is less likely than for U. Significantly, three of the five uraninite samples
therefore give, within their experimental error, a 0 Ma age.21 In any case, their age of 17001800 Ma for the first
generation of uranium mineralization at Koongarra neither fits the geological criteria for an expected 15501600 Ma age,
nor does their 870 Ma date correlate with any geological event capable of remobilizing U and Pb to produce the presumed
second generation of uranium mineralization.Using Ludwig,22 standard 207Pb/206Pb diagrams were prepared for the
uraninite, galena, and whole-rock data sets, and combinations thereof, to check the regression statistics and possible
derived isochrons using the standard York23 method. In each case the mean square of weighted deviates (MSWD), which
tests the goodness of fit of data to a line, is large to extremely large, which reflects in the derived isochron ages of

841140 Ma (uraninites), 1008420 Ma (galenas), 668330 Ma (whole-rocks), 818150 Ma (uraninites plus galenas) and
863130 Ma (all three data sets combined), all ages being within the 95% confidence limits. It is perhaps fortuitously
significant that the combination of all three data sets yields an isochron age of 863130 Ma, almost identical to Hills and
Richards near-concordant age of 862 Ma, although this was using a line-fitting routine of Ludwig that assigns equal
weights and zero error-correlations to each data point to avoid the mistake of weighting the points according to analytical
errors when it is clear that some other cause of scatter is involved, which is clearly the case here. The normal York algorithm
assumes that the only cause for scatter from a straight line are the assigned errors, and for the combined data set here the
amount of scatter calculated thereby yields an astronomical MSWD of 669,000 and a bad line of fit that yields an isochron
age of 1632410 Ma. This result may make more geological sense, but the regression statistics are such that derivation
of any age information from these data is totally unjustified, even though it can be rightfully argued that these samples form
a cogenetic set (they are all samples of uranium ore or its components from the same primary ore zone at Koongarra).It is
not uncommon to find that ages derived from standard 207Pb/206Pb plots are erroneous, even though the data fit welldefined linear arrays (isochrons). Ludwig, Nash, and Naeser24 found that this was due to migration of both Pb and
radioactive daughters of 238U yielding a 207Pb/206Pb isochron giving superficially attractive results which would nonetheless
be seriously misleading because the derived age (in their example) was more than six times higher than the U-Pb
isochron age. Similarly, Cunningham et al.25 obtained 207Pb/206Pb isochron ages up to 50 times higher than those derived
from more reliable U-Pb isochrons for wholerock uranium ore samples, even though the apparent slight degree of scatter
is almost entirely a misleading artifact. Ironically, at Koongarra the U-Pb isochron using Ludwig 26 yields an age of
857149 Ma (with an MSWD of 13,400, tolerably large compared to that obtained with the Pb-Pb isochron), almost identical
to the fortuitous Pb-Pb isochron age obtained using Ludwigs modified algorithm on the combined three data sets
(863130 Ma), as well as Hills and Richards single near-concordant 862 Ma age.Snelling and Dickson27 demonstrated
that there is significant radiometric disequilibrium in the primary ore and surrounding host rocks at Koongarra due to the
redistribution of both U and its Ra decay product. That Ra mobility at depth in the primary ore zone is currently more
significant than U migration was confirmed by Dickson and Snelling,28which of course results ultimately in the redistribution
of 206Pb, the end-member of the whole 238U decay chain. Dickson, Giblin, and Snelling29 and Dickson, Gulson, and
Snelling30 demonstrated that Ra is transported through the unweathered rocks in this area in the ground waters, while
Davey, Dudaitis, and OBrien31 determined the emanation rate of radon gas from the Koongarra No. 1 orebody, an everpresent hazard in uranium ore mining operations. The radon gas is known to migrate along fractures and rise through the
ground over considerable distances to form a halo in the air above, while radon is also transported in ground waters.These
observations alone demonstrate the open-system behavior of the U-Th-Pb system that renders meaningless any age
information derived. However, both Hills32 and Snelling33 34 have recognized that U also has migrated on a considerable
scale in the primary ore zone, since supergene uraninites, often with colloform banding, are found as fracture and cavity
infillings, and between quartz and gangue grain boundaries. The unit cell dimensions of these uraninites, plus this textural
evidence, supports the conclusion that these uraninites have precipitated after dissolution of earlier formed uraninite and
transportation in low-temperature ground waters. With such wholesale migration of U also, all attempts at dating must be
rendered useless, especially when whole-rock samples, in which different generations of uraninites are lumped together, are
used.In contrast to the poor-fitting linear arrays produced from the Pb-Pb data of minerals and whole-rocks from the primary
ore zone, that all appear to give an apparent (false) isochron age grouped around 857863 Ma, both Carr and Dean35 and
Dickson, Gulson, and Snelling36 found that weathered whole-rock and soil samples produced good fitting linear arrays that
would normally represent isochrons that yield ages of 1270 Ma and 1445 Ma respectively. The weathered whole-rock
samples all of course come from Koongarra itself, and consist of secondary ore samples from the weathered schist zone,
plus weathered schist samples that contain uranium dispersed down-slope by ground waters moving through the weathered
rock. Because these whole-rock samples come from a volume of rock through which U is known to be migrating, leading to
redistribution not only of U but of its decay products, it is therefore very surprising to find that these whole-rock samples
define a good enough linear array to yield an isochron. Even the observed scatter calculated using Ludwig37 is much less
than that associated with fitting an isochron to the 207Pb-206Pb data from the primary ore zone samples, which is again
surprising given U migration in the weathered zone, the data from which one would expect to show considerable scatter and
thus no age consensus. Furthermore, it is baffling as to why the isochron-derived age should be so much older than
the age of the primary ore, which of course is ultimately the source through weathering and ground-water transport of the
U, decay products and the stable Pb isotopes. Perhaps the only explanation is that the isochron represents the mixing of
radiogenic Pb from the mineralization with the common or background Pb in the surrounding schists.The idea of such an
isochron being a mixing line was suggested by Dickson, Gulson, and Snelling.38 They were however, dealing with the Pb
isotopic data obtained from soil samples collected from depths of only about 3040 cm, the majority of which represented
sandy soils consisting of detritus eroded from the Kombolgie sandstone. For this mixing explanation to be feasible there
should be some other evidence of mobilization of Pb in the area. Dickson, Gulson, and Snelling found that not only were
there high 206Pb/204Pb ratios in three of their soil samples from the near-surface (01 m) zone south of the No. 1 orebody, but
there was a lack of any other Useries daughter products in the same samples. This near-surface zone is inundated for
approximately six months of the year as a result of the high monsoonal rainfall in this tropical area. Towards the end of the
ensuing six-month dry season, the water table has been known to drop in some cases more than ten meters from its wet
season high. This means that the top of the weathered schist zone is regularly fluctuating between wet and dry conditions,
so that any trace elements such as Pb leached from the weathered ore and transported by ground water in the weathered
schist zone would also be dispersed vertically up into the thin surficial sand cover on top of the weathered schistthe sandy
soils that were sampled by Dickson, Gulson, and Snelling.39, 40 Snelling41 found that Pb was a significant pathfinder
element for uranium ore in the Koongarra environment, anomalous Pb being present in the surficial sand cover above the
zone of weathered primary ore, and that there was even hydrodynamic dispersal of Pb at a depth of 0.51.5 m. Dickson,
Gulson, and Snelling42 found a similarity between the isotopic ratios for Pb extracted from their soil samples by either a mild
HCI-hydroxylamine (pH 1) or a strong 7M HCI-7M HNO3 leach, which indicates that Pb is loosely attached to sand grain
surfaces in the samples rather than tightly bound in silicate or resistate mineral lattices. This in turn suggests Pb is adsorbed
from ground waters, meaning that radiogenic Pb is being added to the common or background Pb in the sand by both
vertical and lateral ground-water dispersion.However, not all of Dickson, Gulson and Snellings soil samples came from the
immediate area to the Koongarra orebodies, nor were they all the samples of Kombolgie sandstone detritus. That this mixing
line explanation for the apparent isochron is clearly demonstrated for these samples from the immediate Koongarra area is
not in question, although it is somewhat surprising that these soil samples should give an apparent isochron age
somewhat higher than that obtained from the weathered schist samples beneath. Indeed, the common or background Pb in
the respective samples should reflect an older apparent age in the schists compared to the sandstone, due to their relative
ages based on geological relationships between them. However, the apparent ages are the other way around, the sandy
soils yielding an older apparent age compared to that yielded by the weathered schists. Perhaps this difference is a

reflection of the extent of mixing in each type of sample at their respective levels in the weathering profile. Nevertheless,
what is astounding is that Dickson, Gulson, and Snelling43 found that even though several of their soil samples consisted of
weathered schist or basement granite (containing accessory zircon) up to 17 km from the known uranium mineralization,
they still plotted on the same apparent isochron. Indeed, the fit is comparatively good, as indicated by the MSWD of only
964 using Ludwig,44 yet much of this observed scattered can be attributed to two samples out of the 113, one of which was
subsequently known to be probably contaminated by cuttings from an adjacent drill hole.45 If that sample is removed from
the regression analysis the MSWD drops to 505, indicating that almost half of the observed scatter is due to that one data
point alone. If the data point that is the next worst for fitting to the apparent isochron is removed, then the MSWD drops by
a further 315 to a mere 190. Yet in both cases the apparent isochron or mixing line still has lying on or close to it the
samples from up to 17 km away from the known uranium mineralization and the samples that are not Kombolgie sandstone
detritus. The final isochron fitted to the remaining 111 samples still yields an age of 142018 Ma.
While Carr and Deans46 nine weathered whole-rock samples are not strictly cogenetic with Dickson, Gulson and Snellings
113 soil samples, the two sample sets are obviously related because the source of the radiogenic Pb in the majority of the
soil samples from the immediate Koongarra area is the same as that in the weathered rocks. Not surprisingly, when the
regression analysis was performed on Carr and Deans nine weathered wholerock samples using Ludwig,47 the MSWD for
the observed scatter was 24,100, indicating a poor fit to an isochron which yielded an age of 1287120 Ma. Yet when
these nine samples were added to the 113 soil samples the MSWD dropped substantially to 1210, and not surprisingly the
fitted isochron yielded an age of 134627 Ma, an isochron age intermediate between those of the two data sets being
combined. However, when the two soil samples responsible for the majority of the scatter in that data set were removed the
MSWD dropped to 430 and yielded an isochron age of 133617 Ma.As with all the other apparent isochron ages this
result has no apparent geological meaning, because there is no geological event to which these ages might correlate.
Indeed, even in the evolutionary time frame the weathering of the Koongarra uranium mineralization is extremely recent,
and in any case these ages derived from Pb-Pb isochrons from the weathered rock and soil samples are much older
than the supposedly more reliable U-Pb isochron age of the Koongarra primary ore. But since that latter result has no
apparent geological meaning, because it also cannot be correlated with any known geological event, nothing then is certain
at all from any of these U-Th-Pb isotopic studies of the Koongarra ores, rocks, and surrounding soils. Indeed, it is just as
certain that the primary ore is 0 years old, based on three 232Th/208Pb single sample ages, as is the claim that one nearconcordant result means that there was formation of Pb-free uraninite at 870 Ma. After all, this postulated formation of Pbfree uraninite is supposed to have occurred in an environment where there was Pb left over from an earlier 17001800 Ma
original uranium mineralization for which we no longer have any evidence, textural or otherwise, apart from a rather tenuous
interpretation of Pb isotopic evidence that has otherwise shown itself to be devoid of any capability of providing any age
information.All these results raise serious fundamental questions about the claimed validity of the U-Th-Pb dating method.
It may seem reasonable to regard an apparent isochron as a mixing line within the restricted area close to the known
source of radiogenic Pb, which can be shown by independent evidence to be migrating into rocks and soils that contain
common or background Pb in the immediate environs. However, it strains all credulity to suggest that a false isochron
through a data set derived from samples representing a variety of rock types, of significantly different evolutionary ages,
over an area of up to 17 km lateral extent from the known radiogenic Pb source, can still represent mixing! One can only
conclude that all assumptions used to derive the estimates of common or background Pb, including models for the
supposed evolution of the stable Pb isotopes through earth history, from their presumed commencement on the protoearth
with its claimed original Pb isotope content some 4.6 billion or so years ago, cannot be valid. Equally, we cannot be sure
what the U-Th-Pb systems isotopic ratios really mean, because the basic assumptions that are foundational to the
interpretation of these isotopic ratios are fatally flawed. Not only has open system behavior of these isotopes been
demonstrated as the norm, but even where there is an apparent isochron with an excellent goodness of fit the derived
age is invariably geologically meaningless. Thus creationists need not be hindered in their building of the creation-Flood
young-earth model for the geological record by the many claims in the open geological literature that U-Th-Pb radiometric
dating has proved the presumed great antiquity of the earth, and the strata and fossils of the so-called geological column.
Conclusion
The concerns raised by Zheng48 regarding U-Pb isochrons are warranted. At Koongarra a 207Pb/206Pb isochron produced
from 11 hand-picked uraninite and galena grains, plus four whole-rock samples, yields an age of 863 Ma, the same as a
near-concordant age from one of the uraninite grains. Nine weathered wholerock samples yield an isochron age of 1270
Ma, while 113 soil samples produce an excellent isochron with an age of 1445 Ma. All of these ages are geologically
meaningless. While the apparent isochron produced by the soil samples may be identified as a mixing line, produced by the
mixing of radiogenic Pb with common or background Pb in the surrounding rocks and soils, even this explanation strains
credulity because the samples come from up to 17 km away from known uranium mineralization, and a few of the soil
samples represent different rock types. Not only then has open system behavior of these isotopes been demonstrated, but
apparent isochrons and their derived ages are invariably geologically meaningless. Thus none of the assumptions used
to interpret the U-Th-Pb isotopic system to yield ages can be valid. If these assumptions were valid, then the 232Th/208Pb
age of 0 Ma for three of the five uraninite samples should be taken seriously. Creationists should therefore not be
intimidated by claims that U-Th-Pb radiometric dating has proved the presumed great antiquity of the earth, and the
strata and fossils of the so-called geological column.

The Failure of U-Th-Pb Dating at Koongarra, Australia


by Dr. Andrew A. Snelling on April 1, 1995
Originally published in Journal of Creation 9, no 1: 71-92.
Abstract
As with other radiometric dating methods, the U-Pb and Pb-Pb isochron methods have been questioned in the open
literature, because often an excellent line of best fit between ratios obtained from a set of good cogenetic samples gives a
resultant isochron and yields a derived age that has no geological meaning. At the Koongarra uranium deposit, Australia,
there is ample evidence of open system behaviour, or repeated migration, of U and Pb ore textures, mineral chemistry,
supergene alteration, uranium/daughter disequilibrium, and groundwater and soil geochemistry. Yet U-Th-Pb isotopic studies
of the uranium ore, host rocks and soils have produced an array of false isochrons that yield ages which are geologically
meaningless. Even a claimed near-concordant U- Pb age of 862 Ma (million years) on one uraninite grain is identical to a
false Pb-Pb isochron age but neither can be connected to any geological event. The open system behaviour of the U-ThPb system is clearly the norm, as is the resultant mixing of radiogenic Pb with common or background Pb, even in soils in
the surrounding region, apparently even up to 17 km away! Because no geologically meaningful results can be interpreted

from the U-Th-Pb data at Koongarra (three uraninite grains even yield a 232Th/208Pb age of 0 Ma), serious questions must
be asked about the validity of the fundamental/foundational basis of the U-Th-Pb dating method. This makes the task of
creationists building their model for the geological record much easier, since claims of U-Th-Pb radiometric dating having
proven the claimed great antiquity of the earth, its strata and fossils can be justifiably ignored.
Introduction
Radiometric dating has now been used for almost 50 years to establish beyond doubt the multi-billion year age of the
earths geological column. Although this column and its age was firmly settled well before the advent of radiometric dating,
the latter has been used to quantify the, ages of the strata and the fossils in the column, so that in many peoples minds
today radiometric dating has proved the presumed antiquity of the earth.However, it is important to remember that all
radiometric dating methods are based on three main assumptions:The physico-chemical system must have always been closed. Thus no parent, daughter or other decay products within the
system can have been removed, and no parent, daughter or other decay products from outside the system can have been
added.The system must initially have contained none of its daughter elements or decay products, or at the very least we
need to know the starting conditions/state of the decay system.
The decay rate, referred to as the half-life of the radioactive parent element, must have always been the same, that is,
constant.
The highly speculative nature of all radiometric dating methods becomes apparent when one realizes that none of the above
assumptions is either valid or provable. Put simply, none of these assumptions can have been observed to have always
been true throughout the supposed millions of years the radioactive elements have presumed to have been decaying.
Of the various radiometric methods, uranium-thorium- lead (U-Th-Pb) was the first used and it is still widely employed today,
particularly when zircons are present in the rocks to be dated. But the method does not always give the expected results,
leading to fundamental questions about its validity. Indeed, the U- Th-Pb system is well known to be prone to open system
behaviour, with U being particularly geochemically mobile, meaning that U is readily lost from the crystal lattices of the
minerals used for dating, including zircons. Pb is also prone to diffusion from minerals. Thus it is questionable as to why
this radiometric dating method is still used. Instead, it is increasingly being applied in more sophisticated ways to
geological dating problems.
In the conclusion to a recent paper exposing shortcomings and criticising the validity of the popular rubidium-strontium (RbSr) isochron method, Zheng wrote:
. . . some of the basic assumptions of the conventional Rb-Sr isochron method have to be modified and an observed
isochron does not certainly define a valid age information for a geological system, even if a goodness of fit of the
experimental data points is obtained in plotting 87Sr/86Sr vs. 87Rb/86Sr. This problem cannot be overlooked, especially in
evaluating the numerical time scale. Similar questions can also arise in applying Sm-Nd and U-Pb isochron methods1
Amongst the concerns voiced by Zheng were the problems being found with anomalous isochrons, that is, where there is an
apparent linear relationship between 87Sr/86Sr and 87Rb/86Sr ratios, even an excellent line of best fit between ratios obtained
from good cogenetic samples, and yet the resultant isochron and derived age have no distinct geological meaning. Zheng
documented the copious reporting of this problem in the literature where various names had been given to these anomalous
isochrons, such as apparent isochron, mantle isochron and pseudoisochron; secondary isochron, inherited isochron, source
isochron, erupted isochron, mixing line, and mixing isochron.Similar anomalous or false isochrons are commonly obtained
from U- Th-Pb data, which is hardly surprising given the common open system behaviour of the U- Th-Pb system. Yet in the
literature these problems are commonly glossed over or pushed aside, but their increasing occurrence from a variety of
geological settings does seriously raise the question as to whether U-Th-Pb data ever yields any valid age information.
One such geological setting that yields these false U -Th -Pb ages and isochrons is the Koongarra uranium deposit and
the surrounding area (Northern Territory, Australia).

Figure 1. Regional geology map showing the location of the Koongarra uranium deposit

Figure 2. Local geology map showing the location of the Koongarra No. 1 and No. 2 orebodies. Because of surficial cover
the geological units and outline of the mineralisation are projected to the surface from the base of weathering.
The Koongarra Area
The Koongarra area is 250 km east of Darwin (Northern Territory, Australia) at latitude 1252S and longitude 13250E. The
regional geology has been described in detail by Needham and Stuart-Smith 2 and by Needham3,4 (see Figure 1), while
Snelling5 describes the Koongarra uranium deposit and the areas local geology (see Figure 2).
The Koongarra uranium deposit occurs in a metamorphic terrain that has an Archaean basement consisting of domes of
granitoids and granitic gneisses (the Nanambu Complex), the nearest outcrop being 5 km to the north (see Figure 1). Some
of the lowermost overlying Lower Proterozoic metasediments were accreted to these domes during amphibolite grade
regional metamorphism (estimated to represent conditions of 5-8 kb and 550-630 C) at 1800- 1870 Ma (million years ago,
according to conventional evolutionary dating). Multiple isoclinal recumbent folding accompanied metamorphism. The Lower
Proterozoic Cahill Formation flanking the Nanambu Complex has been divided into two members. The lower member is
dominated by a thick basal dolomite and passes transitionally upwards into the psammitic upper member, which is largely
feldspathic schist and quartzite. The uranium mineralisation at Koongarra is associated with graphitic horizons within
chloritised quartz-mica (feldspar garnet) schists overlying the basal dolomite in the lower member (see Figures 2 and 3).
A 150 Ma period of weathering and erosion followed metamorphism. A thick sequence of essentially flat-lying sandstones
(the Middle Proterozoic Kombolgie Formation) was then deposited unconformably on the Archaean-Lower Proterozoic
basement and metasediments. At Koongarra subsequent reverse faulting has juxtaposed the lower Cahill Formation schists
and Kombolgie Formation sandstone.

Figure 3. Simplified cross section through the No. 1 orebody, Koongarra, showing geology, distribution of uranium minerals
and alteration, and present groundwater flow.
Owing to the isoclinal recumbent folding of metasedimentary units of the Cahill Formation, the typical rock sequence
encountered at Koongarra is probably a tectono-stratigraphy (see Figure 3):Hanging Wall

-muscovite-biotite-quartz-feldspar schist (at least 180m thick)


-garnet-muscovite-biotite-quartz schist (9-100 m thick)
-sulphide-rich graphite-mica-quartz schist (garnet) (about 25 m thick)
-distinctive graphite-quartz-chlorite schist marker unit (5-8 m thick)

Mineralised
Zone

-quartz-chlorite schist (illite, garnet, sillimanite, muscovite) (50 m thick)

Footwall

-reverse fault breccia (5-7m thick)


-sandstone of the Kombolgie Formation

Polyphase deformation accompanied metamorphism of the original sediments, that were probably dolomite, shales and
siltstones. Johnston6 identified a D2 event as responsible for the dominant S2 foliation of the schist sequence, which at
Koongarra dips at 55 to the south-east The dominant structural feature, however, is the reverse fault system that dips at
about 60 to the south-east, sub-parallel to the dominant S2 foliation and lithological boundaries, just below the mineralised
zone.
The Uranium Deposit
There are two discrete uranium orebodies at Koongarra, separated by a 100 m wide barren zone (see Figure 2). The main
(No.1) orebody has a strike length of 450 m and persists to 100 m depth. Secondary uranium mineralisation is present in the
weathered schists, from below the surficial sand cover to the base of weathering at depths varying between 25 and 30 m
(see Figure 3). This secondary mineralisation has been derived from decomposition and leaching of the primary mineralised
zone, and forms a tongue-like fan of ore-grade material dispersed down-slope for about 80 m to the southeast. The primary
uranium mineralised zone in cross-section is a series of partially coalescing lenses, which together form an elongated
wedge dipping at 55 to the southeast within the host quartz-chlorite schist unit, sub-parallel to the reverse fault. True widths
average 30 m at the top of the primary mineralised zone but taper out at about 100 m below the surface and along
strike.Superimposed on the primary prograde metamorphic mineral assemblages of the host schist units is a distinct and
extensive primary alteration halo associated, and cogenetic, with the uranium mineralisation (see Figure 3). This alteration
extends for up to 1.5 km from the ore in a direction perpendicular to the host quartz-chlorite schist unit, because the
mineralisation is essentially stratabound. The outer zone of the alteration halo is most extensively developed in the semipelitic schists, and is manifested by the pseudomorphous replacement of biotite by chlorite, rutile and quartz, and feldspar
by sericite. Silicification has also occurred in fault planes and within the Kombolgie Formation sandstone beneath the
mineralisation, particularly adjacent to the reverse fault. Association of this outer halo alteration with the mineralisation is
demonstrated by the apparent symmetrical distribution of this alteration about the orebody. In the inner alteration zone, less
than 50 m from ore; the metamorphic rock fabric is disrupted, and quartz is replaced by pervasive chlorite and phengitic
mica, and garnet by chlorite. Uranium mineralisation is only present where this alteration has taken place.The primary ore
consists of uraninite veins and veinlets (1-10 mm thick) that cross-cut the S 2 foliation of the brecciated and hydrothermally
altered quartz-chlorite schist host. Groups of uraninite veinlets are intimately intergrown with chlorite, which forms the matrix
to the host breccias. Small (10-100 mm) euhedral and subhedral uraninite grains are finely disseminated in the chloritic
alteration adjacent to veins, but these grains may coalesce to form clusters, strings and massive uraninite. Coarse colloform
and botryoidal uraninite masses and uraninite spherules with internal lacework textures have also been noted, but the bulk
of the ore appears to be of the disseminated type, with thin (< 0.5 mm) discontinuous wisps and streaks of uraninite, and
continuous strings both parallel and discordant to the foliation (S 2), and parallel to phyllosilicate (001) cleavage
planes.Associated with the ore are minor volumes (up to 5%) of sulphides, which include galena and lesser chalcopyrite,
bornite and pyrite, with rare grains of native gold, clausthalite (PbSe), gersdorffite-cobaltite (NiAsS-CoAsS) and mackinawite

(Fe, Ni)1.1S. Galena is the most abundant, commonly occurring as cubes (5-10 mm wide) disseminated in uraninite or
gangue, and as stringers and veinlets particularly filling thin fractures within uraninite. Galena may also overgrow
clausthalite, and replace pyrite and chalcopyrite. Chlorite, predominantly magnesium chlorite, is the principal gangue, and its
intimate association with the uraninite indicates that the two minerals formed together.Oxidation and alteration of uraninite
within the primary ore zone has produced a variety of secondary uranium minerals, principally uranyl silicates. 7 Uraninite
veins, even veins over 1 cm wide, have been completely altered in situ. Within the primary ore zone this in situ replacement
of uraninite is most pronounced immediately above the reverse fault breccia, and this alteration and oxidation diminish
upwards stratigraphically. It is accompanied by hematite staining of the schists, the more intense hematite alteration in and
near the reverse fault breccia being due to hematite replacement of chlorite. The secondary mineralisation of the dispersion
fan in the weathered schist above the No.1 orebody is characterised by uranyl phosphates found exclusively in the tail of
the fan. Away from the tail uranium is dispersed in the weathered schists and adsorbed onto clays and iron oxides.The age
of the uranium mineralisation is problematical. The mineralisation, however, must post-date both the Kombolgie Formation
sandstone and the Koongarra reverse fault, since it occupies the breccia zones generated by the post Kombolgie reverse
faulting. The pattern of alteration which is intimately associated with the ore also crosses the reverse fault into the
Kombolgie sandstone beneath the ore zone, so this again implies that the ore was formed after the reverse fault and
therefore is younger than both the Kombolgie sandstone and the reverse fault. Because of these geological constraints,
Page et al.8 suggested the mineralisation was younger than 1600-1688 Ma because of their determination of the timing of
the Kombolgie Formation deposition to that period. Sm-Nd isotopic data obtained on Koongarra uraninites9,10 appears to
narrow down the timing of mineralisation to 1550-1650 Ma. It is unclear as to when deep groundwater circulation began to
cause oxidation and alteration of the primary uraninite ore at depth, but Airey et al.11 suggest that the weathering of the
primary ore to produce the secondary dispersion fan in the weathered schists above the No.1 orebody seems to have begun
only in the last 1- 3Ma.
Evidence Of An Open System
There are five main lines of independent evidence that the mineral-rock systems at Koongarra have been open to diffusion
and migration of U, Th and daughter isotopes including Pb. Such behaviour of these isotopes has crucial implications to all
attempts to date the Koongarra uranium ore using the U- Th-Pb isotopic systems.
(1) Ore Textures
Mineralogical and textural studies of the ore under both optical and scanning electron microscopes 12,13 indicate that there
have been as many as three remobilisations of the uranium during the history of the ore. Pb has likewise been mobile. That
is, both the primary U and Pb minerals, uraninite and galena respectively, have been dissolved and
redeposited/recrystallised, often some distance away from their original locations. This is shown diagrammatically in Figure
4 as several generations of uraninite and galena.

Figure 4. Paragenesis diagram showing the stages of formation and development of the minerals comprising the Koongarra
uranium deposit.
Figures 5-10 illustrate examples of the ore textures under the microscopes, the accompanying descriptions indicating how
the textures have been interpreted.

Figure 5. Remobilisation and redeposition of uraninite (white mineral). Photomicrograph shows uraninite veins (left and
right) partially destroyed by dissolution of uranium which has been redeposited as scattered veinlets and shapeless masses
of a new generation of uraninite (middle). (Magnification 10X).
Figure 6. Uraninite (light grey) has been dissolved and redeposited as thin veinlets and shapeless masses within a chlorite
(dark grey) matrix which is also replacing the main uraninite grain. (Magnification 120X).
Figure 7. Two generations of uraninite grains (lighter grey), and more oxidised supergene veins and patches (darker grey).
The small scattered white grains are galena. (Magnification 200X).
Figure 8. Two generations of uraninite grains (white, left of photomicrograph) and later thin supergene encrustations (mid
grey) around quartz grains (dark grey). The very bright mineral (right) is galena which has similarly dissolved and
redeposited. (Magnification 200X).
Figure
9. Remobilised
uraninite
(light
grey)
deposited
as
scattered
grains with a chlorite (dark
grey) matrix. A remobilised
galena vein (white-grey)
cuts across the uraninitechlorite
association.
(Magnification 50X).
Figure 10. An enlarged view
of uraninite (dark grey) subgrains within a larger vein.
Galena (light grey) veinlets
which both cross-cut and separate the uraninite sub-grains. The Pb in the galena is supposed to have migrated from the
uraninite where it was supposedly produced by radioactive decay. (Magnification 50X).
PS 17860/1

PS 17863/4

UO2

89.17

89.43

89.65

89.86

90.70

91.14

91.27

91.29

92.20

89.77

88.91

PbO

7.67

7.22

6.67

6.14

5.93

5.31

4.92

4.57

5.70

5.65

4.66

CaO

1.64

1.77

1.73

1.82

1.83

1.79

1.80

2.13

0.38

0.38

0.27

SiO2

0.39

0.42

0.43

0.46

0.53

0.57

0.56

0.50

0.24

1.00

2.34

SFe(FeO) 0.45

0.44

0.46

0.49

0.44

0.46

0.45

0.46

l.d.

0.11

0.46

MnO

MgO

l.d.

0.11

l.d.

l.d.

0.11

0.11

l.d.

0.12

0.39

0.94

1.86

P2 O5

0.21

0.21

0.19

0.16

0.23

0.18

0.23

0.30

0.13

0.17

0.13

Total

99.53

99.60

99.13

98.93

99.77

99.56

99.23

99.37

99.04

98.02

98.91

PS 17862/3
1

10

UO2

85.58

86.35

86.45

86.96

87.26

88.04

88.48

89.63

89.81

86.64

PbO

11.29

10.69

10.25

9.86

9.24

8.48

7.93

6.73

6.27

6.79

CaO

1.68

1.51

1.56

1.58

1.64

1.74

1.86

1.83

2.09

1.81

SiO2

0.50

0.41

0.46

0.47

0.45

0.46

0.53

0.60

0.63

0.78

SFe(FeO) 0.56

0.48

0.52

0.49

0.50

0.46

0.45

0.47

0.58

2.09

MnO

0.38

0.35

0.38

0.36

0.36

0.40

0.36

0.30

0.35

0.29

MgO

0.24

0.17

0.13

0.13

0.12

0.10

0.15

0.15

l.d.

0.18

P2 O5

0.16

0.14

0.17

0.13

0.14

0.17

0.12

0.17

0.19

1.14

Total

100.39

100.10

99.92

99.98

99.71

99.85

99.80

99.88

99.92

99.72

PS 17865/6
1

10

11

UO2

85.40

85.97

86.47

86.46

87.07

87.79

88.53

89.14

89.30

90.24

90.52

PbO

12.22

11.21

10.73

10.14

9.43

8.79

8.31

7.83

7.20

6.24

5.93

CaO

1.17

1.45

1.33

1.90

1.79

1.79

1.81

1.99

2.02

2.01

1.95

SiO2

0.33

0.36

0.36

0.49

0.51

0.47

0.52

0.49

0.43

0.58

0.48

SFe(FeO) 0.37

0.39

0.36

0.48

0.53

0.49

0.51

0.47

0.56

0.47

0.45

MnO

0.27

0.31

0.31

0.34

0.37

0.32

0.30

0.35

0.34

0.38

0.35

MgO

0.34

0.26

0.28

0.23

0.16

0.18

0.18

0.13

0.28

0.13

0.18

P2 O5

0.13

0.12

0.15

0.15

0.16

0.14

0.15

0.14

0.16

l.d.

0.16

Total

100.23

100.07

99.63

100.19

99.89

99.97

100.31

100.54

100.29

100.05

100.02

PS 17867/8

PS 17868/9

UO2

84.81

85.13

86.24

89.03

89.54

85.12

86.77

81.34

82.41

PbO

10.49

9.11

8.30

5.19

5.14

8.34

9.36

11.46

10.29

CaO

1.37

1.89

1.86

2.70

3.15

4.68

2.17

3.77

4.06

SiO2

2.38

1.35

1.54

1.20

0.85

0.83

0.70

1.20

0.99

SFe(FeO) 0.33

0.44

0.34

0.43

0.52

l.d.

0.53

l.d.

ll.d.

MnO

MgO

0.54

0.17

0.20

0.10

l.d.

0.19

0.11

0.12

0.16

P2 O5

l.d.

l.d.

0.14

0.14

0.11

0.56

l.d.

0.43

0.50

Total

99.92

98.09

98.62

98.79

99.31

99.72

99.64

98.32

98.41

[_ denotes not measured; l.d. denotes less than detection limits]


Table 1. Analyses of some representative Koongarra uraninites.
(2) Mineral Chemistry
Uraninite compositions in the ore are never uniform. Electron microprobe analyses of uraninite grains and veins, 13 that is,
micro-analyses of volumes of uraninite between 5 and 10 mm in diameter (see Table 1), reveal that uraninite compositions,
particularly U, Pb and Ca contents, vary not only from grain to grain within anyone sample regardless of which generation of
uraninite it is, but even at the microscopic level within uraninite grains themselves. Figure 11 illustrates how Pb and Ca have
both substituted for U in the UO2 cubic lattice in varying amounts across the uraninite veins and grains.

Figure 11. Compositional traverse across a uraninite grain similar to those in Figure 10.
Uranium - Lead Oxides
Curite

2PbO.5UO3.4H2O

Fourmarierite

PbO.4UO3.4H2O

Vandendriesscheite

PbO.7UO3.12H2O

Uranyl Silicates
Kasolite

Pb(UO2)SiO4.H2O

Sklodowskite

Mg(UO2)2Si2O7.6H2O

Uranophane

Ca(UO2)2Si2O7.6H2O

Uranyl Phosphates
Saleeite

Mg(UO2)2(PO4)2.8-10H2O

Sabugalite

HAl(UO2)4(PO4)4.16H2O

Metatorbernite

Cu(UO2)2(PO4)4.8H2O

Torbernite

Cu(UO2)2(PO4)2.8-12H2O

Renardite

Pb(UO2)4(PO4)2(OH)4.7H2O

Dewindtite

Pb(UO2)2(PO4)2.3H2O

Uranyl Sulphate
Cu(UO2)4(SO4)2(OH)2.6H2O

Johannite

Uranyl Vanadates
Carnotite - Tyuamunite

K2(UO2)2(VO4)2.3H2O-Ca(UO2)2(VO4)2.5-8H2O

Table 2. The secondary uranium minerals at Koongarra.


(3) Supergene Alteration
As has already been briefly noted, supergene alteration (principally oxidation) of uraninite has not only occurred where the
zone of surficial weathering has intersected the top of the No.1 orebody, but at depth within the primary ore. Uraninite grains
and veins have been replaced by colourful secondary uranium minerals (see Table 2), their occurrence and compositions
depending on the chemistries of the immediate rock/mineral environments and the circulating ground waters (see Figures 3
and 12). The net result has been the complete destruction of the uraninite in what was the top of the No.1 orebody, with its
replacement (sometimes in situ) by uranyl silicate or uranyl phosphate minerals (usually the latter), and the dispersion of the
rest of the U over distances of up to 50 m or more down-slope by ground waters in the weathered zone. Additionally, at the
same time there has been yet another remobilisation of both U and Pb in the primary ore zones, with in situ replacement of
uraninite (see Figures 13-15) and deposition of supergene uraninite (see Figure 16) and the uranyl silicate minerals
sklodowskite and uranophane (see Figures 17 and 18) from the U in solution from circulating ground waters (see Figure 3
again).7 Electron microprobe analyses (see Table 3) show that the U and Pb contents have decreased as uraninites were
altered to uranyl silicates, while the iron and manganese oxides lining fractures in the host rocks have absorbed the U and
Pb that had been dissolved during the oxidation of the uraninites and migrated in the circulating ground waters (see Table
4).
Figure 12. Schematic diagram showing the paths of secondary uranium mineral from uraninite in the Koongarra uranium
deposit.
Figure 13. Kasolite (white) and uranophane (grey) replacing a former uraninite vein. Note that the former vein shape, even
the sub-grains, have essentially been preserved.
(SEM magnification 210X; scale bar microns.)
Figure 14. Globular uraninite mass (black shape
just to the left of center) being altered marginally
to sklowdowskite (grey concentric sheath).
(Magnification 2X; scale bar 3 mm.)
Figure 15. Kasolite (light grey) and sklodowskite
(dark grey) replacing a former uraninite vein.
(SEM magnification 210X.)
Figure
16. Supergene
colloform
banded
uraninite (grey) deposited in what was originally a void. The banding is produced by a time sequence of uraninite deposition.
(SEM magnification 840X.)
Figure 17. A sklodowskite (white) vein composed of radiating aggregates of
needle-shaped crystals.
(SEM magnification 220X;
scale bar 50 microns.)
Figure
18. Uranophane
(white) veinlets deposited
between quartz (grey)
grain boundaries. (SEM
magnification 220X; scale
bar 50 microns.)

PS 17867/8: Uraninite Uranophane-Sklodowskite


1

UO2

84.81

85.13

86.24

76.74

69.58

66.45

PbO

10.49

9.11

8.30

8.99

1.05

0.15

CaO

1.37

1.89

1.86

2.89

4.89

3.86

SiO2

2.38

1.35

1.54

5.53

12.06

14.83

SFe(FeO) 0.33

0.44

0.34

0.29

0.70

l.d.

MgO

0.54

0.17

0.20

0.75

1.16

4.76

Al2O3

0.11

l.d.

l.d.

0.75

l.d.

0.31

P2 O5

l.d.

l.d.

0.14

0.36

0.35

0.34

V2 O3

l.d.

l.d.

l.d.

0.24

0.31

l.d.

Total

100.03

98.09

98.62

96.54

90.10

90.70

12.00
CAS 195: Uraninite Uranophane-Sklodowskite
1

UO2

82.18

85.49

86.22

88.27

90.53

63.74

68.76

66.50

66.44

PbO

11.55

9.34

7.93

6.39

4.65

9.83

4.48

3.55

1.60

CaO

3.08

2.80

3.15

3.13

3.06

2.34

2.98

2.77

2.86

SiO2

1.48

1.66

1.64

1.50

1.14

11.58

9.95

12.30

SFe(FeO) 0.80

0.40

0.88

0.39

0.41

0.87

0.20

0.23

l.d.

MgO

l.d.

l.d.

l.d.

l.d.

l.d.

0.39

0.19

0.20

1.13

Al2O3

P2 O5

l.d.

0.13

l.d.

l.d.

l.d.

2.38

2.15

2.86

2.11

V2 O3

Total

99.09

99.82

99.82

99.68

99.79

91.13

88.71

88.11

86.44

[- denotes not measured; l.d. denotes less than detection limits]


Table 3. Analyses of alteration sequences of uraninites to uranyl silicates at Koongarra.
CAS 165

CAS 114/1

CAS 114/2

CAS 95/1

CAS 95/2

CAS 95/3

UO2

2.81

1.63

1.05

0.36

2.83

1.91

PbO

12.42

4.41

0.30

5.03

8.16

3.34

CaO

0.20

0.09

l.d.

0.04

0.15

0.12

SiO2

2.49

3.11

6.28

2.87

2.54

3.20

SFe(FeO) 5.50

8.71

81.46

0.47

11.09

58.16

MnO2

77.48

80.35

1.96

88.52

73.53

27.70

MgO

0.12

0.37

2.09

0.29

0.52

0.22

Al2O3

0.15

1.23

2.70

0.82

1.75

P2 O5

0.33

l.d.

l.d.

l.d.

V2 O3

l.d.

0.31

0.65

0.26

Total

101.50

99.90

93.14

100.59

100.29

96.66

[- denotes not measured; l.d. denotes less than detection limits]


Table 4. Analyses of iron and manganese oxides in fractures in the Koongarra primary ore.
(4) Uranium/Daughter Disequilibrium
There are two methods of measuring the grade of a uranium ore sample:by assaying for U directly using standard chemical or related techniques, and
by measuring the radioactivity given off by the ore sample, the quantity of such radioactivity being directly related, and
proportional, to the U content.
However, because the radioactivity measured is actually the gamma radiation given off by the daughter element bismuth214 (214Bi) far down the 238U decay chain, any addition or removal of daughter elements between 238U and214Bi will result in a

discrepancy between the above two measurements of the U content of the ore sample. To assess this possibility the two
measurements are compared:Three possibilities arise:Ratio = 1. The ore sample is said to be in equilibrium since the
two measurements agree, implying that the U and its daughter
elements are in equilibrium; neither have apparently migrated.
Ratio > 1. The ore sample is said to be in disequilibrium, and since the U content is greater than the daughter element
content either U has been added to the sample or daughter elements removed.
Ratio < 1. Again the ore sample is aid to be in disequilibrium, but now the U content is less than the daughter element
content implying either U removal or daughter element addition to the sample.
No.

Group Description

No. of Samples

Average U3O8(%)

Average Ratio

sa

Weathered zone

13

0.275

0.914

0.160

Host wall rocks

19

0.025

0.792

0.151

Massive ore

11

8.074

0.959

0.069

Intermediate
orebodies

0.171

0.971

0.132

1.608

0.925

0.102

Mean =

0.884

0.127

No.
Orebody

No.
Orebody
5

between

and

2
Massive ore

Total number of samples


a

No.

54

Standard deviations of average ratio


Table 5. Summary of disequilibrium patterns in the Koongarra orebodies.
Measurements on ore samples from Koongarra indicate that the ore is in overall disequilibrium (Table 5 and Figure
19).14 High resolution gamma-ray spectroscopy was then used to determine which daughter elements of 238U have been
mobilised.15 These investigations showed that even
though the high grade uraninite (massive) ore is near
equilibrium, radium-226 (226Ra) and radon-222
(222Rn), and the immediate host rocks being relatively
enriched in U, having been precipitated from the
circulating groundwaters that had dissolved it from
the orebody. Figure 20 schematically illustrates these
movements of isotopes caused by the present day
circulation of groundwaters.
Figure 19. Frequency histogram of disequilibrium
ratios measured on Koongarra ore and host rock
samples.
Figure 20. Uranium (U) and (Ra) migration and
precipitation
(ppt)
caused
by
present-day
groundwater circulation and chemistry.
(5) Groundwater and Soil Geochemistry
Because of the tropical, monsoonal climate, the
ground waters in the Koongarra area are fast moving, annually recharged and low in salinity, the water table rising and
falling by as much as 10 m between the wet and the dry
seasons. However, U is dissolved by the ground waters from
the mineralised aquifer rocks, the level of dissolved U
depending on the prevailing pH, Eh, salinity and degree of
adsorption. A survey of the chemistry of the ground waters in
open drill holes in and near the Koongarra orebodies revealed
that a hydrogeochemical halo exists in and around the ore
zones reflecting the alteration chemistry of the host rocks and
ore, with U levels up to 4100 .16 Such measurements
confirm the other observations already cited that indicate U is
being dissolved from the ore minerals by present day
circulating ground waters, dispersed and partly redeposited.
Furthermore, the ground waters are also dispersing U- Th
decay products such as helium (He) from the ore zone, with
measured levels up to 14.2 ml/l.17
It is hardly surprising, therefore, that the soils overlying the ore
zones and the immediate areas of host rocks carry anomalous
U concentrations compared to background levels. 18 That the
ground waters have been responsible for dispersing U ( and
Pb) into the surrounding soils is also clearly demonstrated by
analyses down through the soil profile. Furthermore, Dickson et
al.19,20 found the Pb isotopic signature of the U ore in the soils

above the No.2 orebody, which is concealed by about 40 m of barren overburden, and in the soils to the south of the No.1
orebody within the hydrogeochemical halo.
Concentration (Wt%)

Atomic Ratios

Ages

Lead Isotope Ratios

Sample
No.

%U

%Pb

%Th

J804/1

62.38

8.07

0.30

0.142

1.312

0.0673

861

862

864

21330

1450

7.10

J804/b

38.21

4.45

0.28

0.126

1.264

0.0727

774

841

1025

9875

731.9

34.84

J801

55.07

3.64

0.34

0.071

0.810

0.0826

447

610

1282

16870

1408

54.20

J807

44.08

5.35

0.33

0.130

1.259

0.0703

796

838

954

12920

921.9

35.49

J809

52.61

5.45

0.39

0.114

1.061

0.0679

699

744

882

105800

7200

62.64

16.11

15.61

36.72

t206 m.y. t207 m.y.

Common lead correction


Mt Isa lead

Table 6. U-Th-Pb concentrations and isotopic compositions of Koongarra uraninites.


Table 7. Isotopic compositions of Koongarra galenas.
Dating of the Primary Ore
Sample No.
Hills and Richards21,22 isotipically analysed individual grains of uraninite and
galena that had been hand-picked from drill core (see Table 6 and 7). Only one of
J801
10290 1016
55.81
the five uraninite samples gave a near-concordant age of 862 Ma, that is, the
sample plotted almost on the standard concordia curve, and Hills and
J803
41240 3258
143.9
Richards22 interpreted this as recording fresh formation of Pb-free uraninite at 870
Ma (see Figure 21). The other four uraninite samples all lay well below concordia
J804
11530 883
8.539
and did not conform to any regular linear array. Hills and Richards were left with
two possible interpretations. On the one hand, preferential loss of the
J809
10540 1261
47.41
J820

4824

709.2

35.15

J821

3399

461.0

43.24

intermediate daughter products of 238U (that is,


escape of radon, a gas) would cause vertical
displacement of points below an episodic-loss line,
but this would only produce a significant Pb isotopic
effect if the loss had persisted for a very long
proportion of the life of the uraninite (which is
incidentally not only feasible but likely). Alternatively,
they suggested that contamination by small amounts
of an older (pre-900 Ma) Pb could cause such a
pattern as on their concordia plot, to which they
added mixing lines that they postulated arose from
the restoration to each uraninite sample of the
galena which separated from it (see Figure 21
again).
Figure
21. Conventional 206Pb/238U
concordia
diagram of uraninites from Koongarra. The insert
shows the hypothetical directional shift in uraninite
data points supposedly explained by contamination from associated galena.
This of course assumes that the Pb in the galenas was also derived predominantly from U decay. They plotted their Pb
ratios in all their uraninite samples on a standard 207Pb/206Pb diagram, and contended that the pattern of data points did not
conform to a simple age interpretation (see Figure 22). Instead, they contended that the scatter of points could be contained
between two lines radiating from the diagrams origin, lines that essentially represented isochrons for uraninites and galenas
from the Ranger and Nabarlek uranium deposits, similar orebodies in the same geological region. From the positions of the
Koongarra uraninites and galenas on these diagrams they claimed that the galenas contained left-over radiogenic Pb from
earlier uraninites as old as 1700-1800
Ma (the age of the Ranger uranium
mineralisation), these earlier uraninites
being obliterated by the U having
remobilised at 870 Ma, the age of the
lone Pb-free uraninite sample.
Figure
22. Conventional 207Pb/204Pb
vs. 206Pb/204Pb plots of galenas and
uraninites from Koongarra. Limiting
fields
of
anomalous-lead
lines
corresponding to ages of 1800 Ma
and 860 Ma.
In a separate study Carr and
Dean23 isotopically
analysed
unweathered whole- rock samples from
the Koongarra primary ore zone (see
Table 8). These were samples of drill
core that had been crushed. Their

isotopic data on four samples were plotted on a U-Pb isochron diagram and indicated a non-systematic relationship between
the 238U parent and the 206Pb daughter. In other words, the quantities of 206Pb could not simply be accounted for by
radioactive decay of 238U, implying open system behaviour. They also plotted their four results on a standard 207Pb/206Pb
isochron diagram (see Figure 23) and found that these samples fell on a poorly defined linear array whose apparent age
they did not quantify.

Figure 23. Conventional 207Pb/204Pb vs. 206Pb/204Pb plot of the weathered and unweathered whole-rock samples from
Koongarra. The weathered and unweathered samples fall on separate isochrons.
Sample

Pb (ppm)

U (ppm)

80

590.0

Primary Ore
1

0.0233

0.0752

2438.350

183.370

56.708

0.0682

0.0908

1162.990

105.594

79.351

0.0110

0.0692

6845.720

473.718

75.415

112

154.0

0.0346

0.0649

5719.990

371.474

198.191

19

17.0

168.0

Weathered Zone Ore


5

0.1785

0.1192

387.664

46.210

69.205

413.0

0.3804

0.2028

124.773

25.310

47.465

861.0

0.5029

0.2790

72.814

20.315

36.616

50

0.9277

0.4118

44.155

18.184

40.964

10

0.1608

0.1403

248.526

34.859

39.963

30

10

0.1650

0.1420

241.053

34.225

39.772

30

11

1.0477

0.3534

55.190

19.502

57.822

12

0.1213

0.1252

363.622

45.537

44.119

58

13

0.1233

0.1250

357.688

44.709

44.106

10

Table 8. Results of Pb isotopic, U concentration and Pb concentration analyses for Koongarra whole-rock samples.
Dating of Weathered Rocks and Soils
Carr and Dean23 also isotopically analysed a further nine whole-rock
samples from the weathered schist zone at Koongarra (see Table 8).
Some of these samples were again crushed drill core, but the
majority were crushed percussion drill chips. When their isotopic
data were plotted on a U-Pb isochron diagram, six of the nine
samples plotted close to the reference 1000 Ma isochron, while the
other three were widely scattered (see Figure 24). However, on
the207Pb/206Pb diagram all nine weathered rock samples plotted on a
linear array which gave an apparent isochron age of 1270
(see Figures 23 and 25).

50Ma

Figure 24. A U-Pb (238U/204Pb vs. 206Pb/204Pb) isochron diagram with


the weathered whole-rock samples plotted on it. Most fall on the
1000 Ma reference isochron while the 10 Ma reference iisochron is
also drawn in as a guide to the two outliers.

Figure 25. A conventional 207Pb/204Pb vs. 206Pb/204Pb isochron


diagram showing all the weathered whole-rock samples plotted as a
linear array which gives an apparent isochron age of 1270
50Ma. (This diagram is an expansion of the lower left hand corner of
Figure 23.)
In unrelated investigations, Dickson et al.19,20 collected soil samples
from above the mineralisation at Koongarra and from surrounding
areas, and these were analysed for Pb isotopes to see if there was
any Pb isotopic dispersion halo around the mineralisation sufficiently
large enough to warrant the use of Pb isotopic analyses of soils as
an exploration technique to find new uranium orebodies. The
technique did in fact work, Pb isotopic traces of the deeply buried
No.2 orebody mineralisation being found in the soils above, as
mentioned earlier. This mineralisation, 40 m below the surface, is
blind to other detection techniques.Dickson et al.20 found that all 113
soil samples from their two studies were highly correlated (r =
0.99986) on a standard 207Pb/206Pb diagram, yielding an apparent
(false) isochron representing an age of 1445
20 Ma for the
samples (see Figure 26). However, most of the soil samples consisted of detritus eroded from the Middle Proterozoic
Kombolgie sandstone, so because the samples from near the mineralisation gave a radiogenic Pb signature Dickson et al.
interpreted the false isochron as being due to mixing of radiogenic Pb from the uranium mineralisation with the common
Pb from the sandstone.
Figure 26. Plot of 207Pb/206Pb vs. 206Pb/204Pb for all 113 soil samples from the Koongarra area analysed by Dickson et al.,
indicating the high correlation of r = 0.99986 between the two variables with a fitted regression line yielding an
apparent isochron age of 1445
20 Ma. The insert shows the distribution of samples about a threshold dividing
radiogenic Pb and country rock Pb along this proposed mixing line.
Discussion
Primary Ore Samples
Snelling24 has already highlighted a telling omission by Hills
and Richards.22 Having included all the Pb isotopic ratios they
had obtained on their five uraninite samples, they tabulated
also the derived ages, except for those obtainable from 208Pb
(see Table 6 again). Since their data table lists the necessary
ingredients for 208Pb age calculations - %Th,208Pb proportion,
and 208Pb/207Pb and 208Pb/204Pb ratios - their omission of
the 208Pb ages is both conspicuous and significant. These
Th-derived dates should normally be regarded as the most
reliable, since Th is less mobile in geochemical environments
and therefore open system behaviour is less likely than for U.
The 204Pb content of the uraninite is regarded as common or
original Pb since it is not derived from any parent element via
radioactive decay. Because this so-called common Pb is
also believed to carry a significant quantity of the 206Pb, 207Pb
and 208Pb isotopes, a common Pb correction has to be
applied to the raw data before calculation of the U- Th-Pb
ages. This, of course, is an admission that not all the
quantities of these Pb isotopes are derived by radioactive
decay, some being with the U and Th in the beginning. The
standard used to correct the data in Table 6 was the Mt Isa Pb
standard with an isotopic composition:1.44% 204Pb

23.20% 206Pb

22.48% 207Pb

52.88% 208Pb

It should be noted in passing also that the choice of this standard is based on one of several theories of element
nucleogenesis and Pb isotopic evolution,25,26 making the whole age calculation procedure rather subjective, based on
further assumptions.When this common Pb correction is applied to the data in Table 6, 27 most of the 208Pb has resulted
from common Pb contamination. In fact, in samples J804/1, J804/b and J807 all the 208Pb is due to contamination and
none to 232Th decay, thus resulting in 208Pb ages of 0 Ma (within the experimental/analytical errors) for these samples. The
remaining two samples yield 208Pb ages27 of 275 Ma (J801) and 61 Ma (J809), both considerably less than all other Pb
ages. Since they are as valid as any of the other resultant ages calculated, these 232Th/208Pb ages should have been at
least reported (one suspects they were left out of the tabulated results because of the uncomfortable implications). After all,
the 232Th/208Pb age of 0 Ma is the only Pb isotopic date from that study supported directly by a majority of samples (three
out of the five), and Th-derived , dates should be reliable as the 232Th decay chain is a standard isotopic clock, but a 0 Ma
age makes little more sense than their 870 Ma age from the U- Pb data. In any case, Hills and Richards age of 17001800 Ma for the first generation of U mineralisation at Koongarra neither fits the geological criteria for an expected 15501600 Ma age, nor does their 870 Ma date correlate with any geological event capable of remobilising U and Pb to
produce the presumed second generation of U mineralisation.Using the procedure of Ludwig, 28 standard 207Pb/206Pb
diagrams were prepared for the uraninite, galena and whole-rock data sets, and combinations thereof, to check the
regression statistics and possible derived isochrons using the standard York 29 method. In each case the mean square of
weighted deviates (MSWD), which tests the goodness of fit of data to a line, is large to extremely large, which reflects in
the derived isochron ages of 841140 Ma (uraninites), 1008

420 Ma (galenas), 668

330 Ma (whole-rocks), 818

150

Ma (uraninites plus galenas) and 863


130 Ma (all three data sets combined), all ages being within the 95% confidence
limits (see Figures 27-31). It is perhaps fortuitously significant that the combination of all three data sets yields an isochron

age of 863
130 Ma, almost identical to Hills and Richards near-concordant age of 862 Ma, although this was using a
line-fitting routine of Ludwig28 that assigns equal weights and zero error-correlations to each data point to avoid the mistake
of weighting the points according to analytical errors when it is clear that some other cause of scatter is involved, which is
clearly the case here. The normal York 29 algorithm assumes that the only cause for scatter from a straight line are the
assigned errors, and for the combined data set here the amount of scatter calculated thereby yields an astronomical MSWD
of 669000 and a bad line of fit that yields an isochron age of
1632
410 Ma (see Figure 32). This result may make more geological sense, but the regression statistics are such that
derivation of any age information from these data is totally unjustified, even though it can be rightfully argued that these
samples form a cogenetic set (they are all samples of U ore or its components from the same primary ore zone at
Koongarra).

Figure 27. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with all Koongarra uraninites plotted on it using Ludwigs
ISOPLOT program and defining an apparent isochron with a model 2 age of 841

140 Ma.

Figure 28. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with all Koongarra galenas plotted on it using Ludwigs
ISOPLOT program and defining an apparent isochron with a model 2 age of 1008

420 Ma.

Figure 29. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with all the unweathered whole-rock samples from Koongarra
plotted on it using Ludwigs ISOPLOT program and defining an apparent isochron with a model 1 age of 668

330 Ma.

Figure 30. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with both Koongarra uraninites and galenas plotted on it using
Ludwigs ISOPLOT program and defining an apparent isochron with a model 2 age of 818

150 Ma.

Figure 31. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with all Koongarra uraninites, galenas and unweathered
whole-rock samples plotted on it using Ludwigs ISOPLOT program and defining an apparent isochron with a model 2 age
of 863

130 Ma.

Figure 32. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with all Koongarra uraninites, galenas and unweathered
whole-rock samples plotted on it using Ludwigs ISOPLOT program and defining an apparent isochron with a model 1 age
of 1632
410 Ma.
It is not uncommon to find that ages derived from standard 207Pb/206Pb plots are erroneous, even though the data fit welldefined linear arrays (isochrons). Ludwig et al.30 found that this was due to migration of both Pb and radioactive daughters
of 238U yielding a 207Pb/206Pb isochron giving superficially attractive results which would nonetheless be seriously
misleading because the derived age (in their example) was more than six times higher than the U-Pb isochron age.
Similarly, Cunningham et al.31 obtained 207Pb/206Pb isochron ages up to 50 times higher than those derived from more
reliable U-Pb isochrons for whole-rock U ore samples, even though the apparent slight degree of scatter is almost entirely
a misleading artifact. Likewise, at Jabiluka, an almost identical style of uranium deposit in the identical geological setting
only about 60 km due north of Koongarra, Gulson and Mizon 32 had considerable difficulty obtaining Pb-Pb and U-Pb
isochron ages for the U mineralisation due to 238U daughter leakage and diffusion out of the U minerals and ore into the
surrounding host rocks and constituent minerals, that therefore had gained excess radium (Ra) and 206Pb. Ironically, at
Koongarra the U-Pb isochron using Ludwig 28 on Hills and Richards uraninite data yields an age of 857
149 Ma (with an
MSWD of 13400, tolerably large compared to that obtained with the Pb-Pb isochron) (see Figure 33), almost identical to the
fortuitous Pb-Pb isochron age obtained using Ludwigs modified algorithm on the combined three data sets (863
Ma), as well as Hills and Richards single near-concordant 862 Ma age.

130

Figure 33. A conventional 206Pb/238U vs. 207Pb/235U concordia diagram with all the Koongarra uraninites plotted on it using
Ludwigs ISOPLOT program and defining an apparent isochron that intersects concordia at 857
149 Ma.
As has already been described, Snelling and Dickson 14 have demonstrated that there is significant uranium/daughter
disequilibrium in the primary ore and surrounding host rocks at Koongarra due to the redistribution of both U and its Ra
decay product, just as Gulson and Mizon found at Jabiluka. That Ra mobility at depth in the primary ore zone is currently
more significant than U migration has been confirmed by Dickson and Snelling, 15 which of course results ultimately in the
redistribution of 206Pb, the end-member of the whole 238U decay chain. Dickson et al.33 have demonstrated that Ra is
transported through the unweathered rocks in this area in the ground waters, while Davy et al.34 have determined the
emanation rate of radon (Rn) gas from the Koongarra No.1 orebody, an ever present hazard in uranium ore mining
operations. The Rn gas is known to migrate along fractures and rise through the ground over considerable distances to form
a halo in the air above, while Rn is also transported in ground waters. Thus it is to be expected that the pattern of oxidation
of uraninites and dispersion of U should reflect the present-day circulation of ground waters 7 and that present-day ground
waters should be carrying U and He.16,17 Such groundwater dispersion of U and mobility of Ra has, of course, resulted in U
and Pb dispersion into the surrounding soils,18 where the Pb isotopic signature of the U ore is clearly evident.19,20These
observations alone demonstrate the open system behaviour of the U- Th-Pb system that renders meaningless any age
information derived. However, both Hills12 and Snelling13 have recognised that U and Pb also have migrated several times
and on a considerable scale in the primary ore zone, with the latest redistribution having produced supergene uraninites,

often with colloform banding, found as fracture and cavity infillings (see Figure 16 again), and between quartz and gangue
grain boundaries. The unit cell dimensions of these uraninites, plus this textural evidence, supports the conclusion that these
uraninites have precipitated after dissolution of earlier formed uraninite and transportation in low-temperature ground
waters. With such wholesale repeated migrations of U also, all attempts at dating must be rendered useless, especially
when whole-rock samples, in which different generations of uraninites are lumped together, are used. Indeed, it must surely
be virtually impossible to be certain of the precise status and history of any particular piece of uraninite selected for dating.
Even though every conceivable precaution is taken when selecting grains for dating, how can we be sure that the U and
Pb isotopes and isotopic ratios measured represent the original, unaffected by the gross element movements for which
there is such abundant evidence? The uraninite grains or ore samples dated always contain radiogenic Pb both within
crystal lattices of minerals, and as microscopic inclusions or grains and veins of galena, but how can we be sure all the Pb
was generated by radioactive decay from U in situ? In any case, the uraninite grains and veins do not have uniform
compositions - either between or within grains - so that dating of sub-sections of any grain or vein would be expected to
yield widely divergent U-Pb and Pb-Pb ratios and therefore ages even within that single grain or vein. Thus it is logical to
conclude, as others have already,35-37that U- Th-Pb ratios may have little to do with the ages of many minerals, rocks and
ores.
Weathered Rocks and Soils
In contrast to the poor-fitting linear arrays produced from the Pb-Pb data of minerals and whole-rocks from the primary ore
zone, that all appear to give an apparent (false) isochron age grouped around 857-863 Ma, both Carr and Dean 23 and
Dickson et al.20 found that weathered schist whole-rock and soil samples produced good fitting linear arrays that would
normally represent isochrons that yield ages of 1270 Ma and 1445 Ma respectively (see Figures 25 and 26 again). The
weathered whole-rock samples all of course come from Koongarra itself, and consist of secondary ore samples from the
weathered schist zone, plus weathered schist samples that contain U dispersed down-slope by ground waters moving
through the weathered rock. Because these whole-rock samples come from a volume of rock through which U is known to
be migrating, leading to redistribution not only of U but of its decay products, it is therefore very surprising to find that these
whole-rock samples define a good enough linear array to yield an isochron. Even the observed scatter calculated using
Ludwig28 is much less than that associated with fitting an isochron to the 207Pb-206Pb data from the primary ore zone
samples, which is again surprising given U migration in the weathered zone, the data from which one would expect to show
considerable scatter and thus no age consensus. Furthermore, it is baffling as to why the isochron-derived age (1270
Ma) of the weathered secondary ore zone should be so much older than the isochron-derived age (857-863 Ma) of the
primary ore, which of course is ultimately the source through weathering and groundwater transport of the U, decay products
and the stable Pb isotopes that are in the secondary and dispersed ore. Perhaps the only explanation is that the isochron
represents the mixing of radiogenic Pb from the mineralisation with the common or background Pb in the surrounding
schists, which are even in a relative sense older than the U mineralisation.The idea of such an isochron being a mixing line
was suggested by Dickson et al.20 They were, however, dealing with the Pb isotopic data obtained from soil samples
collected from depths of only about 30-40 cm, the majority of which represented sandy soils consisting of detritus eroded
from the Kombolgie sandstone. For this mixing explanation to be feasible there should be some other evidence of
mobilisation of Pb in the area. Dickson et al. found that not only were there high 206Pb/204Pb ratios in three of their soil
samples from the near-surface (0-1 m) zone south of the No.1 orebody in the hydrogeochemical halo, but there was a lack
of any other U-series daughter products in the same samples. This near-surface zone is inundated for approximately six
months of the year as a result of the high monsoonal rainfall in this tropical area. Towards the end of the ensuing six-month
dry season the water table has been known to drop in some cases more than ten metres from its wet season high. This
means that the top of the weathered schist zone is regularly fluctuating between wet and dry conditions, so that any trace
elements such as Pb leached from the weathered ore and transported by ground water in the weathered schist zone would
also be dispersed vertically up into the thin surficial sand cover on top of the weathered schist - the sandy soils that were
sampled by Dickson et al.19,20 Snelling18 found that Pb was a significant pathfinder element for uranium ore in the Koongarra
environment, anomalous Pb being present in the surficial sand cover above the zone of weathered primary ore, and that
there was even hydrodynamic dispersal of Pb at a depth of 0.5-1.5 m. Dickson et al.19 found a similarity between the isotopic
ratios for Pb extracted from their soil samples by either a mild HCI-hydroxylamine (pH 1) or a strong 7M HCI- 7M
HNO3 leach, which indicates that Pb is loosely attached to sand grain surfaces in the samples rather than tightly bound in
silicate or resistate mineral lattices. This in turn suggests Pb is adsorbed from ground waters, meaning that radiogenic Pb is
being added to the common or background Pb in the sand by both vertical and lateral groundwater dispersion.However,
not all of Dickson et al.s soil samples came from the area immediate to the Koongarra orebodies, nor were they all samples
of Kombolgie sandstone detritus. That this mixing line explanation for the apparent isochron is clearly demonstrated for
these samples from the immediate Koongarra area is not in question, although it is somewhat surprising that these soil
samples should give an apparent isochron age (1445 Ma) somewhat older than that obtained from the weathered schist
samples beneath (1270 Ma). Indeed, the common or background Pb in the respective samples should reflect an older
apparent age in the schists compared to the sandstone, due to their relative ages based on the geological relationship
between them. (Remember, the schists are supposed to be the product of regional metamorphism at 1800-1870 Ma, while
the Kombolgie sandstone is regarded as having been deposited around 1600- 1680 Ma.) However, the apparent ages are
the other way around, the sandy soils from the Kombolgie sandstone detritus yielding an older apparent age (1445 Ma)
compared to that yielded by the weathered schists (1270 Ma). Perhaps this difference is a reflection of the extent of mixing
in each type of sample at their respective levels in the weathering profile. Nevertheless, what is astounding is that
Dickson et al.20 found that even though several of their soil samples consisted of weathered schist or basement granite
(containing accessory zircon) up to 17km from the known U mineralisation, they still plotted on the same apparent
isochron. Indeed, the fit is comparatively good (see Figure 34), as indicated by the MSWD of only 964 using Ludwig, 28 yet
much of this observed scatter can be attributed to two samples out of the 113, one of which was subsequently found to be
probably contaminated by cuttings from an adjacent drill hole.19 If that sample is removed from the regression analysis the
MSWD drops to 505, indicating that almost half of the observed scatter is due to that one data point alone. If the data point
that is the next worst for fitting to the apparent isochron is removed, then the MSWD drops by a further 315 to a mere 190.
Yet in both cases the apparent isochron or mixing line still has lying on or close to it the samples from up to 17 km away
from the known U mineralisation and the samples that are not Kombolgie sandstone detritus. The final isochron fitted to
the remaining 111 samples still yields an age of 1420

18 Ma (see Figure 34 again).

Figure 34. A conventional 207Pb/204Pb vs. 206Pb/204Pb


diagram with all Koongarra area soil samples plotted on it
using Ludwigs ISOPLOT program and defining an apparent
isochron with a model 1 age of 1428

33 Ma for all 113

samples and 1420


18 Ma for 111 samples (2 outliers
removed).
While Carr and Deans nine weathered schist whole- rock
samples are not strictly cogenetic with Dickson et al.s 113
soil samples, the two sample sets are obviously related
because the source of the radiogenic Pb in the majority of
the soil samples from the immediate Koongarra area is the
same as that in the weathered schists. Not surprisingly,
when the regression analysis was performed on Carr and
Deans nine weathered schist whole-rock samples using
Ludwig,28the MSWD for the observed scatter was 24100,
indicating a poor fit to an isochron which yielded an age
of 1287
120 Ma (see Figure 35). Yet when these nine
samples were added to the 113 soil samples the MSWD
dropped substantially to 1210, and not surprisingly the fitted isochron yielded an age of 1346
27 Ma, an isochron age
intermediate between those of the two data sets being combined (see Figure 36). However, when the two soil samples
responsible for the majority of the scatter in that data set were removed the MSWD dropped to 430 and yielded an isochron
age of 1336

17 Ma (see Figure 36 again).

Figure 35. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with the weathered whole-rock samples from Koongarra
plotted on it using Ludwigs ISOPLOT program and defining an apparent isochron with a model 1 age of 1287

120 Ma.

Figure 36. A conventional 207Pb/204Pb vs. 206Pb/204Pb diagram with all Koongarra area weathered whole-rock and soil
samples plotted on it using Ludwigs ISOPLOT program and defining an apparent isochron with a model 1 age of 1346
27 Ma for all 122 samples and 1336
17 Ma for 120 samples (2 outliers removed).
General Comments
As with all the other apparent isochron ages, these results from the weathered rocks and soils have no apparent geological
meaning, because there is no geological event to which these ages might correlate. Indeed, even in the evolutionary time-

frame the weathering of the Koongarra U mineralisation is extremely recent, and in any case, these ages derived from PbPb isochrons from the weathered rock and soil samples are much older than the supposedly more reliable U-Pb
isochron age of the Koongarra primary ore. But since that latter result has no apparent geological meaning, because it also
cannot be correlated with any known geological event, nothing then is certain at all from any of these U-Th-Pb isotopic
studies of the Koongarra ores, rocks and surrounding soils. Indeed, it is just as certain that the primary ore is 0 Ma, based
on three 232Th/208Pb single sample ages, as is the claim that one near-concordant result means that there was formation of
Pb-free uraninite at 870 Ma. After all, this postulated formation of Pb-free uraninite is supposed to have occurred in an
environment where there was Pb left over from an earlier 1700-1800 Ma original U mineralisation for which we no longer
have any evidence, textural or otherwise, apart from a rather tenuous interpretation of Pb isotopic evidence that has
otherwise shown itself to be devoid of any capability of providing any age information.
All these results raise serious fundamental questions about the claimed validity of the U-Th-Pb dating method. It may seem
reasonable to regard an apparent isochron as a mixing line within the restricted area close to the known source of
radiogenic Pb, which can be shown by independent evidence to be migrating into rocks and soils that contain common or
background Pb in the immediate environs. However, it strains all credulity to suggest that a false isochron through a data
set derived from samples representing a variety of rock types, of significantly different evolutionary ages, over an area of
up to 17 km lateral extent from the known radiogenic Pb source, can still represent mixing! One can only conclude that all
assumptions used to derive the estimates of common or background Pb, including models for the supposed evolution of
the stable Pb isotopes through earth history, from their presumed commencement on the protoearth with its claimed original
Pb isotope content some 4.6 billion or so years ago, cannot be valid. Equally, we cannot be sure what the U-Th-Pb systems
isotopic ratios really mean, because the basic assumptions that are foundational to the interpretation of these isotopic ratios
are fatally flawed. Not only has open system behaviour of these isotopes been demonstrated as the norm, but even where
there is an apparent isochron with an excellent goodness of fit the derived age is invariably geologically
meaningless.Thus creationists need not be hindered in building their Creation-Flood young-earth model for the geological
record by the many claims in the open geological literature that U-Th-Pb radiometric dating has proved the presumed
great antiquity of the earth, and the strata and fossils of the so-called geological column. Accordingly, all the apparent
isochron and other ages that have been referred to here have been quoted as millions of years (Ma) purely in order to
reveal the shortcomings of the U-Th-Pb dating method. Indeed, even the use of conventional geological era terms such as
Archaean and Lower Proterozoic has been for convenient reference to the rock units under discussion, there being no
absolute age significance attached to these terms here - only a relative position within the overall rock record. There is
clearly a real sequence of rock units that comprise the total geological record, from the so-called Archaean to the Recent,
the formation of which needs to be understood and coherently modelled within the framework of a recent Creation and
global Flood. Much progress towards this goal has been, and is being, made within the relatively small creationist geological
community. Thus the mounting evidence that the claimed absolute dating methods, such as U- Th-Pb radiometrics, are
unreliable at best, and in reality produce many results that are impressive but geologically meaningless, can only assist in
this quest.
Conclusions
The concerns raised by Zheng1 regarding U-Pb isochrons are warranted. At Koongarra a 207Pb/206Pb isochron produced
from 11 hand-picked uraninite and galena grains, plus four whole-rock samples, yields an age of 863 Ma, the same as a
near-concordant age of 862 Ma from one of the uraninite grains. Nine weathered whole-rock samples yield an isochron
age of 1270 Ma, while 113 soil samples produce an excellent isochron with an age of 1445 Ma. All of these ages are
geologically meaningless. While the apparent isochron produced by the soil samples may be identified as a mixing line,
produced by the mixing of radiogenic Pb with common or background Pb in the surrounding rocks and soils, even this
explanation strains credulity because the samples come from up to 17 km away from known U mineralisation, and a few of
the soil samples represent different rock types. Not only then has open system behaviour of these isotopes been
demonstrated, as confirmed by the independent evidence of ore textures, mineral chemistry, supergene alteration,
uranium/daughter disequilibrium, and groundwater and soil geochemistry, but apparent isochrons and their derived ages
are invariably geologically meaningless. Thus none of the assumptions used to interpret the U- Th-Pb isotopic system to
yield ages can be valid. If these assumptions were valid, then the 232Th/208Pb age of 0 Ma for three of the five uraninite
samples should be taken seriously. Creationists should therefore not be intimidated by claims that U-Th-Pb radiometric
dating has proved the presumed great antiquity of the earth, and the strata and fossils of the so-called geological column.
Determination of the Radioisotope Decay Constants and Half-Lives: Rubidium-87 ( 87Rb)
by Dr. Andrew A. Snelling on September 3, 2014
Abstract
In spite of numerous attempts over the last 60 years to determine the 87Rb half-life and decay constant, there is still no
consensus on the absolute values. Even the more accurate determinations of the last 30 years have resulted in
discrepancies. Determinations based on comparisons of ages of earth minerals and rocks give two different values of
the 87Rb half-life and decay constant to the values using determinations based on comparison of ages of meteorites and
lunar rocks. Yet the most recent direct counting and in-growth experimental determinations only agree with the most recent
determinations based on comparison of the Rb-Sr ages of earth minerals and rocks with their U-Pb gold standard ages.
This has resulted in suggested values of the 87Rb half-life and decay constant that are different from the two values of each
previously championed in the literature. A similar discrepancy in determinations of the 176Lu half-life and decay constant has
been recognized in the literature, as well as variations in the 238U/235U ratio so critical to the U-Pb gold standard against
which the Rb-Sr ages have been calibrated. Since the 87Rb decay rate has thus not been accurately determined, the Rb-Sr
dating method is certainly not absolute and therefore cannot be used to discredit the young-earth creationist timescale.
However, further study is warranted to explore any potential significance in this discrepancy in half-lives and decay
constants between earth materials, and meteorites and lunar rocks.
Keywords: radioisotope dating, decay constant, half-life, rubidium-87,87Rb, decay, direct counting, in-growth experiments,
geological comparisons, discrepancies, meteorites, lunar rocks, earth minerals and rocks, 176Lu, U-Pb gold
standard, 238U/235U.
Introduction
Radioisotope dating of rocks and meteorites is perhaps the most potent claimed proof for the old age of the earth and the
solar system. The absolute ages provided by the radioisotope dating methods provide an apparent aura of certainty to the
claimed millions and billions of years for formation of the earths rocks. Both the scientific community and the general public
around the world (except perhaps in the USA) thus remain convinced of the earths claimed great antiquity.The 19972005
RATE (Radioisotopes and the Age of The Earth) project successfully made progress in documenting some of the pitfalls in
the radioisotope dating methods, and especially in demonstrating that radioisotope decay rates may not have always been

constant at todays measured rates (Vardiman, Snelling, and Chaffin 2000, 2005). Yet much research effort remains to be
done to make further inroads into not only uncovering the flaws intrinsic to these long-age dating methods, but towards a
robust understanding of radioisotopes and their decay during the earths history within a creationist framework.One crucial
area the RATE project did not touch on was the issue of how reliable have been the determinations of the radioisotope
decay rates, which are so crucial for calibrating these dating clocks. Accurate radioisotope age determinations depend on
accurate determinations of the decay constants or half-lives of the respective parent isotopes. The reliability of the other two
assumptions these absolute dating methods rely on, that is, the starting conditions and no contamination of closed systems,
are unprovable. Yet these can be circumvented somewhat via the isochron technique, because it is independent of the
starting conditions and is sensitive to revealing any contamination. Data points that do not fit on the isochron are simply
ignored because their values are regarded as due to contamination. That this is common practice is illustrated with
numerous examples from the literature by Faure and Mensing (2005) and Dickin (2005). On the other hand, it could be
argued that this discarding of data points which do not fit the isochron is somewhat arbitrary and therefore is not good
science, because it is merely assumed their aberrant values are due to contamination rather than that being proven to be
so.The aim of this contribution is to begin to document the methodology behind and history of determining the decay
constants and half-lives of the parent radioisotopes used as the basis for the long-age dating methods. We need to explore
just how accurate these determinations are, whether there really is consensus on standard values for the half-lives and
decay constants, and just how independent and objective the standard values are from one another between the different
methods. We begin here with rubidium-87 (87Rb), which is the basis for the Rb-Sr dating method.
Rubidium-87 Decay
The natural radioactivity of rubidium (Rb) was demonstrated in 1906, but it took more than thirty years for 87Rb to be
identified as a naturally-occurring radioactive isotope (radioisotope) (Hahn, Strassman, and Walling 1937; Mattauch 1937).
The feasibility of dating Rb-bearing minerals by the -decay of 87Rb to 87Sr was first proposed by Hahn and Walling (1938),
but the first determination by this method did not follow until a few years later (Hahn et al. 1943). However, the Rb-Sr
method of radioisotope dating did not come into wide use until the 1950s, when mass spectrometers became available for
routine isotopic analyses of solids (Nier 1940). At the same time the technology was developed to measure the
concentrations of Rb and Sr by isotope dilution, combined with the separation of these elements by cation exchange
chromatography. Faure and Powell (1972), Faure and Mensing (2005), and Dickin (2005) provide these and more details
about the history of the development of the Rb-Sr dating method, and its theoretical basis and applicability.
Rubidium is a group-1 alkali metal. On the periodic table it is listed in the group IA column, which consists of Li, Na, K, Rb,
Cs, and Fr. The ionic radius of Rb+ is 1.48 , which is sufficiently similar to that of K+ (1.33) to allow it to substitute for K+ in
all K-bearing minerals. Rubidium is an element that does not form minerals in which it is a major constituent. Instead, Rb is a
dispersed element in trace amounts in other minerals. For example, it occurs in easily detectable amounts in common Kbearing minerals, such as the micas (muscovite, biotite, phlogopite and lepidolite), K-feldspars (orthoclase and microcline),
certain clay minerals, and the evaporite (precipitite) minerals sylvite and carnallite.Rubidium has two naturally-occurring
isotopes 85Rb and 87Rb, whose abundances are 72.17% and 27.83% respectively (Dickin 2005; Faure and Mensing 2005).
Rubidium-87 is radioactive and decays to stable 87Sr:87
Rb 87Sr + + + Q
A negative -particle () and an anti-neutrino () are both emitted in this process. The decay energy (Q) is shared as
kinetic energy by these two particles. Fortunately this -decay of 87Rb is a single-step process. In effect, it can be regarded
as the transformation of a neutron within the 87Rb nucleus into a proton and an electron, the latter then being expelled as a
negative -particle. The relative simplicity of this decay step therefore makes measuring the decay rate of 87Rb relatively
straight forward. There are two parameters by which the decay rate is measured and expressed, namely, the decay constant
() and the half-life (t). The latter is the time it takes for half of a given number of the parent radionuclide atoms to decay.
The two quantities can be almost used interchangeably, because they are related by the equation:t = ln 2 = 0.693
Determination Methods
Three approaches have so far been followed to determine the -decay constant and half-life of the long-lived
radioactive 87Rb.
Direct counting
In this technique, the beta () activity of 87Rb is counted in a source material, and divided by the total number of radioactive
atoms in the known quantity of Rb, based on Avogadros number and the isotopic abundance of 87Rb. Among the difficulties
of this approach are the self-shielding of finite-thickness solid samples, the low specific activities, imprecise knowledge of
the isotopic composition of the parent Rb, the detection of very low-energy decays, and problems with detector efficiencies
and geometry factors (Begemann et al. 2001).The low decay energy for this 87Rb -decay transformation (0.275MeV) has
always caused problems in the accurate determination of the 87Rb decay constant (Dickin 2005; Faure and Mensing 2005).
Because the decay energy is divided between the -particles and the anti-neutrinos, the -particles have a smooth
distribution of kinetic energy from the total energy down to zero. When attempting to determine the decay constant by direct
counting the low-energy -particles cause great problems because they may be absorbed by surrounding Rb atoms before
they ever reach the detector. For example, in a thick (>1m) solid Rb sample, attenuation is so severe that a false
frequency maximum is generated at about 10keV.One way this attenuation problem has been avoided is to use a photomultiplier with a liquid scintillation solution doped with Rb. The -particles will be absorbed by molecules of the scintillator
(emitting flashes of light) before they can be absorbed by other Rb atoms. The major problem with this method is that a lowenergy cut-off at about 10keV must be applied to avoid the high background noise associated with liquid scintillation (Dickin
2005). The consequent extrapolation of count-rate curves down to zero energy leads to a large uncertainty in the result.
Hence this method has given values for the 87Rb half-life from 47.01.0Byr (Flynn and Glendenin 1959) to 52.11.5Byr
(Brinkman, Aten, and Veenboer 1965).Another approach to direct counting is to make measurements with progressively
thinner solid Rb sources using a proportional counter. The results are then extrapolated to a theoretical source of zero
thickness to remove the effect of self-absorption. The proportional counter has a much lower noise level, so the energy cutoff can be set as low as 0.185keV. Rb films with thicknesses down to 1m were measured by Neumann and Huster (1974),
and extrapolated to zero thickness by Neumann and Huster (1976) to derive a 87Rb half-life of 48.80.8Byr (equivalent to a
decay constant of 1.4210-11yr-1).Judged from the fact that many of the direct counting experiments have yielded results
that are not compatible with one another within the stated uncertainties (see below), it would appear that not all the
measurement uncertainties are accounted for, and therefore the stated uncertainties are unrealistically small. Many of such
experiments are plagued by unrecognized systematic errors (Begemann et al. 2001). As the nature of these errors is
obscure, it is not straightforward to decide which of the, often mutually exclusive, results of such direct counting experiments
is closest to the true value. Furthermore, the presence of unknown systematic biases makes any averaging dangerous. It is
possible that reliable results of careful workers, listing realistic uncertainties, will not be given the weights they deservethis

aside from the question of whether it makes sense to average numbers that by far do not agree within the stated
uncertainties.
In-Growth Experiments
An alternative approach to determining the 87Rb decay constant and half-life is to measure the amount of 87Sr produced by
the decay of a known quantity of 87Rb in the laboratory over a known period of time. This method was first attempted by
McMullen, Fritze, and Tomlinson (1966) on an Rb sample that was purified in 1956, and was repeated on the same sample
batch by Davis et al. (1977). Unfortunately, McMullen, Fritze, and Tomlinson (1996) omitted to measure the small but
significant level of residual 87Sr present in their Rb sample before they put it away on the shelf. Hence, the accuracy of their
determination was compromised. However, this problem contributed less than 1% uncertainty to the later determination of
Davis et al. (1977). Their proposed value for the 87Rb half-life of 48.90.4Byr, equivalent to a decay constant of 1.4210-11
yr-1, would thus seem to support the value of Neumann and Huster (1976) obtained by direct counting.This technique thus
relies on measuring the -decay product 87Sr of a well-known amount of a radioactive 87Rb accumulated over a well-defined
period of time. Where feasible, this is the most straightforward technique (Begemann et al. 2001). In-growth measurement
overcomes the problems encountered with direct counting of large fractions of low-energy 87Rb-emitted -particles. It also
comprises the direct 87Sr product of any radiation-less decays (which otherwise cannot be measured at all).
Among the drawbacks of this approach is that the method is obviously not instantaneous. The experiment must be started
long before the first results can be obtained because long periods of time (typically decades) are required for sufficiently
large amounts of the decay products to accumulate. In-growth experiments further require an accurate determination of the
ratio of the two chemical elements (parent/daughter, 87Rb/87Sr) as well as an accurate determination of the isotopic
composition of parent and daughter elements Rb and Sr at the start of the accumulation (Begemann et al. 2001).
Geological comparisons of methods
The third approach to the determination of the 87Rb decay constant (and half-life) has been to date geological samples
whose ages have also been measured by other methods with presumably more reliable decay constants (Dickin 2005;
Faure and Mensing 2005). This method has the disadvantage that it involves geological uncertainties, such as whether all
isotopic systems closed at the same time and remained closed. However, it is claimed to still provide a useful check on the
direct laboratory determinations. In this respect it is worth noting that Pinson et al. (1963) proposed a 87Rb half-life of 48.8
Byr on the basis of Rb-Sr dating of stony meteorites (chondrites) that had also been U-Pb dated, well before Neumann and
Huster (1974, 1976) arrived at the essentially the same 87Rb half-life value via direct counting.
This approach entails multi-chronometric dating of a rock and cross-calibration of different radioisotopic age systems by
adjusting the decay constant of one system so as to force agreement with the age obtained via another dating system
(Begemann et al. 2001). In essence, because the half-life of 238U is the most accurately known of all relevant radionuclides,
this usually amounts to expressing ages in units of the half-life of 238U.
Results of the Rubidium-87 Decay Determinations
During the last 60 years numerous determinations of the 87Rb decay constant and half-life have been made using these
three methods. The results are listed with details in Table 1. The year of the determination versus the value of the decay
constant is plotted in Fig. 1, and the year of the determination versus the value of the half-life is plotted in Fig. 2. In each
diagram the data points plotted have been color-coded the same to differentiate the values as determined by the three
approaches that have been useddirect counting, in-growth experiments, and geological comparisons with other
radioisotope dating methods.
Table 1. Determinations of the 87Rb decay rate expressed in terms of the half-life and decay constant using direct counting,
in-growth experiments and comparisons of radioisotope ages.
Determination of the 87Rb Decay Rate
Decay
Constant
(x10-11yr -1)

Date

Half-Life
(x1010yr)

1948

6.0

0.6

1.155

0.2

Liquid
scintillationHaxel, Houtermans,
direct/absolute counting
and Kemmerich 1948

1951

6.15

0.30

1.127

0.09

Liquid scintillation/spectrometerCurran, Dixon, and


direct counting
Wilson 1951

1952

5.90

0.30

1.175

0.09

Liquid scintillation/spectrometer
direct counting
Lewis 1952

1954

6.2

0.3

1.118

0.09

Spectrometer direct counting

1955

4.3

0.25

1.612

0.07

Scintillation/spectrometer directGeese-Bhnisch
counting
1955

1956

5.0

0.2

1.39

0.06

U-Pb
age
minerals

1959

4.70

0.10

1.475

0.031

Liquid
scintillationFlynn and Glendenin
direct/absolute counting
1959

1961

5.25

0.10

1.32

0.03

Scintillation direct counting

1961

5.82

0.1

1.191

0.03

Scintillation/spectrometer directEgelkraut and Leutz


counting
1961

1961

5.53

0.10

1.253

0.03

Scintillation/spectrometer direct
counting
Beard and Kelly 1961

1962

5.80

0.12

1.195

0.04

Scintillation/spectrometer directLeutz,
Wenninger,
counting
and Ziegler 1962

Method

Source

comparison

MacGregor
and
Wiedenbeck 1954

of
Aldrich et al. 1956

McNair and Wilson


1961

Comparison of Rb-Sr and other


ages for stony meteorites
Pinson et al 1963

1963

4.88

1.420

1964

4.77

0.10

1.453

0.03

Liquid
scintillation
direct/absolute counting
Kovch 1964

1965

5.21

0.15

1.33

0.05

Liquid
scintillationBrinkman, Aten, and
direct/absolute counting
Veenboer 1965

1966

4.72

0.04

1.468

0.012

In growth experiment by massMcMullen, Fritze, and


spectrometry
Tomlinson 1966

1974,
1976

4.88

0.06/0.10

1.420

0.030/0.017

Neumann and Huster


Absolute counting experiments 1974, 1976

1977

4.89

0.04

1.419

0.012

In growth
counting

1977

4.88

0.03

1.420

0.010

K-Ar and U-Pb age comparisonsSteiger


of minerals
1977

1982

4.94

0.03

1.402

0.008

U-Th-Pb age comparison with H,Minster. Birck,


E, and LL chondrites
Allgre 1982

1982

4.88

0.03

1.420

0.010

K-Ar comparison
Australian granites

1985

4.94

0.04

1.402

0.011

Sm-Nd + K-Ar age comparisons


of lunar rocks
Shih et al 1985

2001

4.94

0.03

1.402

0.008

Comparisons of all methods

2002

4.95

0.02

1.400

0.007

Comparison of Rb-Sr and U-Th-Amelin and Zaitsev


Pb mineral ages
2002

2003

4.967

0.032

1.395

0.009

Liquid
scintillation
direct/absolute counting
Kossert 2003

2011

4.975

0.010

1.393

0.004

Comparison of Rb-Sr and U-Th-Nebel, Scherer, and


Pb mineral ages
Mezger 2011

2012

4.962

0.008

1.397

0.0015

In growth experiment by mass


spectrometry
Rotenberg et al 2012

experiment

and
Davis et al. 1977

with

and

Jager
and

SE
Williams et al 1982

Begemann et al 2001

Discussion
The values for the 87Rb decay constant used in Rb-Sr age calculations for rocks over the last sixty years have varied
between 1.4710-11yr-1 and 1.3910-11yr-1 (equivalent to half-lives of 46.8Byr and 50.0Byr respectively). The most
commonly used value of 1.4210-11yr-1 (48.8Byr) was adopted by international convention (Steiger and Jger 1977). Two
papers reporting direct determinations of the 87Rb decay constant and half-life that had appeared before preparation of the
report of the 1976 IUGS Subcommission on Geochronology were cited by that 1976 meeting of this Subcommission (Steiger
and Jger 1977) as influencing their recommendation of the 87Rb decay constant of 1.4210-11 yr-1, corresponding to a halflife of 48.8Byr. These were Neumann and Huster (1974), who reported measuring the specific activity of thin sources of
rubidium chloride (RbCl) using a 4 proportional counter to obtain a 87Rb half-life of 4.88+0.06/0.10Byr, and Davis et al.
(1977), who reported continuing the direct mass spectrometric in-growth experiment initiated by McMullen, Fritze, and
Tomlinson (1966) to obtain a 87Rb half-life of 4.890.04Byr.

Fig. 1. Plot of each 87Rb half-life determination versus the year of its determination, color-coded according to the method of
its determination.
Fig. 2. Plot of each 87Rb decay constant determination versus the year of its determination, color-coded according to the
method of its determination.
A third line of evidence cited by Steiger and Jger (1977) as influencing their choice of the recommended value of the 87Rb
decay constant was a comparison of Rb-Sr ages with K-Ar ages presented by Tetley et al. (1976). However, it was later
noted that in Tetley et al. (1976) there was a discernible tendency towards higher calculated 87Rb decay constant values for
lower K-Ar ages, as would be expected if some of the calculated K-Ar ages were too low due to Ar loss (Begemann et al.

2001). Consideration of the more extensive data set of Williams et al. (1982), which contained the data of Tetley et al. (1976)
as a subset, suggested that the average 87Rb constant value provided by this comparison may have been biased high by 1
to 2% 40Ar loss from many of the analyzed samples. This loss may have been due to alteration, as all micas show substoichiometric K concentrations. Shih et al. (1985) compared Rb-Sr ages of lunar KREEP basalts to 39Ar-40Ar ages
calculated with the 40K decay parameters recommended by Steiger and Jger (1977), and found the best agreement for
the 87Rb decay constant was 1.4020.01110-11yr-1 (corresponding to a half-life of 49.4Byr). Although Shih et al. (1985)
compared significantly fewer ages, according to Begemann et al. (2001) the correction for 40Ar loss inherent in the 39Ar-40Ar
technique probably made the value of the 87Rb decay constant they obtained more reliable than that derived by Tetley et al.
(1976). Note, however, that the large uncertainties in the 40K decay constants of Steiger and Jger (1977) undermine the
usefulness of any normalization to this system (Begemann et al. 2001).Thus the value recommended by the
Subcommission did not gain universal acceptance. For example, Minster, Birck, and Allgre (1982) summarized the Rb-Sr
data for chondrites (stony meteorites) from their Paris laboratory, and found the best agreement between the whole-rock RbSr age of those chondrites and their average U-Pb ages of 4.5550.010Ga was for a 87Rb decay constant of 1.4020.008
10-11yr-1(equivalent to a half-life of 49.4Byr). Conversely, the whole-rock isochron for the chondrites yielded an
initial 87Sr/86Sr ratio of 0.698850.00010 and an age of 4.4980.015Ga for an 87Rb decay constant of 1.4210-11yr1
(corresponding to a half-life of 48.8Byr). Since that initial 87Sr/86Sr ratio is similar to other estimates of the initial87Sr/86Sr for
the solar system (for example, BABI, Papanastassiou and Wasserburg 1969), Begemann et al. (2001) concluded there was
no reason to suspect the Rb-Sr isotopic systematics of these chondrites had been reset to an age lower than the U-Pb
age.Thus Begemann et al. (2001), in issuing a call for new revised values, contended that therefore it was totally reasonable
to expect the Rb-Sr and U-Pb ages for these chondrites to be the same. Thus they argued for the 87Rb decay constant of
1.4020.00810-11yr-1 and the equivalent half-life of 49.4Byr proposed by Minster, Birck, and Allgre (1982) to be the new
values. They also pointed out that the most commonly encountered bias factors in determining the 87Rb decay constant all
operated in a single direction, increasing the apparent value of the 87Rb decay constant. Begemann et al. (2001) thus
concluded that there was therefore a significant probability that the value recommended by Steiger and Jger (1977) was
too high by 1 to 2%. Furthermore, they suggested that investigators seeking a precise comparison of Rb-Sr ages to those
determined by other methods should consider the effect on the comparison of using an 87Rb decay constant of 1.40210-11
yr-1, as suggested originally by Minster, Birck, and Allgre (1982), and recommend that the value of the 87Rb decay constant
should be re-determined with improved accuracy.This recommendation for new determinations of the 87Rb decay constant
and half-life with improved accuracy was immediately pursued by Kossert (2003) by liquid scintillation direct counting, and
later by Rotenberg et al. (2012) with in-growth experiments (see table 1). However, rather than confirming the values
Begemann et al. (2001) preferred, Kossert (2003) found from his measurements that the 87Rb half-life should be 49.67
0.32Byr (equivalent to a decay constant of 1.3950.00910-11yr-1). Similarly, Rotenberg et al. (2012) concluded from their
measurements of the 87Sr accumulation in RbClO4 salt over 32 years that the 87Rb half-life should be 49.624 +0.065/-0.095
Byr (equivalent to a decay constant of 1.3968 +0.0027/-0.001810-11yr-1). Furthermore, two more recent studies reported
comparisons between Rb-Sr and U-Pb ages for minerals (Amelin and Zaitsev 2002) and rocks (Nebel, Scherer, and Mezger
2011), and both also suggested lower values for the 87Rb half-life and decay constant than those championed by Begemann
et al. (2001) (see table 1).In order to more clearly see the differences between the determinations made in the last 30 years
by all three methods, the scatter of the early years of determinations was ignored and only the last 30 years results were
plotted using the same color coding for the different methods, the half-life determinations in Fig. 3 and the decay constant
determinations in Fig. 4. For the purpose of this discussion, the two competing values for the half-life (48.8 and 49.4 Byr)
and decay constant (1.4210-11yr-1 and 1.40210-11yr-1 respectively) have been marked on the diagrams by vertical
dashed lines.

Fig. 3. Comparison of the 87Rb half-life determinations obtained in the last 30 years, grouped and color-coded according to
the methods of determination. The vertical dashed lines represent the two competing values for the 87Rb half-life in the
literature, while the vertical dotted line represents the proposed new best value based on the most recent determinations, as
discussed in the text.

Fig.4. Comparison of the 87Rb decay constant determinations obtained in the last 30 years, grouped and color-coded
according to the methods of determination. The vertical dashed lines represent the two competing values for the 87Rb decay
constant in the literature, while the vertical dotted line represents the proposed new best value based on the most recent
determinations, as discussed in the text.The grouping of determinations around a half-life of 48.8Byr corresponding to a
decay constant of 1.4210-11yr-1includes the earlier Neumann and Huster (1974, 1976) direct counting determination and
Davis et al. (1977) in-growth experiment, and the earlier Steiger and Jger (1977) and Williams et al. (1982) geological
comparisons of the whole-rock Rb-Sr and K-Ar ages of rapidly cooled igneous rocks and granites (figs 3 and 4). In stark
contrast, the grouping of determinations around a half-life of 49.4 Byr corresponding to a decay constant of 1.40210-11yr1
championed by Begemann et al. (2001) is only supported by the earlier comparison of ages studies of Minster, Birck, and
Allgre (1982) on Rb-Sr and U-Pb ages of chondrites (stony meteorites) and Shih et al. (1985) on Rb-Sr and Ar-Ar ages of
lunar KREEP basalts (figs 3 and 4).What is also evident in Figs 3 and 4 is that the most recent half-life and decay constant
determinations by Kossert (2003) by direct counting and Rotenberg et al. (2012) by in-growth experiments do not support
the half-life of 49.4Byr and decay constant of 1.40210-11yr-1. And neither do the most recent geological comparisons of
ages studies by Amelin and Zaitsev (2002) and Nebel, Scherer, and Mezger (2011). This observation was not lost on Nebel,
Scherer, and Mezger (2011) and Rotenberg et al. (2012) either, who thus both championed a different longer half-life and
lower decay constant, which are shown in their own compilation diagrams in Figs 5 and 6 respectively.
Indeed, when Amelin and Zaitsevs (2002) fast and slow cooling determinations are separated from the averaging of them
shown in table 1, and plotted in Figs 3 and 4 separately, as they are by Rotenberg et al. (2012) (fig. 6), the slow cooling
determinations plot with the Begemann et al. (2001) championed half-life and decay constant, whereas the fast cooling
determinations plot with the most recent determinations by Kossert (2003) by direct counting, Rotenberg et al. (2012) by ingrowth experiments, and Nebel, Scherer, and Mezger (2011) by comparisons of geological ages. This grouping is shown in
Figs 3 and 4 by dotted lines corresponding to a half-life of 49.67 0.05Byr and a decay constant of 1.395 0.00510-11yr1
. But there is more here in these data than what has been observed by Nebel, Scherer and Mezger (2011) and Rotenberg
et al. (2012). It could be significant that the comparisons of geological ages performed by Amelin and Zaitsev (2002) and
Nebel, Scherer, and Mezger (2011) were based on minerals (baddeleyite, zircon, apatite, phlogopite, biotite, feldspar, and
amphibole) from earth rocks (phoscorites, carbonatites, and syenites) using U-Pb ages as the gold standard, whereas
those of Minster, Birck, and Allgre (1982) and Shih et al. (1985) were based on chondrites (stony meteorites) and lunar
KREEP basalts respectively. Indeed, this same observation that there are differences in the half-life and decay constant
between determinations using earth materials and determinations using meteorites has already been documented for 176Lu
by Albarde et al. (2006). They found that using the earth-based 176Lu decay constant the Lu-Hf ages of chondrites were
consistently 4% higher than their U-Pb ages, so the reconciliation of the Lu-Hf and U-Pb ages required using a different
meteorite-based 176Lu decay constant. Ironically, but significantly, Albarde et al. (2006) proposed that this discrepancy
between decay constants could be because of accelerated 176Lu decay during the first few million years of the existence of
the solar nebula due to
irradiation of the 176Lu by
-rays emitted by one or more
supernova(e) exploding in the
vicinity of the solar nebula.
Fig. 5. Comparison of the 87Rb
decay constant determinations
of the last thirty years
according to Nebel, Scherer,
and Mezger (2011).
Fig. 6. Comparison of the 87Rb
half-life and decay constant
determinations of the last 30
years according to Rotenberg
et al. (2012).
In the related context of
nucleosynthesis
of
the
elements inside stars, Zhao
and Kppeler (1990) found the
absorption cross-section to
produce 176Lu to be small and
hence
found
problems
explaining their data. They
concluded that to reestablish

the 176Lu clock for determining the age of the s-process elements would, in any case, require a quantitative description of all
processes feeding the ground state as well as a reliable model for the s-process environment, in particular for the
temperatures to which 176Lu was exposed during its production and ever since. However, even though these effects of
ionization on the nuclear half-lives can be enhanced inside stars due to the extreme ionization prevalent there, such
contexts are not relevant in the creationist framework for the history of the earth and the universe. The earth and its rocks,
and their contained elements and isotopes, were all created before the stars, the moon, and the asteroids from which the
meteorites subsequently came. And then in their history since their creation neither the earth nor the asteroids have been
subjected to the intense ionization inside stars.There would thus appear to be similar discrepancies in the 87Rb and 176Lu
decay constants and half-lives between determinations using earth materials and determinations using meteorites and lunar
rocks. This is the first time this discrepancy has been recognized for the 87Rb half-life and decay constant. It also may be
relevant that both 87Rb and176Lu decay via the same single-step -decay process, both producing anti-neutrinos and decay
energy. Such discrepancies are a conundrum to conventional wisdom, but may provide new insights for creationist thinking.
Furthermore, even the U-Pb gold standard has come under similar scrutiny in the last decade. The U-Pb method depends
on the crucial 238U/235U ratio, but discrepancies and variations have been found between the 238U/235U ratio in U-bearing
earth-based (terrestrial) minerals and rocks and the 238U/235U ratio in meteorites (Brennecka and Wadhwa 2012; Hiess et al
2012). The potential significance of these discrepancies and variations from a creationist perspective awaits further study,
especially in the context of examinations of the determinations of the 176Lu, 238U, and 235U decay constants and half-lives.
Nevertheless, it is still obvious that in spite of this discrepancy, the most recent earth-based direct counting and in-growth
experimental determinations of the 87Rb half-life and decay constant (Kossert 2003; Rotenberg et al. 2012) coincide with
the 87Rb half-life and decay constant determined via comparisons of Rb-Sr and U-Pb ages of minerals from earth rocks
(Amelin and Zaitsev 2002; Nebel, Scherer, and Mezger 2011), and not with the currently accepted 87Rb half-life and decay
constant (Begemann et al. 2001). And the fact that there are also variations in the crucial 238U/235U ratio in earth minerals and
rocks on which the U-Pb dating gold standard depends, which is so obviously used to calibrate the 87Rb half-life and
decay constant, only underscores that these radioisotope methods cannot provide the absolute invariable dates they are
so confidently proclaimed to provide. It is thus surely significant that the Rb-Sr method has not been used to provide any of
the absolute dates for the GSSPs (global boundary stratotype sections and points) or golden spikes in the official and
approved Geologic Time Scale of the International Union of Geological Sciences (IUGS) (International Union of Geological
Sciences 2014; Schmitz 2012a, b).Not only does the potential significance of these discrepancies in the 87Rb and 176Lu halflives and decay constants and the variations in the 238U/235U ratio need further investigating, but also the effect on the Rb-Sr
ages of meteorites in relation to their ages derived by the other radioisotope dating systems, when those Rb-Sr ages are
recalculated using the revised 87Rb half-life and decay constant proposed here (49.67 0.05Byr and 1.395 0.00510-11yr1
). It is possible that such systematically recalculated Rb-Sr ages may introduce a systematic difference between the
meteorites Rb-Sr age ages relative to their ages derived by the other radioisotope dating systems. However, since the
changes in the meteorites Rb-Sr ages introduced by such systematic recalculations would only amount to an increase of
about 0.61.8%, the effect in comparison to their ages derived by the other radioisotope dating systems would likely be
negligible, especially given the error margins in many meteorite age determinations. Thus such recalculated meteorite Rb-Sr
ages would not likely reveal any of the trends that would evidence a past accelerated radioisotope decay event (Vardiman,
Snelling, and Chaffin 2005). This is confirmed by Snelling (2014a, b), who has shown that at least for the chondrites (the
carbonaceous, ordinary and enstatite stony meteorites) there are no indications of any systematic differences in their
isochron ages with respect to their ages derived by the other radioisotope dating systems that would evidence a past
accelerated radioisotope decay event. Instead, their Rb-Sr ages either occasionally agree with the dominant strong clusters
of Pb-Pb ages, or more often are widely scattered either side of those clusters.Finally, it would remain prudent to be very
careful with these geological comparison methods for two reasons. First, there are significant flaws in the basic assumptions
that all of them embrace. Second, the U-Pb method relies primarily on decay whereas the Rb-Sr method relies on
decay. As observed by both Austin (2005) and Snelling (2005), these different decay modes seem to yield different ages for
earth rocks using the same samples with essentially the same methodology. Furthermore, as described earlier, direct
counting by liquid scintillation is fraught with pitfalls when measuring low-level emitters.
Conclusions
There have been numerous attempts to determine the 87Rb half-life and decay constant in the last sixty years by three
methodsdirect counting of -particles, in-growth experiments in which the daughter 87Sr accumulated over a given time is
measured, and comparisons of Rb-Sr ages of earth minerals and rocks, lunar rocks, and meteorites with ages derived by
other radioisotope systems, especially U-Pb. The estimates of the 87Rb half-life are really consistent since 1985. Going back
in time the first inconsistent measurement was by comparison of radioisotope ages for some southeast Australian granites,
and that was just barely inconsistent. Thus the data seem to be simply paralleling the improvements in measurement
technologies with poorly estimated errors.However, with the improved accuracy in such determinations in the last thirty
years, a discrepancy has been revealed which has been debated in the literature, but with no attempted explanation as to
the cause. The discrepancy is between the 87Rb half-life and decay constant (48.8Byr and 1.42 x 10-11 yr-1 respectively)
determined by comparing the Rb-Sr ages of earth rocks against their K-Ar and Ar-Ar ages, and the 87Rb half-life and decay
constant (49.4Byr and 1.40210-11yr-1 respectively) determined by comparing the Rb-Sr ages of meteorites and lunar rocks
with their U-Pb and Ar-Ar ages respectively. The direct counting and in-growth experimental determinations of the 1970s
only agree with the 87Rb half-life of 48.8Byr determined by comparing the Rb-Sr ages of earth rocks against their K-Ar and
Ar-Ar ages. Yet the 87Rb half-life of 49.4Byr determined by comparing the Rb-Sr ages of meteorites and lunar rocks with
their U-Pb and Ar-Ar ages respectively have been championed since the mid-1980s as the best to use in calculating Rb-Sr
ages.Nevertheless, now a further discrepancy has been revealed in the last decade by the most recent 87Rb half-life
determinations. Both new determinations by direct counting and in-growth experiments do not agree with either of the two
previously derived 87Rb half-lives and decay constants. Instead, they only agree with the two new determinations derived
from comparing Rb-Sr ages of further earth minerals and rocks with their U-Pb ages, the so-called gold standard. It is thus
proposed here that these new data define a new and better estimate of the 87Rb half-life and decay constant of 49.67 0.05
Byr and 1.395 0.00510-11yr-1 respectively. Furthermore, this highlights the disparity between determinations using earth
materials and extraterrestrial materials, which is similar to the already documented similar disparity in determinations of
the 176Lu half-life and decay constant. These discrepancies only serve to highlight that the Rb-Sr dating method cannot be
absolute when the 87Rb decay rate has not been accurately determined. When it is calibrated against the U-Pb gold
standard with its own uncertainties in the critical 238U/235U ratio it still cannot be regarded as absolute. Therefore, Rb-Sr
dating cannot be used to discredit the young-earth creationist timescale. However, further study is needed to explore
whether this discrepancy between earth and extraterrestrial materials has any significance.

CARBON DATING
Carbon-14 Dating
Understanding the Basics
by Dr. Andrew A. Snelling on October 1, 2010; last featured March 30, 2011
Many people assume that rocks are
dated at millions of years based
on radiocarbon (carbon-14) dating.
But thats not the case. The reason
is simple. Carbon-14 can yield
dates of only thousands of years
before it all breaks down.
Shop Now
The most well-known of all the
radiometric dating methods is
radiocarbon dating. Although many
people think radiocarbon dating is used to date rocks, it is limited to dating things that contain the element carbon and were
once alive (like fossils).
Carbon-14 Dating
Part 1 Understanding the Basics
Part 2 An Evolution DilemmaC-14 in Fossils and Diamonds
Part 3 A Creationist Puzzle50,000-year-old Fossils?
How Radiocarbon Forms
Unlike radiocarbon (14C), the other radioactive elements used to date rocksuranium ( 238U), potassium (40K), and rubidium
(87Rb)are not being formed on earth, as far as we know. Thus it appears that probably those elements were created when
the original earth was formed.In contrast, radiocarbon forms continually today in the earths upper atmosphere. And as far as
we know, it has been forming in the earths upper atmosphere since the atmosphere was made back on the Creation event
So how does radiocarbon form? Cosmic rays from outer space are continually bombarding the upper atmosphere of the
earth, producing fast-moving neutrons (subatomic particles carrying no electric charge) (Figure 1a).1 These fast-moving
neutrons collide with atoms of nitrogen-14, the most abundant element in the upper atmosphere, converting them into
radiocarbon (carbon-14) atoms.
CARBON-14 IS CREATED (Figure
1a): When cosmic rays bombard the
earths atmosphere, they produce
neutrons. These excited neutrons
then collide with nitrogen atoms in the
atmosphere, changing them into
radioactive
carbon-14
atoms.
CARBON-14 IS ABSORBED (Figure
1b): Plants absorb this carbon-14
during photosynthesis. When animals
eat the plants, the carbon-14 enters
their bodies. The carbon-14 in their
bodies breaks down to nitrogen-14
and escapes at the same rate as new
carbon-14 is added. So the level of
carbon-14
remains
stable.
CARBON-14 IS DEPLETED (Figure
1c): When an animal dies the carbon14 continues to break down to
nitrogen-14 and escapes, while no
new carbon-14 is added. By
comparing the surviving amount of
carbon-14 to the original amount,
scientists can calculate how long ago the animal died.Since the atmosphere is composed of about 78% nitrogen, 2 a lot of
radiocarbon atoms are producedin total about 16.5 pounds (7.5 kg) per year. These rapidly combine with oxygen atoms
(the second most abundant element in the atmosphere, at 21%) to form carbon dioxide (CO 2).This carbon dioxide, now
radioactive with carbon-14, is otherwise chemically indistinguishable from the normal carbon dioxide in the atmosphere,
which is slightly lighter because it contains normal carbon-12. Radioactive and non-radioactive carbon dioxide mix
throughout the atmosphere, and dissolve into the oceans.Through photosynthesis carbon dioxide enters plants and algae,
bringing radiocarbon into the food chain. Radiocarbon then enters animals as they consume the plants (Figure 1b). So even
we humans are radioactive because of trace amounts of radiocarbon in our bodies.
Determining the Rate of Radiocarbon Decay
After radiocarbon forms, the nuclei of the carbon-14 atoms are unstable, so over time they progressively decay back to
nuclei of stable nitrogen-14.3 A neutron breaks down to a proton and an electron, and the electron is ejected. This process
is called beta decay. The ejected electrons are called beta particles and make up what is called beta radiation.Not all
radiocarbon atoms decay at the same time. Different carbon-14 atoms revert to nitrogen-14 at different times, which
explains why radiocarbon decay is considered a random process.To measure the rate of decay, a suitable detector records
the number of beta particles ejected from a measured quantity of carbon over a period of time, say a month (for illustration
purposes). Since each beta particle represents one decayed carbon-14 atom, we know how many carbon-14 atoms decay
during a month.Chemists have already determined how many atoms are in a given mass of each element, such as
carbon.4 So if we weigh a lump of carbon, we can calculate how many carbon atoms are in it.If we know what fraction of the

carbon atoms are radioactive, we can also calculate how many radiocarbon atoms are in the lump. Knowing the number of
atoms that decayed in our sample over a month, we can calculate the radiocarbon decay rate.The standard way of
expressing the decay rate is called the half-life.5 Its defined as the time it takes half a given quantity of a radioactive
element to decay. So if we started with 2 million atoms of carbon-14 in our measured quantity of carbon, then the half-life of
radiocarbon would be the time it takes for half, or 1 million, of those atoms to decay. The radiocarbon half-life or decay rate
has been determined at 5,730 years.
Using Radiocarbon for Dating
Next comes the question of how scientists use this knowledge to date things. If carbon-14 has formed at a constant rate for
a very long time and continually mixed into the biosphere, then the level of carbon-14 in the atmosphere should remain
constant.If the level is constant, living plants and animals should also maintain a constant carbon-14 level in them. The
reason is that, as long as the organism is alive, it replaces any carbon molecule that has decayed into nitrogen.
After plants and animals perish, however, they no longer replace molecules damaged by radiocarbon decay. Instead, the
radiocarbon atoms in their bodies slowly decay away, so the ratio of carbon-14 atoms to regular carbon atoms will steadily
decrease over time (Figure 1c).Lets suppose we find a mammoths skull and we want to date it to determine how long ago it
lived. We can measure in the laboratory how many carbon-14 atoms are still in the skull. If we assume that the mammoth
originally had the same number of carbon- 14 atoms in its bones as living animals do today (estimated at one carbon-14
atom for every trillion carbon-12 atoms), then, because we also know the radiocarbon decay rate, we can calculate how long
ago the mammoth died. Its really quite simple.This dating method is similar to the principle behind an hourglass.6 The sand
grains that originally filled the top bowl represent the carbon-14 atoms in the living mammoth just before it died. Its assumed
to be the same number of carbon-14 atoms as in elephants living today. With time those sand grains fall to the bottom bowl,
so the new number represents the carbon-14 atoms left in the mammoth skull when we found it.The difference in the
number of sand grains represents the number of carbon-14 atoms that have decayed back to nitrogen-14 since the
mammoth died. Because we have measured the rate at which the sand grains fall (the radiocarbon decay rate), we can then
calculate how long it took those carbon-14 atoms to decay, which is how long ago the mammoth died.Thats how the
radiocarbon method works. And because the half-life of carbon-14 is just 5,730 years, radiocarbon dating of materials
containing carbon yields dates of only thousands of years, not the dates over millions of years that conflict with the
framework of earth history.
Carbon-14 in Fossils and Diamonds
An Evolution Dilemma
by Dr. Andrew A. Snelling on January 1, 2011
If the radioactive element carbon-14 breaks
down quicklywithin a few thousand years
why do we still find it in fossils and
diamonds? Its a dilemma for evolutionists,
who believe the rocks are millions of years
old.
Shop Now
Many people think that scientists use
radiocarbon to date fossils. After all, we
should be able to estimate how long ago a
creature lived based on how much radiocarbon is left in its
body, right?
Carbon-14 Dating
Part 1 Understanding the Basics
Part 2 An Evolution Dilemma
Part 3 A Creationist Puzzle
Why Isnt Radiocarbon Used to Date Fossils?
The answer is a matter of basic physics. Radiocarbon
(carbon-14) is a very unstable element that quickly changes
into nitrogen. Half the original quantity of carbon-14 will
decay back to the stable element nitrogen-14 after only
5,730 years. (This 5,730-year period is called the half-life of
radiocarbon, Figure 1).1 2 At this decay rate, hardly any
carbon-14 atoms will remain after only 57,300 years (or ten
half-lives).So if fossils are really millions of years old, as
evolutionary scientists claim, no carbon-14 atoms would be
left in them. Indeed, if all the atoms making up the entire
earth were radiocarbon, then after only 1 million years
absolutely no carbon-14 atoms should be left!
The Power of Radiocarbon Detection Technology
Most laboratories measure radiocarbon with a very
sophisticated instrument called an accelerator mass
spectrometer, or AMS. It is literally able to count carbon-14
atoms one at a time.3 This machine can theoretically detect
one radioactive carbon-14 atom in 100 quadrillion regular
carbon-12 atoms!However, theres a catch. AMS instruments
need to be checked occasionally, to make sure they arent
also reading any laboratory contamination, called
background. So rock samples that should read zero are
occasionally placed into the instruments to test their
accuracy. What better samples to use than fossils, coals, and
limestones, which are supposed to be millions of years old
and should have no radiocarbon?
Radiocarbon Found!

Imagine the surprise when every piece of ancient carbon tested has contained measurable quantities of radiocarbon!
4 Fossils, coal, oil, natural gas, limestone, marble, and graphite from every Flood-related rock layerand even some preFlood depositshave all contained measurable quantities of radiocarbon (Figure 2). All these results have been reported in
the conventional scientific literature.
Figure 1 Radiocarbon has a very short half-life. At current decay rates, the number of radiocarbon atoms is halved every
5,730 years. Because of this exponential decay, carbon-14 atoms cant survive millions of years.
Figure 2 Radiocarbon shouldnt be found in old rocks, but it is! Once creatures die, the radiocarbon in their bodies should
quickly break down. After millions of years, their remains would be completely free of radiocarbon. But samples of organic
materials taken from every rock layer, such as fossils, coal, limestone, natural gas, and graphite, all have measurable
radiocarbon. These findings are reported in the secular scientific literature (but they are usually rejected as measurement
errors).This chart shows the percentage of radiocarbon that remains in 40 samples from various layers throughout the
geologic column. (This percentage, technically known as percent modern carbon [pMC], shows the ratio of radiocarbon in
the rocks and fossils compared to the amount we find in living things).This finding is consistent with the belief that rocks are
only thousands of years old, but the specialists who obtained these results have definitely not accepted this conclusion. It
does not fit their presuppositions. To keep from concluding that the rocks are only thousands of years old, they claim that the
radiocarbon must be due to contamination, either from the field or from the laboratory or from both. However, when the
technician meticulously cleans the rocks with hot strong acids and other pre-treatments to remove any possible
contamination, these ancient organic (once-living) materials still contain measurable radiocarbon.Since a blank sample
holder in the AMS instrument predictably yields zero radiocarbon, these scientists should naturally conclude that the
radiocarbon is intrinsic to the rocks. In other words, real radiocarbon is an integral part of the ancient organic materials.
But these scientists presuppositions prevent them from reaching this conclusion.
Radiocarbon in Fossils Confirmed
Photo courtesy of Dr. Andrew Snelling
Figure 3 Sample from Marlstone Rock Bed, a muddy limestone
in one wall of the Hornton Quarries at Edge Hill, west of Banbury
in England. Pieces of fossilized wood in Jurassic rocks,
supposedly millions of years old, yielded radiocarbon ages of
only 20,70028,820 years.For some years creation scientists
have been doing their own investigation of radiocarbon in fossils.
Pieces of fossilized wood in Oligocene, Eocene, Cretaceous,
Jurassic, Triassic, and Permian rock layers supposedly 32250
million years old all contain measurable radiocarbon, equivalent
to ages of 20,700 to 44,700 years (Figures 3
5).5 6 7 8 9 10 11 (Creation geologists believe that with careful
recalibration, even these extremely young time periods would
be fewer than 10,000 years.)Similarly, carefully sampled pieces
of coal from ten U.S. coal beds, ranging from Eocene to
Pennsylvanian and supposedly 40320 million years old, all contained similar radiocarbon levels equivalent to ages of
48,000 to 50,000 years.12 Even fossilized ammonite shells found alongside fossilized wood in a Cretaceous layer,
supposedly 112120 million years old, contained measurable radiocarbon equivalent to ages of 36,400 to 48,710 years
(Figure 5).13
Radiocarbon is Even in Diamonds
Photo courtesy of Dr. Andrew Snelling
Figure 4 Sample from mudstone on top of the Great
Northern Seam in the upper Permian Newcastle Coal
Measures in the Newvale No. 2 Coal Mine north of Sydney,
Australia. A fossilized tree stump, found in Permian layers,
supposedly hundreds of millions of years old, yielded
coalified bark with a radiocarbon age of 33,700 years.
Photo courtesy of Dr. Andrew Snelling
Figure 5 These fossils were in mudstone of the lower
Cretaceous Budden Canyon Formation near Redding,
California. A fossilized ammonite (a marine shellfish) was
discovered with a piece of fossilized wood (from a land
plant) embedded next to it. Located in Cretaceous layers
that were supposedly millions of years old, the fossilized
shell and wood yielded radiocarbon ages of 48,710 and
42,390 years respectively.
Just as intriguing is the discovery of measurable radiocarbon
in diamonds. Creationist and evolutionary geologists agree
that diamonds are formed more than 100 miles (161 km)
down, deep within the earths upper mantle, and do not
consist of organic carbon from living things. Explosive
volcanoes brought them to the earths surface very rapidly in
pipes.As the hardest known natural substance, these
diamonds are extremely resistant to chemical corrosion and
external contamination. Also, the tight bonding in their
crystals would have prevented any carbon-14 in the
atmosphere from replacing any regular carbon atoms in the
diamond.Yet diamonds have been tested and shown to
contain radiocarbon equivalent to an age of 55,000
years.14 15 These results have been confirmed by other
investigators.16 So even though these diamonds are
conventionally regarded by evolutionary geologists as up to
billions of years old, this radiocarbon has to be intrinsic to

them.This carbon-14 would have been implanted in them when they were formed deep inside the earth, and it could not
have come from the earths atmosphere. This is not such a problem for creationist scientists, but it is a serious problem for
evolutionists.
The Radiocarbon Puzzle
Evolutionary radiocarbon scientists have still not conceded that fossils, coals, and diamonds are only thousands of years
old. Their uniformitarian (slow-and-gradual) interpretation requires that the earths rocks be millions or billions of years old.
They still maintain that the carbon-14 is machine background contaminating all these tested samples.Among their
proposed explanations is that the AMS instruments do not properly reset themselves between sample analyses. But if this
were true, why would the instrument find zero atoms when no sample is in it?It should be noted that radiocarbon ages of
up to 50,000 years dont match the creation time frame, either. The Flood cataclysm was only about 4,350 years ago.
However, these young radiocarbon ages are far more in accord with the creation account than the uniformitarian
timescale. The discovery that diamonds have 55,000-year radiocarbon ages may help us unravel this mystery.The article
in the next issue of Answers magazine will examine how it may be possible to systematically correct radiocarbon ages.
Once radiocarbon is interpreted properly, it should help creationists date archaeological remains from post-Flood human
history.
A Creationist Puzzle
50,000-Year-Old-Fossils
by Dr. Andrew A. Snelling on April 1, 2011; last featured April 25, 2012
Evolutionists arent the only ones who run into
challenges
when
trying
to
reconcile
radiocarbon dating with their view of history.
How do creationists explain dates of 50,000
years?
Conventional geologists claim that fossils,
coals, and diamonds are millions to billions of
years old. Yet it has now been firmly
established that they still contain measurable
amounts of radiocarbon, which has a half-life (decay rate) of only 5,730 years.1This creates a dilemma for conventional
geology, as explained in Part 2 of this series.2 Absolutely no radiocarbon should be left in fossils, coals, and diamonds,
because after just one million years it should have decayed away.
Carbon-14 Dating
Part 1 Understanding the Basics
Part 2 An Evolution Dilemma
Part 3 A Creationist Puzzle
Yet the radiocarbon in these fossils, coals, and diamonds equates to ages of up to 55,000 years. This is much older than
the creation time frame of earth history, which attributes most fossils and coals to the global Flood about 4,350 years ago.
Assumptions Change Estimate of Age
To solve this puzzle it is necessary to review the assumptions on which radiocarbon dating is based. These include:
The production rate of carbon-14 has always been the same in the past as now.
The atmosphere has had the same carbon-14 concentration in the past as now.
The biosphere (the places on earth where organisms live) has always had the same overall carbon-14 concentration as the
atmosphere, due to the rapid transfer of carbon-14 atoms from the atmosphere to the biosphere.3None of these
assumptions is strictly correct, beyond a rough first approximation. Indeed, scientists have now documented that the
atmospheres concentration of carbon-14 varies considerably according to latitude. They have also determined several
geophysical causes for past and present fluctuations in carbon-14 production in the atmosphere.4Specifically, we know that
carbon-14 has varied in the past due to a stronger magnetic field on earth and changing cycles in sunspot activity. So when
objects of known historical dates are dated using radiocarbon dating, we find that carbon-14 dates are accurate back to only
about 400 BC.The conventional scientific community ignores at least two factors that are crucial to recalibrating radiocarbon
(so that it accounts for major changes in the biosphere and atmosphere that likely resulted from the Flood): (1) The earths
magnetic field has been progressively stronger going back into the past, and (2) the Flood destroyed and buried a huge
amount of carbon from the pre-Flood biosphere.
The Effect of a Past Stronger Magnetic Field
The evidence for the earths having a progressively stronger magnetic field in the past is based on reliable historical
measurements5 and fossil magnetism trapped in ancient pottery.6, 7A stronger magnetic field is significant because the
magnetic field partly shields the earth from the influx of cosmic rays, which change nitrogen atoms into radioactive carbon14 atoms. So a stronger magnetic field in the past would have reduced the influx of cosmic rays.This in turn would have
reduced the amount of radiocarbon produced in the atmosphere. If this were the case, the biosphere in the past would have
had a lower carbon-14 concentration than it does today.The best estimates indicate that the earths magnetic field was twice
as strong 1,400 years ago, and possibly four times as strong 2,800 years ago. If this is true, the earths magnetic field would
have been much stronger at the time of the Flood, and the carbon-14 levels would be significantly smaller.So if you
mistakenly assume that the radiocarbon levels in the atmosphere and biosphere have always been the same as they are
today, you would erroneously estimate much older dates for early human artifacts, such as post-Babel wooden statuettes in
Egypt. And that is exactly what conventional archaeology has done.
The Effect of More Carbon in the Pre-Flood Biospere
An even more dramatic effect on the earths carbon-14 inventory would be the destruction and burial of all the carbon in the
whole biosphere at the time of the Flood. Based on the enormous size of todays coal beds, oil, oil shale, natural gas
deposits, and all the fossils in limestones, shales, and sandstones, a huge quantity of plants and animals must have been
alive when the Flood struck. It is conservatively estimated that the amount of carbon in the pre-Flood biosphere may have
been many times greater than the amount of carbon in todays biosphere.8We cannot yet know for certain how much
radiocarbon (carbon-14) was in this pre-Flood carbon (a mixture of normal carbon-12 and carbon-14). Yet if the earths
atmosphere started to produce carbon-14 (14C) at the Fall, then many radiocarbon atoms could have been in the pre-Flood
biosphere by the time of the Flood, about 1,650 years after Creation.However, if there was a whole lot more normal carbon
(carbon-12, or 12C) in the pre-Flood biosphere, then the proportion of14C to 12C would have been much less than the

proportion in todays biosphere.So when scientists fail to account for so many more plants and animals in the pre-Flood
biosphere and wrongly assume that plants buried in coal beds had the same proportion of carbon-14 as plants do today,
their radiocarbon dating yields ages much higher than the true Flood age of about 4,350 years.
A Prediction Fulfilled
Now if this model of the earths past radiocarbon inventory is correct, then a logical prediction follows. Since all pre-Flood
plants would have had the same low radiocarbon levels when they were buried, and they all formed into coal beds during
that single Flood year, then those coal beds should all have the same low radiocarbon content.
They do! Samples from coal beds around the United States, ranging from Eocene to Pennsylvanian deposits, supposedly
40320 million years old, all contain the same low radiocarbon levels equivalent to ages of 48,00050,000 years.9
This makes sense only if these coal beds were all formed out of pre-Flood plants during the year-long Flood, about 4,350
years ago. Carbon-14 dates of the same value are expected in creation theory but contrary to the expectations of
conventional old-earth theory.
The Puzzle Is Being Solved
So the radiocarbon puzzle can be solved, but only in the creation framework for earth history. Research is therefore
underway to find a means of recalibrating the radiocarbon clock to properly account for the Flood and its impact on dates
for the post-Flood period to the present.For example, conventional radiocarbon dating gives an age of 48,000 years for a
coal bed deposited during the Flood, about 4,350 years ago. This could be explained if the 14C/12C ratio at the time of the
Flood was only 1/200th the ratio of the present world.If scientists assume the ratio is 200 times greater than it really was,
then their radiocarbon age estimate would be exaggerated by 43,650 years.10In reality, calculations (described above) have
led to estimates that the pre-Flood biosphere may have had more than 100 times the carbon-12 as the present earth. Using
this information, we may be able to calculate how much carbon-14 was actually on the early earth at the Flood. This, in turn,
would allow us to develop a proper interpretation of all carbon-14 dates.Once the research is completed, one of the many
exciting benefits is that it should be possible to begin more accurately dating any archeological artifact.

Radiocarbon dating of fossils compares the amount of radioactive carbon atoms (C-14) to regular carbon atoms (C-12).
Conventional dating methods assume the past ratio based on current levels. But what if these assumptions are wrong?
Lower Rate of Radiocarbon (C-14) Production
Cosmic rays bombard the earths atmosphere and produce neutrons. These neutrons collide with nitrogen atoms, changing
them into radioactive carbon atoms (C-14).Conventional dating assumes radiocarbon (C-14) production has remained
stable. But the earths magnetic field, which protects the earth from cosmic rays, was once several times stronger than it is
today. So we would expect much less radiocarbon to be produced in the past. That would result in much less C-14
compared to C-12.
Greater Volume of Regular Carbon (C-12)
Plants absorb carbon atoms during photosynthesis (mostly regular C-12 and little radioactive C-14). With a limited amount of
radiocarbon to go around, more plants would mean less radiocarbon per plant.Coventional dating assumes the volume of
plants and animals in the world has remained relatively stable. But the abundance of fossils indicates that the pre-Flood
worlds shallow seas and temperate climate supported much more plants and animals (containing mostly C-12) than today.
Lower Ratio of Radiocarbon (C-14) to Regular Carbon (C-12)
Radiocarbon begins to break down after plants and animals die. The amount of radiocarbon remaining determines the time
that has passed. Conventional dating assumes the ratio of C-12 to C-14 was the same in animals in the past. But if the ratio
was much lower in the animals in the past, then those animals would have much less radiocarbon to break down after they
died. This would result in much younger dates than conventional methods assume.
Measurable 14C in Fossilized Organic Materials: Conrming the Young Earth Creation-Flood Model
by Dr. Russell Humphreys, Dr. Andrew A. Snelling, Dr. John Baumgardner, and Dr. Steve Austin on February 9, 2011
Abstract
Given the short 14C half-life of 5730 years, organic
materials purportedly older than 250,000 years,
corresponding to 43.6 half-lives, should contain
absolutely no detectable 14C. (One gram of modern
carbon contains about 6 1010 14C atoms, and 43.6
half-lives should reduce that number by a factor of 7.3
10-14.) An astonishing discovery made over the past
20 years is that, almost without exception, when
tested by highly sensitive accelerator mass spectrometer (AMS) methods, organic samples from every portion of the
Phanerozoic record show detectable amounts of 14C!14C/C ratios from all but the youngest Phanerozoic samples appear to
be clustered in the range 0.10.5 pmc (percent modern carbon), regardless of geological age. A straightforward conclusion
that can be drawn from these observations is that all but the very youngest Phanerozoic organic material was buried
contemporaneously much less than 250,000 years ago. This is consistent with the creation account of a global Flood that
destroyed most of the air-breathing life on the planet in a single brief cataclysm only a few thousand years ago.
Keywords: radiocarbon, AMS 14C analysis, 14C dead, 14C background,14C contamination, uniformitarianism, young earth,
FloodThis paper was originally published in the Proceedings of the Fifth International Conference on Creationism, pp. 127
142 (2003) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh (www.csfpittsburgh.
org).

Introduction
Giem1 reviewed the literature and tabulated about 70 reported AMS measurements of 14C in organic materials from the
geologic record that, according to the conventional geologic time-scale, should be 14C dead. The surprising result is that
organic samples from every portion of the Phanerozoic record show detectable amounts of 14C. For the measurements
considered most reliable, the 14C/C ratios appear to fall in the range 0.10.5% of the modern 14C/C ratio (percent modern
carbon, or pmc). Giem demonstrates instrument error can be eliminated as an explanation on experimental grounds. He
shows contamination of the 14C-bearing fossil material in situ is unlikely but theoretically possible and is a testable
hypothesis, while contamination during sample preparation is a genuine problem but largely solved by two decades of
improvement in laboratory procedures. He concludes the 14C detected in these samples most likely is from the organisms
from which the samples are derived. Moreover, because most fossil carbon seems to have roughly the same 14C/C ratio,
Giem deems it plausible that all these organisms resided on earth at the same time.Anomalous 14C in fossil material actually
has been reported from the earliest days of radiocarbon dating. Whitelaw,2for example, surveyed all the dates reported in
the journal Radiocarbon up to 1970, and he commented that for all of the over 15,000 specimens reported, All such matter
is found datable within 50,000 years as published. The specimens included coal, oil, natural gas, and other allegedly
ancient material. The reason these anomalies were not taken seriously is because the older beta-decay counting technique
had difficulty distinguishing genuine low levels of 14C in the samples from background counts due to cosmic rays. The AMS
method, besides its inherently greater sensitivity, does not have this complication of spurious counts due to cosmic rays. In
retrospect, it is likely that many of the beta-counting analyses were indeed truly detecting intrinsic 14C.Measurable 14C in preFlood organic materials fossilized in Flood strata therefore appears to represent a powerful and testable confirmation of the
young earth Creation-Flood model. It was on this basis that Snelling3,4,5,6,7 analyzed the14C content of fossilized wood
conventionally regarded as 14C dead because it was derived from Tertiary, Mesozoic, and upper Paleozoic strata having
conventional radioisotope ages of 40 to 250 million years. All samples were analyzed using AMS technology by a reputable
commercial laboratory with some duplicate samples also tested by a specialist laboratory in a major research institute.
Measurable 14C was obtained in all cases. Values ranged from 7.581.11pmc for a lower Jurassic sample to 0.380.04pmc
for a middle Tertiary sample (corresponding to 14C ages of 20,7001200 to 44,700950 yearsBP, respectively). The 13C
values for the samples clustered around 25, as expected for organic carbon in plants and wood. The 14C measured in
these fossilized wood samples does not conform to a simple pattern, however, such as constant or decreasing with
increasing depth in the geologic record (increasing conventional age). On the contrary, the middle Tertiary sample yielded
the least 14C, while the Mesozoic and upper Paleozoic
samples did not contain similar 14C levels as might be
expected if these represent pre-Flood trees. The issue
then of how uniformly the 14C may have been
distributed in the pre-Flood world we concluded would
likely be an important one. Therefore, our RATE team
decided to undertake further 14C analyses on a new set
of samples to address this issue as well as to confirm
the remarkable 14C levels reported in the radiocarbon
literature for Phanerozoic material.
14
C Measured in Samples Conventionally Dated Older
Than 100,000 Years
Fig. 1. Uniformitarian age as a function of14C/C ratio in
percent modern carbon. The uniformitarian approach
for interpreting the 14C data assumes a constant 14C
production rate and a constant biospheric carbon
inventory extrapolated into the indenite past. It does
not account for the possibility of a recent global
catastrophe that removed a large quantity of carbon
from the biospheric inventory.Giem8 compiled a long list
of AMS measurements made on samples that, based on their conventional geological age, should be 14C dead. These
measurements were performed in many different laboratories around the world and reported in the standard peer-reviewed
literature, mostly in the journals Radiocarbon and Nuclear Instruments and Methods in Physics Research B. Despite the fact
that the conventional uniformitarian age for these samples is well beyond 100,000 years (in most cases it is tens to
hundreds of millions of years), it is helpful nonetheless to be able to translate 14C/C ratios into the equivalent
uniformitarian 14C age under the standard uniformitarian assumptions of an approximately constant 14C production rate and
an approximately constant biospheric carbon inventory, extrapolated into the indefinite past. This conversion is given by the
simple formula, pmc=1002t/5730, where t is the time in years. Applying this formula, one obtains values of 0.79pmc for t=
40,000 years, 0.24 for t=50,000 years, 0.070pmc for 60,000 years, 0.011pmc for 75,000 years, and .001pmc for 95,000
years, as shown in graphical form in Fig. 1.
Table 1 contains most of Giems9 data plus data from some more recent papers. Included in the list are a number of
samples from Precambrian, that is, what we consider non-organic pre-Flood settings. Most of the graphite samples
with 14C/C values below 0.05pmc are in this category.We display the published AMS values of Table 1 in histogram format in
Fig. 2. We have separated the source material into three categories, (1) those (mostly graphites) that are likely from
Precambrian geological settings and unlikely to contain biological carbon, (2) those that are clearly of biological affinity, and
(3) those (mostly marbles) whose biological connection is uncertain. We show categories (1) and (2) in Figs. 2(a) and 2(b),
respectively, and ignore for these purposes samples in category (3). Some caution is in order with respect to the sort of
comparison implicit in Table 1 and Fig. 2. In some cases the reported values have a background correction, typically on the
order of 0.07pmc, subtracted from the raw measured values, while in other cases such a correction has not been made. In
most cases, the graphite results do not include such background corrections since they are usually intended themselves to
serve as procedural blanks. Therefore, Fig. 2 is to be understood only as a low precision means for comparing these AMS
results.We draw several observations from this comparison, imprecise as it may be. First, the set of samples with biological
affinity display a mean value significantly different from those without such affinity. In terms of the standard geological
timescale, all these samples should be equally 14C dead. The samples with biological affinity display an unambiguously
higher mean than those without such affinity, 0.29 versus 0.06pmc. A second observation is that the variation in 14C content
for the biological samples is large. Although a peak in the distribution occurs at about 0.2pmc, the mean value is near 0.3
pmc with a standard deviation of 0.16pmc.
Table 1. AMS measurements on samples conventionally deemed 14C dead.

Item

14

C/C
(1 S.D.)

(pmc)

Material

Reference

0.71?*

Marble

Aerts-Bijma, Meijer,and van der Plicht10

0.650.04

Shell

Beukens11

0.610.12

Foraminifera

Arnold et al.12

0.600.04

Commercial graphite

Schmidt, Balsley, and Leach13

0.580.09

Foraminifera (Pyrgo murrhina)

Nadeau et al.14

0.540.04

Calcite

Beukens15

0.520.20

Shell (Spisula subtruncata)

Nadeau et al.16

0.520.04

Whale bone

Jull et al.17

0.510.08

Marble

Gulliksen and Thomsen18

10

0.50.1

Wood, 60Ka

Gillespie and Hedges19

11

0.460.03

Wood

Beukens20

12

0.460.03

Wood

Vogel, Nelson, and Southon21

13

0.440.13

Anthracite

Vogel, Nelson, and Southon22

14

0.420.03

Anthracite

Grootes et al.23

15

0.4010.084

Foraminifera (untreated)

Schleicher et al.24

16

0.400.07

Shell (Turitella communis)

Nadeau et al.25

17

0.3830.045

Wood (charred)

Snelling26

18

0.3580.033

Anthracite

Beukens27

19

0.350.03

Shell (Varicorbula gibba)

Nadeau et al.28

20

0.3420.037

Wood

Beukens29

21

0.340.11

Recycled graphite

Arnold et al.30

22

0.320.06

Foraminifera

Gulliksen and Thomsen31

23

0.3?

Coke

Terrasi et al.32

24

0.3?

Coal

Schleicher et al.33

25

0.260.02

Marble

Schmidt et al.34

26

0.23340.061

Carbon powder

McNichol et al.35

27

0.230.04

Foraminifera (mixed species avg.)

Nadeau et al.36

28

0.2110.018

Fossil wood

Beukens37

29

0.210.02

Marble

Schmidt et al.38

30

0.210.06

CO2

Grootes et al.39

31

0.200.35* (range)

Anthracite

Aerts-Bijma et al.40

32

0.200.04

Shell (Ostrea edulis)

Nadeau et al.41

33

0.200.04

Shell (Pecten opercularis)

Nadeau et al.42

34

0.20.1*

Calcite

Donahue et al.43

35

0.1980.060

Carbon powder

McNichol et al.44

36

0.180.05 (range?)

Marble

Van der Borg et al.45

37

0.180.03

Whale bone

Gulliksen and Thomsen46

38

0.180.03

Calcite

Gulliksen and Thomsen47

39

0.180.01**

Anthracite

Nelson et al.48

40

0.18?

Recycled graphite

Van der Borg et al.49

41

0.170.03

Natural gas

Gulliksen and Thomsen50

42

0.1660.008

Foraminifera (treated)

Schleicher et al.51

43

0.162?

Wood

Kirner et al.52

44

0.160.03

Wood

Gulliksen and Thomsen53

45

0.154?**

Anthracite coal

Schmidt et al.54

46

0.1520.025

Wood

Beukens55

47

0.1420.023

Anthracite

Vogel et al.56

48

0.1420.028

CaC2 from coal

Gurnkel57

49

0.140.02

Marble

Schleicher et al.58

50

0.130.03

Shell (Mytilus edulis)

Nadeau et al.59

51

0.1300.009

Graphite

Gurnkel60

52

0.1280.056

Graphite

Vogel et al.61

53

0.1250.060

Calcite

Vogel et al.62

54

0.120.03

Foraminifera (N. pachyderma)

Nadeau et al.63

55

0.1120.057

Bituminous coal

Kitagawa et al.64

56

0.10.01

Graphite (NBS)

Donahue et al.65

57

0.10.05

Petroleum, cracked

Gillespie and Hedges66

58

0.0980.009*

Marble

Schleicher et al.67

59

0.0920.006

Wood

Kirner, Taylor, and Southon68

60

0.090.18* (range)

Graphite powder

Aerts-Bijma et al.69

61

0.090.13* (range)

Fossil CO2 gas

Aerts-Bijma et al.70

62

0.0890.017

Graphite

Arnold et al.71

63

0.0810.019

Anthracite

Beukens72

64

0.08?

Natural graphite

Donahue et al.73

65

0.0800.028

Cararra marble

Nadeau et al.74

66

0.0770.005

Natural gas

Beukens75

67

0.0760.009

Marble

Beukens76

68

0.0740.014

Graphite powder

Kirner et al.77

69

0.07?

Graphite

Kretschmer et al.78

70

0.0680.028

Calcite (Icelandic double spar)

Nadeau et al.79

71

0.0680.009

Graphite (fresh surface)

Schmidt et al.80

72

0.060.11 (range)

Graphite (200 Ma)

Nakai et al.81

73

0.056?

Wood (selected data)

Kirner et al.82

74

0.050.01

Carbon

Wild et al.83

75

0.05?

Carbon-12 (mass sp.)

Schmidt et al.84

76

0.0450.012 (m0.06)

Graphite

Grootes et al.85

77

0.04?*

Graphite rod

Aerts-Bijma et al.86

78

0.040.01

Graphite (Finland)

Bonani et al.87

79

0.040.02

Graphite

Van der Borg88

80

0.040.02

Graphite (Ceylon)

Bird et al.89

81

0.0360.005

Graphite (air)

Schmidt et al.90

82

0.0330.013

Graphite

Kirner et al.91

83

0.030.015

Carbon powder

Schleicher et al.92

84

0.0300.007

Graphite (air redone)

Schmidt et al.93

85

0.0290.006

Graphite (argon redone)

Schmidt et al.94

86

0.0290.010

Graphite (fresh surface)

Schmidt et al.95

87

0.02?

Carbon powder

Pearson et al.96

88

0.0190.009

Graphite

Nadeau et al.97

89

0.0190.004

Graphite (argon)

Schmidt et al.98

90

0.0140.010

CaC2 (technical grade)

Beukens99

*Estimated from graph


**Lowest value of multiple dates
This large spread in 14C content invites an explanation. A third observation, although weaker that the first two, is that the
distribution of values for non-biogenic material displays a peak offset from zero. This may provide a hint that carbon never
cycled through living organismsin most cases locked away in Precambrian geological settingsmay actually contain a
low level of intrinsic 14C.
Coping with Paradigm Conflict
How do the various 14C laboratories around the world deal with the reality that they measure significant amounts of 14C, far
above the detection threshold of their instruments, in samples that should be 14C dead according to the standard geological
timescale? A good example can be found in a recent paper by Nadeau et al.100 entitled, Carbonate14C background: Does it
have multiple personalities? The authors are with the Leibnitz Laboratory at Christian-Albrechts University in Kiel, Germany.
Many of the samples they analyze are shells and foraminifera tests from sediment cores. It would very useful to them if they
could extend the range for which they could date such biological carbonate material from roughly 40,000 years ago
(according to their uniformitarian assumptions), corresponding to about 1pmc, toward the 0.002pmc limit of their AMS
instrument, corresponding to about 90,000 years in terms of uniformitarian assumptions. The reason they are presently
stuck at this 40,000-year barrier is that they consistently and reproducibly measure 14C levels approaching 1pmc in shells
and foraminifera from depths in the record where, according to the standard geological timescale, there should be no
detectable 14C.

Fig. 2. Distribution of 14C values for (a) non-biogenic samples and (b) biogenic samples from Table 1. Given their position in
the geological record, all these samples should contain no detectable 14C according to the standard geological time scale.
Their paper reports detailed studies they have carried out to attempt to understand the source of this 14C. They investigated
shells from a late Pleistocene coring site in northwestern Germany dated by U/Th methods at 120,000 years. The mean 14C
levels measured in the shells of six different species of mussels and snails varied from 0.1 to 0.5 pmc. In the case of one
species, Spisula subtruncata, measurements were made on both the outside and inside of the shell of a single individual
specimen. The average 14C value for the outside of the shell was 0.3pmc, while for the inside it was 0.67. At face value, this
suggests the 14C/C ratio more than doubled during the lifetime of this organism. Most of their foraminifera were from a
Pleistocene core from the tropical Atlantic off the northwest coast of Africa dated at 455,000 years. The foraminifera from
this core showed a range of 14C values from 0.16 to 0.4pmc with an average, taken over 115 separate measurements, of
0.23pmc. A benthic species of foraminifera from another core, chosen because of its thick shell and smooth surface in the
hope its contamination would be lower, actually had a higher average 14C level of 0.58pmc!The authors then performed a
number of experiments involving more aggressive pre-treatment of the samples to attempt to remove contamination. These
included progressive stepwise acid hydrolization of the carbonate samples to CO2 gas and 14C measurement of each of four
separate gas fractions. They found a detectable amount of surface contamination was present in the first fraction collected,
but it was not large enough to make the result from the final gas fraction significantly different from the average value. They
also leached samples in hydrochloric acid for two hours and cracked open the foraminifera shells to remove secondary
carbonate from inside, but these procedures did not significantly alter the measured 14C values.
The authors summarize their findings in the abstract of their paper as follows,
The resultsshow a species-specific contamination that reproduces over several individual shells and foraminifera from
several sediment cores. Different cleaning attempts have proven ineffective, and even stronger measures such as
progressive hydrolization or leaching of the samples prior to routine preparation, did not give any indication of the source of
contamination.
In their conclusion they state,
The apparent ages of biogenic samples seem species related and can be reproduced measuring different individuals for
larger shells or even different sediment cores for foraminifera. Although tests showed some surface contamination, it was
not possible to reach lower 14C levels through cleaning, indicating the contamination to be intrinsic to the sample.
They continue,
So far, no theory explaining the results has survived all the tests. No connection between surface structure and apparent
ages could be established.
The measurements reported in this paper obviously represent serious anomalies relative to what should be expected in the
uniformitarian framework. There is a clear conflict between the measured levels of 14C in these samples and the dates
assigned to the geological setting by other radioisotope methods. The measured 14C levels, however, are far above

instrument threshold and also appear to be far above contamination levels arising from sample processing. Moreover, the
huge difference in 14C levels among species co-existing in the same physical sample violates the assumption that organisms
living together in the same environment should share a common 14C/C ratio. The position the authors take in the face of
these conflicts is that this 14C, which should not be present according to their framework, represents contamination for
which they currently have no explanation. On the other hand, in terms of the framework of a young earth and a recent global
Flood, these measurements provide important clues these organisms are much younger than the standard geological
timescale would lead one to suspect.This same approach of treating measurable and reproducible 14C values in samples
that ought to be 14C dead, given their position in the geological record, as contamination is found throughout the current
literature. Bird,101 for example, freely acknowledge contamination in old samples leads to a radiocarbon barrier:
Detecting sample contamination and verifying the reliability of the ages produced also becomes more difficult as the age of
the sample increases. In practice this means that many laboratories will only quote 14C ages to about 40ka BP (thousands
of 14C years before present), with ages greater than this generally considered to be infinite, or indistinguishable from
procedural blanks. The so-called radiocarbon barrier and the difficulty of ensuring that ages are reliable at <1% modern
carbon levels has limited research in many disciplines.This statement is in the context of a high precision AMS facility the
authors use, capable of measuring 14C levels in the range of <<0.01pmc.In their paper they describe a strategy for
eliminating various types of genuine contamination commonly associated with charcoal samples. A main component of this
strategy is a stepped combustion procedure in which the sample is oxidized to CO2 in a stepwise manner, at temperatures of
330C, 630C, and 850C, with the resulting CO2 fractions analyzed separately using AMS. Oxidation of most of any
surficial contamination generally occurs at the lowest temperature, and the 14C level of the highest temperature fraction is
generally considered the one representing the least contaminated portion of the sample. The variation among the three
fractions is considered a general indicator of the overall degree of contamination. They apply this approach to analysis of
charcoal from one of the early sites of human occupation in Australia.Included in their paper is considerable discussion of
what is known as a procedural blank, or a sample that represents effectively infinite 14C age. For this they use what they
refer to as radiocarbon-dead graphite from Ceylon. They apply their stepped combustion procedure, using only the highest
temperature fraction, on 14 such graphite samples to get a composite value of 0.040.02 pmc for this background material.
They note that a special pre-treatment they use for charcoal samples applied to four of the 14 samples yielded results
indistinguishable from the other 10 graphite samples that had no pre-treatment. They further note that sample size variation
between 0.1 and 2.2mg among the 14 samples also made no difference in the results. From this they acknowledge, the
few 14C atoms observed may already be present in the Ceylon graphite itself. Indeed, they offer no explanation for the fact
that this graphite displays 14C levels well above the detection threshold of their AMS system other than it might be inherent
to the graphite itself.Measuring notable levels of 14C in samples intended as procedural blanks or background samples is a
phenomenon that has persisted from the earliest days of AMS down to the present time. For example, Vogel, Nelson, and
Southon102 describe their thorough investigation of the potential sources and their various contributions to the 14C
background in their AMS system. The material they used for the blank in their study was anthracite coal from a deep mine in
Pennsylvania. An important part of their investigation was variation of the sample size of the blank by a factor of 2000, from
10mg to 20mg. They found that samples 500mg and larger displayed a 14C concentration of 0.440.13pmc, independent of
sample size, implying this 14C was intrinsic to the anthracite material itself. For samples smaller than 500mg, the
measured 14C could be explained in terms of this intrinsic 14C, plus contamination by a constant amount of modern carbon
that seemed to be present regardless of sample size. After many careful experiments, the authors concluded that the main
source of this latter contamination was atmospheric CO 2 adsorbed within the porous Vicor glass used to encapsulate the
coal sample in its combustion to CO2 at 900C. Another source of smaller magnitude was CO2 and CO adsorbed on the
walls of the graphitization apparatus retained from reduction of earlier samples. It was found that filling the apparatus with
water vapor at low pressure and then evacuating the apparatus before the next graphitization mostly eliminated this memory
effect. Relative to these two sources, measurements showed that storage and handling of the samples, contamination of the
copper oxide used in combustion, and contamination of the iron oxide powder used in the graphitization were effectively
negligible. And when the sample size was greater than 500mg, the intrinsic 14C in the coal swamped all the sources of
real 14C contamination. Rather than deal with the issue of the nature of the 14C intrinsic to the anthracite itself, the authors
merely refer to it as contamination of the sample in situ, not [to be] discussed further.
As it became widely appreciated that many high carbon samples, which ought to be 14C dead given their position in the
geological record, had in fact 14C levels far above AMS machine thresholds, the approach was simply to search for specific
materials that had as low a 14C background level as possible.For example, Beukens,103 at the IsoTrace Laboratory at the
University of Toronto, describes measurements on two samples that, from his experience at that time, displayed
exceptionally low background 14C levels. He reports 0.0770.005pmc from a sample of industrial CO2obtained by
combustion of natural gas and 0.0760.009pmc from Italian Carrara marble. Previously for his blank material he had used
an optical grade calcite (Iceland spar) for which he measured a 14C level of 0.15 to 0.13pmc. He emphasizes that the pretreatment, combustion, and hydrolysis techniques applied to these new samples were identical to those normally applied to
samples submitted for analysis to his laboratory and these techniques had not changed appreciably in the previous five
years. He states,The lower 14C levels in these [more recent] measurements should therefore be attributed entirely to the
lower intrinsic14C contamination of these samples and not to changes in sample preparation or analysis techniques.
Note that he indeed considers the 14C in all these materials to be intrinsic, but he has to call it contamination. In his
search for even better procedural blanks, he tested two standard blank materials, a calcite and an anthracite coal, used by
the Geological Survey of Canada in their beta decay counting 14C laboratory. These yielded 14C levels of 0.540.04pmc for
the calcite and 0.360.03pmc for the coal. Beukens noted with moderate alarm that the background corrections being made
by many decay-counting radiocarbon dating facilities that had not checked the intrinsic 14C content of their procedural blanks
by AMS methods were probably quoting ages systematically older than the actual ages. His AMS analysis of the samples
from the Geological Survey of Canada clearly shows these samples are not 14C-free since these levels were markedly
higher than those from his own natural gas and marble blanks.AMS analyses reveal carbon from fossil remains of living
organisms, regardless of their position in the geological record, consistently contains 14C levels far in excess of the AMS
machine threshold, even when extreme pre-treatment methods are applied. Experiments in which the sample size is varied
argue compellingly that the 14C is intrinsic to the fossil material and not a result of handling or pre-treatment. These
conclusions continue to be confirmed in the very latest peer-reviewed papers. Moreover, even non-organic carbon samples
appear consistently to yield 14C levels well above machine threshold. Graphite samples formed under metamorphic and
reducing conditions in Precambrian limestone environments commonly display 14C values on the order of 0.05pmc. Most
AMS laboratories are now using such Precambrian graphite for their procedural blanks. A good question is what possibly
could be the source of the 14C in this material? We conclude that the possibility this 14C is primordial is a reasonable one.
Finding 14C in diamond formed in the earths mantle would provide support for such a conclusion. Establishing that nonorganic carbon from the mantle and from Precambrian crustal settings consistently contains inherent 14C well above the

AMS detection threshold would, of course, argue the earth itself is less than 100,000 years old, which is orders of magnitude
younger than the 4.56Ga currently believed by the uniformitarian community.
Results of RATE 14C AMS Analyses
Table 2 summarizes the results from ten coal samples prepared by our RATE team and analyzed by one of the foremost
AMS laboratories in the world. These measurements were performed using the laboratorys high precision procedures
which involved four runs on each sample, the results of which were combined as a weighted average and then reduced by
0.0770.005pmc to account for a standard background of contamination believed to be introduced by sample processing.
This standard background value is obtained by measuring the 14C in a purified natural gas. Subtraction of this background
value is justified by the assumption that it must represent contamination. Fig. 3 displays these AMS analysis results in
histogram format.
Table 2. Results of AMS 14C analysis of ten RATE coal samples.
Sample

Coal Seam Name

State

County

Geological Interval

14

DECS-1

Bottom

Texas

Freestone

Eocene

0.300.03

DECS-11 Beulah

North Dakota

Mercer

Eocene

0.200.02

DECS-25 Pust

Montana

Richland

Eocene

0.270.02

DECS-15 Lower Sunnyside

Utah

Carbon

Cretaceous

0.350.03

DECS-16 Blind Canyon

Utah

Emery

Cretaceous

0.100.03

DECS-28 Green

Arizona

Navajo

Cretaceous

0.180.02

DECS-18 Kentucky #9

Kentucky

Union

Pennsylvanian

0.460.03

DECS-21 Lykens Valley #2

Pennsylvania

Columbia

Pennsylvanian

0.130.02

DECS-23 Pittsburgh

Pennsylvania

Washington

Pennsylvanian

0.190.02

DECS-24 Illinois #6

Illinois

Macoupin

Pennsylvanian

0.290.03

C/C (pmc)

Details of RATE Sample Selection and Analysis


The ten samples in Table 2 were obtained from the U.S. Department of Energy Coal Sample Bank maintained at Penn
State University. The coals in this bank are intended to be representative of the economically important coalfields of the
United States. The original samples were collected in 400-pound quantities from recently exposed areas of active mines,
where they were placed in 30-gallon steel drums with high-density gaskets and purged with argon. As soon as feasible after
collection, these large samples were processed to obtain representative 300g samples with 0.85mm particle size (20
mesh). These smaller 300g samples were sealed under argon in foil multilaminate bags and have since been kept in
refrigerated storage at 3C. We selected ten of the 33 coals available with an effort to obtain good representation
geographically as well as with respect to depth in the geological record. Our ten samples include three Eocene, three
Cretaceous, and four Pennsylvanian coals.
Fig. 3. Histogram representation of AMS 14C analysis of
ten coal samples undertaken by RATE 14C research
project.
The 14C analysis at the AMS laboratory we selected
involves first processing the coal samples to make
graphite targets and then counting the relative numbers
of atoms from the different carbon isotopes in the
accelerator mass spectrometer system. The accelerator
generates an intense ion beam that ionizes the graphite
on the target, while the mass spectrometer uses electric
and magnetic fields to separate different atomic species
by mass and charge and counts the numbers of triply
ionized 14C, 13C, and 12C atoms. The sample processing
consists of three steps:combustion, acetylene
synthesis, and graphitization.The coal samples are first
combusted to CO2 and then converted to acetylene
using a lithium carbide synthesis process. The
acetylene is then dissociated in a high voltage AC
electrical discharge to produce a circular disk of
graphite on spherical aluminum pellets that represent the targets for the AMS system. Four separate targets are produced
for each sample. Every target is analyzed in a separate AMS run with two modern carbon standards (NBS I oxalic acid).
Each target is then analyzed on 16 different spots (organized on two concentric circles). The advantage of this procedure
over a single high precision measurement is that a variance check (typically a T-test) can be performed for the 16 spots on
each target. If an individual target fails this variance test, it is rejected. While this has advantages for any kind of sample, it is
particularly useful for samples with very low 14C levels because they are especially sensitive to contamination. While great
care is taken to prevent target contamination after the graphitization step, it nevertheless can happen. Any contaminated
spot or any contaminated target would bias the average. This variance test attempts to identify and eliminate this source of
error.
Table 3 gives the measurements in pmc from the four separate targets for our ten coal samples. The numbers in
parentheses are the percent errors, calculated from the 14C count rate of the sample and the two NBS standards and from
the transmission of errors in the 12C and 13C current measurements of the sample and two standards. The composite results
in Table 2 represent the weighted averages of these numbers in Table 3 and the subtraction of a standard background of
0.0770.005pmc.
Table 3. Detailed AMS 14C measurements for 10 RATE coal samples in pmc.

The background standard of this AMS


laboratory is CO2 from purified natural gas
that provides their background level of
DECS-1 0.398 (12.0%) 0.355 (13.2%) 0.346 (15.1%) 0.346 (15.1%)
0.0770.005pmc. This same laboratory
DECS-11 0.237 (18.2%) 0.303 (14.8%) 0.292 (17.8%) 0.294 (17.2%)
obtains values of 0.0760.009pmc and
0.0710.009pmc, respectively, for Carrara
DECS-25 0.342 (13.3%) 0.359 (15.3%) 0.352 (14.2%) 0.328 (14.8%)
Marble (IAEA Standard Radiocarbon
Reference Material C1) and optical-grade
DECS-15 0.416 (13.1%) 0.465 (12.2%) 0.467 (12.2%) 0.377 (13.6%)
calcite from Island spar. They claim this is
one of the lowest background levels
DECS-16 0.184 (25.0%) 0.233 (21.8%) 0.141 (38.4%) 0.163 (34.0%)
quoted among AMS labs, and they
attribute this low background to their
DECS-28 0.203 (18.3%) 0.379 (14.5%) 0.204 (21.2%) 0.204 (21.2%)
special graphitization technique. They
DECS-18 0.533 (11.8%) 0.539 (11.4%) 0.492 (11.6%) 0.589 (10.0%)
emphasize backgrounds this low cannot
be realized with any statistical significance
DECS-21 0.183 (22.0%) 0.194 (20.0%) 0.230 (18.2%) 0.250 (18.0%)
through only one or two measurements,
but many measurements are required to
DECS-23 0.225 (18.1%) 0.266 (13.8%) 0.246 (18.7%) 0.349 (13.2%)
obtain
a
robust
determination.The
laboratory has carefully studied the
DECS-24 0.334 (19.7%) 0.462 (17.5%) 0.444 (13.4%) 0.252 (25.8%)
sources of error within its AMS hardware,
and regular tests are performed to ensure these remain small. According to these studies, errors in the spectrometer are
very low and usually below the detection limit since the spectrometer is energy dispersive and identifies the ion species by
energy loss. The detector electronic noise, the mass spectrometric inferences (the E/q and mE/q 2 ambiguities), and the
cross contamination all contribute less than 0.0004pmc to the background. Ion source contamination as a result of previous
samples (ion source memory) is a finite contribution because 5080% of all sputtered carbon atoms are not extracted as
carbon ions and are therefore dumped into the ion source region. To limit this ion source memory effect, the ion source is
cleaned every two weeks and critical parts are thrown away. This keeps the ion source contamination at approximately
0.0025pmc for the duration of a two-week run. Regular spot checks of these contributions are performed with a zonerefined, reactor-grade graphite sample (measuring 14C/12C ratios), and blank aluminum target pellets (measuring 14C only).
The laboratory claims most of their quoted system background arises from sample processing. This processing involves
combustion (or hydrolysis in the case of carbonate samples), acetylene synthesis, and graphitization. Yet careful and
repeated analysis of their methods over more than 15 years have convinced them that very little contamination is associated
with the combustion or hydrolysis procedures and almost none with their electrical dissociation graphitization process. By
elimination they conclude that the acetylene synthesis must contribute almost all of the system background. But they can
provide little tangible evidence it actually does. Our assessment from the information we have is that the system background
arises primarily from 14C intrinsic to the background standards themselves. The values we report in Table 2 and Fig. 3
nevertheless include the subtraction of the laboratorys standard background. In any case, the measured 14C/C values are
notably above their background value.
Making Sense of the 14C Data
How does one make sense of these 14C measurements that yield a uniformitarian ages of 40,00060,000 years for organic
samples, such as our coal samples, that have uniformitarian ages of 40350 million years based on long half-life isotope
methods applied to surrounding host rocks? Clearly there is an inconsistency. Our hypothesis is that the source of the
discrepancy is the interpretational framework that underlies these methods. Could the proposition, articulated 180 years ago
by Charles Lyell, that the present is the key to the past be suspect? Could the standard practice employed all these years
by earth scientists and others of extrapolating the processes and rates observed in todays world into the indefinite past not
be reliable after all? As authors of this paper we are convinced that there is abundant observational evidence in the
geological record that the earth has experienced a global tectonic catastrophe of immense magnitude that is responsible for
most of the Phanerozoic geological record. We are persuaded it is impossible any longer to claim that geological processes
and rates observable today can account for the majority of the Phanerozoic sedimentary record. To us the evidence is
overwhelming that global scale processes operating at rates much higher than any observable on earth today are
responsible for this geological change.104,105,106,107 Not only are the 14C data at odds with the standard geological
timescale, but the general character of the sedimentary and tectonic record is as well. We realize for many such a view of
the geological data is new, or at least controversial. For those new to this possibility we urge reading of some of our papers
on this topic (for example, Austin108, Baumgardner109,110,111). We are convinced that not only do the observations
strongly support this interpretation of the geological record, but the theoretical framework also now exists to explain it
(Baumgardner112,113,114). Our approach for making sense of these 14C data, therefore, is to do so in the light of a major
discontinuity in earth history in its not so distant past, an event we correlate with the Flood described in the in many ancient
documents.
What was the Pre-Flood 14C Level?
What sorts of 14C/C ratios might we expect to find today in organic remains of plants and animals buried in a single global
cataclysm correlated with all but the latter part of the Phanerozoic geological record (that is, Cambrian to middle-upper
Cenozoic)? Such a cataclysm would have buried a huge amount of carbon from living organisms to form todays coal, oil,
and oil shale, probably most of the natural gas, and some fraction of todays fossiliferous limestone. Estimates for the
amount of carbon in this inventory are at least a factor of 100 greater than what currently resides in the
biosphere.115,116,117 This implies the biosphere just prior to the cataclysm would have had at least 100 times the total
carbon relative to our world today. Living plants and animals would have contained most of this biospheric carbon, with only
a tiny fraction of the total in the atmosphere. The vast majority of this carbon would have been 12C, since even today only
about one carbon atom in a trillion is 14C.To estimate the pre-cataclysm 14C/C ratio we of course require an estimate for the
amount of 14C. As a starting point we might assume the total amount was similar to what exists in todays world. If that were
the case, and this 14C were distributed uniformly, the resulting 14C/C ratio would be about 1/100 of todays level, or about 1
pmc. This follows from the fact that 100 times more carbon in the biosphere would dilute the available 14C and cause the
biospheric14C/C ratio to be 100 times smaller than today. But this value of 1pmc is probably an upper limit because there are
reasons to suspect the total amount of 14C just prior to the cataclysm was less, possibly much less, than exists today. Two
important issues come into play here in regard to the amount of pre-Flood 14Cnamely, the initial amount of 14C after
creation and the 14C production rate in the span of time between creation and the Flood catastrophe. We have seen already
there are hints of primordial 14C in non-biogenic Precambrian materials at levels on the order of 0.05pmc. This provides a
clue that the 14C/C ratio in everything containing carbon just after creation might have been on the order of 0.1pmc. But it is
Sample

Target 1

Target 2

Target 3

Target 4

also likely 14C was added to the biosphere between creation and the Flood. The origin of 14C in todays world is by cosmic
ray particles in the upper atmosphere changing a proton in the nucleus of a 14N atom into a neutron to yield a 14C atom. Just
what the 14C production rate prior to the cataclysm might have been is not easily constrained. It could well have been lower
than today if the earths magnetic field strength were higher and resulting cosmic ray flux lower. But perhaps it was not. In
any case, given the 5,730-year half-life of 14C, it is almost certain the less than 2,000 year interval between creation and the
Flood was insufficient for 14C to have reached an equilibrium level in the biosphere. If the 14C production rate itself was
roughly constant, then the 14C/C ratio in the atmosphere would have been a steadily increasing function of time across this
interval. Hence, we conclude the pre-Flood 14C/C ratios were likely no greater than 1pmc but also highly variable, especially
in the case of plants, depending on when during the interval they generated their biomass.
In addition to the preceding considerations, we must also account for the 14C decay that has occurred since the cataclysm.
Assuming a constant 14C half-life of 5,730 years, the 14C/C ratio in organic material buried, say, 5,000 years ago would be
reduced by an additional factor of 0.55. When we combine all these factors, we conclude it is not at all surprising organic
materials buried in the cataclysm should display the roughly 0.050.5pmc we actually observe. We note that when these
considerations are included, especially the larger pre-cataclysm carbon inventory, a 14C/C value of 0.24pmc, for example, is
consistent with an actual age of 5,000 years. By contrast, when these considerations are not taken into account, the
uniformitarian formula, pmc=1002t/5730, displayed in graphical form in Figure 1, yields an age of 50,000 years. Yet
in either case, the 14C ages are still typically orders of magnitude less than those provided by the long half-life radioisotope
methods.In this context it is useful to note that 14C/C levels must have increased dramatically and rapidly just after the
cataclysm, assuming near modern rates of 14C production in the upper atmosphere, due to the roughly hundredfold
reduction in the amount of carbon in the biospheric inventory. The large variation in 14C levels between species as well as
from the outside to the inside of a single shell as reported by Nadeau et al.118 indeed seems to suggest significant spatial
and temporal variations in this dynamic period during which the planet was recovering from the cataclysm.
Effect of Accelerated Decay on Pre-Flood 14COther RATE projects are building a compelling case that episodes of
accelerated nuclear decay must have accompanied the creation of the earth as well as the Global Flood.119,120,121 We
believe several billions of years worth of cumulative decay at todays rates must have occurred for isotopes such as 238U
during the creation of the physical earth, and we now suspect a significant amount of such decay likely also occurred during
the Flood cataclysm. An important issue then arises as to how an episode of accelerated decay during the Flood might have
affected a short half-life isotope like 14C. The fact that significant amounts of 14C are measured routinely in fossil material
from organisms alive before the cataclysm argues persuasively that only a modest amount of accelerated 14C decay
occurred during the cataclysm itself. This suggests the possibility that the fraction of unstable atoms that decayed during the
acceleration episode for all of the unstable isotopes might have been roughly the same. If the fraction were exactly the
same, this would mean that the acceleration in years for each isotope was proportional to the isotopes half-life. In this case,
if 40K, for example, underwent 400Ma of decay during the Flood relative to a present half-life of 1250Ma, then14C would
have undergone (400/1250)*5730 years=1,834 years of decay during the Flood. This amount of decay represents 12
(1834/5730)
=20% reduction in 14C as a result of accelerated decay. This is well within the uncertainty of the level of 14C in the
pre-Flood world so it has little impact on the larger issues discussed in this paper.
Discussion
The initial vision that high precision AMS methods should make it possible to extend 14C dating of organic materials back as
far as 90,000 years has not been realized. The reason seems to be clear. Few, if any, organic samples can be found
containing so little 14C! This includes samples uniformitarians presume to be millions, even hundreds of millions, of years
old. At face value, this ought to indicate immediately, entirely apart from any consideration of a Flood catastrophe, that life
has existed on earth for less than 90,000 years. Although repeated analyses over the years have continued to confirm
the 14C is an intrinsic component of the sample material being tested, such 14C is still referred to as contamination if it is
derived from any part of the geological record deemed older than about 100,000 years. To admit otherwise would fatally
undermine the uniformitarian framework. For the creationist, however, this body of data represents obvious support for the
recent creation of life on earth. Significantly, the research and data underpinning the conclusion that 14C exists in fossil
material from all portions of the Phanerozoic record are already established in the standard peer-reviewed literature. And the
work has been performed largely by uniformitarians who hold no bias whatever in favor of this outcome. The evidence is
now so compelling that additional AMS determinations by creationists on samples from deep within the Phanerozoic record
can only make the case marginally stronger than it already is.Indeed, the AMS results for our ten coal samples, as
summarized in Table 2 and Fig. 3, fall nicely within the range for similar analyses reported in the radiocarbon literature, as
presented in Table 1 and Fig. 2(b). Not only are the mean values of the two data sets almost the same, but the variances are
also similar. Moreover, when we average the results from our coal samples over geological interval, we obtain mean values
of 0.26pmc for Eocene, 0.21 for Cretaceous, and 0.27 for Pennsylvanian that are remarkably similar to one another. These
results, limited as they are, indicate little difference in 14C level as a function of position in the geological record. This is
consistent with the young-earth view that the entire fossil record up to somewhere within the middle-upper Cenozoic is the
product of a single recent global catastrophe. On the other hand, an explanation for the notable variation in 14C level among
the ten samples is not obvious. One possibility is that the 14C production rate between creation and the Flood was
sufficiently high that the14C levels in the pre-Flood biosphere increased from, say, 0.1pmc at creation to perhaps as much as
1pmc just prior to the Flood. Plant material that grew early during this period and survived until the Flood would then contain
low levels of 14C, while plant material produced by photosynthetic processes just prior to the cataclysm would contain much
higher values. This situation would prevail across all ecological zones on the planet, and so the large variations in 14C levels
would appear within all stratigraphic zones that were a product of the Flood.Moreover, in contrast to the uniformitarian
outlook that 14C in samples older than late Pleistocene must be contamination and therefore is of little or no scientific
interest, such 14C for the creationist potentially contains vitally important clues to the character of the pre-Flood world. The
potential scientific value of these 14C data in our opinion merits a serious creationist research effort to measure the 14C
content in fossil organic material from a wide variety of pre-Flood environments, both marine and terrestrial. Systematic
variations in 14C levels, should they be discovered, conceivably could provide important constraints on the time history of 14C
levels and 14C production, the pattern of atmospheric circulation, the pattern of oceanic circulation, and the carbon cycle in
general in the pre-Flood world.Furthermore, a careful study of the 14C content of carbon that has not been cycled through
living organisms, especially carbonates, graphites, and diamonds from environments believed to pre-date life on earth,
could potentially place very strong constraints on the age of the earth itself. The data already present in the peer-reviewed
radiocarbon literature suggests there is indeed intrinsic 14C in such materials that cannot be attributed to contamination. If
this conclusion proves robust, these reported 14C levels then place a hard limit on the age of the earth of less than 100,000
years, even when viewed from a uniformitarian perspective. We believe a creationist research initiative focused on this issue
deserves urgent support.
Conclusion

The careful investigations performed by scores of researchers in more than a dozen AMS facilities in several countries over
the past 20 years to attempt to identify and eliminate sources of contamination in AMS 14C analyses have, as a by-product,
served to establish beyond any reasonable doubt the existence of intrinsic 14C in remains of living organisms from all
portions of the Phanerozoic record. Such samples, with ages from 1500Ma as determined by other radioisotope methods
applied to their geological context, consistently display 14C levels that are far above the AMS machine threshold, reliably
reproducible, and typically in the range of 0.10.5pmc. But such levels of intrinsic 14C represent a momentous difficulty for
uniformitarianism. A mere 250,000 years corresponds to 43.6 half-lives for 14C. One gram of modern carbon contains about
61010 14C atoms, and 43.6 half-lives worth of decay reduces that number by a factor of 71014. Not a single atom of 14C
should remain in a carbon sample of this size after 250,000 years (not to mention one million or 50 million or 250 million
years). A glaring (thousand-fold) inconsistency that can no longer be ignored in the scientific world exists between the AMSdetermined 14C levels and the corresponding rock ages provided by 238U, 87Rb, and 40K techniques. We believe the chief
source for this inconsistency to be the uniformitarian assumption of time-invariant decay rates. Other research reported by
our RATE group also supports this conclusion.122,123,124 Regardless of the source of the inconsistency, the fact that 14C,
with a half-life of only 5,730 years, is readily detected throughout the Phanerozoic part of the geological record argues the
half billion years of time uniformitarians assign to this portion of earth history is likely incorrect. The relatively narrow range
of 14C/C ratios further suggests the Phanerozoic organisms may all have been contemporaries and that they perished
simultaneously in the not so distant past. Finally, we note there are hints that 14C currently exists in carbon from
environments sealed from biospheric interchange since very early in the earth history. We therefore conclude the 14C
evidence provides significant support for a model of earths past involving a recent global Flood cataclysm and possibly also
for a young age for the earth itself.
Radiocarbon Ages for Fossil Ammonites and Wood in Cretaceous Strata near Redding, California
by Dr. Andrew A. Snelling on December 10, 2008
Abstract
Fossil ammonites from lower Cretaceous mudstones in northern California, which are supposedly 112120 million years old
and biostratigraphic index fossils, were sampled along with fossil wood buried with them. Fragments of two fossil ammonite
shells and four pieces of fossil wood yielded easily measurable radiocarbon (14C) equivalent to apparent 14C ages of
between 36,400350 and 48,710930 years for the ammonites, and between 32,780230 and 42,390510 years for the
wood. Any contamination with modern 14C due to the sample environment and handling was eliminated by the laboratorys
severe pre-treatment procedure. Any alleged contamination due to sample combustion or AMS instrument background was
more than compensated for by the laboratory background of 0.077 pMC already having been subtracted from the reported
results. The ammonite shells could not have been contaminated in the ground by replacement with modern carbonate 14C
either, because they yielded almost identical 14C apparent ages as the wood buried and fossilized with them. It was
concluded that the measured 14C is in situ radiocarbon intrinsic to the ammonites and wood when they were buried and
fossilized. So once past conditions in the atmosphere and biosphere are taken into account, their true ages are consistent
with their burial during the Global Flood only about 4,300 years ago, when the ocean waters washed sediments and
ammonites onto the continents.
Keywords: Fossil ammonites, Fossil wood, Lower Cretaceous, Radiocarbon, Apparent ages, Contamination, Pre-treatment
procedure, Laboratory background, In situ radiocarbon, True ages, Flood
Introduction
Measurable 14C (radiocarbon) has been detected in fossils from the earliest days of radiocarbon dating. In many instances,
according to their supposed uniformitarian ages those fossils should be completely 14C-dead, that is, all 14C originally in them
should have decayed, so they should not have any 14C left in them.When these data are put in perspective, their deadly
significance to the uniformitarian timescale is readily apparent. 14C has a half-life of 5,730 years. If an organism when it was
buried and fossilized contained the level of 14C currently in plants and animals, then after one million years, corresponding to
174.5 14C half-lives, the fraction of the original 14C remaining would be 3 10-53. However, a mass of 14C equal to the entire
mass of the earth (6 1024 kg) contains only about 3 1050 14C atoms (Baumgardner 2005). Thus, not a single atom of 14C
formed even 1 million years ago anywhere in or on the earth should conceivably still exist. Therefore, there should be
absolutely no measurable 14C able to be detected in fossils claimed to be a million or more years old.Whitelaw (1970)
surveyed all the 14C dates reported in the journal Radiocarbon up to 1970 and found that for more than 15,000 samples, all
such matter was datable within 50,000 years as published. These samples included coal, oil, natural gas, and other
allegedly very ancient (> one million years old) material. The scientific community never took these anomalies seriously,
because these measurements were obtained using the -decay counting technique, by which it was difficult to distinguish
between the decay of 14C atoms and background cosmic rays. Thus, for samples such as these whose location in the
geological record mandated they were so old they should be 14C dead, the measurable 14C levels detected were simply
attributed to measurement errors that had incorporated the cosmic ray background.The accelerator mass spectrometer
(AMS) method was developed in the early 1980s. It superseded the -decay counting technique, because it counts 14C
atoms directly and so is not compromised by background cosmic rays. Snelling (1997, 1998, 1999, 2000a, b, 2008) had
samples of fossilized woods, from Tertiary, Mesozoic and upper Paleozoic strata conventionally dated as 30 to 250 million
years old, 14C tested using the AMS method at two laboratories. In all cases 14C levels were obtained that are well above the
AMS detection threshold. Values ranged from 7.581.11 pMC (percent modern carbon, that is, percent of the modern
atmospheric 14C/C ratio) for a lower Jurassic sample to 0.380.04 pMC for a middle Tertiary sample. This range in the 14C/C
ratio implies radiocarbon ages of between 20,7001200 and 44,700950 years, respectively, assuming the modern
atmospheric 14C/C ratio existed when these trees were alive. Such 14C contents in these fossilized woods limit their ages to
only thousands of years, contrary to their conventional uniformitarian ages of tens to hundreds of millions of years.
Giem (2001) tabulated about seventy AMS measurements, published between 1984 and 1998 in the standard radiocarbon
literature, all of organic materials that yielded significant levels of 14C when according to their conventional uniformitarian
ages they should be entirely 14C dead. These organic materials included not only fossilized wood, but natural graphite, coal,
natural gas, oil, fossilized shells and bones, and even marble, from every portion of the Phanerozoic (Cambrian to Recent)
geologic record. All contained detectable 14C levels well above the AMS threshold, with the 14C/C ratios from those
measurements considered the most reliable falling in the 0.1-0.5 pMC range. Giem (2001) argued that instrument error
could be eliminated as an explanation on experimental grounds. He further showed that contamination of the 14C-bearing
fossil material in situ was unlikely, but theoretically possible, and was a testable hypothesis. While contamination during
sample preparation was a genuine problem, the conventional literature showed that it could be reduced to low levels by
proper laboratory procedures. He concluded, therefore, that the detected 14C in these samples most likely originates from
the original organisms themselves.Baumgardner et al. (2003) and Baumgardner (2005) selected ten coal samples from the
U. S. Department of Energy Coal Sample Bank maintained at Pennsylvania State University. These ten samples provided

good representation of U.S. coals geographically as well as with respect to depth in the geologic record. Three were Eocene
coals, three Cretaceous, and four Pennsylvanian, spanning 40300 million years on the conventional uniformitarian
timescale. The AMS 14C/C results ranged from 0.100.03 pMC in a Utah Cretaceous coal to 0.460.03 pMC in a Kentucky
Pennsylvanian coal (with the laboratorys standard background of 0.0770.005 pMC subtracted). Of added significance
were the averages of the results for each geologic time intervala mean value of 0.26 pMC for the Eocene coals, 0.21 pMC
for the Cretaceous coals, and 0.27 pMC for the Pennsylvanian coals. These data were interpreted as implying that all these
coal beds are, in reality, the same 48,00050,000 radiocarbon years age, which is consistent with their deposition in the
Flood year of the creation timescale.Baumgardner (2005) also had twelve diamonds tested for 14C by the AMS method. Five
of these diamonds were from underground mining of volcanic kimberlite pipes at Kimberley (South Africa), and at Orapa and
Letlhakane (Botswana), while the sixth diamond was from the Kankan alluvial deposit in Guinea (West Africa), and the other
six diamonds were from alluvial deposits in Namibia. In conventional terms, all these diamonds are regarded as hundreds of
millions to billions of years old, having been sourced from the earths mantle. The AMS 14C/C results ranged from 0.100.03
pMC for the Kimberley diamond to 0.390.02 pMC for one of the Namibian alluvial diamonds (with the laboratorys standard
background of 0.0770.005 pMC not subtracted). The average of the values for the five diamonds from kimberlite mines
was 0.04 pMC, and for the seven alluvial diamonds 0.12 pMC, after the laboratorys standard background was subtracted.
Similar detectable levels of 14C have been confirmed more recently in other diamonds (Taylor and Southon 2007). This may
suggest this 14C is intrinsic to the diamonds, which being extremely resistant to contamination almost certainly have not
experienced any recent exchange of their carbon atoms with those in the atmosphere.For more than two decades the
conventional radiocarbon specialists have struggled to understand and explain these significant amounts of 14C, intrinsic to
all these organic materials and diamonds, that they have measured well above the threshold of their AMS instruments, when
these materials should be 14C dead according to their standard uniformitarian geologic timescale. Invariably they have been
forced by their paradigm to conclude that this 14C principally represents intrinsic contamination in the samples when they
arrived at the laboratories, with the minor possibilities of background 14C added to the samples in the laboratories during
processing, and of instrument background due to the detectors and ion beam (Beukens 1990; Bird et al. 1999; Nadeau et al.
2001; Taylor and Southon 2007; Vogel, Nelson, and Southon 1987). However, contamination due to laboratory processing of
samples can be effectively ruled out, due to the harsh chemical treatment of samples designed to remove all possible
contamination, even that from handling of samples during field collection, storage and dispatch to the laboratory.
Furthermore, the14C levels being measured in these ancient organic materials after sample preparation swamps any 14C
attributable to the supposed laboratory processing and/or instrument backgrounds.On the other hand, if the long-ages
uniformitarian timescale is rejected, then these ancient organic materials would not necessarily be 14C dead, particularly if
they really were only 6,000 or so years old since the time of Creation. In this timescale the carbon cycle suffered a
catastrophic upheaval at the time of the Flood cataclysm about 4,500 years ago, so 14C ages of 48,00050,000 years for
coal beds deposited during that event cannot be absolute or real time ages. Ultimately, what is needed is a method of
recalibrating these radiocarbon years with the realtime, creation framework of earth history, which takes into account the
effects of the Flood, the post-Flood Ice Age, and the decreasing strength of the earths magnetic field. In the meantime, it is
important to continue increasing the database of radiocarbon levels in ancient organic materials from the geologic record.
To-date there appears to be no 14C data for two different types of ancient organic materials fossilized together in the geologic
record. It would surely be expected that co-fossilized organic materials should contain similar 14C levels, because the
organisms would have originally lived at the same time. However, if in situ contamination is the problem claimed by the
conventional uniformitarian view for these elevated 14C levels in ancient organic materials, the different co-fossilized organic
materials might be expected to have different 14C levels. This would be due to the different modes of supposed
contamination of the different organic materials. A test of just this situation is available in the Cretaceous strata of Ono
Quadrangle near Redding in northern California, where wood has been found fossilized with ammonites.
Click to enlarge.
Fig. 1. Map showing the location of the Ono Quadrangle straddling the
boundary between Shasta and Tehama Counties near Redding in
northern California.
The Cretaceous Strata and Fossils of Ono Quadrangle
The location of the Ono Quadrangle, near Redding in northern
California, is shown in Fig. 1. It straddles the boundary between Shasta
and Tehama Counties west of Interstate 5 between Red Bluff and
Redding. Fig. 2 (after Rodda and Murphy 1987) is a geologic map of
the Ono Quadrangle. Of interest are the lower Cretaceous strata of the
Budden Canyon Formation.
Previous Work
The geology and paleontology of the area, known as the Cottonwood
District, have been described in numerous scientific publications
spanning more than 140 years, including some of the earliest
geological studies in California (Whitney, 1864, 1865). Diller (1889, 1893) and Diller
and Stanton (1884) provided more information on the geology of the Cretaceous
rocks of the Cottonwood District than any paper for the subsequent 60 years. The
latter included a generalized geologic section along the North Fork of Cottonwood
Creek with lists of fossils collected at several generalized stratigraphic
horizons.Diller (1894) was the first to describe the late Cenozoic non-marine
deposits of the Cottonwood District. Anderson (1902, 1938, 1958) was responsible
for describing the Cretaceous fossils from the Pacific coast, mostly Californian
ammonites, accompanied by syntheses and generalized discussions of the
stratigraphy and historical geology. However, his geologic and locality data were an
inadequate basis for subsequent investigations. Instead, Anderson and Russell
(1939) remains the basic documentation of the geology and stratigraphy of the late
Cenozoic, nonmarine deposits of the northern Sacramento valley, including the
Tehama and Red Bluff Formations of Shasta County. In the area which this study
focuses on, it has been supplanted by later more detailed publications.
Click to enlarge.
Fig. 2. Geologic map of the Ono Quadrangle, showing the exposed area of the lower Cretaceous Budden Canyon
Formation (after Rodda and Murphy 1981).

The principal sources of geologic data for the Cottonwood District of the Ono Quadrangle are detailed geologic maps,
detailed stratigraphic sections, and detailed information on the stratigraphic relationships at fossil collecting sites provided by
subsequent papers. Murphy, Rodda, and Morton (1969) provides a detailed geologic study of the area. Stratigraphic studies
with data on fossil collecting include Murphy (1956), which provides data for lower Cretaceous rocks, and Murphy and
Rodda (1959, 1960), which provide data on the areas upper Cretaceous rocks. Jones, Murphy, and Packard (1965), and
Hill (1975), provide data for the Huling Creek-North Fork of Cottonwood Creek area.Many other publications discuss the
general distribution and relationships of the geological and fossiliferous units exposed in the Cottonwood District. These
include Hacket (1966), Ingersoll (1979), Ingersoll, Rich, and Dickinson (1977), Lachenbruch (1962), Murphy, Peterson, and
Rodda (1964), Ojakangas (1968), Olmstead and Davis (1961), Peterson (1966, 1967), Pierce (1983), Popenoe, Imlay, and
Murphy (1960), and Repenning (1960).
The shells of Cretaceous marine mollusks, including gastropods (snails), bivalves (clams, oysters, etc.), scaphopods (tusk
shells), and cephalopods (ammonites, nautiloids), are by far the most abundant and most extensively described fossils in
the Cottonwood District. Of these, the most widely distributed are the ammonites. They are also the most important
scientifically, because they serve as the standard index fossils for determining biostratigraphic ages through the Mesozoic
(the Triassic, Jurassic and Cretaceous Periods) and for establishing correlations with rock units in distant areas.
For more than 145 years the ammonites and other fossil invertebrates from the Cottonwood District have thus been the
subject of numerous papers. Nearly 300 named species of invertebrate fossils have been recognized in the Cretaceous
rocks of the area, including 180 species of ammonites, 50 species of gastropods, 50 species of bivalves, 1 species of
scaphopod, and 3 species of brachiopods. Of these, 50% were described as new species on the basis of the specimens
collected, or probably collected, in the area, mostly from the North Fork of Cottonwood Creek and the Huling Creek
drainages.The majority of the fossil species recorded from the Cottonwood District were described in reconnaissancelevel
reports, based on highly generalized, commonly inaccurate, locality and stratigraphic data. One of the first fossils described
in California was an ammonite from the Cottonwood District, which provided the first recognition of Cretaceous rocks in
California (Trask 1855). Gabb (1864, 1869), papers issued by the Second Geological Survey of California (the Whitney
Survey), are the first major paleontological reports to describe Californian Cretaceous fossils. Many species were recorded
from the Cottonwood District, including some 35 new species based on specimens probably collected along the North Fork
of Cottonwood Creek or along Huling Creek. Stewart (1926, 1930) reviewed and redescribed Gabbs Californian species of
fossil gastropods and bivalves, without providing any new geologic data.Stanton (1893) and Diller and Stanton (1894)
provide lists of fossil species (mostly mollusks) collected from generalized stratigraphic horizons along the North Fork of
Cottonwood Creek and Huling Creek. Other early paleontological reports that mention the Cottonwood District include White
(1885) and Cooper (1888). Anderson (1902, 1931, 1938, 1958) made extensive collections from the Cottonwood District,
describing almost 100 new species (mostly ammonites) based on specimens from, or probably from, the drainage areas of
the North Fork of Cottonwood Creek and Huling Creek.Murphy (1956), though he did not extensively describe or discuss the
fossils, provided the first detailed geologic map, locality data, and accurate stratigraphic placement of the fossils of the
Cottonwood District. That paper, and similar work by Rodda (1959) on the upper Cretaceous rocks, provided the initial
stratigraphic and locality data on which several later paleontological studies were partly based. These papers, which include
the descriptions of many new species and the redescription of old ones, are Hill (1975), Jones, Murphy, and Packard (1965),
Matsumoto (1959a, b, 1960), Murphy (1967a, b, c), Murphy and Rodda (1959, 1960, 1977), Popenoe (1957), and Saul
(1973). Matsumoto (1959a, b, 1960) represent a major review of the Californian upper Cretaceous ammonites based
principally on a compilation of collections and locality data from many sources.Other descriptive paleontological papers that
include fossils from the Cottonwood District are Saul and Popenoe (1962) and Wiedman (1962). Murphy, Rodda, and
Morton (1969) provide additional important geologic and biostratigraphic data for the Ono Quadrangle, while Murphy (1975)
is a detailed study of the stratigraphy and paleontology of the lower Cretaceous strata, including those within the
Cottonwood District. Since the compilation of Rodda and Murphy (1985) there have been only two further papers that
directly concern the invertebrate fossils of the Cottonwood District, namely, Saul (1978) and Popenoe, Saul, and Susuki
(1987), both of which describe bivalves and gastropods.The few, non-molluscan, Cretaceous invertebrates that have been
described or recorded from the Cottonwood District include brachiopods and shelled microfossils. Brachiopods are
uncommon, and only three species have been recorded (Anderson 1938; Murphy 1956). Rodda and Murphy (1985)
reported a diverse assemblage of marine microfossils, including 41 species of foraminifera, 23 species of radiolarian, and 2
unidentified species of ostracods. They also reported three shark teeth and a single fish vertebra from the area. More
recently, a further single shark tooth, two fish teeth, one fish scale, and two sections of fish vertebra have been found
(Franklin, personal communication). No other vertebrate fossils have been recorded from the areas Cretaceous rocks.
Petrified and carbonized wood are common fossils in many parts of the Cottonwood District. Twigs, leaves and other plant
parts have been noted, but there are few published records for the area. Ward (1905) mentions dicotyledonous leaves
from near the mouth of Huling Creek, but those fossils were never specifically identified or described. Eight species of
Cretaceous plants were identified by Fontaine (1905) from specimens collected from
a tributary of the North Fork of Cottonwood Creek. Some of those specimens were
redescribed by Miller (1975).Chandler and Axelrod (1961) described one of the
earliest fossil angiosperms from a specimen collected by Murphy and Rodda in
lower Cretaceous strata along the North Fork of Cottonwood Creek. However,
Wolfe, Doyle, and Page (1975) questioned the angiosperm affinities of this
specimen, so its significance is equivocal. Nevertheless, Rodda and Murphy (1985)
reported undetermined angiosperm seeds, conifer twigs, and seed fern parts from
two localities in the same area. More recently, a number of plant seed pods as well
as tree cones with seeds have been found in the Cottonwood area (Franklin,
personal communication).
Geology
Two principal geological units are exposed in the Cottonwood District in the northern
part of Ono Quadranglethe Cretaceous (late Flood) marine sedimentary units of
the Budden Canyon Formation, and the overlying non-marine, undifferentiated
Tehama-Red Bluff Formations of Pliocene and Pleistocene (post-Flood) age (Rodda
and Murphy, 1978, 1985, 1987). Even younger are the thin river terrace deposits
and modern channel and floodplain deposits along the major streams. The general
distribution of these major geological units is indicated in Fig. 2. They are part of the
extensive suite of Mesozoic and Cenozoic rocks that crop out on the western side of
the Sacramento River valley. Only the Cretaceous rocks and their fossils are of
significance to this study.

Click to enlarge.
Fig. 3. Diagrammatic classification of the members of the Budden Canyon formation plotted against the biostratigraphic
stages of the Cretaceous and the geologic timescale as determined by radioisotope dating (after Gradstein et al. 2004 and
Rodda and Murphy 1987). The approximate stratigraphic positions of the local strata columns from which each samples was
collected is also shown.The Budden Canyon Formation is composed of predominantly gray-colored, marine mudstone,
siltstone, sandstone, and conglomerate, and has been subdivided into seven members (fig. 3). In the Cottonwood District
the maximum thickness of the Budden Canyon Formation is about 6,700 meters (22,000 feet) along the Dry Creek-Budden
Canyon type section. To the north the Formation thins. Based on its fossil content the Formation ranges in geologic age from
the early part of the Lower Cretaceous (Hauterivian) at more than 130 million years ago, to the middle part of the Upper
Cretaceous (Turonian) at less than 90 million years ago (Gradstein, Ogg, and Smith 2004), a supposed total time span of at
least 40 million years (fig. 3).
In the area the strata of the Budden Canyon Formation strike north to northeast and dip to the east and southeast (fig. 2).
The oldest (lower) part of the strata sequence is to the west and north where the Formation laps onto the older igneous and
metamorphic rocks of the Klamath Mountains. The principal streams flow eastward and southward across the strike of the
Formation, and expose the strata in regular sequence from oldest to youngest (fig. 2). The Budden Canyon Formation
consists mostly of fine-grained mudstones and siltstones which form rounded hills mantled by deep, mostly grass-covered
soil. Outcrops are virtually confined to the bottoms and banks of streams and gullies.Within the study area (figs. 2 and 4)
only the lower half of the Budden Canyon Formation is exposed. This portion is 900 1,200 meters (3,0004,000 feet) thick
along the North Fork of Cottonwood Creek and Huling Creek drainages. Rodda and Murphy (1987) acquired more detailed
and comprehensive data on the Cretaceous stratigraphy exposed in these drainages, substantiating the accuracy of earlier
mapping, and providing increased documentation of the density and patterns of faulting. They reported that field
observations and examination of air photos indicated numerous small faults in zones parallel to and extending away from
previously known major faults. Thus this extensive faulting, in these Cretaceous strata where there are few distinctive
marker beds, renders most measured stratigraphic sections discontinuous by an unknown, but presumably small, amount.
From their field observations, Rodda and Murphy (1987) maintained that the displacements on these small faults are almost
all less than the thickness of one ammonite zone. Furthermore, this recognition of the abundance of small faults in some
areas (for example, near the confluence of Bee Creek with the North Fork of Cottonwood Creekfig. 4) had required
modification or replacement of some previously measured and described stratigraphic sections.
Click to enlarge.
Fig. 4. Map of the North Fork of Cottonwood Creek drainage area near Ono in the
northern section of Ono Quadrangle, Shasta County, showing the measured
stratigraphic sections from which the fossil samples were collected (after Rodda and
Murphy 1987).
Another feature Rodda and Murphy (1987) reported as now known to be more
extensive than was previously recognized is the early post-depositional submarine
slumping of the Cretaceous strata. This slumping, well exposed in some stream
bluffs, has locally folded, contorted, or disrupted the bedding, and is another cause
of discontinuous measured stratigraphic sections. They also recognized that the
primary sedimentary structures of most of the mudstone and siltstone strata (the most common rock types), and some of the
sandstone layers, have been obliterated by the intensive burrowing of the diverse infauna that must have densely populated
the seafl oor when these Cretaceous sediments were deposited and buried them. These burrowing animals essentially
homogenized the fi ne-grained sediments, except where early-formed calcareous nodules locally stabilized the mud and silt.
Rodda and Murphy (1987) also significantly increased the available data on paleocurrent directions for these Cretaceous
strata, supporting previous conclusions of generally southward-flowing bottom-currents being responsible for their
deposition. Furthermore, they interpreted the presence of two previously unrecognized depositional environments in the
Cretaceous rocks exposed along the North Fork of the Cottonwood Creek. An interval in the upper Albian Stage is marked
by greensands and phosphatic sand grains that accumulated in broad shallow channels. Another interval is marked by
shallow channels filled with a mixture of shallow-water shells, including many oysters, which contrast with the apparently
deep-water aspect of these Cretaceous strata. They claim that the apparent shallow-water interpretation of these deposits is
supported by the recognition of shallow-water foraminifera in this interval.
The Fossils
Well preserved fossil invertebrates, mostly ammonites, are found occasionally in most parts of the Budden Canyon
Formation, but the outcrops along the North Fork of Cottonwood Creek and its tributaries are especially rich collecting sites.
According to Rodda and Murphy (1987) they comprise the finest stratigraphic sequence of Cretaceous fossils in North
America, and possibly the world. Some of the richest outcrops include the exposures along the North Fork of Cottonwood
Creek and the branches of Huling Creek (fig. 4). Some 300 named species of Cretaceous invertebrates have been
described or identified from the numerous fossil localities in this area.Typically, the fossils, especially the ammonites, occur
in limestone nodules scattered in the mudstone and siltstone outcrops. A collecting locality commonly is a single fossiliferous
nodule, and when that is collected the locality no longer exists. Later collecting may discover additional fossils in the same
outcrop after a few years of additional weathering and erosion, but not necessarily at the same spot. Some fossil localities
are in the coarser-grained sandstones and conglomerates, but these mostly contain thickershelled non-ammonite
mollusks.Rodda and Murphy (1987) accurately located the fossils they collected from measured stratigraphic sections along
Bee Creek, Huling Creek and the North Fork of Cottonwood Creek (fig. 4). They found that most collecting localities from
earlier paleontological studies were too generalized or poorly described to be plotted on detailed stratigraphic sections like
theirs. However, they endeavored to carefully locate the stratigraphic horizons where many of the fossil species described in
early publications had been found, but not previously located on detailed measured sections.Rodda and Murphy (1987)
found that within the Cretaceous rocks of this area (fig. 4), invertebrates, mostly mollusks, are the most abundant and
important fossils. Of these, ammonites are the most significant. Other groups they recovered included decapod crustaceans,
bivalves, gastropods, scaphopods, barnacles, and foraminifera. A few specimens of fossil vertebrates and plants (other than
fossil wood) were recovered also. They thus increased the number of fossil collecting sites in this area by at least 50%.
Their study also provided much additional control on stratigraphic-geographic ranges and, by assumption, time relationships
of the known fossil species, especially the ammonites. In addition, the number and diversity of fossil species known to occur
in this area was increased significantly, providing improved documentation of the supposed paleogeography and
evolutionary history. A particular example was their discovery of an ammonite genus, Pictetia, not previously recorded from
the western hemisphere.Rodda and Murphy (1987) recovered approximately 1500 fossil specimens from 609 collecting
localities. From these collections they recognized some 175 species, including about 70 species of ammonites, 40 species

of other invertebrates, and 64 species of microfossils. Of special significance are the ammonites, the extinct group of coiled,
shelled cephalopods, that are found around the world wherever Mesozoic rocks are found. They have thus been studied
intensively for their presumed theoretical and practical values for well over a century, being the principal fossils for
determining the supposed precise geological (biostratigraphic) ages of Mesozoic rocks. The area studied exposes a
remarkable sequence of ammonite-bearing strata, which is especially rich and regarded as complete.Of the 70 species of
ammonites that were recognized in their collections, certain groups were especially well representedthe genera Puzosia,
Lytoceras, Pseudouhligella, Acanthoplites, and heteromorphs (irregularly coiled ammonites) of great variety (Rodda and
Murphy 1987). Their collection also contained genera and species not previously recorded in California
Pictetia sp., Lechites sp., Stoliczkaia sp., Myloceras sp., Lyelliceras sp., and Zelandites sp., of supposed late Albian
age; Colombiceras sp., of supposed Aptian age; and Crioceratites sp. of supposed Barremian age.
An important result related to the evolutionary ammonite biostratigraphy was the discovery by Rodda and Murphy (1987)
that the entire sequence of ammonite zones below the Huling Tongue Member of the Budden Canyon Formation is of
supposed Barremian age (fig. 3). Additionally, underlying older strata of supposed Barremian age (the lower part of the
Chickabally Member) along the North Fork of Cottonwood Creek yielded many and varied fossils, including ammonites,
gastropods, bivalves, nautiloids, belemnites, and decapod crustaceans.
Other molluscan fossils, especially bivalves and gastropopds, are locally common, with richer occurrences of these fossils in
the coarser grained rockssandstone and conglomerate. Some forms discovered by Rodda and Murphy (1987) were new
records for the Cretaceous of California, and several new species were recognized as well. Additionally, they confirmed the
wealth of shrimp-like decapod crustaceans previously reported in these Cretaceous rocks of the Cottonwood District by
Rodda and Murphy (1985). Numerous specimens, mostly isolated claws, representing at least four different species, were
collected from limestone nodules in the drainage of the North Fork of Cottonwood Creek.Trace fossils are abundant and
diverse, especially the filled burrows of benthic infaunal invertebrates such as mud-shrimp and various worms. Most of the
mudstones and siltstones were thoroughly burrowed and homogenized by the burrowing activities of these invertebrate
seafloor animals. Few samples of these could be collected easily by Rodda and Murphy (1987), but they photographically
recorded in the field the varied types and densities. Two basic types are permanent dwelling burrows, mostly vertical, and
errant feeding burrows.
Click to enlarge.
Fig. 5. The local stratigraphic columns from which the fossil samples were collected
(after Rodda and Murphy 1987). The location of these measured stratigraphic
sections is shown on Fig. 4.
The presence of diverse, abundant microfossils in the study area (fig. 4) was
confirmed by Rodda and Murphy (1987). Of these, only the foraminifera and
radiolarian were identifiable. The ostracods and micro-mollusks had not then been
studied in lower Cretaceous rocks of California. They did not recover any coccoliths
from any of their samples. Micromollusks and microscopic parts of other
invertebrates are locally abundant.The only vertebrate fossils recovered from these
Cretaceous rocks were mostly fragmentary fish remains (teeth, scales, and bone).
The only identifiable specimens were shark teeth, one of the family Lamnidae, and
the other from a cow shark, Notidanodon sp., a group not previously reported from
Cretaceous rocks in California.Few fossil plant specimens were collected by Rodda
and Murphy (1987), other than fossil wood which is abundant in several stratigraphic
sections in the study area (fig. 4). A fossil seed, probably from a conifer, was
collected from supposed Barremian age mudstone along the upper drainage of the
North Fork of Cottonwood Creek. Other plant fossils included twigs, conifer leaves,
and a reed fragment.
The Present Study
The purpose of this study was to test the radiocarbon contents of wood and
ammonite shells buried and fossilized together in the same sedimentary rock layers. The literature summarized above
clearly indicates that fossilized wood is abundant within several stratigraphic sections of the Cretaceous Budden Canyon
Formation strata exposed along the North Fork of Cottonwood Creek and its tributaries in the northern part of Ono
Quadrangle near Redding, California. Furthermore, the same stratigraphic sections would undoubtedly also contain
abundant marine invertebrate fossils, especially ammonites. Therefore, a field search in this area would be highly likely to
find marine shells and wood buried and fossilized together. However, ammonite fossils were to be specifically targeted,
because of their stated global significance as index fossils for evolutionary biostratigraphic age determinations of these
strata as lower Cretaceous.Field trips into the study area were undertaken in 2006 by, and with, Pastor Al Franklin, of Grace
Baptist Church in Redding, because of his knowledge of the area and its fossil localities, and because of the permission he
has to enter the private lands of the area, granted him by the landholders. Four suitable ammonite fossils were successfully
found that each had fossilized wood very closely associated with it, the ammonite shells and the wood usually being in direct
physical contact with one another.

Click to enlarge.
Fig. 6. The locations and strata from which the fossilized pieces of wood and ammonites were collected for this study, as
shown in Fig. 4. (a) (b) Sample 1; (c) (d) Sample 2; (e) (f) Sample 3; and (g) (h) Sample 4.
The locations where each of these fossils were found are indicated on Fig. 4. The relevant portions of local stratigraphic
sections, as measured by Rodda and Murphy (1987), with the approximate stratigraphic positions of the ammonite and
wood fossils indicated on them, are provided in Fig. 5. Where those local stratigraphic sections fit in relation to the
stratigraphy of the Budden Canyon Formation is indicated on Fig. 3, along the approximate stratigraphic positions of the
collected fossils. Fig. 6 photographically documents the locations from which the four ammonite with wood fossils were
collected, while Fig. 7 shows the fossils themselves.The four samples were first photographed (fig. 7) before any work was
performed on them. Then, using a scalpel and tweezers, pieces of the ammonite shells and the fossilized wood were
separately broken off of each sample. Pieces were broken off until there appeared to be sufficient material for each
radiocarbon analysis, and then suitably sealed and stored. The fossil samples were numbered 14, so the broken off pieces

of ammonite shell were labelled RNCS-1 to RNCS-4 respectively, and the broken off pieces of fossilized wood were similarly
labelled RNCW-1 to RNCW-4. Sample 4 contained both petrified and coalified wood, so pieces were broken off separately,
and sealed, stored and labelled as RNCW-4A and RNCW-4B respectively. All nine so-labelled sub-samples were then
submitted to Professor Roelf Beukens at the IsoTrace Radiocarbon Laboratory at the University of Toronto, Ontario,
Canada, each sub-sample being accompanied by this laboratorys required Radiocarbon Sample Submission Form
carefully filled out with location and estimated age details.At this laboratory, the petrified and coalified wood sub-samples
were prepared for radiocarbon analysis with a modified AAAOx pre-treatment (Beukens 2007a, b), the standard procedure
developed to guarantee the elimination of any contamination. First, though, the fossilized wood sub-samples were
demineralized to remove any contaminant inorganic minerals. This involved drenching the sub-samples in hot and strong
hydrochloric acid to dissolve away any calcium, barium, or strontium salts (which is done to avoid producing insoluble
fluorides in the next step), and then soaking the sub-samples for at least a week in a hot and strong mixture of hydrochloric
and hydrofluoric acids. After this, the acid-soluble humics were removed from what remained of the sub-samples with an
extended hot and strong hydrochloric acid treatment. This was followed by an extended cold and fresh alkali extraction.

Click to enlarge.
Fig. 7. The fossilized pieces of wood and ammonites collected for this study from the localities shown in Fig. 4. (a) (b)
Sample 1; (c) (d) Sample 2; (e) (f) Sample 3; and (g) (h) Sample 4.
The laboratory (Beukens 2007 a, b) reported that the dried residues of the petrified
wood sub-samples (RNCW-1, 2, 3, 4A) at this point in the procedure did not have
any wood structure and resembled detrital material. However, it was concluded
that it was possible part of these residues consisted of acid and alkali insoluble
humic or tannic compounds, as their carbon contents were low (typically 10%).
These residues were then partially oxidized using a chlorite bleach. In the case of the coalified wood sub-sample (RNCW4B), however, the dried residue consisted of needles with a carbon content normal for organic material. An acid chlorite
bleach was also used on this residue to partially oxidize it. These five fossilized wood residues were then all degasified
under vacuum before subsequent combustion. Nevertheless, the combustion of petrified wood residue RNCW-4A produced
less than 0.1 mg of datable carbon, which was too small for analysis.The fossilized ammonite shell sub-samples were
prepared by the laboratory for analysis via a different process. However, those labeled RNCS-1 and 3 were deemed to
contain insufficient shell material to be processed and analyzed. The remaining two shell sub-samples were first washed in
cold deionized water. Then the outer surfaces of the shell pieces were removed by leaching with hydrochloric acid, the outer
16% of RNCS-2, and the outer 14% of RNCS-4 (Beukens 2007 a, b). An XRD analysis of the pre-leach residue of RNCS-2
was then undertaken to determine the calcite content of RNCS-2, which was estimated at between 0.5 and 2.0%, the rest of
the shell material consisting of aragonite. An estimate of the calcite content of RNCS-4 could not be made as that subsample material was considered too small for an XRD analysis.The remainder of the ammonite shell material in RNCS-2
and 4 was then processed according to the regular treatment for carbonates, namely, hydrolysis of the carbonate to carbon
dioxide (CO2). To achieve this, phosphoric acid in used because it has a low vapor pressure. First though, the phosphoric
acid has to be degassed to remove any water or atmospheric CO2 which might be dissolved in it. The most common
method of doing this is via vapor distillation, but at the IsoTrace laboratory this technique is not used. Instead, each subsample was mounted in a reactor vessel, containing the phosphoric acid and, in a separate small ampoule, the crushed
shell material, connected to a vacuum pump, which pumps the vessel down to below 10 -4 Tor. By this means the water and
most of the CO2 was first pumped away at around 10-3 Tor, the pressure plateauing at that level until all the water was
removed. When the pressure dropped from around 10-3 Tor to below 10-4 Tor any remaining dissolved gases were removed.
The reactor vessel was then placed on a shaker to initiate and complete the hydrolysis, without ever exposing the contents
to the atmosphere again. The evolved CO2 was then collected on the same collection system which handled the combustion
of the fossilized wood residues. This system is regularly tested by this laboratory with Icelandic spar carbonate and the IAEA
(International Atomic Energy Agency) Carrara marble reference material.The resultant graphite from this treatment and
processing of each of the fossilized wood and ammonite shell fragments was then analyzed for radiocarbon ( 14C) using the
IsoTrace laboratorys state-of-theart AMS (accelerator mass spectrometry) system. Four separate high-precision analyses of
each subsample were averaged and corrected for natural and sputtering isotope fractionation, using the measured 13C/12C
ratios. The averaged radiocarbon analysis for each sub-sample reported by the laboratory, after the laboratorys
background correction of 0.077 percent modern carbon (pMC) was subtracted, was quoted as an apparent uncalibrated
radiocarbon age in years before present (BP), using the Libby meanlife of 8,033 years. The quoted errors represent the
68.3% confidence limits.
Results
The results of the radiocarbon analyses are listed in Table 1 (from Beukens 2007 a,b). Note that the laboratory was not able
to analyze sub-samples RNCS-1 and 3 (fossilized ammonite shells) and RNCW-4A (petrified wood) due to insufficient
material. However, the six listed radiocarbon analyses would still seem to provide an adequate database for the purpose of
this study. Note also that the percent modern carbon (pMC) values listed were reported by the laboratory after subtraction of
their background correction of 0.077 pMC. Therefore, if this background correction had not been subtracted, the listed
apparent radiocarbon ages would only be slightly, to negligibly, younger.
Table 1. Radiocarbon (14C) analytical results (pMC) and calculated apparent ages (year BP) for the fossilized wood and
ammonite shells.
Sample

Description

Weight Used (mg)

Percent Modern Carbon (pMC)

Apparent Age (years BP)

RNCW-1

petrified wood

1080

1.8960.013

33,490240

RNCS-2

fossilized shell fragment

386

1.2240.012

36,400350

RNCW-2

petrified wood

4400

1.1180.010

37,150330

RNCW-3

petrified wood

729

1.7400.012

32,780230

RNCS-4

fossilized shell fragment

510

0.2760.005

48,710930

RNCW-4A petrified wood

297

RNCW-4B coalified wood

382

0.5930.007

42,390510

Discussion
Fig. 5 compares the local measured stratigraphic sections for fossil samples 1, 3 and 4 (after Rodda and Murphy 1987) and
indicates the approximate stratigraphic locations of the collected fossil samples 1, 3, and 4. Using the position of the Huling
Tongue Member (sandstone) as a biostratigraphic marker, the approximate biostratigraphic postions of the stratigraphic
sections 1 and 4 have been plotted on Fig 3. (last column). Fossil sample 2 was not collected from a measured stratigraphic
section, but from its location marked on Fig. 4 it can be seen that it lies stratigraphically just below fossil sample 3, and
stratigraphically well above fossil sample 4. These observations thus enable the biostratigraphic position of measured
stratigraphic section 3 relative to measured stratigraphic section 4 (figs. 4 and 5) to be plotted on Fig. 3. Furthermore, the
relative biostratigraphic position of stratigraphic section 2, though not measured, can also be plotted on Fig. 3.The
biostratigraphic positions of the stratigraphic sections from which the four fossil samples were collected as plotted on Fig. 3
enable the approximate biostratigraphic positions, and therefore ages, of the four fossil samples to be determined. The four
fossilized ammonite and wood samples would all appear to be Aptian in biostratigraphic age (lower Cretaceous), which in
terms of the absolute chronology of the geologic timescale equates to approximately 112120 million years old (Gradstein et
al. 2004). However in stark contrast, the fossilized ammonite shells yield radiocarbon ages of only 36,400350 years and
48,710930 years, while the fossilized wood found buried with them (literally, as seen in fig. 7) yielded radiocarbon ages of
32,780230 years to 42,390510 years (table 1).The usual response to such a glaring and enigmatic discrepancy in
absolute ages is to claim that the ammonite shells and wood have obviously been contaminated with modern carbon,
making them date very young when in fact they really are extremely old. After all, as noted earlier, if the entire mass of the
earth consisted only of 14C atoms, then because the decay of 14C is so rapid, after only 1 million years not a single atom
of 14C should conceivably still exist. Therefore, there should be absolutely no measurable 14C able to be detected in these
fossilized ammonite shells and wood if they really are, as claimed to be, 112120 million years old.Four sources of potential
contamination could be invoked in this instance. First, any contamination in the laboratory can be immediately ruled out,
because extreme handling and preparation measures were used in this highly respected academic laboratory, measures
that have proved effective in removing any potential contamination. These included extended use of strong acid at elevated
temperatures to guarantee removal of any carbonate and other minerals that might have contributed modern radiocarbon to
the fossilized wood, and leaching of 1416% of the outer surfaces of the ammonite shells with hydrochloric acid to remove
any modern calcite from the shells that would have substituted for their aragonite and thus contributed modern radiocarbon
to them.However, the only possible contamination that might have been introduced into these fossilized wood and ammonite
shells during laboratory processing to prepare them for analysis could have been at the stage where the evolved CO 2via
combustion was then reduced to graphite. It is quite possible that the acetylene used for this reduction could introduce some
trace modern radiocarbon contamination, or more correctly, the laboratory background. This, however, would be taken care
of by the standard background correction of 0.077 pMC subtracted routinely from the raw analytical results, a procedure that
would thus appear to be legitimate. Nevertheless, it can be readily shown that for radiocarbon ages on the order of 33,000
years or more, the effect of subtracting this standard laboratory background (and therefore the possible contamination due
to the acetylene) is small, almost negligible, on the order of only 1%.That potential trace radiocarbon contamination
introduced at the reduction to graphite by acetylene step equates to the graphitization background suggested by Taylor
and Southon (2007). However, they also suggest four other sources of modern radiocarbon contamination in the laboratory
procedures. These, they claim, are due to the acidification and combustion during the sample preparation, the transfer of the
graphite to the AMS (accelerator mass spectrometer) sample holder, during any storage, and within the AMS instrument if
the detector registers a 14C signal when a 14C-ion is not present in the beam. In order to quantify these potential sources of
modern radiocarbon contamination in the laboratory, Taylor and Southon (2007) analysed natural diamonds for 14C by
placing them directly within the AMS instruments sample holder, thus eliminating the acidification, combustion and
graphitization steps. Of course, they were assuming the diamonds (being in excess of 100 million years old) contained no
radiocarbon, so that any measured by the AMS instrument would be the sum total of their claimed transfer, storage and
instrument backgrounds (that is, modern radiocarbon contamination). Nevertheless, the only 14C they measured ranged from
0.0050.001 to 0.0690.04 pMC. At such a trivial background level, even if such potential laboratory contamination with
modern radiocarbon were conceded, it would make no difference whatsoever to the very much higher radiocarbon levels
within these ammonite shell and wood fossils listed in Table 1. The 0.077 pMC laboratory background already subtracted by
the IsoTrace laboratory would more than adequately cover such contingencies.The second potential source of any
contamination would be any contamination due to handling of the fossil samplesfor example, from human hands or plastic
storage bags. This can also be definitively ruled out, because any such contamination would only have been on the surfaces
of the fossil samples and would have been immediately eliminated by the laboratorys extreme sample preparation
techniques, described earlier. Indeed, in the case of the ammonite shells, significant portions of their outer surfaces were
deliberately leached by hydrochloric acid to remove any such surface contamination. And the fossilized wood samples were
soaked in hot strong acids.The third potential source of contamination would have been the source area in the ground from
where the fossil samples were taken, when they were buried deep below the surface. Here there were definitely two relevant
factors. At the time of deposition of the host sediments and burial of the ammonites and wood in them, some of the water
which transported the sediments would have been trapped in them. Such connate waters and basinal fluids, especially if the
sediments were deposited in a marine environment, could have been quite saline and thus chemically reactive, aided by the
raised temperatures due to the depth of burial. Diagenesis of the host sediments occurred around these fossils, primarily
resulting in the cementing together of the sediment grains. The internal body cavities of the ammonite shells were infilled
with sediments that underwent diagenesis, and the chemical-laden connate waters percolated through the cellular structures
of the buried wood to petrify it. In some cases the wood was coalified, indicative of a depth of burial of thousands of meters
and elevated temperatures of 100C or more (Diessel 1992, Stach et al. 1982), though the presence of clay minerals such
as montmorillonite can significantly lower the temperatures and the time required for coalification (Hayatsu et al. 1984). The
result of these fluids circulating around these ammonite and wood fossils would not have only assisted fossilization,
petrification and coalification, but may have also caused secondary calcite replacement of aragonite in the ammonite shells
and added contaminant carbonate and silicate minerals, and silica (quartz), to the wood.However, no carbonate minerals or
silica were in any way visibly evident within or clinging to the coalified wood when it was sent to the laboratory, although the
petrified wood was obvious impregnated with silica as a result of the petrification process. In any case, such minerals would
have been removed from the fossilized wood, even from within it, by the severe demineralizing treatment in the laboratory.
For example, the hot and strong hydrofluoric acid would have removed the silica impregnating the wood. This is confirmed
by the laboratorys description of the fossilized wood residues (Beukens 2007a, b). Furthermore, the connate waters and
basinal fluids buried deeply with these fossils at the time of introducing dissolved minerals, and impregnating the wood with

silica while possibly replacing some of the aragonite in the ammonite shells with calcite, supposedly more than 100 million
years ago and since, would have only contained old carbon. Thus if such old carbon were added to these fossils, such as in
the calcite replacement of some aragonite in the ammonite shells, if anything it would have swamped any radiocarbon in the
shells and wood, so that they would have yielded infinite radiocarbon ages, consistent with them being supposedly 112120
million years old.This only leaves, finally, the fourth potential source of contaminationnamely, the ground and surface
waters that washed over and percolated through these fossils in their host sedimentary rocks right up until the present when
sampled. The surface waters particularly would likely contain some dissolved carbonate, and the fluctuating water levels in
the streams would have resulted in the alternating wetting and drying of the fossils, the latter depositing whatever salts the
water contained. Furthermore, both the surface waters and the fossils when exposed have been in contact with the modern
radiocarbon in the atmosphere, making contamination of these fossils with modern radiocarbon seem a likely possibility,
which might explain the young apparent radiocarbon ages for these fossils.However, any soluble inorganic carbonate
carbon in the ground and surface waters would not have exchanged with the insoluble organic carbon in the wood, because
the two forms of carbon are incompatible. Also, as noted previously, any carbonate mineral deposited within or onto the
wood by ground or surface waters would have also been removed by the severe demineralizing treatment with hot and
strong acids in the laboratory prior to the radiocarbon analyses. Indeed, the laboratorys reports (Beukens 2007a, b)
described the fossilized wood residues as possibly consisting of humic or tannic compounds and needles with a carbon
content, all of which were insoluble in the hot strong acids and alkalis used in their demineralizing treatment. As it
was only these insoluble organic carbon residues that were subsequently analysed for radiocarbon, there would not have
been any contamination of them with modern radiocarbon from either ground or surface waters or even the atmosphere.
Such exchanges with the wood could only have occurred when the original trees were alive.The only question that remains
is whether the fossil ammonite shells may have been contaminated with modern radiocarbon in any carbonate dissolved in
the ground or surface waters that replaced aragonite in the shells. The laboratory did report (Beukens 2007a) that the preleach residue of the fossilized ammonite shell RNCS-2 had an XRD estimated calcite content of between 0.5 and 2.0%.
While the laboratory did not say that this 0.52.0% calcite content was modern calcite and therefore contamination, it did
indicate that a 0.5% contamination of modern calcite would be equivalent to an apparent age of 43,000 years. Clearly, with
an estimated 0.52.0% calcite content, if this were all modern calcite, then the 36,400350 years apparent radiocarbon age
for the RNCS-2 fossil ammonite shell could all be due presumably to modern calcite contamination. However, Webb et al.
(2007) found that if groundwaters have high Mg:Ca ratios, then modern aragonite could be added to fossil bivalve shells,
which could also adversely affect radiocarbon dating of such shells. Nevertheless, in respect of this RNCS-2 fossil ammonite
shell, the fossilized wood RNCW-2 buried with it (fig. 7c & d), which has been demonstrated already not to have been
contaminated with any modern radiocarbon, yielded an almost identical apparent radiocarbon age of 37,150330 years
(table 1). Similarly, the fossilized ammonite shell RNCS-4 yielded an apparent radiocarbon age of 48,710930 years, and
the coalified wood RNCW-4B buried with it (fig. 7g & h) yielded a similar apparent radiocarbon age of 42,390510 years
(table 1). If both these fossilized ammonite shells were contaminated with modern calcite containing modern radiocarbon,
then they should both have yielded very young apparent radiocarbon ages, much younger than the apparent radiocarbon
ages for the fossilized wood buried with them. Quite clearly, if these fossilized woods are not contaminated with modern
radiocarbon, then the similar apparent radiocarbon ages for the fossilized ammonite shells buried with them strongly implies
that these fossilized ammonite shells are not contaminated with modern radiocarbon either, even from ground or surface
waters.These second to fourth sources of potential contamination are all outside the radiocarbon laboratories and out of
their control, so collectively they have been termed in situ contamination (Vogel et al. 1987) or pseudo 14C-free sample
background (Taylor and Southon 2007). This term has been defined as 14C present in carboniferous material that should
not contain 14C because of its geologic age (Taylor and Southon 2007). That definition unashamedly reveals the bias in
assuming such carboniferous material is indeed of geologic age, that is, millions of years old. Moreover, this terminology
was introduced by the radiocarbon laboratories over twenty years ago because it was embarrassing that they routinely
found radiocarbon levels of 0.25 pMC or more in ancient carboniferous materials (such as coal, calcite, and fossilized wood)
that should be truly 14C free due to their supposed geologic ages (Brown and Southon 1997). To put this into perspective, a
measured value of 0.25 pMC is some 450 times higher than the measurement threshold for the AMS instrument. Indeed, all
the fossilized wood and ammonite shells analysed in this study yielded pMC levels >0.25 (table 1). Furthermore, as has
been forcefully argued already, the severe hot strong acid treatment on the fossilized wood had to have removed any and all
potential pre-laboratory in situ contamination. Also, the ammonite shells buried and fossilized with them yielded similar
measured high radiocarbon levels, suggesting that the radiocarbon measured in both the fossilized wood and ammonite
shells was not pre-laboratory in situ contamination or pseudo 14C-free sample background.It can only be concluded,
therefore, that the radiocarbon measured by the laboratory must be real in situ radiocarbon intrinsic to the original wood and
ammonite shells, and not contamination of any sort. This does not imply that this radiocarbon is a reliable measure of the
true age of the wood and the ammonites. In fact, other fossilized woods analyzed for radiocarbon have yielded various
apparent radiocarbon ages from 22,730170 years (Snelling 2000a) to 44,700950 years (Snelling 1997, 2000b). However,
these measured radiocarbon levels do indicate that the wood and ammonites are young, and not 112120 million years old.
Clearly, the long-age biostratigraphic dating method, using the ammonites and their relative position in the supposed
evolutionary history of life on earth calibrated against the absolute ages provided by radioisotope dating (Gradstein et al
2004), is totally unreliable. Furthermore, the unproven assumptions on which radioisotope dating is based, and the
numerous problems associated with it, are well documented (Snelling 2000c, 2005; Woodmorappe 1999).It should also be
noted that these radiocarbon ages for the fossilized wood and ammonites were calculated on the assumption that these
fossilized ammonite shells and pieces of wood had similar radiocarbon contents, when they lived and then were buried
together, to the radiocarbon contents of modern marine invertebrate shells and terrestrial trees. However, this assumption
can be shown to be false for at least two reasons. First, the Flood removed so much carbon from the biosphere and buried
it. There was all the carbon in both marine and land animals, and all the plants, which were destroyed and buried as fossils
in Flood-deposited sediments, such as limestones and coal beds. Second, the earths magnetic field was much stronger in
the recent past back to the Flood, and perhaps stronger also back to the Creation Week (Humphreys 1983, 1986, 1990). A
stronger magnetic field, relative to the present field, would shield the earths atmosphere from more cosmic rays, thus
resulting in a much lower radiocarbon production rate in the atmosphere through the pre-Flood period when these
ammonites and trees lived. These two factors alone thus would have meant that there was much less radiocarbon in ancient
organic materials when they were buried during the Flood about only 4,300 years ago. Therefore, if it is wrongly assumed
these ammonite shells and pieces of wood contained todays much higher radiocarbon level when they were buried
together, then the measured low levels of radiocarbon in them yield apparent ages that are much too old.What is now
required is a recalibration of the apparent radiocarbon ages for these supposedly ancient organic materials that would
significantly reduce their true ages to make them compatible with the creation timescale of earth history. However, so far
there doesnt appear to be a discernible systematic pattern of radiocarbon levels (and apparent ages) in the ancient organic

materials tested, such as fossilized woods and coal beds, with respect to their relationship to the timing of their burial during
or soon after the Flood. For example, the radiocarbon testing of ten U. S. coal beds spanning a significant portion of the
fossil-bearing strata record, from the Carboniferous (Pennsylvania), Cretaceous and Eocene, yielded apparent ages of
48,00050,000 years (Baumgardner 2005; Baumgardner et al 2003). If the Eocene is considered to be very early post-Flood
(Austin et al 1994), then why do these Eocene coal beds yield the same apparent radiocarbon ages as the Pennsylvanian
(Flood) coal beds? Perhaps the Eocene coal beds also consist of pre-Flood plant debris that floated through the Flood and
then was buried very soon after. In contrast, the fossilized woods buried as a result of what may be post-Flood (Eocene and
Oligocene) volcanic eruptions apparently decimating post-Flood forests around Crinum, Queensland (Snelling 1997, 2000b)
and Cripple Creek, Colorado (Snelling 2008) yielded apparent radiocarbon ages of 44,700950 years and 41,260540 years
respectively. These latter apparent ages might then be consistent with an expected early post-Flood radiocarbon buildup
which could have produced younging upwards apparent ages. However, in this study the Cretaceous (late Flood buried) four
fossilized woods yielded apparent radiocarbon ages of 32,780230 years to 42,300510 years, younger than the
Cretaceous coal beds. Similar younger apparent radiocarbon ages of 33,700400 years, 33,720430 years, and
22,730170 years to 28,820350 years were obtained for Floodburied Permian (Snelling 1998), Triassic (Snelling 1999),
and Jurassic (Snelling 2000a) fossilized woods respectively. Perhaps the low radiocarbon levels in the pre-Flood world were
unevenly distributed in the biosphere, according to varying abilities of organisms for radiocarbon uptake or rejection.
Continuing investigations are needed.Nevertheless, Baumgardner (2005) has suggested, based on earlier studies (Brown
1979; Giem 2001; Morton 1984; Scharpenseel and Becker-Heidmann 1992), that the pre-Flood biosphere and atmosphere
just prior to the Flood could have had, conservatively, 300700 times the total carbon relative to our present worlds
biosphere and atmosphere. Then if we assume the total number of 14C atoms was similar to what exists in todays world,
and these were uniformly distributed throughout the pre-Flood biosphere which had 500 times more total carbon than
todays biosphere, then the resulting 14C/C ratio would be 1/500 of todays level, or about 0.2 pMC, which is equivalent to an
apparent radiocarbon age of more than 50,000 years. However, Baumgardner (2005) also took into account the amount
of 14C potentially generated in situ from an episode of accelerated nuclear decay during the Flood catastrophe (Vardiman,
Snelling, and Chaffin 2005), which could have been comparable to, or even larger than, the amount of 14C that existed in the
tissues of pre-Flood organisms. Furthermore, with a stronger magnetic field in the pre-Flood world reducing the 14C
production rate in the atmosphere, in the 1,650 or so years between Creation and the Flood it is likely that only a smaller
amount of 14C would have been generated compared to todays total amount of 14C atoms. Therefore, the pre-Flood 14C/C
ratio would have been much lower, and more or less uniform, say 0.050.1 pMC, while the in situ component would have
been much more variable and dependent on the local U concentrations and amounts of N during the subsequent
accelerated nuclear decay episode. This might then account for the wide range of low radiocarbon levels measured in the
pieces of pre-Flood wood (referred to above) that were buried and fossilized in sediments during the Flood only about 4,300
years ago while an episode of accelerated nuclear decay was also occurring.Finally, one further circumstance warrants a
comment, namely, the fact that these ammonites and pieces of wood are found buried and fossilized together. The host
mudstones have always been designated as having been deposited in a marine environment, in water deep enough for
ammonites to have lived, presumably a continental shelf or shallow sea environment. However, woody trees do not live in
such environments. And even pieces of wood washed out to sea today float, rather than sinking to be buried on the sea floor
with marine invertebrates similar to ammonites. Thus these pieces of terrestrial wood were more likely buried with these
marine ammonite shells as a result of sediments (primarily muds) being wash rapidly by ocean waters onto this continental
land, consistent with what would be expected to have happened during the Flood.
Summary and Conclusions
The lower Cretaceous Chickabally Member mudstones of the Budden Canyon Formation have been extensively studied for
more than 140 years, and the ammonite fossils in them are significant as index fossils for biostratigraphy. Four samples of
ammonites and wood buried and fossilized together in these mudstones were collected in the Cottonwood District near
Redding, northern California. Fragments from two of the fossilized ammonite shells and four of the fossilized pieces of wood
were analyzed by the IsoTrace Radiocarbon Laboratory at the University of Toronto, Canada. At this laboratory these
fragments were prepared for radiocarbon analyses with the standard pre-treatment procedure developed to guarantee the
elimination of any contamination. The wood residues were combusted to graphite, while the carbonate shells were
hydrolyzed to carbon dioxide and then reduced to graphite. The resultant graphite was then analyzed for radiocarbon using
a state-of-the-art accelerator mass spectrometer (AMS) system.The radiocarbon results ranged from 0.2760.005 to
1.2240.012 pMC or apparent ages of 36,400350 to 48,710930 years for the fossil ammonite shells, and from
0.5930.007 to 1.8960.013 pMC or apparent ages of 32,780230 to 42,390510 years for the fossilized wood. Yet the
biostratigraphic ages of these fossils, based on the ammonites as index fossils, is Aptian (lower Cretaceous), and the
absolute geochronologic age 112120 million years. Therefore, if these fossils were truly that old, then there should have
been absolutely no measurable 14C detected in them.However, the 14C measured in these fossils is well above the detection
limit of the AMS instrument. Therefore the usual response to such a glaring and enigmatic discrepancy in absolute ages is to
claim that the ammonite shells and wood were contaminated with modern carbon, either in the ground, or during sampling
and in the laboratory. Four such sources of potential contamination were examined. In the laboratory the severe pretreatment of the samples guarantees that any contamination from sampling and handling is totally eliminated. Then even if
there were some contamination, as claimed by some, during the combustion and due to instrument background, the
estimated 14C level involved would only be in the range 0.0050.069 pMC. This is a trivial amount of 14C that if conceded
would make no difference whatsoever to the very much higher radiocarbon levels measured in these ammonite shells and
wood fossils, particularly as the results reported by the IsoTrace Laboratory already have a laboratory background of 0.077
pMC subtracted from them.Furthermore, potential contamination of the fossils by ground or surface waters while they were
still entombed in the mudstones can be ruled out. Any such environmental contamination of the fossil wood would be
removed by the severe sample pre-treatment in the laboratory. On the other hand, environmental contamination of the
ammonite shells by replacement with modern carbonate 14C can be discounted, because the ammonite shells yielded
almost identical 14C levels and apparent ages as the wood buried and fossilized with them.Therefore, it was concluded that
the measured 14C is in situ radiocarbon intrinsic to the ammonites and wood when they were buried and fossilized, so that
they are very young, not 112120 million years old. Furthermore, because the earths stronger magnetic field in the recent
past reduced the atmospheric 14C production rate, and because the recent Flood removed so much carbon from the
biosphere and buried it, the measured apparent radiocarbon ages are still much higher than the true ages of the fossil
ammonites and wood. So their true ages are consistent with their burial during the Flood only, about 4,300 years ago, when
the ocean waters washed sediments and ammonites onto this continental land.
Acknowledgments
This project would not have been possible without the encouragement and support of Pastor Al Franklin and the
congregation of the Grace Baptist Church in Redding, CA. Pastor Franklin conceived the project and assisted with the field

work, sampling, and documentation, while the funding for my visit and all the radiocarbon analyses were provided by the
church congregation. Dr John Baumgardner is also acknowledged for his help with obtaining the radiocarbon analyses,
while Mark Armitage helped photograph some of the samples.
Radiocarbon in Diamonds Confirmed
by Dr. Andrew A. Snelling on November 7, 2007
Abstract
During the RATE (Radioisotopes and the Age of The Earth) research project, some of the research effort was focused on
investigating radiocarbon (carbon-14) dating.
Keywords: radiocarbon, diamonds, RATE, dating methods, carbon-14, in situ, Flood, radioisotopes, age of the earth,
University of California
During the RATE (Radioisotopes and the Age of The Earth) research project at the Institute for Creation Research, cosponsored by theCreation Research Society, some of the research effort was focused on investigating radiocarbon (carbon14) dating. This is one of theradioactive dating methods, but because carbon-14 decays relatively rapidly it only provides
ages in the range of tens of thousands of years. In fact, if every atom making up the earth was carbon-14, even after just 1
million years there would be absolutely no atoms of carbon-14 left, because they would have all decayed away, based on
todays measured half-life! Thats why radiocarbon dating isnt used to date rocks at millions of years.
The RATE radiocarbon research first focused on demonstrating that significant detectable levels of carbon-14 are present in
ancient coal beds.1,2 Ten samples from U.S. coal beds, conventionally dated at 40320 million years old, were found to
contain carbon-14 equivalent to ages of around 48,00050,000 years. The laboratory did repeat analyses and confirmed
that this carbon-14 in the coals was not due to any contamination either in situ in the samples or added to the samples in the
laboratory. Of course, these would not be the true ages of these coal beds, because these 48,00050,000 year ages are
calculated at the present-day level and production rate of radiocarbon. The fact that all these coal beds yield radiocarbon
ages in the same ballpark is consistent with them all having been formed at the same time in a recent catastrophic event.
This is, of course, consistent with masses of pre-Flood vegetation being swept away and buried on a huge scale globally
during the cataclysmic Flood.Buoyed by this success, the RATE radiocarbon research next checked for carbon-14 in
diamonds. Diamonds are the hardest known natural substance and resist physical abrasion. Also, the chemical bonding of
the carbon in diamonds makes them highly resistant to chemical corrosion and weathering. Diamonds also repel and
exclude water from adhering to their surfaces, which would eliminate any possibility of the carbon in the diamonds becoming
contaminated. Sure enough, the diamonds submitted for radiocarbon analyses did contain detectable, significant levels of
carbon-14, equivalent to an age of around 55,000 years. Again, the laboratory did repeat analyses and discounted any
possibility that this carbon-14 was due to contamination, in situ to the diamonds or added in the laboratory. At 12 billion
years old, these diamonds, which are formed deep inside the earth, are regarded as being related to the earths early
history. Therefore, it was concluded that carbon-14 in these diamonds was consistent with a young age for the earth itself.
Confirmation that there is in situ carbon-14 in diamonds has now been reported in the conventional literature.3 R.E. Taylor of
the Department of Anthropology at the University of CaliforniaRiverside and of the Cotsen Institute of Archaeology at the
University of CaliforniaLos Angeles teamed with J. Southon at the Keck Accelerator Mass Spectrometry Laboratory of the
Department of Earth System Science at the University of CaliforniaIrvine to analyze nine natural diamonds from Brazil. All
nine diamonds are conventionally regarded as being at least of early Paleozoic age, that is, at least several hundred million
years old. So, if they really are that old they should not have any intrinsic carbon-14 in them. Eight of the diamonds yielded
radiocarbon ages of 64,900 years to 80,000 years. The ninth diamond was cut into six equal fragments, which were each
analyzed. They yield essentially identical radiocarbon ages ranging from 69,400 years to 70,600 years. This suggests the
carbon-14 was evenly distributed through this diamond, which is consistent with it being intrinsic carbon-14, and not
contamination. Interestingly, samples of Ceylon graphite from Precambrian metamorphic rock (conventionally around 1
billion years old) were analyzed at the same time and yielded radiocarbon ages of from 58,400 years to 70,100 years.
These results, from a different radiocarbon laboratory to that used by the RATE group, confirm that there is intrinsic carbon14 in natural diamonds. Therefore, they cannot be hundreds of millions or billions of years old, as there is no other current
credible explanation for the presence of this carbon-14. Less carbon-14 was found in the diamonds in this study reported in
the conventional literature. That was because the diamonds were mounted directly in the beam within the analytical
instrument, whereas in the RATE study the diamonds were combusted to convert the carbon to carbon dioxide, which was
then converted to graphite that was analyzed in the instrument. That process may have introduced some more carbon-14 to
the analyses.The University of California scientists, of course, did not conclude that the diamonds they analyzed are
evidence that the earth is young. Instead, they interpreted these 64,90080,000 year age to represent one component of
machine background in the analytical instrument. Yet this begs the question as to why then did the Precambrian graphite
contain on average more carbon-14 to yield younger ages than the diamonds? And why did the diamonds have such
different carbon-14 contents to yield different apparent radiocarbon ages? Because the same instrument was used to
analyze all the diamonds and the graphite, the results should surely have all been affected by the same machine
background. Rather, these results may further confirm the conclusions of the RATE radiocarbon project that natural
diamonds, which are related to the earths early history, show evidence of being only thousands of years old and provide
noteworthy support that the earth is young.

Geological Conflict
Young Radiocarbon Date for Ancient Fossil Wood Challenges
Fossil Dating
by Dr. Andrew A. Snelling on March 1, 2000
Originally published in Creation 22, no 2 (March 2000): 44-47.
Many do not realise that index fossils are still crucial to the millions-ofyears geological dating, in spite of the advent of radioactive 'dating'
techniques.
For most people, the discovery of fossilised wood in a quarry would
not be newsworthy. However, some pieces recently found embedded

in limestone alongside some well-known index fossils (see aside below) for the Jurassic period (supposedly 142205.7
million years ago) have proved highly significant.It is not generally realised that index fossils are still crucial to the millionsof-years geological dating, in spite of the advent of radioactive dating techniques. Not all locations have rocks suitable for
radioactive dating, but in any case, if a radioactive date disagrees with a fossil date then it is the latter which usually has
precedence.
Figure 1. Locality map showing the outcrop pattern of the Marlstone Rock Bed across southern and central England (ref. 1,
main article).
Finding this fossil wood in Jurassic limestone suggested the possibility of testing for the presence of radiocarbon ( 14C). Most
geologists, however, would not bother with such tests because they wouldnt expect any 14C to still exist. With a half-life of
only 5,570 years, no 14C should be detectable after about 50,000 years, let alone millions of years, even with the most
sensitive equipment. So this fossilised wood from the Marlstone Rock Bed of Jurassic age had potential for testing the
validity of the fossil dating technique underpinning modern geology.
The Marlstone Rock Bed
The Marlstone Rock Bed is a distinctive limestone unit that outcrops from Lyme Regis on the Dorset coast of southern
England, north-eastwards to just west of Hull near the North Sea coast (Figure 1).1 In many places, the top 530 cm (212
inches) or more of this bed has been weathered and altered, the original green iron minerals2 being oxidized to limonite
(hydrous iron oxides), and also in a few areas the sand content is higher. In the past, the outcrop has been quarried
frequently for iron ore or building stone.Evolutionary geologists consider that the top three metres (10 feet) of the Marlstone
Rock Bed represent the whole of the Tenuicostatum Zone, the basal zone of the Toarcian Stage,3 the last stage of the Early
Jurassic. This dating is based on the presence of the ammonite index fossil Dactylioceras tenuicostatum.4Thus the bed is
said to be about 189 million years old according to the geological time-scale.5Amongst the remaining quarries still working
the top of the Marlstone Rock Bed are the Hornton Quarries at Edge Hill near the village of Ratley, on the north-western
edge of the Edge Hill plateau, some 10 km (6 miles) north-west of the town of Banbury (Figures 2 and 3). Building stone,
known as Hornton Stone, has been quarried there since medieval times.6,7
A dating test at Hornton Quarries
During two visits to the Hornton Quarries, it was established that fossil wood occurs alongside ammonite and belemnite
index fossils (see aside below) in the Hornton Stone, the oxidized silty top of the Marlstone Rock Bed. The ammonite
recovered in the quarries is Dactylioceras semicelatum (Figure 4), abundant in a subzone of the Tenuicostatum Zone.1
Fossil wood was actually found sitting on top of a fossilised belemnite (Figure 5), probably belonging to the genus
Acrocoelites, a Toarcian Stage index fossil in north-west Europe.8 Many such belemnite fossils had been found during
quarrying operations (Figure 6). Together these index fossils have, in evolutionary reckoning, established the rock containing
them as being Early Jurassic and about 189 million years old.9,10 Logically, the fossil wood must be the same age.
Figure 2. Locality map showing the distribution of the
Marlstone Rock Bed west of Banbury, and the Hornton
Quarries at Edge Hill near the village of Ratley.
Three samples of fossil wood were collected from the
south wall of Hornton Quarries, one from immediately
adjacent to the belemnite fossil (Figure 5) during the
first visit, and two from locations nearby during the
second visit. All the fossil wood samples were from
short broken lengths of what were probably branches of
trees fossilised in situ. The woody internal structure was
clearly evident, thus the samples were not the remains
of roots that had grown into this weathered rock from
trees on the present land surface. When sampled, the
fossil wood readily splintered, diagnostic of it still being
woody in spite of its impregnation with iron minerals
during fossilisation.Pieces of all three samples were
sent for radiocarbon (14C) analyses to Geochron
Laboratories in Cambridge, Boston (USA), while as a
cross-check, a piece of the first sample was also sent to
the Antares Mass Spectrometry Laboratory at the
Australian Nuclear Science and Technology Organisation (ANSTO), Lucas Heights near Sydney (Australia). Both
laboratories are reputable and internationally recognised, the former a commercial laboratory and the latter a major research
laboratory.The staff at these laboratories were not told exactly where the samples came from, or their supposed evolutionary
age, to ensure that there would be no resultant bias.Both laboratories used the more sensitive accelerator mass
spectrometry (AMS) technique for radiocarbon analyses, recognised as producing reliable results even on samples with
minute quantities of carbon.

Figure 3(a) General view of the south wall of the Hornton Quarries at Edge Hill near Ratley,
north-west of Banbury.
(b) Closer view of the quarry face of the south wall showing the oxidized limestone of the
top of the Marlstone Rock Bed which is quarried as Hornton Brown building stone.
The results
The radiocarbon (14C) results are listed in Table 1. Obviously, there was detectable
radiocarbon in all the fossil wood samples, the calculated 14C ages ranging from 20,700
1,200 to 28,820 350 years BP (Before Present).
For sample UK-HB-1, collected from on top of the belemnite index fossil (Figure 5), the
results from the two laboratories are reasonably close to one another within the error

margins, and when averaged yield a 14C age almost identical (within the error margins) to the 22,730 170 years BP of
sample UK-HB-2.
Alternatively, if all four results on the three samples are averaged, the 14C age is almost identical (within the error margins)
to the Geochron result for UK-HB-1 of 24,005 600 years BP. This suggests that a reasonable estimate for the 14C age of
this fossil wood would be 23,00023,500 years BP.
Quite obviously this radiocarbon age is drastically short of the age of 189 million years for the index fossils found with the
fossil wood, and thus for the host rock.
Of course, uniformitarian geologists would not even test this fossil wood for radiocarbon. They dont expect any to be in it,
since they would regard it as about 189 million years old due to the age of the index fossils. No detectable 14C would
remain in wood older than about 50,000 years. Undoubtedly, they would thus suggest that the radiocarbon, which has been
unequivocally demonstrated to be in this fossil wood, is due somehow to contamination. Such a criticism is totally unjustified
(see aside two).
Conclusions
The fossil wood in the top three metres of the Marlstone Rock Bed near Banbury, England, has been 14C dated at 23,000
23,500 years BP. However, based on evolutionary and uniformitarian assumptions, the ammonite and belemnite index
fossils in this rock date it at about 189 million years. Obviously, both dates cant be right!
Furthermore, it is somewhat enigmatic that broken pieces of wood from land plants were buried and fossilised in a limestone
alongside marine ammonite and belemnite fossils. Uniformitarians consider limestone to have been slowly deposited over
countless thousands of years on a shallow ocean floor where wood from trees is not usually found.

Figure 4. The ammonite index fossil Dactylioceras semicelatum recovered from the top section of the Marlstone Rock Bed in
the Hornton Quarries at Edge Hill.

Figure 5. Fossil wood in the top section of the Marlstone Rock Bed exposed in the south wall of the Hornton Quarries at
Edge Hill. The pen is not only for scale, but points to an end-on circular profile of a belemnite fossil sitting directly
underneath the fossil wood (sampled as UK-HB-1).
However, the radiocarbon dating of the fossil wood has emphatically demonstrated the complete failure of the evolutionary
and uniformitarian assumptions underpinning geological dating.
A far superior explanation for this limestone and the mixture of terrestrial wood and marine shellfish fossils it contains is
extremely rapid burial in a turbulent watery catastrophe that affected both the land and ocean floor, such as the recent global
Flood.

Figure 6. Four belemnite fossils, probably Acrocoelites, recovered from the top section of the Marlstone Rock Bed in the
Hornton Quarries at Edge Hill (pen for scale). These cylindrical skeletal shells of the belemnites which taper to apices are
called rostrums (ref. 2, of Index fossils and geologic dating, aside below).The 23,00023,500 year BP 14C date for this fossil
wood is not inconsistent with it being buried about 4,500 years ago during the Flood, the original plants having grown before
the Flood.A stronger magnetic field before, and during, the Flood would have shielded the earth more effectively from
incoming cosmic rays,11 so there would have been much less radiocarbon in the atmosphere then, and thus much less in
the vegetation. Since the laboratories calculated the 14C ages assuming that the level of atmospheric radiocarbon in the
past has been roughly the same as the level in 1950, the resultant radiocarbon ages are much greater than the true
age.12,13Thus, correctly understood, this fossil wood and its 14C analyses cast grave doubts upon the index fossil dating
method and its uniformitarian and evolutionary presuppositions.
On the other hand, these results are totally consistent with the details of the recent global Flood,.
Index fossils and geologic dating
To evolutionary geologists, fossils are still crucial for dating strata, but not all fossils are equally useful. Those fossils that
seem to work well for identifying and dating rock strata are called index fossils.To qualify as an index fossil, a particular
fossil species must be found buried in rock layers over a very wide geographical area, preferably on several continents. On
the other hand, the same fossil species must have a narrow vertical distribution, that is, only be buried in a few rock layers.
The evolutionist interprets this as meaning that the species lived and died over a relatively short time (perhaps a few million
years). Therefore, the rock layers containing these fossils supposedly only represent that relatively short period of time, and
thus a date can be assigned accordingly on every continent to the rock layers where these fossils are found. The date
relative to other index fossils and rock layers is, of course, determined by the species position in the evolutionary tree of
life.14Among well-known index fossils are ammonites (extinct, coiled-shell cephalopods, marine molluscs similar to todays

Nautilus), and the belemnites (extinct, straight-shell cephalopods).15 Both are fossils of squid-like creatures, common to
abundant in so-called Mesozoic rocks. They are very important index fossils for dating and correlation of rock layers, for
example, across Europe, particularly for the so-called Cretaceous and Jurassic periods of the geological timescale,16,17 which are claimed to span 65142 and 142205.7 million years ago respectively.18 However, these index fossils
have not been dated directly by radioactive techniques.
Could the radiocarbon be due to contamination?
Four reasons why not
Pieces of the same sample were sent to the two laboratories and they both independently obtained similar results.
Furthermore, three separate samples were sent to the same laboratory in two batches and again similar results were
obtained. This rules out contamination.
The radiocarbon dates depend on the amounts of radiocarbon, originally in the living plants, now left in the fossil wood
samples. In these samples, the 14C left was between about 2.5% and 7.5% of the amount in living plants today. Any
unavoidable contamination (e.g., dust, fungal spores) would be minuscule and would amount to at most 0.2%, which would
have a negligible effect on these radiocarbon dates.19
The last column in Table 1 lists the d13CPDB results,20 which are consistent with the analysed carbon in the fossil wood
representing organic carbon from the wood of land plants.21
Such a claim would, by implication, cast a slur on the Ph.D. scientific staff of two radiocarbon laboratories, who, as qualified
routine practitioners, understand the potential for contamination and how to avoid it in sample processing.
Return to text.
Table 1.
14

C 'AGE' (YEARS BP)

d13CPDB

SAMPLE LAB

LAB CODE

Geochron
UK-HB-1 ANSTO

GX-21666-AMSOZC201

24,005

20,700 1,200

UK-HB-2 Geochron

GX-21611-AMS

22,730 170

-24.0

UK-HB-3 Geochron

GX-21612-AMS

28,820 350

-25.3

600-22.9
-16.6

14

Radiocarbon ( C) analytical results for fossil wood samples, Marlstone Rock Bed, Hornton Quarries, England.
Stumping Old-Age Dogma
Radiocarbon in an ancient fossil tree stump casts doubt on traditional rock/fossil dating
by Dr. Andrew A. Snelling on September 1, 1998
Originally published in Creation 20, no 4 (September 1998): 48-51.
When it comes to the dating of sedimentary rocks, the fossils in them are of paramount importance.
Shop Now
In the words of the late Derek Ager, fossils have been and still are the best and most accurate method of dating and
correlating the rocks in which they occur 1
Why is this so?
Dating rocks and fossils
Figure 1. Click here for a larger image. Location map for the Newvale No. 2 Coal Mine, north of
Sydney on the east coast of Australia.
Uniformitarian geologists and evolutionary paleontologists believe that as countless creatures lived and
died over millions of years of Earth history, some were buried in slowly accumulating sediments and
then fossilised. Accumulated genetic changes over millions of years supposedly resulted in
the evolution of new species, genera and families. So when fossils are found in sedimentary rock
layers, they are identified within the context of where they fit in the evolutionary tree of life, and a
millions-of-years age is therefore assigned to the fossil and the rock accordingly.2
In recent years a variety of techniques have been developed to date some rocks and minerals using
the decay of radioactive elements in them. These methods include potassium-argon, rubidiumstrontium, uranium-thorium-lead and samarium-neodymium dating. They are used, for example, on
layers of volcanic rocks above and below fossil-bearing sedimentary rock layers. Thus these methods, though not directly
dating the fossils, have often confirmed the millions of years ages assigned to the rocks and fossils by their interpretation
within the uniformitarian and evolutionary framework.
The only radioactive dating method that could be directly applied to many fossils is radiocarbon or carbon-14 ( 14C) dating.
However, because radiocarbon decays relatively rapidly, it is only useful in practice up to about 50,000 years. 3 Thus most
fossils, being regarded as millions of years old, are never tested for radiocarbon, because they are not supposed to have
any left.
A fossilised tree stump
Figure 2. General view of the fossilised tree stump (scale bar in cm). The shiny coalified
bark around the perimeter can be clearly seen.
Figure 3. Click here for a larger image. The local geological column for the upper
portion of the Newcastle Coal Measures showing where the Great Northern coal seam
occurs.
Among the sedimentary rock layers, some of the most significant are coal beds. These
can be tens of metres thick and stretch for many hundreds of square kilometres. They
consist of the broken remainsleaves, twigs, bark, logs, etc.of countless millions of
trees of many varieties. Usually the process of fossilising this vegetation debris
obliterates most of the recognisable features of these individual components, as the
whole buried mass is transformed into coal.However, sometimes components can be
identifiedfor example, fossilised tree stumps sitting on top of coal beds. Such a
fossilised tree stump was found by miners in the Newvale No. 2 (underground) Coal
Mine north of Sydney, Australia (Figure 1). A portion of it was saved by one of the
miners (Figure 2).One of the major coal beds exploited in the Newvale No. 2 Coal

Mine is the Great Northern Seam, near the top of the sequence of rock units known collectively as the Newcastle Coal
Measures (Figure 3) within the Sydney Basin. Based on the plant fossils found in them, these coal beds (including the
associated mudstone in which the stump was found) have been designated Upper Permian, which uniformitarian geologists
would therefore assign to a period of Earths history around 250 million years ago.4,5
Figure 4 shows the relative position of the fossilised tree stump when it was found, surrounded by a 150 mm (almost 6
inches) thick layer of mudstone sitting directly on top of the coal (Great Northern Seam). That portion of the fossilised tree
stump recovered has a diameter of 110 mm (almost 4 inches) and stands 100 mm (about 4 inches) high (Figure 2 and
Figure 5). A shiny thin skin encompasses the outer perimeter (Figure 2, Figure 5 and Figure 6) and represents the original
tree bark, which upon burial was coalified. In contrast, the former wood has been silicified (literally turned to stone by
impregnation with silica), though it is dull black from still being carbon-rich (Figure 6).
Sample Type

Lab Code

14

coalified bark

GX-21867

33,000 400

-27.2

silicified wood

GX-22613

>48,000

-26.7

C Age (years BP)

13CPDB

Table 1. Radiocarbon (14C) analyses of samples from the fossilised tree stump
Radiocarbon (14C) analyses
Figure 4. Click here for a larger image. Diagram to illustrate the relative position of the fossilised tree
stump sitting on top of the Great Northern coal seam (not drawn to scale).
Small pieces of the coalified bark and the silicified wood immediately underneath it were sent for
radiocarbon (14C) analyses to Geochron Laboratories in Cambridge, Boston (USA), a reputable,
internationally-recognized commercial laboratory. The laboratory staff were not told exactly where the
samples came from, or their supposed evolutionary age, to ensure that there would be no resultant
bias. This laboratory uses the more sensitive accelerator mass spectrometry (AMS) technique for
radiocarbon analysis, now recognized as producing the most reliable results, even on minute quantities
of carbon in samples.The radiocarbon results are listed in Table 1. There was detectable radiocarbon in
the coalified bark, yielding a supposed 14C age of 33,700 400 years BP (before present). On the
other hand, the small quantity of carbon extracted from the silicified wood sample was insufficient to
yield a finite 14C age, so the result could only be reported as >48,800 years BP, beyond the detection limits. Of course, the
wood inside a tree stump would not be >15,100 years older than the bark enclosing it. So the 14C age of the bark places an
age limit on the immediately underlying silicified wood.
Of course, if the wood really were 250 million years old as is supposed, one should not be able to obtain a finite age from
radiocarbonall detectable 14C should have decayed away in a fraction of that alleged time.
Objections
Figure 5. Another view of the fossilised tree stump (scale bar in cm). The shiny
coalified bark can again be clearly seen, as can the dull silicified wood within the
stump.The most obvious objection that might be raised against these radiocarbon
results by sceptics uncomfortable with the implications is that the minute quantity of
radiocarbon detected in this fossilised tree stump is due to contamination. 6 Such a
criticism is unjustified, and by implication casts a slur on the radiocarbon laboratorys
Ph.D. scientific staff. As qualified routine practitioners, they understand the problem
with contamination, and how to avoid it in sample preparation. Yet they reported these
analyses as genuine in situ radiocarbon (14C). Furthermore, the last column of Table 1
lists the d13CPDB results, which are consistent with the analysed carbon in the
fossilised tree stump representing organic carbon from wood, not from
contamination. 7Another objection is that acceptance of these results as genuine 14C
ages is based on bias, incompetence or ignorance.8 However, those who would
make such accusations in reality reject these results primarily because such 14C
ages cannot possibly be obtained from a fossilised tree stump sitting in a layer of
250 million years old Upper Permian mudstone. Of course, such pronouncements
are based solely on a rock-and-fossil dating scheme derived from evolutionary and
uniformitarian beliefs, not from some independent, objective scientific standard.
Conclusions
Figure 6. A close-up of the fossilised tree stump (scale bar in cm). Not only can the
shiny coalified bark be seen, but also the growth rings in the dull silicified interior wood.
Within the Creation/Flood framework of Earth history, the Flood occurred about 4,500
years ago. Therefore, even though this tree stump must have grown before the Flood (to
be then buried and fossilised in sediments laid down by the Flood) there cannot have
been more than about 5,000 years at most since it died.
However, a 33,700 400 years BP radiocarbon age for this fossilised tree stump is
neither inconsistent nor unexpected. A stronger magnetic field before, and during, the
Flood would have shielded the Earth more strongly from incoming cosmic rays,9 so there
would have been much less radiocarbon in the atmosphere then, and thus much less in
the vegetation. Since the laboratory calculated the 14C age based on the assumption that
the level of atmospheric radiocarbon in the past has been roughly the same as the level in
1950, the resultant radiocarbon age is much greater than the true age.10
On the other hand, a 33,700 400 years BP radiocarbon age emphatically conflicts with,
and casts doubt upon, the evolutionary fossil and uniformitarian rock age of 250 million
years for this fossilised tree stump. Clearly, the radiocarbon dating method, although
demonstrating that the specimen cannot be millions of years old, has not provided its true
age.11 However, correctly understood, this radiocarbon analysis is totally consistent with
the account of a young Earth and a recent global Flood.

***
Determination of the Radioisotope Decay Constants and Half-Lives: Lutetium-176 ( 176Lu)
by Dr. Andrew A. Snelling on December 3, 2014
Abstract
Over the last 75 years numerous determinations have been made of the 176Lu half-life and decay constant. However, even
though the measurement technology has improved, the determinations over the last 30 years have resulted in unexplained
and unresolved discrepancies. Direct physical counting experiments have resulted in two separate groupings of results with
different 176Lu mean half-life values. Early determinations based on comparisons of ages of meteorites gave different 176Lu
half-life and decay constant values than determinations based on comparisons of ages of terrestrial minerals and rocks. But
more recent determinations using meteorites have yielded 176Lu half-life and decay constant values that agree with both
determinations based on comparisons of ages of terrestrial minerals and rocks, and with one of the groupings of
determinations by direct physical counting experiments. Thus the 176Lu half-life and decay constant values of 37.12 Byr and
1.867 10-11 per year respectively have now been generally adopted for standard use by the geological community, based
particularly on the determinations using comparisons of ages of terrestrial minerals and rocks. The more recent meteorite
determinations used chondrite meteorites rather than the eucrite meteorites used for the earlier determinations, because it
was realized the latter have suffered thermal and shock metamorphism that resulted in significant disturbance of the Lu-Hf
systematics among mineral phases due to open geochemical system behavior, such as leakage of 176Hf. Yet the
discrepancies in the determinations by the direct physical counting experiments remain unexplained and unresolved.
Furthermore, all the determinations using age comparisons on terrestrial minerals and rocks and most of the more recent
determinations using age comparisons on chondrite meteorites have been calibrated against the U-Pb method, but even
this gold standard has unresolved uncertainties due to variations measured in terrestrial rocks and minerals and
meteorites of the 238U/235U ratio which is so critical to the method. So the U-Pb method should not be used as a standard to
determine other decay constants. This only serves to highlight that if the Lu-Hf dating method has been calibrated against
the U-Pb gold standard with its own uncertainties, then it cannot be absolute, and therefore it cannot be used to reject the
young-earth creationist timescale. Indeed, current radioisotope dating methodologies are at best hypotheses based on
extrapolating current measurements and observations back into an assumed deep time history for the cosmos.
Shop Now
Keywords: radioisotope dating, decay constant, half-life, lutetium-176,176Lu, decay, direct physical counting, -rays,
geological age comparisons, terrestrial minerals and rocks, meteorites, eucrites, chondrites, discrepancies, U-Pb gold
standard, 238U/235U ratio.
Introduction
Radioisotope dating of rocks and meteorites is perhaps the most potent claimed proof for the supposed old age of the earth
and the solar system. The absolute ages provided by the radioisotope dating methods provide an apparent aura of certainty
to the claimed millions and billions of years for formation of the earths rocks.However, accurate radioisotopic age
determinations require that the decay constants of the respective parent radionuclides be accurately known. Ideally, the
uncertainty of the decay constants should be negligible compared to, or at least be commensurate with, the analytical
uncertainties of the mass spectrometer measurements entering the radioisotope age calculations (Begemann et al. 2001).
Clearly, based on the ongoing discussion in the conventional literature this is not the case at present. The stunning
improvements in the performance of mass spectrometers during the past four or so decades, starting with the landmark
paper by Wasserburg et al. (1969), have not been accompanied by any comparable improvement in the accuracy of the
decay constants (Begemann et al. 2001; Steiger and Jger 1977), in spite of ongoing attempts (Miller 2012). The
uncertainties associated with direct half-life determinations are, in most cases, still at the percent level at best, which is still
significantly better than any radioisotope method for determining the ages of rock formations. The recognition of an urgent
need to improve the situation is not new (for example, Min et al. 2000; Renne, Karner, and Ludwig 1998). It continues to be
mentioned, at one time or another, by every group active in geo- or cosmochronology (Schmitz 2012).From a creationist
perspective, the 19972005 RATE (Radioisotopes and the Age of The Earth) project successfully made progress in
documenting some of the pitfalls in the radioisotope dating methods, and especially in demonstrating that radioisotope
decay rates may not have always been constant at todays measured rates (Vardiman, Snelling, and Chaffin 2000, 2005).
Yet much research effort remains to be done to make further in-roads into not only uncovering the flaws intrinsic to these
long-age dating methods, but towards a thorough understanding of radioisotopes and their decay during the earths history
within a creationist framework.One crucial area the RATE project did not touch on was this issue of how accurate and
reliable are the determinations of the radioisotope decay rates (decay constants and half-lives), which are so crucial for
calibrating these dating clocks. The reliability of the other two assumptions these absolute dating methods rely on, that is,
the starting conditions and no contamination of closed systems, are unprovable. Yet these can be circumvented somewhat
via the isochron technique, because it is independent of the starting conditions and is sensitive to revealing any
contamination. This is especially the case when mineral isochrons are used. Data points that do not fit on the isochron are
simply ignored because their values are regarded as due to contamination. That this is common practice is illustrated with
numerous examples from the relevant literature by Faure and Mensing (2005) and Dickin (2005). On the other hand, it could
be argued that this discarding of data points which do not fit the isochron is somewhat arbitrary and therefore is not good
science, because it is merely assumed aberrant values are due to contamination rather than that being proven to be so.
In order to rectify this deficiency, Snelling (2014) documented the methodology behind and history of determining the decay
constant and half-life of the parent radioisotope 87Rb used as the basis for the Rb-Sr long-age dating method. He showed
that there is still some uncertainty in what the values for these measures of the 87Rb decay rate should be, and that the
determined values differ when Rb-Sr ages are calibrated against the U-Pb ages of either the same terrestrial minerals and
rocks or the same meteorites and lunar rocks. Ironically it is the slow decay rates of isotopes such as 87Rb used for deep
time dating that makes precise measurements of their decay rates so difficult. Thus it could be argued that direct
measurements of their decay rates should be the only acceptable experimental evidence, especially because
measurements which are calibrated against other radioisotope systems are already biased by the currently accepted
methodology employed by the secular community in their rock dating methods. We thus need to further explore just how
accurate these determinations are for other parent radioisotopes, whether there really is consensus on standard values for
their half-lives and decay constants, and just how independent and objective their standard values are from one another
between the different long-age dating methods. Of course, it is to be expected that every long-lived radioactive isotope is
likely to show similar variation and uncertainty in half-life determinations because these are difficult measurements to make.
However, even small variations and uncertainties in the half-life values result in large variations and uncertainties in the
calculated ages for rocks, and the question remains as to whether the half-life values for each long-lived parent radioisotope

are independently determined. Here we continue these investigations by exploring the determinations of the lutetium-176
(176Lu) decay rate, which is the basis for the Lu-Hf method.
Lutetium and Lutetium-176 Decay
With an atomic number of 71, lutetium is therefore element 71, which places it in the sixth period of the periodic table. It is
the last of the lanthanide series and thus is the heaviest of the rare earth elements (REEs). It has two naturally occurring
isotopes whose abundances are 97.4% 175Lu and 2.6% 176Lu, so that the 175Lu/176Lu ratio is 37.46 (Dickin 2005). There has
been some disagreement over this ratio, but the value of 37.701 0.028 experimentally determined by Patchett and
Tatsumoto (1980b) has been used exclusively in all subsequent research. However,the differences in the values obtained
between that and the few other determinations remain unexplained.The geochemical properties of Lu are similar to those of
the REE Sm (samarium). Lutetium has a valence of 3+ and an ionic radius of 0.93 , the latter being similar to that of
Ca2+ at 0.99 . This causes Lu3+ to be captured by crystals in place of Ca2+. Lutetium is thus present and widely dispersed in
all types of rocks, but concentrations rarely exceed 0.5 ppm and so it does not form its own minerals in most geological
environments. The average concentrations of Lu in ordinary rock-forming silicate minerals are generally low, so the average
Lu concentrations in igneous rocks increase very little with the increasing degree of differentiation from basalt to granite
(Faure and Mensing 2005). The most important mineral carriers of Lu in the common rock types are apatite, garnet, and
biotite, plus the rarer alkali-rich pyroxene aegirine and amphibole arfvedsonite. Elevated Lu concentrations are present in Zrbearing minerals such as zircon, baddeleyite and eudialyte, but the highest Lu concentrations are found in rare-earth oxides
such as euxenite, carbonates such as bastnaesite, phosphates such as xenotime and monazite, and silicates such as
gadolinite and allanite. All such rare-earth minerals are relatively rare and occur primarily in complex pegmatites, alkali-rich
intrusives, and carbonatites. Thus these rare-earth minerals are not important for dating by the Lu-Hf method, but their
presence as accessory minerals makes their host rocks suitable for such dating.Lutetium-176 is radioactive and displays
branched isobaric decay by beta () emission to stable 176Hf and by electron capture to stable 176Yb. The frequency of
electron captures is of the order of 3% or less (Dixon, McNair, and Curran 1954; Glover and Watt 1957), so in view of the
long half-life of 176Lu and the current estimate of the frequency at only 0.095% (http://www. nucleonica.net/unc.aspx) the
slow decay to 176Yb can be usually ignored. Thus current estimates for the decay constant of 176Lu are solely based on its decay to 176Hf, ignoring any potential electron capture. 176Hf is left in an excited state after the -emission, and decays to the
ground state by -emission. The relevant decay scheme is therefore depicted as:
176
Lu 176Hf + - + + Q
where - is a -particle, is an anti-neutrino, and Q is the decay energy.
The first Lu-Hf geochronological measurement was made by Herr et al. (1958), who attempted to determine the 176Lu halflife by analyzing the isotopic composition of Hf in the heavy-REE-rich mineral gadolinite. Boudin and Deutsch (1970) were
the first to determine the 176Lu half-life by dating Lu-bearing minerals of known age, while Owen and Faure (1974)
attempted to use the method to date common rocks and minerals, but had trouble with the isotopic analysis of Hf due to the
difficulties of its chemical separation and its poor ionization efficiency during thermal-ionisation mass spectrometry (TIMS).
However, these problems were finally overcome by Patchett and Tatsumoto (1980a) with a modified analytical technique, so
after this breakthrough the Lu-Hf method became useful in the dating of terrestrial rocks and minerals, and of meteorites.
There are two parameters by which the decay rate is measured
and expressed, namely, the decay constant () and the half-life
(t). The decay constant can be defined as the probability per unit
time of a particular nucleus decaying, whereas the half-life is the
time it takes for half of a given number of the parent radionuclide
atoms to decay. The two quantities can be almost used
interchangeably, because they are related by the equation:t = ln 2 / = 0.693 /
Thus here we will simply focus on the determinations of the 176Lu
half-life.
Determination Methods
Two approaches have so far been followed to determine the decay half-life of the long-lived radioactive 176Lu.
Direct counting
Except for the very first attempts, direct determinations of the halflife of 176Lu have been by - and --coincidence counting.
Fig. 1. The
decay
scheme
of 176Lu as
initially
McNair, and
ground
state.In the
activity
material
McNair, and
Flynn, and
radioactive
Avogadros
same time

proposed (after Arnold and Sugihara 1953; Dixon,


Curran 1954). In this figure there is a 10 next to the
state of 176Lu, but this is a spin 7, negative parity
--coincidence counting technique, the beta ()
of 176Lu was counted in a suitable solid Lu2O3 source
using a proportional tube spectrometer/counter (Dixon,
Curran 1954; Donhoffer 1964; McNair 1961; Prodi,
Glendenin 1969), and divided by the total number of
atoms in the known quantity of Lu, based on
number and the isotopic abundance of 176Lu. At the
a low-level NaI (sodium iodide) scintillation
spectrometer/counter doped with Tl was used to detect
the
-ray
spectrum and to measure the energy peaks of the rays
produced by the energy transitions of the
daughter 176Hf as it decays from its excited state to its
ground
state (Arnold 1954; Brinkman, Aten, and Veenboer 1965; Glover and Watt 1957) (fig. 1). Thus the detection of the number of
daughter 176Hf atoms produced can be compared with the number of -particles counted from the parent 176Lu decay. Among
the difficulties of this approach, and with just -particle counting, are problems with detector efficiencies and geometry
factors, the self-shielding or self-absorption of the finite-thickness solid samples, the stopping power of the crystal source

and the fraction of pulses appearing in the photo peak (typically included in the detection efficiency for scintillation and
solid state detection systems). Furthermore, the escape of I (iodine) K electron shell x-rays is an important factor in the size
of the lowest -ray energy peak, which would only be a secondary check against the two primary -rays, that is, the 201 and
306.9 keV -rays, for determining the amount of 176Lu present (Arnold 1954; Begemann et al. 2001; Brinkman, Aten, and
Veenboer 1965; Dixon, McNair, and Curran 1954; Flammersfeld and Mattauch 1943).In the --coincidence counting
technique, one or more Ge detectors or a -ray spectrometer is used to simultaneously measure the -ray energy peaks
corresponding to those produced by the cascading energy states of the daughter 176Hf atom (fig. 2). Since each
daughter 176Hf atom produced by -decay from each parent 176Lu atom cascades down through these energy levels to reach
its ground state, measuring each -ray energy peak should effectively count the same number of daughter 176Hf atoms
produced in the given time of the experiment, which only after correction for the number of decays which produce a given ray, namely the branching ratio for each , is equivalent to the 176Lu decay rate. Both Sguigna, Larabee, and Washington
(1982) and Grinyer et al. (2003) set out the mathematical principles by which the 176Lu half-life is calculated from the -ray
counts in each energy peak. In order to increase the counting rate, Sguigna, Larabee, and Washington (1982) used three
Ge-Ge detectors placed at 120 to each other around their Lu2O3 sample and shielded from each other to prevent scattered
coincidences, whereas Grinyer et al. (2003) used an 8 -ray spectrometer consisting of 20 HPGe -ray detectors. Because
of the better resolution of the solid state detectors, these determinations would tend to be trusted more than those
determinations made with the combination of proportional chamber/NaI (Tl doped) detectors (see below in the discussion).
Fig. 2.The decay scheme of 176Lu as currently proposed (after Firestone and Shirley 1996; Grinyer et al. 2003).Gehrke,
Casey, and Murray (1990), Dalmasso, Barci-Funel, and Ardisson (1992), and Nir-El and Lavi (1998) have all highlighted the
difficulties in this approach, which has produced a wide range of half-life results. These difficulties include the calibration of
the detector efficiencies, variations in response of detectors to different parts of the source sample, and corrections for -ray
self-attenuation in the Lu solid source material, true-coincidence summing depending on the source samples distance from
the detector, and internal conversion. However, in their experiment Grinyer et al. (2003) arranged the twenty detectors in the
8 -ray spectrometer so each viewed approximately 13% of the solid angle, and so that angular correlation effects were
minimized. Furthermore, they introduced to their mathematical treatment of their data a lumped efficiency parameter, which
was the probability per decay that a -ray photo-peak event was detected, and included the effects of internal conversion,
solid angle coverage, photo-peak efficiency, and self-attenuation of the source sample. This lumped efficiency parameter
thus appears to have been a model that was applied to the real experimental data. Grinyer et al. (2003) also applied another
small correction factor to represent the probability that another -ray (or the x-ray following internal conversion) entered the
detector at the same time, thus destroying a photopeak event that should have been counted. These are generally termed
peak summing corrections.Nevertheless, judged from the fact that many of the direct counting experiments have yielded
results that are not compatible with one another within the stated uncertainties (see below), it would appear that not all the
measurement uncertainties are accounted for in whatever correction factors have been used, and therefore the stated
uncertainties are unrealistically small and typically are underestimated. It can therefore be argued that many of such
experiments are plagued by unrecognized systematic errors (Begemann et al. 2001). As the nature of these errors is
obscure, it is not straightforward to decide which of the, often mutually exclusive, results of such direct counting experiments
is closest to the true value. Furthermore, the presence of unknown systematic biases makes any averaging dangerous. It is
possible that reliable results of careful workers, listing realistic uncertainties, will not be given the Fig. 2. The decay scheme
of 176Lu as currently proposed (after Firestone and Shirley 1996; Grinyer et al. 2003). weights they deservethis aside from
the question of whether it makes sense to average numbers that by far do not agree within the stated uncertainties.
Geological comparisons of methods
The second approach to the determination of the 176Lu decay half-life has been to Lu-Hf date geological samples whose
ages have also been measured by other methods with presumably more reliable decay constants, particularly the U-Pb and
Pb-Pb methods (Dickin 2005; Faure and Mensing 2005). This approach has the disadvantage that it involves geological
uncertainties, such as whether all isotopic systems closed at the same time and remained closed. However, it is claimed to
still provide a useful check on the direct laboratory determinations. In this respect it is worth noting that Boudin and Deutsch
(1970) proposed a 176Lu half-life of 33 Byr on the basis of Lu-Hf dating of two minerals that had also been U-Pb dated,
essentially the same as the 176Lu half-life value of 32.7 Byr determined via direct counting at the same time by Prodi, Flynn,
and Glendenin (1969).
This approach entails multi-chronometric dating of terrestrial rocks or a mineral (or minerals) from them (for example,
Scherer, Mnker, and Mezger 2001; Sderlund et al 2004), or meteorites and a mineral from them (for example, Amelin
2005; Patchett et al. 2004), and cross-calibration of different radioisotopic age systems by adjusting the decay constant of
one system so as to force agreement with the age obtained via another dating system (Begemann et al. 2001). In essence,
because the half-life of 238U is regarded as the most accurately known of all relevant radionuclides, this usually amounts to
expressing ages in units of the half-life of 238U. This has increasingly become the preferred method for determining the halflife of 176Lu.
Results of the Lutetium-176 Decay Determinations
During the last 75 years numerous determinations of the 176Lu decay half-life have been made using these two methods.
The results are listed with details in Table 1. The year of the determination versus the value of the half-life is plotted in Fig. 3.
The data points plotted have been color-coded to differentiate the values as determined by the two approaches that have
been useddirect - and --coincidence counting, and geological comparisons with other radioisotope dating methods.
Table 1. Determinations of the 176Lu decay rate expressed in terms of the half-life using direct physical counting
experiments, and comparisons of radioisotope ages of terrestrial minerals and rocks, and meteorites.
Determination of the 176Lu Decay Rate
Date Half-Life (Byr) Error

Method

Source

1938 40

direct counting

Heyden and Wefelmeier 1938

1943 24

direct counting

Flammersfeld and Mattauch


1943
Arnold 1954

1954 21`.5

direct counting

1954 45.6

direct
(scintillation/spectrometer)

1957 21

direct counting (mass spectrometer)

Notes

countingDixon, McNair, and Curran


1954
Glover and Watt 1957

1958 21.7

3.5

direct counting

Herr et al. 1958

1961 36

direct counting (spectrometer)

McNair 1961

1964 21.8

0.6

direct counting (liquid scintillation)

Donhoffer 1964

1965 35

1.4

direct counting (liquid scintillation)

Brinkman, Aten, and Veenboer


1965

1965 35.4

0.5

direct counting (liquid scintillation)

Brinkman, Aten, and Veenboer


1965

1965 36.8

0.6

direct counting (liquid scintillation)

Brinkman, Aten, and Veenboer


1965

1967 50

direct counting (spectrometer)

Sakamoto 1967

1969 32.7

0.5

direct counting (liquid scintillation)

Prodi, Flynn, and Glendenin


1969

1970 33

comparison of Lu-Hf and U-Pb ages


of two minerals
Boudin and Deutsch 1970

1972 37.9

0.3

direct counting

Komura,
Tanaka
1972

1980 40.8

2.4

direct counting

Norman 1980

1.4

comparison of Lu-Hf isochron of


eucrite meteorites with Rb-Sr, U-Pb,
and Sm-Nd ages
Patchett and Tatsumoto 1980b

1.4

Tatsumoto,
comparison of Lu-Hf and U-Pb ages Patchett
of Antarctic meteorites
1981

1980 35.3

1981 35.7

Sakamoto,

Unruh,

and

and

comparison of Lu-Hf, Rb-Sr, and


U-Pb ages of the Amtsoq Gneisses
Pettingill and Patchett 1981

1982 35.7
1982 35.9

0.5

direct counting

Sguigna,
Larabee,
Waddington 1982

1983 37.8

0.1

direct counting

Sato, Ohoka, and Hirose 1983

1990 40.5

0.9

direct counting (spectrometer)

Gehrke, Casey, and Murray


1990

1992 37.3

0.5

direct counting

Dalmasso, Barci-Funel, and


Ardisson 1992

1998 36.9

0.2

direct counting

Nir-El and Lavi 1998

comparison of methods
Tatsumoto et al 1981)

2001 35

and

(adjusting
Begemann et al. 2001

0.46

comparison of Lu-Hf and


isochron ages of four minerals

2002 35.9

0.5

comparison of Lu-Hf and Sm-Nd


ages of eucrite meteorites (based on
Sguigna, Larabee, and Waddington
1982)
Blichert-Toft et al. 2002

2003 34.95

0.21

comparison
of
meteorites
Allende Pb-Pb ages

2003 40.8

0.3

direct counting (spectrometer)

Grinyer et al. 2003

2003 36.77

0.75

direct counting (spectrometer)

Nir-El and Haquin 2003

2004 37.12

0.09

comparison of Lu-Hf and U-Pb ages


of Precambrian mafic intrusions
Sderlund et al. 2004

2004 37.12

0.09

comparison of Lu-Hf and


ages of chondrite meteorites

2005 37.18

0.43

comparison of Lu-Hf and U-Pb ages


of phosphates from meteorites
Amelin 2005

for Richardton (H5)

2005 37.83

0.24

comparison of Lu-Hf and U-Pb ages


of phosphates from meteorites
Amelin 2005

for Acapulco

2001 37.16

U-PbScherer, Mnker, and Mezger


2001

with
Bizzarro et al. 2003

Sm-Nd
Patchett et al. 2004

2006 38.3

0.4

mean values

Albarde et al. 2006

for physical counting


experiments

2006 37.12

0.05

mean values

Albarde et al. 2006

for age comparisons of


terrestrial minerals

2006 35.39

0.16

mean values

Albarde et al. 2006

for age comparisons of


meteorites

0.12

Bouvier,
comparison of Lu-Hf and U-Pb ages Patchett
of chondrite meteorites
2008

2008 36.78

Vervoort,

and

Discussion
Since the early 1960s the reported half-life determinations have scattered around 37 Byr, although there are extreme
outliers on the low side at 21.8 Byr (Donhoffer 1964), as well as on the high side at 50 Byr (Sakamoto 1967) (table 1 and fig.
3). However, during the last three decades since 1980 the scatter has been very much reduced. So that the differences and
trends in the data can be more easily seen, the relevant section of Fig. 3 has been expanded in Fig. 4. The values of 40.8
2.4 Byr of Norman (1980) and 40.5 0.9 Byr of Gehrke, Casey, and Murray (1990), the latter obtained on a sample enriched
in 176Lu to 44%, are more than 3 above each of the measurements preceding and following this work (table 1). Nir-el and
Lavi (1998) suggested the adoption of a half-life of 37.3 0.1 Byr, which is the weighted mean of the half-life determinations
of 37.8 0.01 Byr (Sato, Ohoka, and Hirose 1983), 37.3 0.05
Byr (Dalmasso, Barci-Funel, and Ardisson 1992), and their own
36.9 0.02 Byr. Nir-el and Lavi (1998) did not explain their
selection criteria, except that the adopted value should be
calculated from the grouping of values in the range 3738 Byr.
Taken at face value, this criterion disqualifies their own result of
36.9 Byr, but makes eligible that of Komura, Sakamoto, and
Tanaka (1972) of 37.9 Byr which they did not consider. These
half-life differences do not at first glance appear to be really
significant in view of the value being in excess of 30 billion
years. However, even a small difference in the half-life can
mean a huge difference when it is used to calculate the age of
a rock, on the order of tens of millions of years,

Fig. 3.Plot of each 176Lu half-life determination versus the year


of its determination, color-coded according to the method of its
determination.
Fig. 4.Enlarged plot of the 176Lu half-life determinations since
1980 versus the year of their determination, color-coded
according to the method of their determination.
Patchett and Tatsumoto (1980b) produced a Lu-Hf isochron for
eucrite meteorites, which based on several lines of good
evidence are regarded to have all come from the same
asteroid 4-Vesta that supposedly differentiated at about 4.55
Ga (McSween et al. 2013). Using this known age, the Lu-Hf
isochron yielded a decay constant of 1.96 0.08 10 -11per year, with uncertainty at the 95% confidence level, which is
equivalent to a half-life of 35.3 1.4 Byr. This was subsequently updated by Tatsumoto et al. (1981) to 1.94 0.07 10 11
per year (a half-life of 35.7 1.4 Byr) by the addition of three more eucrite meteorite analyses to their Lu-Hf isochron.
However, Begemann et al. (2001) argued that because some of the eucrites, notably those at the higher end of their Lu-Hf
isochron, have been interpreted as having an age of formation that is 0.1 Ga younger than the main population (Mittlefehldt
et al. 1998), it could be argued that
a 176Lu decay constant of 1.98 10 11
per year (a half-life of 35 Byr) is
the best value resulting from the
Tatsumoto, Unruh, and Patchett
(1981) study.
In subsequent cosmochronology,
geochronology and corresponding
chemical evolutionary studies, the
decay constant of 1.94 10-11 per
year (half-life of 35.7 Byr) from
Tatsumoto, Unruh, and Patchett
(1981) was used from 1981 to 1997.
This half-life of 35.7 Byr is 4% lower
than the optimum value of 37.3 0.1
Byr suggested by Nir-el and Lavi
(1998). Blichert-Toft and Albarde
(1997) analyzed a number of
chondritic meteorites for their Lu-Hf
isotope systematics, and redefined
the meteoritic reference parameters

for Hf isotope evolution in rocky planets and asteroids. They used the decay constant of 1.93 10 -11 per year (a half-life of
35.9 Byr) from Sguigna, Larabee, and Waddington (1982). This value is so similar to that of Tatsumoto, Unruh, and Patchett
(1981) that this switch had only a small effect on Hf isotopic studies of the earth and other planetary samples. However, the
discrepancy of ~4% between half-lives from physical measurements and from meteorite radioisotope ages remained, as
well as the dispersion in all determinations, so this problem still needed to be addressed in future investigations.Because
naturalists postulate the origin of the elements via nucleosynthesis in stars so that they comment on Lu decay in that
context, some explanations of terms they use are warranted here. The s-process (or slow-neutron-capture-process) is the
nucleosynthesis process that evidently occurs at relatively low neutron density and intermediate temperature conditions in
stars. Under these conditions heavier nuclei are created by neutron capture, increasing the atomic weight of the nucleus
by one. A neutron in the new nucleus then decays by - decay to a proton, creating a nucleus of higher atomic number. This
s-process is thus secondary, meaning that it requires preexisting heavy isotopes as seed nuclei to be converted into other
heavy nuclei. The rate of neutron capture by atomic nuclei is slow relative to the rate of radioactive - decay, hence the
name. Although considerable variability exists, it is estimated that the current time between successive neutron captures is
about 100 years, whereas the time for decay is about one minute. Thus if decay can occur at all, it almost always occurs
before another neutron can be captured. On the other hand, the r-process, which differs from the s-process by its faster rate
of neutron capture of more than one neutron, entails a succession of rapid neutron captures (hence the name r-process) by
heavy seed nuclei, typically 56Fe or other more neutron-rich heavy isotopes, before decay takes place. The r-process is
thus responsible for the creation of approximately half of the neutron-rich atomic nuclei heavier than iron, whereas the sprocess produces approximately the other half of the isotopes of the elements heavier than iron. Taken together these two
processes are claimed to account for a majority of the supposed galactic chemical evolution of elements heavier than
iron.In any case, it should be emphasized that any differences in the 176Lu half-life between those values determined by
direct counting experiments and those values determined by comparisons of radioisotope ages of the same meteorites and
terrestrial materials cannot be accounted for by the branching in the decay of 176Lu because both methods measure only the
partial decay to 176Hf (fig. 1), which accounts for >99% of the 176Lu decay. Furthermore, as pointed out by Begemann et al.
(2001), 176Lu is also claimed to be important as the only long-lived nuclide that is close to 100% s-process in terms of its
supposed stellar origin, being shielded from r-process contributions by stable 176Yb. Thus, according to naturalistic
theory, 176Lu could be used to calculate the supposed age of the galactic s-process. However, a consistent level of early
interest (for example, Audouze, Fowler, and Schramm 1972; Beer et al. 1981; McCulloch, De Laeter, and Rosman 1976)
became tempered by the realization that an excited isomer of 176Lu that decays rapidly to176Hf would have its abundance
enhanced by typical stellar temperatures. This thermal effect would thus prevent any s-process age calculation (Beer et al.
1984). In any case, this is not relevant to these problems involving terrestrial, lunar, or meteoritic materials, because at
temperatures up to 10 million degrees the 176Lu half-life would apparently be shortened by only a miniscule amount. On the
other hand, if the 176Lu atoms were highly ionized or buffeted by the high pressure waves which occur due to cavitation, for
example, during the rapid flow of water (as would have been occurring during the Flood), then changes to the decay rate
could have been significantly large, even by several orders of magnitude (Cardone, Mignani, and Petrucci 2009). It only
stands to reason that any process which reduces the coulomb barrier of the nucleus can significantly affect the decay
process, especially a radionuclide which decays.The discrepancies between different determinations of the 176Lu half-life
became even more apparent when in 2003 two independent results of direct counting experiments were published. Grinyer
et al. (2003) measured a half-life of 40.8 0.03 Byr, whereas Nir-El and Haquin (2003) reported a half-life of only 36.77
0.75 Byr from their experiment. Both of these published uncertainties were claimed to be the actual uncertainties from the
presumed correct values for the half-life, given the claimed elimination of measurement uncertainties in the methods used in
each case. Furthermore, further radioisotope age comparisons had determined the 176Lu half-life as 37.16 0.46 Byr based
on the U-Pb and Lu-Hf isochron ages of minerals extracted from four earth rocks (Scherer, Mnker, and Mezger 2001), as
35.9 0.09 Byr based on Sm-Nd and Lu-Hf isochron ages of eucrite meteorites (Blichert- Toft et al. 2002), and as 34.95
0.21 Byr based on the Lu-Hf isochron age of a large group of chondrite and eucrite meteorites compared to the Pb-Pb
isochron age of 4.56 Ga for the naturalistic formation of chondrites (Amelin et al. 2002; Bizzarro et al. 2003) (table 1). The
subtle discrepancies between determinations were thus highlighted by Scherer, Mezger, and Mnker (2003) in a diagram
which is reproduced here as Fig. 5. They added further 176Lu decay constant determinations based on the U-Pb and Lu-Hf
isochron ages of minerals extracted from more earth rocks, plus their own 176Lu decay constant determination based on
comparing the radioisotope ages of eucrite meteorites, to their previous study (Scherer, Mnker, and Mezger 2001) to show
that there are differences in the 176Lu decay rate as determined by physical counting experiments, as determined by age
comparisons of terrestrial minerals, and as determined by age comparisons of meteorites.
Fig. 5.Comparison of the 176Lu decay constant determinations since 1980, grouped according to the determination method
by Scherer, Mezger, and Mnker (2003). The vertical shaded areas indicate the two standard deviations (2) from the mean
values (vertical lines) for each method.Albarde et al. (2006) continued the recognition and discussion of this same
observation. Their diagram plotting the differences in the decay constant between determinations using terrestrial materials
and determinations using meteorites, as well determinations resulting from physical counting experiments was based on the
Scherer, Mezger, and Mnker (2003) diagram, but also included the Nir-El and Haquin (2003) physical counting result and
determinations by Sderlund et al. (2004) from comparing the Lu-Hf and U-Pb radioisotope ages of Precambrian mafic
intrusions. Their diagram is reproduced here as Fig. 6. Albarde et al. (2006) found that using the earth-based 176Lu decay
constant the Lu-Hf ages of chondrites were consistently 4% higher than their U-Pb ages, so the reconciliation of the Lu-Hf
and U-Pb ages required using a different meteorite-based 176Lu decay constant. Ironically, but significantly, Albarde et al.
(2006) proposed that this discrepancy between the terrestrial and meteoritic decay constants could be because of
accelerated 176Lu decay during the first few million years of the supposed existence of the solar nebula due to irradiation of
the 176Lu by -rays emitted by one or more supernova(e) exploding in the vicinity of the solar nebula.
Fig. 6.Comparison of the 176Lu decay constant determinations since 1980, grouped according to the determination method
by Albarde et al. (2006). The vertical shaded areas indicate the two standard deviations (2) from the mean values (vertical
lines) for each method.

However, in the related context of nucleosynthesis of the elements inside stars, Zhao and Kppeler (1990) had found the
absorption cross-section to produce 176Lu to be small and hence had found problems explaining their data. On the other
hand, 176Lu is produced at rather high rates in high energy accelerators, which of course means relatively high crosssections at energies above 1 GeV. Nevertheless, they concluded that to reestablish the 176Lu clock for determining the
supposed age of the s-process elements would, in any case, require a quantitative description of all processes feeding the
ground state as well as a reliable model for the s-process environment, in particular for the temperatures to which 176Lu was
exposed during its production and ever since. However, even though these effects of ionization on the nuclear half-lives can
be enhanced inside stars due to the extreme ionization prevalent there, such contexts are not relevant in the creationist
framework for the history of the earth and the universe. The earth and its rocks, and their contained elements and isotopes,
were all created before the stars, the other planets, the moons, and the asteroids from which the meteorites subsequently
came. And then in their history since their creation, neither the earth nor the asteroids have been subjected to the intense
ionization inside stars.In any case, Wimpenny, Amelin, and Yin (2013) have pointed out that using the 176Lu decay constant
determined from radioisotope age comparisons of terrestrial rocks to calculate the Lu-Hf ages of meteorites supposedly
older than ~4.556 Ga results in apparent ages that are older than the claimed age of the solar system. They noted the
suggested possible explanation that the 176Lu decay rate had been enhanced (or accelerated) by photo-excitation of 176Lu by
-rays (Albarde et al. 2006) or by cosmic rays (Thrane et al. 2010) shortly after supposed accretion began in the early solar
system. But they then sought to test that explanation by looking for the depleted 175Lu/176Lu ratios and excess176Hf such
irradiation would have produced in meteorites. They found no systematic differences in the Lu isotopic compositions
between terrestrial
basalts
and
achondrite
meteorites (eucrites
and
angrites).
Furthermore,
the
Lu-Hf systematics
in whole rock and
mineral separates
from
angrite
meteorites (Amelin,
Wimpenny, and Yin
2011;
Sanborn,
Carlson,
and
Wadhwa
2012)
produce Lu-Hf ages
concordant
with
their Pb-Pb and
Sm-Nd radioisotope
ages rather than
expected
older
ages. Thus they
concluded
that
together these data do not support the hypothesis of accelerated decay of 176Lu by -ray or cosmic ray irradiation during
supposed solar system accretion, which of course is an evolutionary concept that simply does not work.
Now while Albarde et al. (2006) included in their diagram (fig. 6) the latest determinations from physical counting
experiments and from the comparison of ages for more earth rocks, they missed including further more recent
determinations using the comparison of ages for more meteorites, and for minerals extracted from them. Patchett et al.
(2004) analyzed the Lu-Hf and Sm-Nd isotopic systematics for 19 chondrite meteorites and found agreement between the
calculated Sm-Nd and Lu-Hf isochron ages by using a 176Lu decay constant of 1.867 10-11 per year (a half-life of 37.12 Byr)
similar to that obtained by Scherer, Mnker, and Mezger (2001) and Sderlund et al. (2004) from comparison of Lu-Hf and
U-Pb isochron ages in terrestrial rocks and minerals. Similarly, Amelin (2005) obtained Lu-Hf and U-Pb isotope data for
phosphate minerals separated from the Acapulco (primitive achondrite) and Richardton (ordinary chondrite) meteorites and
concordant Lu-Hf and Pb-Pb isochron ages yielded 176Lu decay constants of 1.832 10-11 per year and 1.864 10-11 per
year (half-lives of 37.83 Byr and 37.18 Byr) respectively (table 1). Then Bouvier, Vervoort, and Patchett (2008) similarly
obtained Lu-Hf isotopic data for another 28 pristine chondrite meteorites and added to them the Lu-Hf data of Bizzarro et al.
(2003) and Patchett et al. (2004) to calculate a Lu-Hf isochron age concordant with the presumed Pb-Pb isochron age for
the start of the supposed accretion of the solar system (Bouvier et al. 2007), by which they determined a 176Lu decay
constant of 1.884 10-11 per year (a half-life of 36.78 Byr).All the 176Lu half-life determinations since 1980 are thus plotted in
Fig. 7, color-coded according to the different determination methods. At this expanded scale for the half-life values it is
apparent that there are two groupings of values among those obtained by the physical counting experiments, and another
two groupings of values among those obtained by age comparisons using meteorites, each with their own mean value and
two standard deviation spread. On the other hand, the half-life values obtained by age comparison using terrestrial minerals
and rocks cluster very tightly around a mean value of 37.12 Byr (a decay constant of 1.867 10 -11 per year) with a two
standard deviation spread of 0.43 Byr, which would be much less if one samples value was ignored. The mean values and
two standard deviation spreads for all the groupings depicted in Fig. 7 are listed in Table 2, including the grouping of all the
determinations using meteorites for age comparisons (the black vertical line and cross-hatched area in fig. 7). For
comparison, the mean (average) half-life value for the 14 direct counting determinations made prior to 1980 (from table 1) is
listed in Table 2. At 32.8 Byr it is much lower than the means of both groups of post- 1980 direct counting determinations.
This is of course a reflection of the better resolution of the solid state detectors used for the post-1980 determinations, in
contrast to the combination of proportional chamber/ NaI (Tl doped) detectors used for the pre-1980 determinations (as
already noted above). In any case, what is immediately evident from Fig. 7 and Table 2 is that the mean half-life values for
the groupings of shorter half-life physical counting experiments, age comparisons using terrestrial minerals and rocks, and
longer half-life age comparisons using meteorites are virtually identical within their two standard deviation spreads. That is
why a 176Lu decay constant of 1.867 10-11 per year (a half-life of 37.12 Byr), based primarily on the age comparisons using
terrestrial minerals and rocks, has now generally been adopted by the geological community as the agreed value for
standard use in Lu-Hf age calculations (for example, Amelin, Wimpenny, and Yin 2011; Debaille, Yin, and Amelin 2011;
Sanborn, Carlson, and Wadhwa 2012; Thrane et al. 2010).

Table 2.Mean half-life values and two standard deviation spreads for the groupings of determinations by the different
methods as plotted in Fig. 7.
Determination Method

Grouping

Mean Value

Two Standard Deviation

physical counting

shorter half-life

36.93 Byr

0.41 Byr

physical counting

longer half-life

40.7 Byr

1.2 Byr

physical counting

prior to 1980

32.8 Byr

age comparisons (terrestrial minerals and rocks)

37.12 Byr

0.43 Byr

age comparisons (meteorites)

shorter half-life

35.39 Byr

0.88 Byr

age comparisons (meteorites)

longer half-life

37.23 Byr

0.22 Byr

age comparisons (meteorites)

all determinations

36.35 Byr

0.65 Byr

Fig. 7. Comparison of all the 176Lu half-life determinations since 1980 grouped and color-coded according to the
determination method. The vertical color shaded areas indicate the two standard deviations from the mean values (colored
vertical lines) for the groupings of determinations in each method. The uncertainties are at the 2 level.
Three issues stemming from this choice need to be resolved. First, why do a few of the physical counting experiments yield
a longer half-life mean value of 40.7 Byr, especially when the most recent of these experiments (Grinyer et al. 2003) was
reported at the same time as the most recent determination in the shorter half-life grouping with a half-life mean value of
36.93 Byr in Fig. 7 (Nir-El and Haquin 2003)? There was a difference in the methodology of these two physical counting
experiments. Grinyer et al. (2003) used the --coincidence method, whereas Nir- El and Haquin (2003) simply used counting. The --coincidence method has more potential difficulties if not done correctly, so the direct -counting method
may give a better result. However, that must not explain this difference in the determined half-life values, because the -coincidence method was also used by Sguigna, Larabee, and Waddington (1982), but they obtained a half-life value of only
35.9 Byr, whereas Norman (1980) used the -counting method and obtained a half-life value of 40.8 Byr, the same as
Grinyer et al. (2003). And the different half-life values obtained cannot be explained by the Lu source material used,
because both Grinyer et al. (2003) and Nir- El and Haquin (2003) used Lu 2O3powder, whereas Norman (1980) used Lu
metal foil. Of course, the grain size of the powder can make a difference in the effective density of the powder and therefore
the absorption correction. So these disparities in the determined half-lives between these physical counting experiments
remain unresolved.Second, why do the earlier determinations using age comparisons of meteorites yield a shorter half-life
value than the more recent determinations using age comparisons of meteorites? Bouvier, Vervoort, and Patchett (2008)
seem to have the answer when they concluded that thermal- or shock-metamorphosed meteorites (such as eucrites)
cannot be used to reliably determine the 176Lu decay constant and this is especially true when multiple meteorites or
meteorite groups with distinct thermal and shock metamorphic histories, and parent bodies are included in the regression to
determine the 176Lu decay constant. Patchett and Tatsumoto (1980b), Tatsumoto, Unruh, and Patchett (1981), and BlichertToft et al. (2002) all used eucrite meteorites, and Bizzarro et al. (2003) used both eucrite and chondrite meteorites, whereas
Patchett et al. (2004) and Bouvier, Vervoort, and Patchett (2008) only used chondrite meteorites, and Amelin (2005) used
phosphate minerals from chondrite and primitive achondrite meteorites, not eucrite meteorites. As Amelin, Wimpenny, and
Yin (2011) and Sanborn, Carlson, and Wadhwa (2012) found, there is evidence of significant disturbance of the Lu-Hf
systematics among the mineral phases in some meteorites after initial closure of the isotopic system due to open
geochemical system behavior, such as leakage of 176Hf.Third, the favored radioisotope method used for age comparisons to

most reliably determine the 176Lu decay constant and half-life is the U-Pb method. All the age comparisons on terrestrial
rocks and minerals (Scherer, Mnker, and Mezger 2001; Scherer, Mezger, and Mnker 2003; Sderlund et al. 2004), and
the age comparisons on meteorites by Amelin (2005) and Bouvier, Vervoort, and Patchett (2008), used the U-Pb method to
determine the 176Lu decay constant and half-life. So the values of 1.867 10 -11 per year and 37.12 Byr now routinely
adopted for Lu-Hf age calculations are firmly calibrated against the U-Pb method. Yet even this U-Pb gold standard has
come under similar scrutiny in the last decade. The U-Pb method depends on the crucial 238U/235U ratio, but discrepancies
and variations have been found between the 238U/235U ratio in U-bearing terrestrial minerals and rocks and the 238U/235U ratio
in meteorites (Brennecka and Wadhwa 2012; Hiess et al. 2012). These discrepancies and variations remain unexplained, so
their potential significance from a creationist perspective awaits further study, especially in the context of examination of the
determinations of the 238U and 235U decay constants and half-lives. Furthermore, the fact that there are these variations in
the crucial 238U/235U ratio in terrestrial minerals and rocks on which the U-Pb dating gold standard depends, which has been
used to calibrate Lu-Hf isochron ages to determine the 176Lu half-life and decay constant, only underscores that these
radioisotope methods cannot provide the absolute invariable dates they are so confidently proclaimed to provide. Thus the
U-Pb method should not be used as a standard to determine other radioisotope decay constants.Finally, it would remain
prudent to be very careful with these geological comparison methods for two other reasons. First, there are significant flaws
in the basic assumptions on which all the radioisotope dating methods depend. Second, the U-Pb method relies primarily on
decay, whereas the Lu-Hf method relies on decay. As observed by both Austin (2005) and Snelling (2005), these
different decay modes seem to yield different ages for some earth rocks using the same samples with essentially the same
methodology. Furthermore, as described earlier, the differences in the 176Lu decay constant and half-life values determined
by physical counting and --coincidence counting experiments remain unexplained and unresolved, when these direct
counting methods should be expected to yield reliable and consistent 176Lu decay constant and half-life determinations.
However, at the low counting rates which are involved even small corrections to the data can easily cause the observed
differences. So without an accurately known 176Lu decay constant and half-life, accurate Lu-Hf radioisotope ages cannot be
determined. Therefore, Lu-Hf dating cannot be used to reject the young-earth creationist timescale. Indeed, current
radioisotope dating methodologies are at best hypotheses based on extrapolating current measurements and observations
back into an assumed deep time history for the cosmos. Nevertheless, although the products from the multi-billion half-life of
Lu are a seeming challenge to a young age for the earth, they have been already explained by the evidence for accelerated
radioisotope decay during a past catastrophic event, such as the Flood (Vardiman, Snelling, and Chaffin 2000, 2005).
Conclusions
There have been numerous attempts to determine the 176Lu half-life and decay constant in the last 75 years by two methods
direct counting of the -rays emitted by the daughter 176Hf as it falls from an excited state to its ground state after being
derived by the decay of its parent 176Lu, and comparisons of Lu-Hf ages of terrestrial minerals and rocks, and meteorites,
with ages derived by other radioisotope systems, especially U-Pb. The estimates of the 176Lu half-life have not been
consistent, even since 1980 as improvements have been made in measurement technologies. Indeed, there are two distinct
separate groupings of determinations obtained by physical direct counting experiments with mean half-life values of 36.93
Byr and 40.7 Byr, with these differences apparent in two experiments performed in the same year (2003). This discrepancy
has still neither been explained nor resolved.The initial use of meteorites to determine the 176Lu decay constant (and halflife) via age comparisons also suggested a disparity with the 176Lu decay constant (and half-life) determined by the physical
counting experiments and with the176Lu decay constant (and half-life) determined by age comparisons of terrestrial minerals
and rocks. However, more recent determinations using meteorites for age comparisons have produced a distinct separate
grouping with 176Lu decay constant and half-life mean values virtually identical to those obtained by age comparisons of
terrestrial minerals and rocks, and to those in the shorter half-life grouping of physical counting experiments. It was thus
realized that the eucrite meteorites used for not as suitable because of significant disturbance of the Lu-Hf systematics
among mineral phases due to open geochemical system behavior, such as leakage of 176Hf. Consequently a 176Lu decay
constant of 1.867 10-11 per year (a half-life of 37.12 Byr), based primarily on the age comparisons using terrestrial minerals
and rocks, has now generally been adopted by the geological community as the agreed value for standard use in Lu-Hf age
calculations.Nevertheless, all the age comparisons on terrestrial rocks and minerals, and the more recent age comparisons
on meteorites, used the U-Pb method to determine the 176Lu decay constant of 1.867 10-11 per year and half-life of 37.12
Byr, and thus Lu-Hf age calculations are ultimately calibrated against the U-Pb method. However, this U-Pb gold standard
depends on the crucial 238U/235U ratio, and yet discrepancies and variations have been found between the 238U/235U ratio in Ubearing terrestrial minerals and rocks and the 238U/235U ratio in meteorites. These discrepancies and variations remain
unexplained. This only serves to highlight that if the Lu-Hf dating method has been calibrated against the U-Pb gold
standard with its own uncertainties, then it cannot be absolute, especially as the176Lu decay rate has also not been
accurately determined by physical direct counting experiments. Without an accurately known 176Lu decay constant and halflife, accurate Lu-Hf radioisotope ages cannot be determined. Therefore, Lu-Hf dating cannot be used to reject the youngearth creationist timescale, especially as current radioisotope dating methodologies are at best hypotheses based on
extrapolating current measurements and observations back into an assumed deep time history for the cosmos.
Determination of the Radioisotope Decay Constants and Half-Lives: Rhenium-187 ( 187Re)
by Dr. Andrew A. Snelling on February 25, 2015
Abstract
Over the last 66 years numerous determinations have been made of the 187Re half-life and decay constant. With the
measurement technology having improved, the determinations over the last 20 years have resulted in close agreement
between the four determination methodsdirect physical counting and in-growth experiments, and radioisotope age
comparisons using molybdenites and groups of meteorites. Thus the 187Re half-life and decay constant values of 41.577
0.12 Byr and 1.6668 0.0034 10-11 per year respectively have now been adopted for standard use by the uniformitarian
geological community. These values are based on determinations recalibrating Re-Os model ages of molybdenites by
forcing them (essentially by circular reasoning) to agree with the U-Pb concordia-Pb-Pb intercept ages of zircons from the
same 11 magmatic-hydrothermal systems dating from ca. 90 Ma to ca. 2670 Ma. This value is essentially identical to the
best of the recent determinations using groups of iron meteorites. It is also within the uncertainty range of the half-life
values obtained by the best of the physical direct counting and in-growth experiments. Yet, in spite of such experiments
directly measuring 187Re decay and its decay product 187Os respectively, preference has been given to the half-life value
determined by forcing the Re-Os data to agree with U-Pb dates. But many unprovable assumptions are also involved, not
least being that the radioisotope systems closed at the same time and subsequently remained closed. Furthermore, even
this gold standard has unresolved uncertainties due to the U decay constants being imprecisely known, and to measured
variations of the 238U/235U ratio in terrestrial rocks and minerals and in meteorites. Both of these factors are so critical to the
U-Pb method, as well as the additional factor of knowing the initial concentrations of the daughter and index isotopes, so it

should not be used as a standard to determine other decay constants. In any case, the half-life of 187Re has been shown to
be dependent on environmental conditions such as the degree of ionization. Fluctuations in nuclear decay rates have also
been observed to correlate with solar activity for a few other isotopes, while there is evidence decay rates of the
radioisotopes used for rock dating have not been constant in the past. This only serves to highlight that if the Re-Os dating
method has been calibrated against the U-Pb gold standard with all these uncertainties, then it cannot be absolute, and
therefore it cannot be used to reject the young-earth creationist timescale. Indeed, current radioisotope dating
methodologies are at best hypotheses based on extrapolating current measurements and observations back into an
assumed deep time history for the cosmos.
Keywords: radioisotope dating, decay constants, half-lives, rhenium-187, 187Re, decay, direct counting, in-growth
experiments, geological comparisons, molybdenites, iron meteorites, U-Pb gold standard, U decay constants, 238U/235U
ratio, environmental conditions
Introduction
Radioisotope dating of rocks and meteorites is perhaps the most potent claimed proof for the supposed old age of the earth
and the solar system. The absolute ages provided by the radioisotope dating methods provide an apparent aura of certainty
to the claimed millions and billions of years for formation of the earths rocks. Many in both the scientific community and the
general public around the world thus remain convinced of the earths claimed great antiquity.However, accurate radioisotopic
age determinations require that the decay constants of the respective parent radionuclides be accurately known and
constant in time. Ideally, the uncertainty of the decay constants should be negligible compared to, or at least be
commensurate with, the analytical uncertainties of the mass spectrometer measurements entering the radioisotope age
calculations (Begemann et al. 2001). Clearly, based on the ongoing discussion in the conventional literature this is not the
case at present. The stunning improvements in the performance of mass spectrometers during the past four or so decades,
starting with the landmark paper by Wasserburg et al. (1969), have not been accompanied by any comparable improvement
in the accuracy of the decay constants (Begemann et al. 2001; Steiger and Jger 1977), in spite of ongoing attempts (Miller
2012). The uncertainties associated with direct half-life determinations are, in most cases, still at the 1% level, which is still
significantly better than any radioisotope method for determining the ages of rock formations. However, even uncertainties
of only 1% in the half-lives lead to very significant discrepancies in the derived radioisotope ages. The recognition of an
urgent need to improve the situation is not new (for example, Min et al. 2000; Renne, Karner, and Ludwig 1998). It continues
to be mentioned, at one time or another, by every group active in geo- or cosmochronology (Schmitz 2012).From a
creationist perspective, the 19972005 RATE (Radioisotopes and the Age of The Earth) project successfully made progress
in documenting some of the pitfalls in the radioisotope dating methods, and especially in demonstrating that radioisotope
decay rates may not have always been constant at todays measured rates (Vardiman, Snelling, and Chaffin 2000, 2005).
Yet much research effort remains to be done to make further inroads into not only uncovering the flaws intrinsic to these
long-age dating methods, but towards a thorough understanding of radioisotopes and their decay during the earths history
within a biblical creationist framework.One crucial area the RATE project did not touch on was the issue of how reliable have
been the determinations of the radioisotope decay rates, which are so crucial for calibrating these dating clocks. Indeed,
before this present series of papers (Snelling 2014a, b) there have not been any attempts in the creationist literature to
review how the half-lives of the parent radioisotopes used in long-age geological dating have been determined and to collate
their determinations to discuss the accuracy of their currently accepted values. After all, accurate radioisotope age
determinations depend on accurate determinations of the decay constants or half-lives of the respective parent isotopes.
The reliability of the other two assumptions these absolute dating methods rely on, that is, the starting conditions and no
contamination of closed systems, are unprovable. Yet these can supposedly be circumvented somewhat via the isochron
technique, because it is independent of the starting conditions and is sensitive to revealing any contamination, which is still
significantly better than any radioisotope method for determining the ages of rock formations. Data points that do not fit on
the isochron are simply ignored because their values are regarded as due to contamination. That this is common practice is
illustrated with numerous examples from the literature by Faure and Mensing (2005) and Dickin (2005). On the other hand, it
could be argued that this discarding of data points which do not fit the isochron is arbitrary and therefore is not good
science, because it is merely assumed their aberrant values are due to contamination rather than that being proven to be
so. Indeed, in order to discard such outliers in any data set, one must establish a reason for discarding those data points
which cannot be reasonably questioned.In order to rectify this deficiency, Snelling (2014a,b) has documented the
methodology behind and history of determining the decay constants and half-lives of the parent radioisotopes 87Rb and 176Lu
used as the basis for the Rb-Sr and Lu-Hf long-age dating methods respectively. He showed that there is still some
uncertainty in what the values for these measures of the 87Rb and 176Lu decay rates should be. This is especially the case
with determinations of the 176Lu decay rate by physical direct counting experiments. Furthermore, the determined values
differ when Rb-Sr ages are calibrated against the U-Pb ages of either the same terrestrial minerals and rocks or the same
meteorites and lunar rocks. Ironically it is the slow decay rate of isotopes such as 87Rb used for deep-time dating that makes
a precise measurement of that decay rate so difficult. Thus it could be argued that direct measurements of their decay rates
should be the only acceptable experimental evidence, especially because measurements which are calibrated against other
radioisotope systems are already biased by the currently accepted methodology employed by the secular community in their
rock dating methods. Indeed, both the 87Rb and 176Lu decay half-lives have ultimately been calibrated against the U-Pb
radioisotope systems, yet there are now known measured variations in the 238U/235U ratio that is critical to that method
(Brennecka and Wadhwa 2012; Hiess et al 2012).Therefore, the aim of this contribution is to further document the
methodology behind and history of determining the present decay constants and half-lives of the parent radioisotopes used
as the basis for the long-age dating methods. We need to explore just how accurate these determinations are, whether there
really is consensus on standard values for the half-lives and decay constants, and just how independent and objective the
standard values are from one another between the different methods. Of course, it is to be expected that every long-lived
radioactive isotope is likely to show similar variation and uncertainty in half-life measurements because these are difficult
measurements to make. However, even small variations and uncertainties in the half-life values result in large variations and
uncertainties in the calculated ages for rocks, and the question remains as to whether the half-life values for each long-lived
parent radioisotope are independently determined. We continue here with determinations of the rhenium-187 ( 187Re) decay
rate, which is the basis for the Re-Os dating method.
Rhenium and Rhenium-187 Decay
With an atomic number Z of 75, rhenium (Re) is element number 75 and appears in the sixth period on the periodic table.
Although Re is chemically related to Mn and Tc in the group or column VIIB, its chemical properties are more similar to those
of Mo (Z = 42), which it replaces in Mo-bearing minerals such as molybdenite (MoS 2) (Faure and Mensing 2005). Rhenium
radioactively decays to osmium (Z = 76), its neighboring element, yet Os is a member of the platinum group elements
(PGEs) in group VIII of the periodic tableruthenium (Ru), rhodium (Rh), palladium (Pd), osmium (Os), iridium (Ir), and
platinum (Pt). Both Re and Os have several valences consistent with their electron configurationsRe (7+, 6+, 4+, 2+, and

1-) and Os (8+, 6+, 4+, 3+, and 2+). Additionally, their ionic radii in the 4+ valence state are identical at 0.71 , and both
elements are strongly siderophile (iron-loving). The ionic radius of Re 4+ (0.71 ) is also similar to that of Mo 4+ (0.68 ), and
their electronegativities are virtually identical (1.9 for Re and 1.8 for Mo), which explains why Re 4+ can replace Mo4+ in
molybdenite and other Mo-bearing sulfide minerals (Faure and Mensing 2005). Furthermore, Re 4+ tends to occur in Cu
sulfide minerals, presumably because Re4+ is captured in place of Cu2+(ionic radius of 0.69 and electronegativity of 1.9).
Re and Os concentrations are generally much higher in chondrite and iron meteorites, in which they are highly correlated,
than in most terrestrial rocks (Faure and Mensing 2005). Iron meteorites have average concentrations of 987 ppb Re and
15,260 ppb Os, whereas ordinary chondrites contain on average only 62.6 ppb Re and 690 ppb Os. Thus it has been
feasible to use the Re-Os dating method on iron meteorites. On the other hand, the average concentrations of Re and Os in
terrestrial igneous rocks are less than 1 ppb for Re and less than 3 ppb for Os. However, granitic gneisses and shales have
elevated Re concentrations ranging up to 200 ppb, making it possible to use the Re-Os dating method on organic-rich black
shales with their high Re concentrations (Schmitz 2012), the only example of radioisotope dating methods having been used
directly on sedimentary rocks. Whereas most rock-forming silicate minerals have low Re and low Os concentrations,
molybdenite (MoS2) and Cu-Ni-PGE sulfides (for example, pentlandite, chalcopyrite, and pyrrhotite) have high Re and low
Os concentrations of 100100,000 ppb and 6900 ppb respectively, which has enabled their use for Re-Os dating of sulfide
ore deposits. Re typically exists at the ppm level in molybdenite, whereas no Os is believed to be incorporated into
molybdenite during its formation (Stein et al. 2001). Thus measured Os is assumed to be radiogenic 187Os accumulated in
situ from 187Re decay, and thus model molybdenite Re-Os ages can presumably be calculated from the Re-Os isotopic
analyses of single samples. For multiple molybdenite samples assumed to have the same paragenesis the Re-Os isochron
approach is employed to supposedly verify zero initial or common 187Os and document presumed isochroneity of the
supposedly related samples. The minerals magnetite, ilmenite, and garnet with average Re concentrations of 413 ppb
have been suggested as having future potential for radioisotope dating by the Re-Os method. Of course, this implies that
the technique used has to accurately measure very miniscule amounts of Os in small samples supposedly hundreds of
millions of years old and then separate and measure from that miniscule amount the radiogenic Os compared to the
primordial and cosmogenic Os.Rhenium has two naturally occurring isotopes 185Re and 187Re whose abundances are 37.4%
and 62.6% respectively (Shirey and Walker 1998). 187Re is radioactive and emits --particles as it decays to 187Os, which is
one of the seven naturally occurring isotopes of Os. The relevant decay scheme is depicted as:
187
Re 187Os + - + v + Q
where - is a negatively charged -particle, is an anti-neutrino, and Q is the total decay energy (0.00265 MeV).
There are two parameters by which the decay rate is measured and expressed, namely, the decay constant () and the halflife (t). The decay constant can be defined as the probability per unit time of a particular nucleus decaying, whereas the
half-life is the time it takes for half of a given number of the parent radionuclide atoms to decay. The two quantities can be
almost used interchangeably, because they are related by the equation:t = ln 2 / = 0.693 /
The decay rate of 187Re has been difficult to determine because the endpoint energy of the -spectrum is only about 2.6 keV.
Similarly, the total decay energy is extremely low at 2.65 keV, even compared with that for 87Rb at 275 keV. This makes
measurement of the decay rate by direct counting very difficult, just as it has always caused problems in the accurate
determination of the 87Rb decay rate (Dickin 2005; Faure and Mensing 2005).
Determination Methods
Three approaches have been followed to determine the -decay constant and half-life of the long-lived radioactive 187Re
since Naldrett and Libby (1948) made the first attempt.
Direct counting
In this technique, the beta () activity of 187Re is counted in a source material over a designated time period, and divided by
the total number of radioactive atoms in the known quantity of Re, based on Avogadros number and the isotopic abundance
of 187Re. However, detectors with an entrance window are not suitable for the detection of the spectral shape and half-life
of 187Re because of the long half-life of 187Res unique first-forbidden -decay and its endpoint energy of only 2.6 keV, the
lowest known in nature (Ashktorab et al. 1993). Furthermore, accurate counting of solid samples has proved to be almost
impossible, due to the absorption of -particles by surrounding Re atoms (Dickin 2005). An alternative technique was thus
developed which uses either a gaseous Re compound to replace the gas filling of a proportional counter, or a liquid Re
compound in a scintillation detector. In both cases it was difficult to find compounds with suitable properties, but Brodzinski
and Conway (1965) using biscyclopentadienylrhenium hydride vapor obtained a 187Re half-life of 66 13 Byr by the former
method, and Naldrett (1984) using oxotrichlorobis (tri-phenylphosphine) rhenium V obtained a value of 35 4 Byr by the
latter method.The precision of such direct counting studies was subsequently improved by the use of microcalorimetric
determinations in which all the energy released in the -decay of 187Re is measured except that of the emitted neutrino
(Alessandrello et al. 1999; Arnaboldi et al. 2003; Ashktorab et al. 1993; Galeazzi et al. 2001). Ashktorab et al. (1993) used a
high-temperature quartz proportional counter with metallo-organic biscyclopentadienylrhenium hydride vapors (containing
natural rhenium) added to the counting gas (90% argon 10% methane) to determine a half-life for 187Re of 43.5 1.3 Byr. On
the other hand, Alessandrello et al. (1999), Galeazzi et al. (2001), and Arnaboldi et al. (2003) used cryogenic
microcalorimeters or low-temperature bolometers, which consisted of a rhenium-containing absorber that converted the
energy of the incident radiation into heat, a sensor that detected the temperature variations of the absorber, and a weak
thermal link between the detector and the heat sink. All three experiments used neutron transmutation doped germanium
thermistors with silicon chips, but whereas Galeazzi et al. (2001) used a metallic rhenium crystal for the absorber,
Alessandrello et al. (1999) and Arnaboldi et al. (2003) used silver perrhenate (AgReO 4) crystals instead. However, in spite of
using similar protocols, even all these experiments using modern methods yielded results for the half-life of 187Re with
uncertainties at the several percent level (Galeazzi et al. 2001) that are significantly different43.0 3.5 Byr (Alessandrello
et al. 1999), 41.2 0.6 Byr (Galeazzi et al. 2001), and 43.2 0.15 Byr (Arnaboldi et al. 2003). This makes these half-life
values unsuitable for high resolution Re-Os geochronology, particularly when compared with other radioisotopic systems in
use in the earth sciences.Judged from the fact that these direct counting experiments have yielded results that are not
compatible with one another within the stated uncertainties (see below), it would appear that not all the measurement
uncertainties are accounted for, and therefore the stated uncertainties are unrealistically small and typically are
underestimated. Many of such experiments are likely plagued by unrecognized systematic errors (Begemann et al. 2001).
As the nature of these errors is obscure, it is not straightforward to decide which of the, often mutually exclusive, results of
such direct counting experiments is closest to the true value. Furthermore, the presence of unknown systematic biases
makes any averaging dangerous. It is possible that reliable results of careful workers, listing realistic uncertainties, will not
be given the weights they deservethis aside from the question of whether it makes sense to average numbers that by far
do not all agree within the stated uncertainties.
In-Growth Experiments

An alternative approach to determining the 187Re decay constant and half-life is to measure the amount of 187Os produced by
the decay of a known quantity of 187Re in the laboratory over a known period of time. This method has only been attempted
by Lindner et al. (1986, 1989), who used five batches of a concentrated solution of perrhenic acid (HReO 4) for which the
precise quantity of contained Re was carefully determined and from which all residual contaminant Os was removed. To
three of the five batches were added precise measured quantities of independent spikes of the stable Os isotopes 190Os
and 192Os in an effort to improve the precision of their determination of the 87Re half-life. To one of the other batches only
a 192Os spike was added, and then a 190Os spike was only added to selected aliquots as they were withdrawn after the
respective growth periods. To the last batch only a 190Os spike was added, and then a 192Os spike was only added just
before testing after the respective growth periods. This procedure was adopted as an internal test of chemical equilibrium
and isotopic exchange. Since one spike was added at the onset of 187Os growth and the other at the end of the growth
period, any significant lack of isotopic exchange between the radiogenic 187Os and either spike would have caused a clear
discrepancy between the results based on the two spikes. The time for the start of in-growth accumulation of daughter 187Os
was measured from the end of the removal of the contaminant Os. The first batch was removed for testing after 400 days
and the other batches were progressively removed for testing in intervals of approximately 200 days, the last batch being
removed after almost five years. The 187Re/187Os, 187Re/190Os, and 187Re/192Os ratios in the final solutions at each time interval
were measured by isotope dilution using inductively coupled plasma mass spectrometry. The unweighted average of all
these determinations of the 187Re half-life, stated at the 95% confidence level, was 42.3 1.3 Byr, with the stated
uncertainties amounting to 3% and allowing for the known systematic uncertainties in the calibrations of the spikes and
standard solutions, as well as the random uncertainties (Lindner et al. 1989).
This technique thus relies on measuring the -decay product 187Os of a carefully measured amount of radioactive 187Re
accumulated over a well-defined period of time. Where feasible, this is the most straightforward technique (Begemann et al.
2001). In-growth measurement overcomes the problems encountered with direct counting of large fractions of lowenergy 187Re-emitted -particles. It also comprises the direct 187Os product of any radiation-less decays (which otherwise
cannot be measured at all). Among the drawbacks of this approach is that the method is obviously not instantaneous. The
experiment must be started long before the first results can be obtained because long periods of time are required for
sufficiently large amounts of the decay products to accumulate. In this case the period was only five years, which it could be
argued might have been too short a length of time for sufficient daughter 187Os to have accumulated for the most accurate
measurements to be made. In-growth experiments further require an accurate determination of the ratio of the two chemical
elements (parent/daughter, 187Re/187Os) as well as an accurate determination of the isotopic composition of parent and
daughter elements Re and Os at the start of the accumulation (Begemann et al. 2001). And Lindner et al. (1989) did not
take into account the effect of the neutron capture cross section of 187Os on 187Re decay, which we now know would have
been high during this five-year measurement period at the experiments temperature and thus facilitating the decay process
(Segawa et al. 2007).
Geological Comparisons of Methods
The third approach to the determination of the 187Re decay constant (and half-life) has been to date geological samples
whose ages have also been measured by other methods with presumably more reliable decay constants (Dickin 2005;
Faure and Mensing 2005). This essentially involves circular reasoning, because it is being assumed the other radioisotope
dating methods, principally the U-Pb method, gives the reliable dates to which the 187Re half-life can be calibrated to bring
the Re-Os radioisotope ages into agreement. It should be noted, however, that this is hardly objective, because all the
radioisotope ages of rocks could be wrong due to the underlying unprovable and suspect assumptions on which all the
radioisotope methods are based. Nevertheless, geological determinations of the 187Re half-life have been made over the
past 60 years using the terrestrial mineral molybdenite (MoS2) and groups of iron meteorites.
The first attempt to determine the 187Re half-life using Re-Os molybdenite dating studies could only estimate the value as
being between 5 and 250 Byr because the age of the molybdenite was unknown (Herr, Hintenberger, and Voshage 1954;
Hintenberger, Herr, and Voshage 1954). Further work by Herr and Merz (1955, 1958) on molybdenites of known ages, that
is, by assuming the ages of the molybdenites were identical to their host rocks, narrowed the estimate of the 187Re half-life
value to between 55 and 68 Byr. Then only several years later, Hirt, Herr, and Hoffmeister (1963) and Hirt et al. (1963)
determined a 187Re half-life value of 43 Byr using molybdenites of known ages, which is in good agreement with the value of
41.577 Byr via a similar determination more than 40 years later (Selby et al. 2007), a remarkable achievement given the
poor age constraints for the studied molybdenites and the ore deposits from which they came, and the very low precision of
the analytical techniques available at the time.
The 187Re decay constant and half-life have also been defined by the generation of Re-Os isochrons for groups of iron
meteorites of presumed age, first attempted by Luck, Birck, and Allgre (1980) and Luck and Allgre (1983), who obtained
half-life values of 42.8 Byr and 45.6 Byr respectively. The value for the 187Re decay rate has been subsequently refined
using new chemical separation and mass spectrometric techniques in order to generate Re-Os isochrons with higher
precision for groups of iron meteorites with presumed ages. Thus Smoliar, Walker, and Morgan (1996), Shen,
Papanastassiou, and Wasserburg (1996), and Shukolyukov and Lugmair (1997) obtained 187Re half-life values of 41.6 Byr,
41.7 Byr, and 41.8 Byr respectively, which are almost identical to the determination of 41.577 Byr by Selby et al. (2007)
using molybdenites of known ages (according to uniformitarian assumptions). This method has the disadvantage that it
involves geological uncertainties, such as whether all isotopic systems closed at the same time and remained closed.
However, it is claimed to still provide a useful check on the laboratory determinations by direct physical counting and ingrowth experiments. Nevertheless, this approach entails multi-chronometric dating of a mineral or groups of meteorites and
cross-calibration of different radioisotopic age systems by adjusting the decay constant of the Re-Os system so as to force
agreement with the age obtained via another dating system, usually U-Pb (Begemann et al. 2001). In essence, because the
half-life of 238U is claimed to be the most accurately known of all relevant radionuclides, this usually amounts to expressing
ages in units of the half-life of 238U.
Results of the Rhenium-187 Decay Determinations
During the last 66 years numerous determinations of the 187Re decay constant and half-life have been made using these
three methods. The results are listed with details in Table 1. The year of the determination versus the value of the half-life is
plotted in Fig. 1, and the year of the determination versus the value of the decay constant is plotted in Fig. 2. In each case
the data points plotted have been color-coded the same to differentiate the values as determined by the three approaches
that have been useddirect counting, in-growth experiments, and geological comparisons with other radioisotope dating
methods.
Discussion
It is obvious from Table 1 that, in spite of the remarkable achievements involved, the earliest determinations of the 187Re
decay rate were grossly inflated, due of course to the rudimentary technology available at the time. That is why most of them
were not plotted on Figs. 1 and 2. Indeed, the only determinations prior to 1963 plotted were those of Herr and Merz (1958)

and Watt and Glover (1962), because they were within the same range as all the later determinations. Those 187Re half-life
values are 61.5 Byr and 30 Byr respectively, and they were obtained via comparison of the Re-Os ages of molybdenites of
known ages determined by other radioisotope systems, and via direct counting with a proportional counter, respectively.
However, the error bars on the pre-1963 data (see table 1) should have reflected the uncertainties due to the available
technology, so it may be difficult to justify discarding that pre-1963 data based only on the technology argument. If that early
data were retained, then it may conceivably be possible to postulate that the early measurements are reflecting a real shift in
the decay constant with time.
Table 1. Determinations of the 187Re decay rate expressed in terms of the half-life and the decay constant using direct
physical counting experiments, in-growth experiments, and comparisons of radioisotope ages of terrestrial minerals and
rocks, and meteorites.
Date

Half-Life
(Byr)

Error
(Byr)

Decay
Constant (yr-Error
1)

Method

Source

Direct counting
Geiger counter

by

Direct counting
Geiger counter

by

1948 5000
1952 5000

Direct counting by proportional


counter
Curran (1952)

1953 5000

Direct counting of tracks


photographic emulsion

1954 100

Direct counting
counter

1954 5

Hintenberger,
Herr,
and
Voshage
(1954);
Herr,
Comparison of Re-Os age with Hintenberger, and Voshage
unknown age of molybdenite
(1954)

1948 4000

1000

250

gas-filled
Naldrett and Libby (1948)
gas-filled
Sugarman and Richter (1948)

by

in
Gauthe and Blum (1953)

Geiger
Suttle and Libby (1954)

Direct counting by proportional


counter
Dixon and McNair (1954)

1954 1,300,000

Direct counting by Geiger


counter of internal gas-tube
Walton (1957)

1957 240

50

1958 61.5

6.5

1958 320

70

Direct counting
counter

1962 120

40

Direct counting by Geiger


counter of internal gas-tube
Wolf and Johnston (1962)

1.13 10-11

Comparison of Re-Os ages of


0.12 10-11 molybdenites with known ages Herr and Merz (1958)
by

Geiger
Naldrett (1958)

Direct counting by proportional


counter
Watt and Glover (1962)

1962 30

Comparison of Re-Os ages of Hirt, Herr, and Hoffmeister


molybdenites of known ages
(1963); Hirt et al. 1963)

1.61 10-11

1963 43

1965 66

13

Direct counting of internal gas by


proportional counter
Brodzinski and Conway (1965)

1965 47

Direct counting of internal gas by


proportional counter
Payne and Drever (1965)

0.03 10-11

Comparison of Re-Os ages of


iron meteorites with Rb-Sr, U-Pb,
and Sm-Nd ages
Luck, Birck, and Allgre (1980)
Comparison of Re-Os ages of
molybdenites of known ages
(Rb- Sr, K-Ar, U-Pb)
Luck and Allgre (1982)

1980 42.8

1.62 10-11

1982 45.5

1.8

1.52 10-11

0.06 10-11

1983 45.6

1.2

1.53 10-11

Comparison of Re-Os and Pb-Pb


0.08 10-11 ages of meteorites
Luck and Allgre (1983)

1984 35

1986 43.5

1.3

1989 42.3

1.3

Direct
counting
scintillation counter

by

liquid
Naldrett (1984)

1.59 10-11

In-growth experiment by liquid


0.05 10-11 scintillation counter
Lindner et al. (1986)

1.64 10-11

In-growth experiment by isotope


dilution
and
ICP
mass
spectrometer
Lindner et al. (1989)

0.05 10-11

1992 42

0.4

1.65 10-11

Comparison of Re-Os ages of


0.011 10-11 iron meteorites with Pb-Pb ages Horan et al. (1992)

1993 45

1.54 10-11

Direct counting of internal gas by


0.10 10-11 proportional counter
Ashktorab et al. (1993)

1996 41.6

0.13

1.666 10-11

Comparison of Re-Os ages of Smoliar, Walker, and Morgan


0.017 10-11 iron meteorites with Pb-Pb ages (1996)

1996 41.7

0.9

1.66 10-11

Comparison of Re-Os ages of Shen, Papanastassiou,


0.027 10-11 iron meteorites with Pb-Pb ages Wasserburg (1996)

1997 41.8

0.1

1.658 10-11

Comparison of Re-Os ages of Shukolyukov


0.003 10-11 iron meteorites with Pb-Pb ages (1997)

1999 43

3.5

1.61 10-11

Direct
counting
0.26 10-11 microbolometer detectors

2001 41.2

0.6

1.682 10-11

Direct counting
0.045 10-11 microcalorimeter

2003 43.2

0.15

1.604 10-11

Direct
counting
0.004 10-11 microbolometer detectors

1.6668 10-11

Comparison of Re-Os ages of


0.0034 10-molybdenites and U-Pb ages of
11
zircons from same orebodies
Selby et al. (2007)

2007 41.577

0.12

by

and

and

Lugmair

by
Alessandrello et al. (1999)

cryogenic
Galeazzi et al. (2001)
by
Arnaboldi et al. (2003)

Fig. 1. Plot of each 187Re half-life determination versus the year of its determination, color-coded according to the method of
its determination.
Fig. 2. Plot of each 187Re
decay
constant
determination versus the
year of its determination,
color-coded according to
the method of its
determination.
Fig. 3 is a plot of all the
post-1980
determinations
of
the 187Re half-life, colorcoded the same to
differentiate the values
as determined by the
three approaches that
have been used. All the half-life values fall within a narrow range between 41.2 Byr (Galeazzi et al. 2001) and 45.6 Byr (Luck
and Allgre 1983), except for the value of 35 Byr obtained by Naldrett (1984) (see table 1). The values of 43 Byr obtained by
Hirt, Herr, and Hoffmeister (1963) and Hirt et al. (1963) via comparison of the Re-Os ages of molybdenites of known ages
and 47 Byr obtained by Payne and Drever (1965) via direct counting of internal gas by proportional counter are also within
the narrow range of values obtained from the post-1980 determinations, although their uncertainties of 5 Byr make them
too imprecise to be seriously considered. Yet by the same standard the uncertainties of 3 Byr and 3.5 Byr in the direct
counting determinations by Ashktorab et al. (1993) and Alessandrello et al. (1999) respectively would render their half-life
values of 45 Byr and 43 Byr similarly imprecise.

Fig. 3. Enlarged plot of


the 187Re
half-life
determinations since 1980
versus the year of their
determination, color-coded
according to the method of
its determination.
Therefore, to be able to
better assess the post1980
half-life
determinations, only those
values that fell within the
narrow 4146 Byr range
were plotted with their
uncertainties in Fig. 4
according to their method
of
determination.
A
distinction was also made
between
determinations
obtained by radioisotope
age
comparisons
of
molybdenites and those
obtained by radioisotope
age
comparisons
of
groups of iron meteorites.
Fig. 4. Comparison of all the 187Re half-life determinations since 1980 in the range of 4146 Byr grouped and color-coded
according to the determination method. The vertical color-shaded areas indicate the two standard deviations (2) from the
mean values (colored vertical lines) for the groupings of determinations in each method.Among the physical direct counting

determinations, the large uncertainties in the Ashktorab et al. (1993) and Alessandrello et al. (1999) experiments causes the
four results to overlap throughout the 3948 Byr range. Yet the Alessandrello et al. (1999) result of 43 Byr falls between the
Galeazzi et al. (2001) and Arnaboldi et al. (2003) results of 41.2 Byr and 43.2 Byr respectively, but the uncertainties of
these latter results are small so that their uncertainty ranges do not overlap. Nevertheless, the three most recent physical
direct counting results were grouped to produce a mean 187Re half-life value of 42.47 1.42 Byr, which is marked by a
vertical dashed line in Fig. 4 together with the uncertainty range shaded. This mean 187Re half-life value is virtually identical
to the value of 42.3 1.3 Byr obtained by Lindner et al. (1989) in their in-growth experiment, which was a refinement of their
initial experiment (Lindner et al. 1986) and so it is ignored in this discussion of the best determination results.The 45.5 1.8
Byr 187Re half-life value obtained in the earlier Luck and Allgre (1982) determination via radioisotope age comparisons of
molybdenites is almost completely outside the range of the mean half-life value from the most recent physical direct
counting experiments and the range of the half-life value from the Lindner et al. (1989) in-growth experiment (fig. 4), so for
this discussion is ignored. On the other hand, the very precise Selby et al. (2007) 187Re half-life value of 41.577 0.12 Byr

plots on Fig. 4 within the left side of the range envelopes of the mean half-life value from the most recent physical direct
counting experiments and of the half-life value from the Lindner et al. (1989) in-growth experiment. It is also almost identical
to the Galeazzi et al. (2001) physical direct counting value for the 187Re half-life of 41.2 0.6 Byr and within its uncertainty
range.Among the determinations obtained by radioisotope age comparisons of groups of iron meteorites, the Luck and
Allgre (1983) 187Re half-life value of 45.6 1.2 Byr plots in Fig. 4 well outside the range envelopes of the mean half-life
value from the most recent physical direct counting experiments and of the half-life value from the Lindner et al. (1989) ingrowth experiment. Furthermore, it does not in any way overlap with any other determination obtained by radioisotope age
comparisons of groups of iron meteorites, so in this discussion it has been ignored. Similarly, the Luck, Birck, and Allgre
(1980) 187Re half-life value of 42.8 Byr falls outside the uncertainties range envelopes of the four post-1990 determinations
by the same method, even though the left-hand limit of its uncertainty range of 1 Byr overlaps with some of the post-1990
determinations and their uncertainties ranges. The post-1990 determinations utilized improved chemical separation and
mass spectrometric techniques, so those four determinations plot close together in Fig. 4 and together yield a mean 187Re
half-life value of 41.78 0.38 Byr. This mean value is very close to and between the 187Re half-life values of 41.577 Byr of
Selby et al. (2007) obtained by radioisotope age comparisons of molybdenites and 42.3 Byr of Lindner et al. (1989) obtained
by an in-growth experiment. Its uncertainty envelope of 0.38 Byr also fully encloses the Selby et al. (2007) 187Re half-life
value of 41.557 0.12 Byr (fig. 4), and is itself fully enclosed within the uncertainty envelopes of the Lindner et al. (1989) ingrowth experiment half-life value and of the mean half-life value of the three most recent physical direct counting
experiments. Thus unlike the determinations of the 87Rb and 176Lu half-lives (Snelling 2014a, b), here there is essential
agreement on the 187Re half-life value obtained by these four different determination methods, but only if the pre-1963
determinations are ignored.But how was such agreement obtained? In particular, how objective were the determinations by
the radioisotope ages comparisons? The Selby et al. (2007) half-life value of 41.557 0.12 Byr and decay constant value of
1.6668 0.034 10-11 per year are now regarded as well-established and thus used by the geochronology community
(Schmitz 2012). Yet these are derived values subject to geological uncertainties and cross-calibration with other radioisotope
systems that are equally subject to similar geological uncertainties and assumptions, as already noted (Begemann et al.
2001). Furthermore, direct physical counting of decay events should be the best measure of the decay rate, followed closely
by in-growth measurements of the resultant decay products over a given time period. However, these more direct methods
are relegated to only potentially affirming the determinations obtained by the radioisotope age comparisons of molybdenites
and groups of iron meteorites, whereas due to the direct nature of the physical counting methods the age comparison
methods with their inherent assumptions involved should be subject to confirmation by them.
If we consider only the post-1990 determinations via radioisotope age comparisons which utilized improved chemical
separation and mass spectrometric techniques, then the first such determination was by Horan et al. (1992). They
performed Re-Os isotopic analyses on eleven iron meteorites and constructed a Re-Os isochron for a group of six iron IIA
meteorites, obtaining a Re-Os isochron age for them of 4596 31 Ma using the Lindner et al. (1989) 187Re decay half-life of
42.3 Byr. However, according to the conventional naturalistic model for the formation of the solar system (even though there
are several good reasons to not accept this model) these iron meteorites could not be older than the age of the solar
system, as defined by Chen and Wasserburg (1981) at 4559 15 Ma based on the mean207Pb/206Pb model age of five
coarse-grained Ca-Al inclusions (CAIs) in the Allende carbonaceous chondrite meteorite. So Horan et al. (1992) adjusted
the 187Re decay half-life to 42.0 Byr in order to force the Re-Os isochron age of these iron IIA meteorites to agree with the
Chen and Wasserburg (1981) Pb-Pb model age. This then was how the Horan et al. (1992) determination of the 187Re halflife was obtained, with several accompanying assumptions, not least being the presumed age of the solar system, the
presumed identical age of the Allende CAIs and these iron IIA meteorites, and the reliability of the Pb-Pb radioisotope
systems, all of which cannot be proven.Subsequently, Smoliar, Walker, and Morgan (1996) used Re-Os isotopic data for six
IIIAB magmatic iron meteorites, which they assumed formed as part of the primary sequence of crystallization of an
asteroidal core, to generate a Re-Os isochron with a slope of 0.07887 22 (2). They then assumed that the age of these
IIIAB iron meteorites was 4557.8 0.4 Ma, identical to the Pb-Pb model age of pyroxene separates from two angrite
meteorites as determined by Lugmair and Galer (1992). Note that the uncertainty in the Pb-Pb age for the angrites did not,
however, take into account uncertainties in the decay constants for 235U and 238U and the critical 238U/235U ratio. The
assumption of a similar age for both these IIIAB iron meteorites and the two angrite meteorites was made because of the
similar 53Mn-53Cr ages of IIIAB iron and angrite meteorites (Hutcheon et al. 1992; Hutcheon and Olsen 1991). However, that
assumption is based on the further assumption that the 53Mn-53Cr ages of those IIIAB iron and angrite meteorites can be
calibrated against the Pb-Pb model ages of two eucrite meteorites and Allende CAIs (Chen and Wasserburg 1985, Manhs,
Gpel, and Allgre 1986, 1988, Birck and Allgre 1988). The high-precision Re-Os iron IIIAB isochron obtained by Smoliar,
Walker, and Morgan (1996), combined with the assumed Pb-Pb model age for the IIIAB irons, permitted them to force the
Re-Os isochron to agree with the Pb-Pb model age in order to determine a more precise and accurate decay constant
for 187Re of 1.666 10-11 per year, corresponding to a half-life of 41.6 Byr.Smoliar, Walker, and Morgan (1996) calculated a
0.31% uncertainty for the derived decay constant value from the regression statistics of their Re-Os isotopic data for those
iron IIIAB meteorites. However, calibration of the Os spike they used included a systematic uncertainty of up to 1.2% (2),
associated with the non-stoichiometry of the Os salt used to prepare the standard solution (Morgan et al. 1995), so this
0.31% uncertainty based just on the regression statistics is likely much lower that it really should be. Furthermore, in
assuming that the iron IIIAB meteorites have an identical Re-Os isochron age to the Pb-Pb model age for the angrite
meteorites, Smoliar, Walker, and Morgan (1996) were assuming that the Re-Os, Pb-Pb, and Mn-Cr radioisotope systems all
isotopically closed at the same time and have not been subsequently disturbed. Those assumptions cannot be proved.
Nevertheless, the uncertainties in the decay constants for the U isotopes, and the assumption of identical closure ages for
Pb-Pb in angrites and Re-Os in IIIAB irons, perhaps contribute less to the overall uncertainty in the decay constant than the
stoichiometry issue.The results from the Smoliar, Walker, and Morgan (1996) study are in good agreement (within 0.4%)
with results published subsequently by Shen, Papanastassiou, and Wasserburg (1996), who obtained a 187Re decay
constant of 1.66 10-11 per year and a half-life of 41.7 Byr, but who used a different Os standard for spike calibration. Shen,
Papanastassiou, and Wasserburg (1996) produced a Re-Os isochron for several iron IIAB meteorites on which other iron
meteorites they analyzed also plotted. They estimated a 1.6% uncertainty in their tracer Os calibration and a combined
uncertainty of about 4% in converting an isochron slope to an absolute age. Nevertheless, in order to determine the 187Re
decay constant of 1.66 10-11 per year and half-life of 41.7 Byr they assumed that the Os tracer calibration was correct, and
assumed that the IIAB irons whole-rock Re-Os isochron age should be adjusted to correspond to the same age as the 4566
Ma Pb-Pb model age for Allende CAIs (Chen and Wasserburg 1985; Manhs, Gpel, and Allgre 1988), the same age as
that for the oldest Pb-Pb model ages obtained for phosphates from ordinary chondrite meteorites, and the same age as that
for angrite and eucrite meteorites (Chen and Wasserburg 1985; Gpel, Manhs, and Allgre 1994; Lugmair and Galer
1992).Shen, Papanastassiou, and Wasserburg (1998) reported that their Re-Os isotopic data for pallasite meteorites plotted
along their same well-defined Re-Os isochron for the IIAB iron meteorites. So using their preliminary Re-Os data for

pallasite meteorites, Shukolyukov and Lugmair (1997) had thus calculated a 187Re decay constant of 1.658 10-11 per year
and a half-life of 41.8 Byr, by adjusting the Re-Os isochron to a Mn-Cr isochron age of 4558 Ma for the Omolon pallasite
meteorite obtained by calibrating that age against the Pb-Pb model age of pyroxene separates from the angrite meteorite
LEW86010 (Lugmair and Galer 1992). This calculation also involved the assumptions that the Omolon pallasite meteorite
formed contemporaneously with other pallasite meteorites, and that the Mn-Cr and Re-Os radioisotope systems closed at
the same time, as well as the assumption that the Omolon pallasite meteorite is the same age as the Pb-Pb model age as
the angrite meteorite LEW86010. Once again, all those assumptions cannot be proved.Subsequently, Birck and Allgre
(1998) obtained a Re-Os isochron age of 4624 Ma for a mixed collection of iron meteorites, which they interpreted as too
high compared to the age of the solar system of 4566 Ma as determined by the U-Pb systematics of Allende CAIs (Manhs,
Gpel, and Allgre 1986, 1988). They thus concluded that a better measurement of the 187Re decay constant and half-life
was highly desirable, but in the meantime the value of the 187Re decay constant should be adjusted to a value between 1.66
10-11 per year and 1.666 10-11 per year.It is interesting to note that these groups of iron meteorites have sometimes
yielded Re-Os isochron ages older than the presumed age of the solar system, so that the 187Re decay constant has had to
be adjusted based on recalibrating the Re-Os isochron ages with the accepted age of the solar system based on Pb-Pb
model ages of other meteorites and their components. However, there is also the possibility that instead this disparity
between Re-Os isochron ages and Pb-Pb model ages is perhaps related to the extreme sensitivity of the 187Re half-life to
the strength of the nuclear force, and to variations in fundamental couplings and the fine-structure constant, as first
discussed by Dyson (1972), and more recently discussed by Olive et al. (2002, 2004) and Olive and Qian (2004). This is
obviously worthy of further investigation, but is beyond the scope of this present contribution.Consequently, in order to
resolve this lingering disagreement over the 187Re decay constant and half-life and to obtain a better measurement, Selby et
al. (2007) embarked on a different approach, using cross calibration of the Re-Os molybdenite and U-Pb zircon
chronometers from magmatic-hydrothermal ore systems. This approach was taken because the Re-Os molybdenite
geochronometer has been claimed to be highly robust (Stein et al. 2001), and because the U-Pb zircon method is widely
acclaimed by uniformitarians as the most accurate and precise method of geochronology as a result of the claimed highly
robust nature of zircon, the dual U-Pb decay scheme utilized, and the claimed most precisely determined decay constants
(Bowring and Schmitz 2003; Heaman and Parrish 1991; Jaffey et al. 1971; Schoene et al. 2006). Selby et al. (2007) thus
assembled the Re-Os model ages for molybdenites and U-Pb concordia-Pb-Pb intercept ages for zircons from 11 magmatichydrothermal ore deposits apparently spanning from ca. 90 Ma to ca. 2670 Ma. The analytical uncertainties were at least 0.2
Ma for both the Re-Os molybdenite and U-Pb zircon age determinations, and yet they assumed that the timing of
crystallization of magmatism causing these magmatic-hydrothermal systems as recorded by the U-Pb zircon ages must be
slightly older than the sulfide mineralization (which includes the molybdenites) in the magmatic-hydrothermal systems
themselves. Nevertheless, they still found agreement between the U-Pb zircon and Re-Os molybdenite ages with an
excellent correlation of 1.0004 0.0023 (2) and probability of 0.99, using the 187Re and U decay constants of Smoliar,
Walker, and Morgan (1996) and Jaffey et al. (1971) respectively. This was statistically indistinguishable from a correlation of
0.9984 0.0021 (2) with a probability of 0.98, using the 238U decay constant of Jaffey et al. (1971) and the 235U decay
constant of Schoene et al. (2006). Then using the Re-Os molybdenite data together with the U-Pb zircon ages for each
deposit, new individual 187Re decay constant values were calculated for each of the 11 magmatic-hydrothermal systems, by
forcing the Re-Os ages to agree with the U-Pb ages. In this way Selby et al. (2007) declared the weighted average of these
11 values of 1.6668 0.0034 10-11 per year (2, with a probability of 0.97) to be the new 187Re decay constant (equivalent
to a half-life of 41.577 Byr), which is now in standard use by uniformitarians in geochronology (Schmitz 2012).
Selby et al. (2007) claimed their cross calibration between the U-Pb zircon and Re-Os molybdenite chronometers provides
an important independent test of the previously used 187Re decay constant, because the methodology, sample materials,
and preparation of the Os standard solution differ greatly from the study of Smoliar, Walker, and Morgan (1996). Indeed, in
many cases the molybdenites and zircons analyzed were found together in the same rock specimens, supposedly
eliminating the need to infer genetic relationships that was necessary in analyses of different meteorite groups. However, it
is well-established that minerals even found side-by-side in the same rock specimens may not have formed at the same
time (for example, Bucher and Frey 2002; Cox, Bell, and Pankhurst 1995), yet that is the assumption that was necessary for
this U-Pb/Re-Os cross calibration study. Nevertheless, Selby et al. (2007) did show that their weighted average 187Re halflife value of 41.557 0.12 Byr using the U decay constants of Jaffey et al. (1971) is identical within the calculated
uncertainty limits to the value of 41.6 0.13 Byr obtained by Smoliar, Walker, and Morgan (1996), which is also based on
the U decay constants of Jaffey et al. (1971) (fig. 5).
Fig. 5. Plot comparing the 187Re half-life values derived from
Selby et al. (2007) to the value of Smoliar, Walker, and Morgan
(1996). The uncertainties are at the 2 level.
Furthermore, their weighted average 187Re half-life value of
41.524 0.12 Byr using the 238U decay constant of Jaffey et al.
(1971) but the 235U decay constant of Schoene et al. (2006) is
also within the calculated uncertainty to the value of 41.6 0.13
Byr obtained by Smoliar, Walker, and Morgan (1996) (fig. 5).
Moreover, they also showed that calculating the Re-Os
molybdenite ages using the 187Re half-life values obtained by
other meteorite radioisotope age comparison studies resulted in
Re-Os molybdenite ages widely divergent from the U-Pb zircon
ages for the same magmatic-hydrothermal systems.
However, this U-Pb gold standard dating method has come
under similar scrutiny in the two last decades. Even in their
study, Selby et al. (2007) admitted that in many of the
magmatic-hydrothermal systems they U-Pb zircon dated, the
data were discordant and showed scatter, which likely impacted
the accuracy and precision of the U-Pb concordia-Pb-Pb
intercept ages used in the cross calibration with the Re-Os
molybdenite ages. Indeed, most of the reported U-Pb ages
were obtained by analyses of multi-grain zircon fractions, which can accentuate the effects of any Pb loss and/or
inheritance. Furthermore, Ludwig (2000) has demonstrated that, although almost universally ignored, the effect of the errors
in the U decay constants on U-Pb concordia-Pb-Pb intercept ages are significant, being a 4.5 Ma error for a 500 Ma age,
which amounts to almost a 1% error. The U-Pb method also depends on the crucial 238U/235U ratio, but discrepancies and

variations have been found between the 238U/235U ratio in U-bearing terrestrial minerals and rocks and the 238U/235U ratio in
meteorites (Brennecka and Wadhwa 2012; Hiess et al 2012). These discrepancies and variations remain unexplained,
especially in the context of the 238U and 235U decay constants and half-lives. Furthermore, the fact that there are these
variations in the crucial 238U/235U ratio in terrestrial minerals and rocks on which the U-Pb dating gold standard depends,
which has been used to re-calibrate Re-Os isochron ages to determine the 187Re half-life and decay constant, only
underscores that these radioisotope methods cannot provide the absolute invariable dates they are so confidently
proclaimed to provide. In any case, there is the additional assumption in all the radioisotope dating methods of having to
know the original concentrations of the daughter and index isotopes which is very significant in the U-Pb method because
the original concentrations are assumed not to be zero, in contrast to the K-Ar method. Yet there must be great uncertainty
as to what those initial values were in the unobserved past, despite the isochron and concordia techniques attempting to
negate the necessity for knowing those initial values. Thus the U-Pb method should not be used as a standard to determine
other radioisotope decay constants.Indeed, it would remain prudent to be very careful with these geological comparison
methods for two other reasons. First, there are significant flaws in the basic assumptions on which all the radioisotope dating
methods depend. Second, the U-Pb method relies primarily on decay, whereas the Re-Os method relies on decay. As
observed by both Austin (2005) and Snelling (2005), these different decay modes seem to yield different ages for some
earth rocks using the same samples with essentially the same methodology. Furthermore, these different ages yielded by
the same earth rocks are often widely divergent, even up 100200% different, which is such a huge divergence that it
renders these dating methods highly suspect, even if the differences in the determinations of the half-lives of the parent
radioisotopes seem miniscule and therefore trivial by comparison. However, it was considered prudent to still document here
these seemingly miniscule differences in half-life values, because they may be indicative of other underlying factors at work
(as already discussed briefly), and they can still lead to very significant discrepancies in the derived radioisotope ages that
might otherwise appear to be acceptably accurate to uniformitarians.In any case, as with 176Lu, there is a dependence of the
half-life of 187Re on environmental conditions such as the degree of ionization (Begemann et al. 2001). And the effects are
not entirely negligible, as is often assumed, so they must significantly impact the dating of terrestrial and meteoritic samples.
Indeed, Bosch et al. (1996) and Bosch (1999) successfully measured by two independent methods the bound state - decay
of fully ionized bare 187Re nuclei and determined a half-life of only 32.9 2.0 yr, compared to the nine orders of magnitude
greater 41.577 0.12 Byr half-life for neutral 187Re atoms. More relevant is the report from Jenkins, Mundy, and Fischbach
(2010) confirming that the previously detected correlations between solar activity and nuclear decay rates were likely not
due to seasonal variations in the sensitivities of the respective laboratory detectors, which were too small to produce the
observed fluctuations. Thus they concluded that the observed fluctuations between solar activity and nuclear decay rates
could arise from a direct interaction between the decaying nuclei and some particles or fields emanating from the sun.
However, calculated solar neutrino absorption cross sections are too small to explain the observed effect, and other
mechanisms have been investigated, such as gravitational effects (Chown 2009; Kostleck and Tasson 2009). With such
environmental influences established, it begs the question of how certain can anyone be that the nuclear decay rates
of 187Re and other parent radioisotopes used in rock dating by uniformitarians have remained constant at todays measured
slow rates through the claimed >4.5 Byr history of the earth and solar system.Nevertheless, all the age comparisons on
terrestrial rocks and minerals, and the age comparisons on meteorites, used the U-Pb method to finally determine the 187Re
decay constant of 1.6668 10-11 per year and half-life of 41.577 Byr, which are within the ranges determined by both
physical direct counting and in-growth experiments. Thus Re-Os age calculations are ultimately calibrated against the U-Pb
method. However, this U-Pb gold standard depends on whether the U decay constants are accurately and precisely
known, and on the crucial 238U/235U ratio. Yet discrepancies and variations have been found between the 238U/235U ratio in Ubearing terrestrial minerals and rocks and the 238U/235U ratio in meteorites. These discrepancies and variations remain
unexplained. This only serves to highlight that if the Re-Os dating method has been calibrated against the U-Pb gold
standard with its own uncertainties, then the claimed accurately-determined 187Re decay rate cannot be absolute, especially
given the environmental influences known to alter nuclear decay rates. Without an accurately known 187Re decay constant
and half-life, accurate Re-Os radioisotope ages cannot be determined. Therefore, Re-Os dating cannot be used to reject the
young-earth creationist timescale, especially as current radioisotope dating methodologies are at best hypotheses based on
extrapolating current measurements and observations back into an assumed deep time history for the cosmos.
Conclusions
There have been numerous attempts to determine the 187Re half-life and decay constant in the last sixty-six years by four
methodsdirect physical counting and in-growth experiments, and radioisotope age comparisons using molybdenites and
groups of iron meteorites. With the measurement technology having improved, the determinations over the last 20 years
have finally resulted in close agreement. Thus the 187Re half-life and decay constant values of 41.577 0.12 Byr and 1.6668
0.0034 10-11 per year respectively have now been adopted for standard use by the uniformitarian geological community.
These adopted values are the weighted averages of eleven determinations in which the Re-Os model ages of molybdenites
were re-calibrated by forcing them to be the same as the U-Pb concordia-Pb-Pb intercept ages of zircons from the same 11
magmatic-hydrothermal systems spanning from ca. 90 Ma to ca. 2670 Ma.This adopted half-life value of 41.577 0.12 Byr
is virtually identical to the 41.6 0.13 Byr value obtained by forcing the Re-Os isochron age of a group of iron meteorites to
agree with the presumed Pb-Pb model age for the solar system obtained from pyroxene separates from two angrite
meteorites. Furthermore, the uncertainty range of the adopted 41.577 0.12 Byr half-life value lies within the uncertainty
range of the 41.78 0.38 Byr mean half-life value of the four post-1990 determinations using Re-Os/Pb-Pb age
comparisons for groups of iron meteorites. And the adopted value also lies within the uncertainty range of the 42.3 1.3
Byr half-life value obtained by an in-growth experiment, and within the uncertainty range of the 42.47 1.42 Byr mean
half-life value of the three most recent physical direct counting experiments. Yet even though there is this close agreement
between these four determination methods, the half-life value obtained by comparison of Re-Os molybdenite ages with UPb zircon ages has been given preference over the values obtained by direct counting and in-growth experiments which
directly measure 187Re decay and the187Os product of 187Re decay respectively, and so are independent of all the
assumptions involved with the radioisotope dating methods. Indeed, model dependent results should not take precedence
over the direct experimental evidence.All the age comparisons on molybdenites and groups of iron meteorites used the UPb method to determine the 187Re decay half-life, and thus all Re-Os age calculations are ultimately calibrated against the UPb method. However, this U-Pb gold standard depends on having precisely known 238U and 235U decay constants, as well
as on the crucial238U/235U ratio. Yet there are still uncertainties in the measured U decay constants, and discrepancies and
variations have been found between the 238U/235U ratio in U-bearing terrestrial minerals and rocks, and the 238U/235U ratio in
meteorites. These discrepancies and variations remain unexplained. This only serves to highlight that if the Re-Os dating
method has been calibrated against the U-Pb gold standard with its own uncertainties, then it cannot be absolute.In any
case, the half-life of 187Re has also been shown to be dependent on environmental conditions. Fluctuations in nuclear decay
rates have been observed to correlate with fluctuations in solar activity, and the degree of ionization of187Re atoms can result

in a change of nine orders of magnitude in the measured decay rate. Furthermore, there is evidence that nuclear decay
rates have not been constant in the past. Thus without an accurately known 187Re decay constant and half-life, accurate ReOs radioisotope ages cannot be determined. Therefore, Re-Os dating cannot be used to reject the young-earth creationist
timescale, especially as current radioisotope dating methodologies are at best hypotheses based on extrapolating current
measurements and observations back into an assumed deep time history for the cosmos.
Acknowledgments
The invaluable help of my research assistant Lee Anderson, Jr. in compiling and plotting the data in the color-coded age
diagrams is acknowledged. The three reviewers are also acknowledged for their helpful comments and input, though the
final content of this paper is solely my responsibility. Our production assistant Laurel Hemmings is also especially thanked
for her painstaking work in preparing this paper for publication.
Determination of the Radioisotope Decay Constants and Half-Lives: Samarium-147 ( 147Sm)
by Dr. Andrew A. Snelling on June 10, 2015
Abstract
Over the last 80 years numerous determinations have been made of the 147Sm half-life. The determinations since 1960 have
converged on close agreement between the two primary determination techniques used in direct physical counting
experimentsionization chambers and liquid scintillation counters, and with radioisotope age comparisons using two
meteorites. Thus the147Sm half-life value of 106 0.8 Byr has now been adopted for standard use by the uniformitarian
geological community. This value is based on the weighted average of four direct counting determinations in the period
19611970 and the recalibration in the 1970s of Sm-Nd model ages of two meteorites by forcing them (essentially by
circular reasoning) to agree with their Pb-Pb isochron and model ages. However, direct counting experiments in 2003
determined the 147Sm half-life value was 10% or more longer at 117 2 Byr. This was achieved by using four standard Sm
solutions with internal -radioactive standards in 19 alpha spectrometer and 24 ionization chamber determinations, making
it the most thorough and comprehensive effort to determine the 147Sm half-life. The thinner counting sources used, while
resulting in low -activities being measured, greatly reduced the counting uncertainty due to self-absorption of the emitted particles. Although rejected or ignored, this 117 2 Byr value for the 147Sm half-life, which agrees with some earlier
determinations, may well be highly significant and more reliable than the adopted value. Yet, in spite of the many
experiments directly measuring 147Sm decay, preference has been given to the half-life value of 106 0.8 Byr determined by
forcing the Sm-Nd data to agree with Pb-Pb dates. But many unprovable assumptions are also involved, not the least being
that the radioisotope systems closed at the same time and subsequently remained closed. Furthermore, even this gold
standard has unresolved uncertainties due to the U decay constants being imprecisely known, and to measured variations
of the 238U/235U ratio in terrestrial rocks and minerals and in meteorites. Both of these factors are so critical to the U-Pb
method, as well as the additional factor of knowing the initial concentrations of the daughter and index isotopes, so it should
not be used as a standard to determine other decay constants. In any case, the determined half-life of 147Sm has been
shown to be dependent on the thicknesses of the Sm counting source and the detector. There is also evidence decay rates
of the radioisotopes used for rock dating have not been constant in the past. This only serves to emphasize that if the SmNd dating method has been calibrated against the U-Pb gold standard with all its attendant uncertainties, then it cannot be
absolute, and therefore it cannot be used to reject the young-earth creationist timescale. Indeed, current radioisotope dating
methodologies are at best hypotheses based on extrapolating current measurements and observations back into an
assumed deep time history for the cosmos.
Keywords: radioisotope dating, decay constants, half-lives, samarium-147, 147Sm, decay, direct counting, emulsions,
liquid scintillation counters, ionization chambers, surface-barrier detectors, Geiger counter, counting efficiencies, geological
comparisons, meteorites, U-Pb gold standard, 238U/235U ratio
Introduction
Radioisotope dating of rocks and meteorites is perhaps the most potent claimed proof for the supposed old age of the earth
and the solar system. The absolute ages provided by the radioisotope dating methods provide an apparent aura of certainty
to the claimed millions and billions of years for formation of the earths rocks. Many in both the scientific community and the
general public around the world thus remain convinced of the earths claimed great antiquity.However, accurate radioisotopic
age determinations require that the decay constants of the respective parent radionuclides be accurately known and
constant in time. Ideally, the uncertainty of the decay constants should be negligible compared to, or at least be
commensurate with, the analytical uncertainties of the mass spectrometer measurements of isotopic ratios entering the
radioisotope age calculations (Begemann et al. 2001). Clearly, based on the ongoing discussion in the conventional
literature this is not the case at present. The stunning improvements in the performance of mass spectrometers during the
past four or so decades, starting with the landmark paper by Wasserburg et al. (1969), have not been accompanied by any
comparable improvement in the accuracy of the decay constants (Begemann et al. 2001; Steiger and Jger 1977), in spite
of ongoing attempts (Miller 2012). The uncertainties associated with most direct half-life determinations are, in most cases,
still at the 1% level, which is still significantly better than any radioisotope method for determining the ages of rock
formations. However, even uncertainties of only 1% in the half-lives lead to very significant discrepancies in the derived
radioisotope ages. The recognition of an urgent need to improve the situation is not new (for example, Min et al. 2000;
Renne, Karner, and Ludwig 1998). It continues to be mentioned, at one time or another, by every group active in geo- or
cosmochronology (Schmitz 2012). This is a key issue especially for very long half-life radioisotopes due to the very slow
accumulation of decay particle counting data, because the statistical error is equal to the square root of the total decay
particle counts.From a creationist perspective, the 19972005 RATE (Radioisotopes and the Age of The Earth) project
successfully made progress in documenting some of the pitfalls in the radioisotope dating methods, and especially in
demonstrating that radioisotope decay rates may not have always been constant at todays measured rates (Vardiman,
Snelling, and Chaffin 2000, 2005). Yet much research effort remains to be done to make further inroads into not only
uncovering the flaws intrinsic to these long-age dating methods, but towards a thorough understanding of radioisotopes and
their decay during the earths history within a creationist framework.One crucial area the RATE project did not touch on was
the issue of how reliable have been the determinations of the radioisotope decay rates, which are so crucial for calibrating
these dating clocks. Indeed, before this present series of papers (Snelling 2014a, 2014b, 2015) there have not been any
attempts in the creationist literature to review how the half-lives of the parent radioisotopes used in long-age geological
dating have been determined and to collate their determinations so as to discuss the accuracy of their currently accepted
values. After all, accurate radioisotope age determinations depend on accurate determinations of the decay constants or
half-lives of the respective parent radioisotopes. The reliability of the other two assumptions these absolute dating methods
rely on, that is, the starting conditions and no contamination of closed systems, are unprovable. Yet these can supposedly
be circumvented somewhat via the isochron technique, because it is independent of the starting conditions and is sensitive
to revealing any contamination, which is still significantly better than any radioisotope method for determining the ages of
rock formations. Data points that do not fit on the isochron are simply ignored because their values are regarded as due to

contamination. Yet there is also no reliable way of determining the difference between isochrons and mixing lines. That this
is common practice is illustrated with numerous examples from the literature by Dickin (2005) and Faure and Mensing
(2005). On the other hand, it could be argued that this discarding of data points which do not fit the isochron is arbitrary and
therefore is not good science, because it is merely assumed the aberrant values are due to contamination rather than that
being proven to be so. Indeed, in order to discard such outliers in any data set, one must establish a reason for discarding
those data points which cannot be reasonably questioned.In order to rectify this deficiency, Snelling (2014a, 2014b, 2015)
has documented the methodology behind and history of determining the decay constants and half-lives of the parent
radioisotopes 87Rb, 176Lu, and 187Re used as the basis for the Rb-Sr, Lu-Hf, and Re-Os long-age dating methods respectively.
He showed that there is still some uncertainty in what the values for these measures of the 87Rb and 176Lu decay rates
should be, in contrast to the apparent agreement on the 187Re decay rate. This uncertainty is especially prominent in
determinations of the 176Lu decay rate by physical direct counting experiments. Furthermore, the determined values of
the 87Rb decay rate differ when Rb-Sr ages are calibrated against the U-Pb ages of either the same terrestrial minerals and
rocks or the same meteorites and lunar rocks. Ironically it is the slow decay rate of isotopes such as 87Rb used for deep time
dating that makes a precise measurement of that decay rate so difficult. Thus it could be argued that direct measurements
of these decay rates should be the only acceptable experimental evidence, especially because measurements which are
calibrated against other radioisotope systems are already biased by the currently accepted methodology that the secular
community uses in their rock dating methods. Indeed, the 87Rb, 176Lu, and 187Re decay half-lives have all ultimately been
calibrated against the U-Pb radioisotope system, yet there are now known measured variations in the 238U/235U ratio that is
critical to that method (Brennecka and Wadhwa 2012; Hiess et al. 2012).Therefore, the aim of this contribution is to further
document the methodology behind and history of determining the present decay constants and half-lives of the parent
radioisotopes used as the basis for the long-age dating methods. It is necessary to explore just how accurate these
determinations are, whether there really is consensus on standard values for the half-lives and decay constants, and just
how independent and objective the standard values are from one another between the different methods. Of course, it is to
be expected that every long-lived radioactive isotope is likely to show similar variation and uncertainty in half-life
measurements because these are difficult measurements to make. However, even small variations and uncertainties in the
half-life values result in large variations and uncertainties in the calculated ages for rocks, and the question remains as to
whether the half-life values for each long-lived parent radioisotope are independently determined. We continue here with
samarium-147 (147Sm), which is the basis for the Sm-Nd dating method.
Samarium and Samarium-147 Decay
Samarium (Sm) and neodymium (Nd) are both rare-earth elements (REEs) with atomic numbers (Z) of 62 and 60
respectively. The rare-earth elements generally form ions with a 3+ charge whose radii decrease with increasing atomic
number from 1.15 in lanthanum (La), atomic number 57, to 0.93 in lutetium (Lu), atomic number 71 (Faure and Mensing
2005). The REEs occur in high concentrations in several economically important minerals such as bastnaesite (CeFCO 3),
monazite (CePO4), and cerite [(Ca,Mg)2(Ce)8(SiO4)7 3H2O]. Furthermore, they occur as trace elements in common rockforming minerals (silicates, phosphates, and carbonates) in which they replace major element ions. They may also reside in
inclusions of certain accessory minerals in the common rock-forming silicates.Minerals exercise a considerable degree of
selectivity in admitting REEs into their crystal structures (Faure and Mensing 2005). Feldspar, biotite, and apatite tend to
concentrate the light REEs (the Ce group), whereas pyroxenes, amphiboles, and garnet concentrate the heavy REEs (the
Gd group). The selectivity of the rock-forming minerals for the light or heavy REEs obviously affects the REE concentrations
of the rocks in which those minerals occur. Sm and Nd both belong to the light REEs, so they tend to concentrate in
feldspar, biotite, and apatite. Thus the Sm and Nd concentrations in calc-alkaline plutonic and volcanic igneous rocks range
from <1 ppm in ultramafic rocks to about 8 ppm Sm and 45 ppm Nd in granite. Alkali-rich igneous rocks have consistently
higher Sm and Nd concentrations than the calc-alkaline suite, ranging up to about 15 ppm Sm and 85 ppm Nd.Sm and Nd
exhibit an unusual geochemical behavior, which arises from what is known as the lanthanide contraction (Faure and
Mensing 2005). This contraction results from the way electrons fill their f shell orbitals. As a consequence, the ionic radius of
Sm (Z = 62) is smaller than that of Nd (Z = 60). Even though the difference in the radii is small (Nd 3+ = 1.08 ; Sm3+ = 1.04
), Nd is preferentially concentrated in the liquid phase during partial melting of silicate minerals, whereas Sm remains in the
residual solids. For this reason, basalt magmas have lower Sm/Nd ratios than the source rocks from which they formed.
Thus this preferential partitioning of Nd into the melt phase has caused the rocks of the continental crust to be enriched in
Nd relative to Sm compared to the residual rocks in the lithospheric mantle.Even though the concentrations of Sm and Nd
reach high values in accessory phosphate minerals such as apatite and monazite and in carbonatites, these minerals and
carbonatites are still more enriched in Nd than in Sm and hence their Sm/Nd ratios are less than 0.32. Among the rockforming silicate minerals, garnet is the only one with a high Sm/Nd ratio (0.54) even though its concentrations of Sm and Nd
are both low (12 ppm). Several other rock-forming silicate minerals, such as K-feldspar, biotite, amphibole, and
clinopyroxene, have higher Sm and Nd concentrations than garnet, but their Sm/Nd ratios are less than 0.32 in most cases.
Sm and Nd each have seven naturally-occurring isotopes. Of these 147Sm, 148Sm, and 149Sm are all radioactive, but the latter
two have such long half-lives (about 10 16 years) that they are not capable of producing measurable variations in the
daughter isotopes 144Nd and 145Nd, even over supposed conventional cosmological intervals (1010 years) (Dickin 2005).
Yet 147Sm only has an abundance of 15.0% in naturally occurring Sm (Lide and Frederikse 1995). Although the half-life
of 147Sm is also very long (currently determined as 106 billion years), it decays by -particle emission to 143Nd, a stable
isotope of Nd. The relevant decay scheme is often depicted as:
147
Sm 143Nd + 4He + E
where 4He is an -particle and E is the total decay energy. The energy of the 147Sm emitted -particles is 2.23 MeV. This
decay scheme has proven useful to uniformitarians for apparently dating terrestrial rocks, stony meteorites (both chondrites
and achondrites), and lunar rocks.There are two parameters by which the decay rate is measured and expressed, namely,
the decay constant () and the half-life (t). The decay constant can be defined as the probability per unit time of a particular
nucleus decaying, though strictly speaking probabilities do not have units associated with them and the decay constant is
derived from a definitive functional relationship. In contrast, the half-life is the time it takes for half of a given number of the
parent radionuclide atoms to decay. The two quantities can be almost used interchangeably, because they are related by the
equation:t = ln 2/ = 0.693/
The decay rate of 147Sm has not been all that difficult to determine once the necessary instrumentation was developed to
accurately count the emitted -particles. However, the Sm-Nd dating method has had its problems.The lanthanide
contraction causes the distribution of Sm and Nd to be opposite to that of Rb and Sr (Faure and Mensing 2005). And
because Sm and Nd have very similar chemical properties (unlike Rb and Sr), large ranges of Sm/Nd ratios in whole-rock
systems are rare, and in particular low Sm/Nd ratios near the vertical y-axis on an isochron dating graph are very rare.
Therefore, because of the difficulty of obtaining a wide range of Sm/Nd ratios from a single rock body, and because of the

greater technical demands of Nd isotope analysis, the Sm-Nd isochron dating method has been generally only applied to
dating rock units for which Rb-Sr isochron dating has proven unsatisfactory. Many of those applications were also made
before the U-Pb zircon dating method had reached its present level of development. Therefore, some of those rock units
have subsequently been dated to apparent greater accuracy and precision by the U-Pb method. Nevertheless, the Sm-Nd
method has continued to be used to date rocks and meteorites and thus the determination of the 147Sm half-life used by the
method requires examination.
Determination Methods
Attempts to measure the -radioactivity of Sm date back to the early 1930s. Famous names, such as Hevesy and Pahl,
Curie and Joliot, Libby, and many more, are among the researchers who studied this phenomenon by various techniques
(Begemann et al. 2001). At that time, before 147Sm had finally been identified as the isotope accountable for the radioactivity of Sm (Weaver 1950), the half-life was calculated in terms of the total element of Sm, with results ranging from
0.63 to 1.4 1012 years (that is, 630 to 1400 Byr). Even in 1949, when 148Sm (Wilkins and Dempster 1938) and 152Sm
(Dempster 1948) had been reported erroneously to be responsible for the -activity of samarium, Picciotto still published his
result in terms of total Sm as 6.7 0.4 10 11 years (670 Byr), this quoted statistical error amounting to approximately 300
observed decays of 147Sm. Almost all attempts to measure the -radioactivity of 147Sm have been by direct counting of particles, although two geological comparisons of radioisotope ages of individual meteorites have been used to confirm the
direct counting measurements of the 147Sm half-life.
Direct counting
Initially direct measurements were done by counting -particles registered on a photographic emulsion. For example,
Picciotto (1949) used solutions of samarium sulfate to deposit layers of calculated weights of samarium across photographic
emulsions which were then left exposed to the samarium -activity for four to more than 19 days. At the end of each
exposure time interval the tracks left by the -particles on the emulsions when developed were counted. The calculated
numbers of -particles per second per gram of Sm were averaged to derive a decay rate in terms of the total Sm (all
isotopes) of 6.7 0.4 1011 years (670 Byr). Even though Beard and Wiedenbeck (1954) obtained a 147Sm half-life value
using an ionization chamber (Geiger counter), they corroborated their result by checking the energy spectrum of their
samarium source by exposing it for three weeks to a nuclear plate, which registered the tracks left by the -particles. The
mean track length compared favorably with previous emulsion experiments, and the energy distribution confirmed the tracks
were produced by -particles emitted by 147Sm.A more recent use of the emulsion method by Martins, Terranova, and
Moreira Correa (1992) involved spreading a calibrated solution of samarium nitrate over a glass plate which was then
heated to leave a stable film of samarium oxide whose uniform thickness was measured. This method enabled accurate
control, within the limits of volumetric error, of the quantity of samarium deposited on the glass plate. This glass plate was
then contacted against a plastic emulsion plate for registration of the spontaneous 147Sm -particle emission. After exposure
for 30 days, stored in an underground laboratory to protect it from cosmic radiation, the emulsion was processed to etch the
tracks left by the -particles, which were then counted. The samarium oxide film was also checked by -spectrometry to rule
out any contribution to the counted tracks from the possible presence of any U or Th atoms. A measurement of the 147Sm
half-life was then calculated from the number of -particle tracks produced from the known quantity of samarium in the
exposure time period of 30 days. Though they did not account for possible -particle energy loss in the samarium or the
glass and plastic plates, this possibility was deemed negligible.However, since the mid-1950s a variety of instruments have
primarily been used for direct counting of 147Sm -particles. In these instruments, the alpha () activity of 147Sm in a source
material is counted over a designated time period, and divided by the total number of radioactive 147Sm atoms in the known
quantity of Sm, based on Avogadros number and the isotopic abundance of 147Sm. Two types of such instruments have
been used to count the -particles emitted from different 147Sm sourcesliquid scintillation spectrometers (Beard and Kelly
1958; Donhoffer 1964; Kinoshita, Yokoyama, and Nakanashi 2003; Kossert et al. 2009; Wright, Steinberg, and Glendenin
1961), and ionization chambers or Geiger counters (Beard and Wiedenbeck 1954; Gupta and MacFarlane 1970; Karras and
Nurmia 1960; Kinoshita, Yokoyama, and Nakanashi 2003; MacFarlane and Kohman 1961). A variety of 147Sm sources have
also been used, namely, samarium octoate (Wright, Steinberg, and Glendenin 1961), samarium oxide (Beard and
Wiedenbeck 1954; Kinoshita, Yokoyama, and Nakanashi 2003; Kossert et al. 2009; MacFarlane and Kohman 1961; Su et al.
2010), and samarium metal (Gupta and MacFarlane1970; Su et al. 2010).In the most recent half-life determination, Su et al.
(2010) deposited 147Sm-enriched metal and oxide on pure quartz glass substrates by vacuum evaporation and sputtering
respectively, and then checked the uniformity of the thicknesses of these samples by exposing them to plastic emulsion
plates in direct contact with them. However, in using this method it is difficult to compensate for -energy loss due to
absorption. After 100 hours exposure the emulsion plates were chemically etched and the -particle tracks were observed
and counted. This does not appear to be a good way of establishing uniformity of thickness, whereas -gauging would be a
better method. Nevertheless, the -activities of the two samples were then measured by silicon surface-barrier detectors
placed in vacuum chambers for a period of 200 hours (8 days 8 hours). Finally the 147Sm half-life was then calculated from
the -activity spectra for each sample (Sm metal and Sm oxide).A liquid scintillation counter or spectrometer detects and
measures ionizing radiation by using the excitation effect of incident -particles on a scintillator material, and detecting the
resultant light pulses. It consists of a scintillator which generates photons of light in response to incident -particles, a
sensitive photomultiplier tube which converts the light to an electrical signal, and electronics to process this signal.Liquid
scintillation counting measures the -activity of a sample, prepared by mixing the -active material with a liquid scintillator,
and counting the resultant photon emissions. This allows for more efficient counting due to the intimate contact of the activity with the scintillator. Samples are dissolved or suspended in a cocktail containing a solvent, typically some form of a
surfactant, and small amounts of other additives known as fluors or scintillators.The radioactive sample is then placed in a
vial containing a premeasured amount of scintillator cocktail and this vial plus vials containing known amounts of 147Sm are
loaded into the liquid scintillation counter. Many counters have two photomultiplier tubes connected in a coincidence circuit.
The coincidence circuit assures that genuine light pulses, which reach both photomultiplier tubes, are counted, while
spurious pulses (due to line noise, for example), which would only affect one of the tubes, are ignored.When a charged
particle strikes the scintillator, its atoms are excited and photons are emitted. These are directed at the photomultiplier tubes
photocathode, which emits electrons by the photoelectric effect. These electrons are electrostatically accelerated and
focused by an electrical potential so that they strike the first dynode of the tube. The impact of a single electron on the
dynode releases a number of secondary electrons which are in turn accelerated to strike the second dynode. Each
subsequent dynode impact releases further electrons, and so there is a current amplifying effect at each dynode stage.
Each stage is at a higher potential than the previous stage to provide the accelerating field. The resultant output signal at the
anode is in the form of a measurable pulse for each photon detected at the photocathode, and is passed to the processing
electronics. The pulse carries information about the energy of the original incident -radiation on the scintillator. Thus both
the intensity and energy of the -particles can be measured.The scintillation spectrometer consists of a suitable scintillator
crystal, a photomultiplier tube, and a circuit for measuring the height of the pulses produced by the photomultiplier. The

pulses are counted and sorted by their height, producing an xy plot of scintillator flash brightness versus number of flashes,
which approximates the energy spectrum of the incident -radiation.The ionization chamber is the simplest of all gas-filled
radiation detectors, and is widely used for the detection and measurement of certain types of ionizing radiation, including particles (fig. 1). Conventionally, the term ionization chamber is used exclusively to describe those detectors which collect
all the charges created by direct ionization within the gas through the application of an electric field. It only uses the discrete
charges created by each interaction between the incident radiation and the gas, and does not involve the gas multiplication
mechanisms used by other radiation instruments, such as the Geiger- Mller counter or the proportional counter.

Fig. 1. Schematic diagram of the cylindrical ionization chamber (alpha spectrometer) used by Gupta and MacFarlane (1970)
in their determination of the 147Sm half-life.
An ionization chamber measures the current from the number of ion pairs created within a gas caused by incident radiation. It consists of a gas-filled chamber with two electrodes, known as the anode and cathode (fig. 1). The electrodes
may be in the form of parallel plates, or a cylinder arrangement with a coaxially located internal anode wire. A voltage is
applied between the electrodes to create an electric field in the fill gas. When gas between the electrodes is ionized by
incident ionizing -radiation (or - or -radiation), ion-pairs are created and the resultant positive ions and dissociated
electrons move to the electrodes of the opposite polarity under the influence of the electric field. This generates an ionization
current which is measured by an electrometer. The electrometer must be capable of measuring the very small output current
which is in the region of femtoamperes (10-15 amps) to picoamperes (10-12 amps), depending on the chamber design, radiation dose and applied voltage. The unique feature of ionization chambers is that the electric field strength is low enough
that no multiplication of ion pairs occurs. Hence the current generated at a given voltage depends on the type and energy of
the incident radiation but is independent over a range of applied voltages, approximately 100300 volts.Each ion pair
creates deposits or removes a small electric charge to or from an electrode, such that the accumulated charge is
proportional to the number of ion pairs created, and hence the -radiation dose. This continual generation of charge
produces an ionization current, which is a measure of the total ionizing dose entering the chamber. The electric field also
enables the device to work continuously by mopping up electrons, which prevents the fill gas from becoming saturated,
where no more ions could be collected, and by preventing the recombination of ion pairs, which would diminish the ion
current. This mode of operation is referred to as current mode, meaning that the output signal is a continuous current, and
not a pulse output as in the cases of the Geiger-Mller tube or the proportional counter. In the ionization chamber operating
region the collection of ion pairs is effectively constant over a range of applied voltage, as due to its relatively low electric
field strength the ion chamber does not have any multiplication effect. This is in distinction to the Geiger-Mller tube or the
proportional counter whereby secondary electrons, and ultimately multiple avalanches, greatly amplify the original ioncurrent charge. In the proportional counter the electric field produces discrete, controlled avalanches such that the energy
and type of radiation can be determined at a given applied field strength.The GeigerMller counter, also called a Geiger
counter, is also used for measuring ionizing radiation. It detects radiation such as -particles, using the ionization produced
in a GeigerMller tube. The processing electronics displays the result. The Geiger-Mller tube is filled with an inert gas
such as helium, neon, or argon at low pressure, to which a high voltage is applied. The tube briefly conducts an electrical
charge when a particle or photon of incident -radiation makes the gas conductive by ionization. The ionization is
considerably amplified within the tube by an avalanche effect to produce an easily measured detection pulse, which is fed to
the processing and display electronics. The electronics also generates the high voltage, typically 10001400 volts, which
has to be applied to the Geiger-Mller tube to enable its operation. The voltage must be high enough to produce avalanche
effects for all incident radiation. Thus a Geiger-Mller tube has no ability to discriminate between incident radiations; all
radiation produces the same current.There are two main limitations of the Geiger counter. Because the output pulse from a
Geiger- Mller tube is always the same magnitude regardless of the energy of the incident -radiation, the tube cannot
differentiate between radiation types. A further limitation is the inability to measure high -radiation intensities due to the
dead time of the tube. This is an insensitive period after each ionization of the gas during which any further incident radiation will not result in a count, and the indicated rate is therefore lower than actual. Typically the dead time will reduce
indicated count rates above about 10 4 to 105 counts per second depending on the characteristic of the tube being used.
Whilst some counters have circuitry which can compensate for this, for accurate measurements ion chamber instruments
are preferred to measure high radiation rates.Counting efficiencies of liquid scintillation counters under ideal conditions
range from about 30% for tritium (a low-energy -emitter), to nearly 100% for phosphorus-32 ( 32P), a high-energy -emitter.
Thus counting efficiencies for -particles are within this range, which is <100%. Some chemical compounds (notably
chlorine compounds) and highly colored samples can interfere with the counting process. This interference, known as
quenching, can be overcome through data correction or through careful sample preparation. In ionization counters the
counting efficiency is the ratio between the number of -particles or photons counted and the number of -particles or
photons of the same type and energy emitted by the -radiation source. Counting efficiencies vary for different isotopes and
sample compositions, and for different scintillation counters. Poor counting efficiency can be caused by an extremely low
energy to light conversion rate (the scintillation efficiency), which, even optimally, will be a small value. It has been

calculated that only some 4% of the energy from a -emission event is converted to light by even the most efficient
scintillation cocktails. Proportional counters and end-window Geiger-Mller tubes have a very high efficiency for all ionizing
particles that reach the fill gas. Nearly every initial ionizing event in the gas will result in avalanches, and thereby an output
signal. However the overall detector efficiency is largely affected by attenuation due to the window or tube body through
which particles have to pass. They are also extremely sensitive to various types of background radiation due to their lack of
discrimination.
Judged from the fact that many of these direct counting experiments, particularly the earlier ones, have yielded results that
are not compatible with one another within the stated uncertainties (see below), it would appear that not all the
measurement uncertainties are accounted for, and therefore the stated uncertainties are likely unrealistically small and
typically are underestimated. Begemann et al. (2001) maintain that many of such experiments are likely plagued by
unrecognized systematic errors. As the nature of these errors is obscure, it is not straightforward to decide which of the,
often mutually exclusive, results of such direct counting experiments is closest to the true value, although most of the postearly-1960s experiments appear to converge on a common value (see below). Furthermore, the presence of unknown
systematic biases makes any averaging dangerous. It is possible that reliable results of careful workers, listing realistic
uncertainties, will not be given the weights they deservethis aside from the question of whether it makes sense to average
numbers that by far do not all agree within the stated uncertainties.
Geological comparisons of methods
A second approach used by secular scientists to determine the 147Sm decay half-life has been to date geological samples
whose ages have also been measured by other methods with presumably more reliable decay constants (Dickin 2005;
Faure and Mensing 2005). This essentially involves circular reasoning, because it is being assumed the other radioisotope
dating methods, principally the U-Pb method, gives the reliable dates to which the 147Sm half-life can be calibrated to bring
the Sm-Nd radioisotope ages into agreement. It should be noted, however, that this is hardly objective, because all the
radioisotope ages of rocks could be wrong due to the underlying unprovable and suspect assumptions on which all the
radioisotope dating methods are based. Nevertheless, a few geological determinations of the 147Sm half-life were made in
the 1970s using components of individual achondrite meteorites (Dickin 2005, 7071; Lugmair 1974; Lugmair, Scheinin, and
Marti 1975; Lugmair and Marti 1977).This method has the disadvantage that it involves geological uncertainties, such as
whether all isotopic systems closed at the same time and remained closed. However, it is claimed to still provide a useful
check on the laboratory determinations by direct physical counting. Nevertheless, this approach entails multi-chronometric
dating of minerals and components in individual meteorites and cross-calibration of different radioisotopic age systems by
adjusting the decay constant of the Sm-Nd system so as to force agreement with the age obtained via another dating
system, usually U-Pb (Begemann et al. 2001). In essence, because the half-life of 238U is claimed to be the most accurately
known of all relevant radionuclides, this usually amounts to expressing ages in units of the half-life of 238U.
Results of the Samarium-147 Decay Determinations
During the last 80 years numerous determinations of the 147Sm decay constant and half-life have been made using these
methods. The results are listed with details in Table 1. The year of the determination versus the value of the half-life is
plotted in Fig. 2. In each case the data points plotted have been color-coded the same to differentiate the values as
determined by the two approaches that have been useddirect counting, and geological comparisons with other
radioisotope dating methods.
Table 1. Determinations of the 147Sm decay rate expressed in terms of the half-life using direct physical counting
experiments, and comparisons of radioisotope ages of terrestrial minerals and rocks, and meteorites.
Determination of the Sm Decay Rate
Date

Half-Life Uncertainty
(Byr)
(Byr)

Method

Instrument

Notes

Source

1936 150

11

direct counting

emulsion plate

Hosemann 1936

1949 100

direct counting

emulsion plate

Picciotto 1949

1954 125

direct counting

ionization chamber

1958 128

direct counting

liquid scintillation

Beard and Kelly 1958

1960 114

direct counting

ionization chamber

Karras
1960

1960 117

direct counting

ionization chamber

Karras 1960

1961 113

direct counting

Graeffe
Nurmia 1961

1961 115

direct counting

ionization chamber

MacFarlane
Kohman 1961

1961 105

direct counting

liquid scintillation

Wright, Steinberg, and


Glendenin 1961

1964 104

direct counting

liquid scintillation

Donhoffer 1964

1965 108

direct counting

liquid scintillation

Valli et al. 1965

1970 106

direct counting

ionization chamber

Gupta and MacFarlane


1970

4 Geiger counter

Beard and Wiedenbeck


1954

and

Nurmia

and
and

1975 106

geological
comparisons

Juvinas
basalticLugmair, Sheinin, and
achondrite meteorite
Marti 1975

1977 106

geological
comparisons

Angra
dos
Reis
achondrite meteorite
Lugmair and Marti 1977

weighted average of
four measures
Lugmair and Marti 1978

1978 106

0.8

direct counting

1987 105

direct counting

1992 123

direct counting

emulsion plate

Martins, Terranova, and


Moreira Correa 1992

2001 106

direct counting

emulsion plate

correction of Martins et
al. 1992
Begemann et al. 2001

2003 117

direct counting

alpha spectrometer

with vacuum chamberKinoshita,


Yokoyama,
(also liquid scintillation) and Nakanashi 2003

2009 107

0.9

direct counting

liquid scintillation

2010 106

direct counting

alpha spectrometer

with vacuum chamber,


Sm metal
Su et al. 2010

2010 107

direct counting

alpha spectrometer

with vacuum chamber,


Sm oxide
Su et al. 2010

Al-Bataina and Jnecke


1987

Kossert et al. 2009

Fig. 2. Plot of each 147Sm half-life determination


versus the year of its determination, color-coded
according to the method of its determination. The
error bars for each determination are also plotted
from the error values listed in Table 1.
Discussion
Results obtained after 1954, and particularly
during the 1960s and the ensuing decade, began
to converge towards a147Sm common half-life
value (see table 1 and fig. 2). In the early 1970s,
when Lugmair and his colleagues began to
develop the decay of 147Sm to 143Nd as a dating
tool (Lugmair 1974), they used only a weighted
average of the last four half-life measurements at
that time, those of Wright, Steinberg, and
Glendenin (1961), Donhoffer (1964), Valli et al.
(1965), and Gupta and MacFarlane (1970)
(Lugmair and Marti 1978). They range from 1.04
to 1.08 1011 years (104 to 108 Byr) with a
weighted mean of 1.060 0.008 1011 years
(106 0.8 Byr) (1 uncertainty). It is worth noting
that the statistical error of this weighted mean
infers that more than 10,000 -decays of 147Sm have occurred to produce this result. Nevertheless, it should be noted that
three of those measurements were made using liquid scintillation counters and one using an ionization chamber (see table
1), all of which have different counting efficiencies, as already discussed. Nevertheless, that 147Sm half-life value has been
adopted by all geochronologists and cosmochronologists since that time (the late 1970s).There have been five more
modern measurements of the 147Sm half-life since the 1970s (table 1). The first by Al-Bataina and Jnecke (1987) with a
value of 1.05 0.04 1011 years (105 4 Byr) agrees very well with the previous direct counting results of Donhoffer (1964),
Gupta and MacFarlane (1970), Valli et al. (1965), and Wright, Steinberg, and Glendenin (1961), as well as with the
geological comparisons with Pb-Pb ages of meteorites by Lugmair (1974), Lugmair, Scheinin, and Marti (1975), and
Lugmair and Marti (1977) (table 1). It is also in close agreement with subsequent determinations by Kossert et al. (2009)
and Su et al. (2010) (table 1). However, the determination by Martins, Terranova, and Moreira Correa (1992) of 1.23 0.04
1011 years (123 4 Byr) is substantially higher than the Al-Bataina and Jnecke (1987) determination of 1.05 0.04
1011 years (105 4 Byr), as is the Kinoshita, Yokoyama, and Nakanashi (2003) determination of 1.17 0.02 10 11 years
(117 2 Byr) (table 1). These variant results can be easily seen in Fig. 2.Begemann et al. (2001) claimed that because there
is good agreement of ages obtained using the generally accepted 147Sm half-life value of 1.06 0.01 1011 years (106 1
Byr) with ages obtained by the U-Pb (Pb-Pb) systems (for example, Lugmair 1974; Lugmair and Marti 1977; Lugmair,
Scheinin, and Marti 1975), and because the 238U and 235U half-lives are more accurately known, the determination by
Martins, Terranova, and Moreira Correa (1992) of 1.23 0.04 10 11 years (123 4 Byr) should be viewed with caution, and
indeed, that discrepant result most likely is an artifact. Begemann et al. (2001) noted that Martins, Terranova, and Moreira
Correa (1992) reported in their experimental procedure the number of -decays was registered on a thin film of natural
samarium oxide Sm2O3 with an overall uniform thickness of (0.207 0.005) mg/cm2. However, in their subsequent
calculation of the decay constant, Martins, Terranova, and Moreira Correa (1992) apparently used the same thickness as for
pure samarium element instead of correcting for oxygen. If this was a mistake as Begemann et al. (2001) claimed, then the
published half-life value should have been multiplied by the weight ratio Sm2/Sm2O3 = 0.8624, provided the Sm2O3 used was
truly stoichiometric. This would have yielded a 147Sm half-life value of 1.06 10 11 years (106 Byr), which is in very good
agreement with the previous five direct counting measurements (table 1). Begemann et al. (2001) reported that according to
a private communication with one of the authors of the Martins, Terranova, and Moreira Correa (1992) paper, it is very likely
that this explanation is correct, although it is not possible to give a definitive answer. For this and other reasons, Begemann
et al. (2001) stated that they do not share the authors opinion that their result may be considered as the most accurate
measurement of the half-life performed up to now (that is, up to 1992). Nevertheless, Su et al. (2010) performed parallel
determinations using Sm metal and Sm 2O3 (as described above) and obtained similar results of 1.06 0.01 10 11 years
(106 1 Byr) and 1.07 0.01 10 11 years (107 1 Byr) respectively (table 1).What then can be said about the Kinoshita,
Yokoyama, and Nakanashi (2003) determination of 1.17 0.02 10 11years (117 2 Byr) (table 1)? Actually, it can be argued
that their careful experimental approach using multiple repeated measurements with multiple 147Sm sources using both an

alpha spectrometer and a liquid scintillation spectrometer to measure the number of emitted -particles per unit time, plus
their detailed analysis of the counting efficiencies of the instrumental techniques they used, makes their determinations of
the 147Sm half-life more robust than any other determinations. Kossert et al. (2009) subsequently questioned the Kinoshita,
Yokoyama, and Nakanashi (2003) determined 147Sm half-life value because it did not agree with many other determined
values. They also incorrectly asserted that Kinoshita, Yokoyama, and Nakanashi (2003) did not measure the isotopic ratio
when their paper clearly reports that they did. These are not valid reasons to dismiss their determined 147Sm half-life value.
Worse still, rather than engage in discussion, Su et al. (2010) simply ignored the Kinoshita, Yokoyama, and Nakanashi
(2003) determined147Sm half-life value. Yet the reality is that the Kinoshita, Yokoyama, and Nakanashi (2003)
determined 147Sm half-life value of 117 2 Byr is in very good agreement with the four direct counting values of 114 5 Byr
of Karras and Nurmia (1960), 117 5 Byr of Karras (1960), and 115 5 Byr of Gupta and MacFarlane (1970) who all used
ionization chambers for their determinations, and of 113 Byr of Graeffe and Nurmia (1961). However, it seems that the
accepted147Sm half-life value of 106 Byr is determined by majority vote because it is supported by the seven direct
counting determinations of Al-Bataina and Jnecke (1987), Donhoffer (1964), Gupta and MacFarlane (1970), Kossert et al.
(2009), Su et al. (2010), Valli et al. (1965), and Wright, Steinberg, and Glendenin (1961) (table 1).Kinoshita, Yokoyama, and
Nakanashi (2003) were well aware that the 147Sm half-life values reported from 1954 to 1970 showed some scatter between
104 Byr and 128 Byr, so they purposed to design their experimental procedures to thoroughly reevaluate the 147Sm half-life
by means of a developed alpha spectrometer with counting sources prepared in a manner different from those adopted in
those earlier determination experiments. Two kinds of commercially available Sm2O3 reagents of 99.9% purity and two kinds
of commercially available Sm standard solutions for atomic absorption spectroscopy were prepared as four standard Sm
solutions (with known concentrations in g-Sm/g-solution). These were labelled as Sm-A, Sm-B, Sm-C, and Sm-D.
Radioactive and non-radioactive impurities in these Sm solutions were measured by low-level gamma spectrometry, neutron
activation, ICP-MS spectrometry, and thermal ionization mass spectrometry (TIMS), and it was confirmed that the solutions
did not contain any detectable impurities, and that the isotopic composition was natural (that is, 15.0% 147Sm).Kinoshita,
Yokoyama, and Nakanashi (2003) used calibrated solutions of the -radioactive nuclides 210Po, 238U, and241Am as internal
standards in preparation of the counting sources for alpha spectrometry. Known aliquots of each of the four Sm standard
solutions were mixed well with known amounts of the calibrated 210Po, 238U, and 241Am standard solutions to produce ten
kinds of combinations. Each mixture was then evaporated on watch glasses and the amount of Sm on each watch glass
was adjusted to ~100 g, which means that there was approximately 15 g of 147Sm on each watch glass. Alpha
spectrometry, using a silicon surface-barrier detector with an active area of 450 mm 2, was carried out for one to two weeks
for each of the counting sources, and the -disintegration rate of the known amounts of 147Sm was determined by reference
to the -activities of the internal standards.
Measurements of the -disintegration rates of the known amounts of 147Sm were also carried out concurrently by Kinoshita,
Yokoyama, and Nakanashi (2003) using a liquid scintillation spectrometer. However, only the 241Am internal standard
solution was added to each of the four Sm standard solutions. Each mixture was prepared for the measurements and then
10 ml of either the Amersham ACS II scintillation cocktail or the p-terphenyl + POPOP + toluene scintillation cocktail was
added. A blank sample for background liquid scintillation measurement was also prepared in the same way. The liquid
scintillation counting was continued for five hours for each of the counting sources. The same cocktail mixture for
establishing the quench curve was used in determining the counting efficiency.Kinoshita, Yokoyama, and Nakanashi (2003)
then calculated a series of 147Sm half-life values based on the 147Sm -activity of each of the prepared counting sources
calibrated against the -activity of each internal standard measured by the alpha spectrometer, and these are plotted in Fig.
3, along with results from several earlier determinations. They also calculated a series of 147Sm half-life values measured
with the liquid scintillation spectrometer, and these are plotted in Fig. 4. The attached error for each plotted half-life value
included the nominal systematic errors in the calibration of the -emitting standard and in the measurement of the isotopic
abundance of 147Sm, in addition to the statistical error of 1 in -counting. It is quite obvious from Figs. 3 and 4 that
the 147Sm half-life values obtained by Kinoshita, Yokoyama, and Nakanashi (2003) using an alpha spectrometer and a liquid
scintillation spectrometer respectively agree very closely with each other within the error bars shown.

Fig. 3. The 147Sm half-life values and their errors determined by alpha spectrometer using a silicon surface-barrier detector
(after Kinoshita, Yokoyama, and Nakanashi [2003]). The values from earlier determinations are indicated by cross symbols,
and values obtained by Kinoshita, Yokoyama, and Nakanashi (2003) are indicated by closed symbols. The prepared
sources are referred to as the sample names with the original solution name, the reference standard, source number, and
the measurement method.

Fig. 4. The 147Sm half-life values and their errors determined by liquid scintillation spectrometer (after Kinoshita, Yokoyama,
and Nakanashi [2003]). The prepared sources are depicted as in Fig. 2. The Amersham ACS II scintillation cocktail is shown
as closed symbols, and the p-terphenyl + POPOP + toluene scintillation cocktail is denoted with open symbols.
Kinoshita, Yokoyama, and Nakanashi (2003) were also careful to discuss the sources of error and their potential adverse
effects on their 147Sm half-life determination. They admitted that, in alpha spectrometry for a deposited counting source, selfabsorption of -particles in the source is a serious problem. However, the energy loss and absorption of -particles in the
window-layer of the silicon surface-barrier detector (~50 nm) and in the vacuum chamber (<4 Pa) are negligible. As to the
extent of absorption in their alpha spectrometry measurements, they argued that since the thickness of the residue of Sm
and the -emitting standard on the watch glass was less than 15 g/cm2, energy loss and absorption of -particles
from 147Sm and the -emitting standard of the counting source was expected to be negligible. On the other hand, while the
counting efficiency of the liquid scintillation spectrometer is usually high (~100%), the spectrometer has the disadvantages
of high background and poor energy-resolution. Hence, they concluded that in spite of the agreement between the
mean 147Sm half-life value of 115 2 Byr measured with the liquid scintillation spectrometer and the mean 147Sm half-life
value of 117 2 Byr measured with the surface-barrier detector, within the error, the former value was regarded as
supporting data for the latter value because of these described disadvantages.Kinoshita, Yokoyama, and Nakanashi (2003)
determined that the arithmetic mean of the 147Sm half-life values obtained in their experimental work for the 19 counting
sources by alpha spectrometry (fig. 3) was 117 2 Byr, the stated associated error being one standard deviation. This
average 147Sm half-life value was thus about 10% longer than the then, and still, currently adopted value of 106 2 Byr.
However, Kinoshita, Yokoyama, and Nakanashi (2003) noted that in the earlier determinations the half-life of 147Sm was
obtained by measuring the 147Sm -activity with a 4 gas-flow counter (Beard and Wiedenbeck 1954), liquid scintillation
spectrometers (Beard and Kelly 1958; Donhoffer 1964; Wright, Steinberg, and Glendenin 1961), and ionization chambers
(Gupta and MacFarlane 1970; MacFarlane and Kohman 1961) from sources with the number of 147Sm atoms in them also
measured. Because the147Sm half-life is then calculated conventionally from the obtained values of the 147Sm -activity and
the number of147Sm atoms, the experimental errors in measuring those two values result in inaccurate determinations of
the 147Sm half-life. For example, impurities in the Sm reagent bring error into the number of 147Sm atoms in the counting
source, and uncertainty in the counting efficiency, self-absorption of the counting source and radioactive impurities in the Sm
reagent bring error into the value of the 147Sm -activity. Kinoshita, Yokoyama, and Nakanashi (2003) thus suggested that
the purity of the Sm reagent was not sufficient in the earlier experimental determinations. They also suspected that the
corrections for counting efficiency and self-absorption were not appropriate in the earlier experimental determinations. High
background and poor energy resolution of the liquid scintillation spectrometers used in the earlier experimental
determinations might also have resulted in inaccurate 147Sm half-life values. Since Kinoshita, Yokoyama, and Nakanashi
(2003) were confident that all these sources of error were excluded from their experimental work, in contrast to the earlier
experimental determinations, they concluded that their result of 117 2 Byr for the 147Sm half-life is reliable.As already noted,
Kossert et al. (2009) subsequently questioned the Kinoshita, Yokoyama, and Nakanashi (2003) result of 117 2 Byr for
the 147Sm half-life. They were confident that they had accurately determined the 147Sm half-life as being 107 0.9 Byr
because they had expended great effort to evaluate the liquid scintillation counting efficiency to be 100%, because the
number of 147Sm atoms in the Sm reagent had been measured by means of ICP-OES (inductively coupled plasma optical
emission spectrometry) using a reference standard from the National Institute of Standards and Technology, and because
their result agreed well with most of the other measurement results and the recommended value. However, as has been
already noted, a significant number of other measurement results also disagree with the recommended 147Sm half-life value
of 106 2 Byr and are closer to the Kinoshita, Yokoyama, and Nakanashi (2003) result of 117 2 Byr. Furthermore, while
Kossert et al. (2009) asserted that Kinoshita, Yokoyama, and Nakanashi (2003) measured quite low activities in their
experiments, they incorrectly accused Kinoshita, Yokoyama, and Nakanashi (2003) of not measuring the Sm isotopic ratio
when they clearly stated that they did. Such an error would cast doubt on the Kossert et al. (2009) assessment of the
experimental work of Kinoshita, Yokoyama, and Nakanashi (2003), especially as Kossert et al. (2009) admitted they did not
have enough information on the Kinoshita, Yokoyama, and Nakanashi (2003) measurement details, in particular on the
treatment of the background and tailing of the -peaks in their measured spectra. But Kinoshita, Yokoyama, and Nakanashi
(2003) did provide measurement details, as already described, and the -peaks in their spectra measured by alpha
spectrometer are very narrow, sharp, and distinct with a short tail and very low background, as illustrated in their report. So
the reality is that Kossert et al. (2009) questioned the Kinoshita, Yokoyama, and Nakanashi (2003) result primarily because it
didnt agree well with most of the other recent measurement results and the recommended value, even though Kinoshita,
Yokoyama, and Nakanashi (2003) did 19 alpha spectrometer determinations using combinations of four standard Sm
solutions and three internal standards (fig. 3) backed up by 24 liquid scintillation spectrometer determinations yielding
essentially the same results (fig. 4).On the other hand, Su et al. (2010) simply ignored the Kinoshita, Yokoyama, and
Nakanashi (2003) determined 147Sm half-life value of 117 2 Byr after an introductory passing reference to it, and simply
asserted that the absolute half-life they had determined, 106 1 Byr (Sm metal) and 107 1 Byr (Sm 2O3), is consistent
with the recommended value of 106 Byr. Yet Su et al. (2010) used a silicon surface-barrier detector coupled to an alpha
spectrometer to similarly obtain their 147Sm -activity peaks, just as Kinoshita, Yokoyama, and Nakanashi (2003) had done.
And Su et al. (2010) tabulated the uncertainty components in both their 147Sm -activity measurements and their
measurements of the number of 147Sm atoms in their counting sources, which together summed to 1.1% of the 147Sm half-life
value, close to the figure of 0.9% uncertainty in the Kossert et al. (2009) determination, but considerably less than the ~2%
obtained by Kinoshita, Yokoyama, and Nakanashi (2003). It would thus seem that the major reason for the difference
between the Su et al. (2010) and Kinoshita, Yokoyama, and Nakanashi (2003) results may be due to the different Sm
reagents used and the different procedures in preparing the counting sources.Su et al. (2010) used 147Sm-enriched metal
and Sm2O3 powder obtained from Oak Ridge National Laboratory and did not analyze them for purity, although they did
perform isotope dilution mass spectrometry (IDMS) to check the 147Sm abundance, whereas Kinoshita, Yokoyama, and
Nakanashi (2003) simply used commercially available Sm 2O3 reagents and analyzed them for both impurities and the 147Sm
abundance. Then Su et al. (2010) used vacuum evaporation (Sm metal) and sputtering (Sm 2O3) to deposit the Sm reagents
onto glass substrates to a thickness of about 250 g/cm2 before checking the uniformity of 147Sm distribution in these
prepared counting sources by exposure to plastic emulsions. On the other hand, Kinoshita, Yokoyama, and Nakanashi
(2003) simply evaporated the prepared mixtures of Sm standard solutions and internal standards onto glass substrates
adjusted to ~15 g of 147Sm to ensure calculated amounts of -particles would be measured from the 147Sm and internal
standards in these prepared counting sources. However, neither the choice of Sm reagents nor the procedures in preparing
the counting sources seem to be significantly different between these two 147Sm half-life determinations, even though
Kinoshita, Yokoyama, and Nakanashi (2003) stated they prepared their counting sources in a manner different from earlier
experiments. Apart from Su et al. (2010) not analyzing their chosen Sm reagents for purity, which would not be expected to
be significant given the source of those reagents, the only major difference appears to have been the thicknesses of the

counting sources and therefore the numbers of 147Sm atoms in them, resulting in only ~150 counts in the central channel of
the -peaks obtained by Kinoshita, Yokoyama, and Nakanashi (2003), the quite low activities in their experiments referred to
by Kossert et al. (2009), compared to the 500600 counts in the central channel of the -peaks obtained by Su et al. (2010).
Yet in spite of the lower -activity, it could be argued that the thinner counting source should have yielded the more accurate
determination of the 147Sm half-life value, because even though the thicker counting source would contain more 147Sm atoms
and therefore emit more -particles, the thicker the counting source the more self-absorption of -particles there could be
within the counting source thus significantly reducing the determined 147Sm half-life value.If indeed the thicknesses of the
counting sources and thus the self-absorption of -particles in them have resulted in the experimentally determined values
of the 147Sm half-life being up to 10% or more different, then this has profound significance on the determinations of the SmNd radioisotope ages of rocks, minerals, and meteorites. Without certainty as to whether the present 147Sm half-life (decay
rate) has been accurately determined experimentally by direct counting, and the assumption of constant radioisotope decay
rates at the currently determined values, the Sm-Nd radioisotope ages calculated by uniformitarians cannot be accurately
known, no matter how accurate are the measured 147Sm/144Nd and 143Nd/144Nd ratios in rocks, minerals, and meteorites.In
any case, the adopted 147Sm half-life value of 106 0.8 Byr had essentially been already settled on the basis of the
geological comparisons done on two meteorites by Lugmair, Scheinin, and Marti (1975) and Lugmair and Marti (1977) who
adjusted these meteorites Sm-Nd ages to agree with their Pb-Pb ages (Begemann et al. 2001; Dickin 2005, 7071).
Lugmair (1974) compared the Sm-Nd isochron age he obtained for the Juvinas eucrite meteorite with the similar Rb-Sr
isochron age obtained on the same meteorite by Allgre, Birck, and Fourcade (1973). However, it was Lugmair, Scheinin,
and Marti (1975) whom Dickin (2005, 7071) credited for the Sm-Nd isochron age of 4.56 0.08 Byr for Juvinas that is in
agreement with the Pb-Pb isochron age for the solar system and thus confirming the adopted value of 106 Byr for the 147Sm
half-life (Dickin 2005, 115118; Patterson 1956). This though begs the question how do they know that these so-called
isochrons are not mixing lines which have no time significance? It should also be noted that Dickin (2005, 71, fig. 4.1
caption) wrote that the Nd isotope ratios are affected by the choice of normalizing factor for mass fractionation, a reminder
that the agreement of this Sm-Nd isochron age with the Pb-Pb isochron age was only achieved by making choices of
suitable factors. Subsequently, Lugmair and Marti (1977) obtained a Sm-Nd isochron age of 4.55 0.04 Byr for the Angra
dos Reis (ADOR) angrite meteorite based on analyses of phosphate and pyroxene mineral separates, which they made
sure agreed with the Pb-Pb model age of 4.555 0.005 Byr for ADOR obtained by Tatsumoto, Knight, and Allgre (1973).
Lugmair and Galer (1992) later refined that Pb-Pb model age for ADOR to 4.55780 0.00042 Byr based on analyzing a
pyroxene mineral separate, which they then used to suggest what the 147Sm half-life should be in order to make the Sm-Nd
isochron age for this meteorite agree exactly with this Pb-Pb model age.Lugmair and Marti (1977) and Lugmair and Galer
(1992) justified their geological comparisons to adjust the 147Sm half-life value so that the Sm-Nd isochron ages of the
Juvinas and Angra dos Reis meteorites agreed with their Pb-Pb isochron and model ages on the basis that these
meteorites Pb-Pb ages were more precise because the decay constants of the parent 238U and 235U are known more
precisely. And Lugmair and Marti (1978) added to the adoption of that 147Sm half-life value of 106 Byr by choosing the most
precise results from only four of the earlier direct counting experiments (see table 1) that would give the desired weighted
average value for the 147Sm half-life of 106 0.8 Byr that has been adopted by all geo- and cosmochronologists since that
time (Begemann et al. 2001). This was also the basis for Begemann et al. (2001) seeking to correct the Martins,
Terranova, and Moreira Correa (1992) determined higher 147Sm half-life value to bring it into agreement with this
adopted 147Sm half-life value calibrated against the U-Pb radioisotope system, and for Kossert et al. (2009) and Su et al.
(2010) dismissing and ignoring respectively the more thoroughly determined but higher 147Sm half-life value of Kinoshita,
Yokoyama, and Nakanashi (2003).However, the U-Pb gold standard dating method has come under much scrutiny in the
two last decades. Ludwig (2000) has demonstrated that, although almost universally ignored, the effect of the errors in the U
decay constants on U-Pb concordia-Pb-Pb intercept ages are significant, being a 4.5 Myr error for a 500 Myr age, which
amounts to almost a 1% error. The U-Pb method also depends on the crucial 238U/235U ratio, but discrepancies and variations
have been found recently between the 238U/235U ratio in U-bearing terrestrial minerals and rocks and the 238U/235U ratio in
meteorites (Brennecka and Wadhwa 2012; Hiess et al 2012). Much earlier, Apt et al. (1978) had reported that the 235U/238U
ratio in uranium ores in Canada, Brazil, Zaire, and Australia varied from 0.7107 to 0.7144 when the recognized value is 0.72.
Such variations in uranium ores have been further documented by Bopp et al. (2009). These discrepancies and variations
remain unexplained, especially in the context of the 238U and 235U decay constants and half-lives. Furthermore, the fact that
there are these variations in the crucial 238U/235U ratio in terrestrial minerals and rocks on which the U-Pb dating gold
standard depends, which has been used to recalibrate Sm-Nd isochron ages to determine the 147Sm half-life and decay
constant, only underscores that these radioisotope methods cannot provide the absolute invariable dates they are so
confidently proclaimed to provide.In any case, there is the additional assumption in all the radioisotope dating methods of
having to know the original concentrations of the daughter and index isotopes which is very significant in the U-Pb method
because the original concentrations are assumed not to be zero, in contrast to the K-Ar method. Yet there must be great
uncertainty as to what those initial values were in the unobserved past, despite the isochron and concordia techniques
attempting to negate the necessity for knowing those initial values, and despite the assumption ever since Patterson (1956)
and Tatsumoto, Knight, and Allgre (1973) that the Pb isotopic composition of the troilite (FeS) in the Canyon Diablo iron
meteorite represents the initial primordial Pb of the earth and the solar system (Dickin 2005; Faure and Mensing 2005).
Thus the U-Pb method should not be used as a standard to determine other parent radioisotope half-lifes and decay
constants.Indeed, it would remain prudent to be very careful with these geological comparison methods for two other
reasons. First, there are significant flaws in the basic assumptions on which all the radioisotope dating methods depend, not
least being the assumption that the decay rates of the parent radioisotopes have always been constant in the past at todays
measured decay rates. Second, the U-Pb method relies primarily on -decay, as does the Sm-Nd method. Yet both Austin
(2005) and Snelling (2005) have reported that the parent U and Sm -decaying radioisotopes seem to yield systematically
different U-Pb and Sm-Nd ages for some earth rocks using the same samples with essentially the same methodology.
Additionally, they suggested the pattern of differences was potentially related to the parent radioisotopes atomic weights
and half-lives, which could be indicative of parent radioisotopes decay rates having not been constant in the past but
instead were substantially faster. Furthermore, these different radioisotope ages yielded by the same earth rocks are often
widely divergent, even up 100200% different, which is such a huge divergence that it renders these dating methods highly
suspect, even if the differences in the determinations of the half-lives of the parent radioisotopes seem miniscule and
therefore trivial by comparison. However, it was considered prudent to still document here these seemingly miniscule
differences in half-life values, because they may be indicative of other underlying factors at work (as already discussed
briefly), and they can still lead to very significant discrepancies in the derived radioisotope ages that might otherwise appear
to be acceptably accurate to uniformitarians.Nevertheless, the age comparisons on meteorites used the U-Pb method back
in the 1970s to settle, apparently beyond any subsequent dispute, the determination of the 147Sm decay half-life at 106 0.8
Byr, which is within the range determined by many of the physical direct counting experiments by several techniques (see

table 1 and fig. 2). Yet the robustness of the 10% higher 117 2 Byr 147Sm half-life value determined by Kinoshita,
Yokoyama, and Nakanashi (2003) might even indicate that the other higher 147Sm half-life values determined in some earlier
experiments (see table 1 and fig. 2) should not be simply dismissed as due to poorer experimental methodology or
equipment. In any case, Sm-Nd age calculations are now ultimately calibrated against the U-Pb method, and thus the 147Sm
half-life value of 106 0.8 Byr has been adopted. However, this U-Pb gold standard depends on whether the U decay
constants are accurately and precisely known, and on the crucial 238U/235U ratio. Yet discrepancies and variations have been
found between the 238U/235U ratio in U-bearing terrestrial minerals and rocks and the 238U/235U ratio in meteorites which
remain unexplained. This only serves to highlight that if the Sm-Nd dating method has been calibrated against the U-Pb
gold standard with its own uncertainties, then the claimed accurately-determined 147Sm decay rate cannot be absolute,
especially given the evidence in some earth rocks of past higher radioisotope decay rates and the evidence that some direct
counting experiments yielded 10% or more higher 147Sm half-life values. Yet even though it is to be expected these half-life
measurements vary by 10% or so because of the difficulties in measuring such a long half-life, the resultant calculated
radioisotope ages end up being an order of magnitude or more different from one another, which is far too inaccurate in
providing the absolute ages required by uniformitarians. Thus without an accurately known 147Sm decay half-life, accurate
Sm-Nd radioisotope ages cannot be accurately determined. Therefore, Sm-Nd dating cannot be used to reject the youngearth creationist timescale, especially as current radioisotope dating methodologies are at best hypotheses based on
extrapolating current measurements and observations back into an assumed deep time history for the cosmos.
Conclusions
There have been numerous attempts to determine the 147Sm decay half-life in the last 80 years by two primary techniques
used in direct physical counting experimentsionization chambers and liquid scintillation counters, and by radioisotope age
comparisons using two meteorites. The determinations since 1960 have converged with close agreement on the 147Sm halflife value of 106 0.8 Byr, which has since the 1970s been adopted for standard use by the uniformitarian geological
community. This adopted 147Sm half-life value is the weighted average of four determinations by direct counting experiments
in the period 19611970, confirmed by geological comparisons in the 1970s in which the 147Sm half-life value was adjusted
in order to recalibrate (or force, essentially by circular reasoning) the Sm-Nd isochron ages of two meteorites to be the same
as their Pb-Pb isochron and model ages.However, even though this 147Sm half-life value of 106 0.8 Byr has been
universally adopted by the geochronology and cosmochronology community since Lugmair and Marti (1978) proposed it
(Begemann et al. 2001), the more recent direct counting experiments by Kinoshita, Yokoyama, and Nakanashi (2003)
determined the 147Sm half-life value was 10% or more longer at 117 2 Byr. They achieved this by using four standard Sm
solutions with internal -radioactive standards in 19 alpha spectrometer and 24 ionization chamber determinations, making it
the most thorough and comprehensive effort to the determine the 147Sm half-life. It would appear that the thinner counting
sources they used, while resulting in low -activities being measured, greatly reduced the counting uncertainty due to selfabsorption of the emitted -particles. So in spite of being rejected or ignored, this 117 2 Byr value for the 147Sm half-life,
which agrees with some earlier determinations, may well be highly significant and more reliable than the adopted value.Yet
even though there is close agreement between many determined values, the 147Sm half-life value obtained by recalibrating
the Sm-Nd isochron ages for two meteorites with their Pb-Pb isochron and model ages has been given preference over the
values obtained by direct counting experiments which directly measure 147Sm decay, and so are independent of all the
assumptions involved with the radioisotope dating methods. Indeed, model dependent results should not take precedence
over the direct experimental evidence.Since the age comparisons on two meteorites used the U-Pb method to determine
the 147Sm decay half-life, all Sm-Nd age calculations are thus ultimately calibrated against the U-Pb method. However, this
U-Pb gold standard depends on having precisely determined 238U and 235U decay constants, as well as on the
crucial 238U/235U ratio being known and constant. Yet there are still uncertainties in the measured U decay constants, and
discrepancies and variations have been found between the 238U/235U ratio in U-bearing terrestrial minerals and rocks, and
the 238U/235U ratio in meteorites. These discrepancies and variations remain unexplained. This only serves to highlight that if
the Sm-Nd dating method has been calibrated against the U-Pb gold standard with its own uncertainties, then it cannot be
absolute.Furthermore, there is evidence that nuclear decay rates have not been constant in the past. Thus without an
accurately known 147Sm decay half-life, accurate Sm-Nd radioisotope ages cannot be accurately determined. Therefore, SmNd dating cannot be used to reject the young-earth creationist timescale, especially as current radioisotope dating
methodologies are at best hypotheses based on extrapolating current measurements and observations back into an
assumed deep time history for the cosmos.
Acknowledgments
The invaluable help of my research assistant Lee Anderson Jr. in compiling and plotting the data in the color-coded age
diagrams is acknowledged. The three reviewers are also acknowledged for their helpful comments and input, though the
final content of this paper is solely my responsibility. Our production assistant Laurel Hemmings is also especially thanked
for her painstaking work in preparing this paper for publication.
DATING RESULTS
Significance of Highly Discordant Radioisotope Dates for Precambrian Amphibolites in Grand Canyon, USA
by Dr. Andrew A. Snelling on April 14, 2010
Abstract
The Brahma amphibolites of the Precambrian
crystalline basement of Grand Canyon were
originally erupted as basalt lavas and
subsequently suffered high-grade regional
metamorphism. Composed predominantly of
hornblende
with
minor
subordinate
plagioclase, the collected samples showed no
signs of post-metamorphic alteration. K-Ar
radioisotope analyses yielded a wide range of
model ages, even for adjacent samples from the same outcrop of the same original lava flow. No statistically viable K-Ar
isochron age could be obtained because of so much scatter in the data, which is most likely due to 40Ar* mobility within
these rocks. By contrast, the Rb-Sr, Sm-Nd, and Pb-Pb radioisotope systems yielded good, statistically consistent, isochron
ages of 1240 84 Ma, 1655 40 Ma, and 1883 53 Ma, respectively. These are obviously discordant with one another and
with published ages, but there are no clear reasons to reject any of them as unreliable or invalid. One explanation for the
discordance is that the decay rates of the parent radioisotopes were different relative to their presently measured rates at

some time during the time interval since these rocks formed. We observe that the a-decaying U and Sm yield older ages
than the -decaying Rb, and the heavier atomic weight U yields a Pb-Pb age older than the Sm-Nd age. This pattern in the
discordances thus may provide clues into the physics responsible for time variations in the decay process. Obviously, if
decay rates have not been constant, radioisotope decay methods do not yield valid absolute ages for rocks.
Shop Now
This paper was originally published in the Proceedings of the Sixth International Conference on Creationism, pp. 407424
(2008) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh.
Keywords: Grand Canyon, Precambrian crystalline basement, Brahma amphibolites, hornblende, plagioclase, whole-rock
chemical analyses, whole-rock K-Ar model ages, whole-rock Rb-Sr, Sm-Nd, and Pb-Pb isochron ages, discordancy,
accelerated radioisotope decay, half-lives, mode of decay, atomic weights
Introduction
The assumption of constant radioisotope decay rates has undergirded the interpretation of all radioisotope data and the
establishment of the absolute dates in the uniformitarian geologic timescale. Anomalous radioisotope dates that do not fit
the chosen timescale are usually explained by open-system behavior and/or inheritance, and then discarded. Because most
anomalous radioisotope dates are not published, it is difficult to know just what proportion of dating analyses in
geochronology laboratories are discarded. Furthermore, rock samples are often only dated using the one radioisotope
method. Thus, it is difficult to quantify just how significant are the few multiple radioisotope concordances published in the
literature, and how reliable and consistent is the apparent overall trend of progressively decreasing dates from lower strata
in the geologic record through to upper strata. However, the impression gained from a detailed examination of the primary
radioisotope dating systems1 is that, if the absolute dates of the uniformitarian timescale were ignored, and both accepted
and anomalous radioisotope dates were considered, where more than one radioisotope system has been utilized to date
specific rock strata, radioisotope discordances would be in the majority. That such discordances are often the case has
already been discussed,2 and has been thoroughly tested and documented on some specific strata.3, 4, 5, 6Furthermore, it is
highly significant that there are no obvious geologic or geochemical explanations evident for these discordances.7 Thus, if it
werent for the assumption that the approved radioisotope dates are acceptable because they correlate with the
conventional uniformitarian timescale, then all the discordant isochron ages could actually be anomalous. Using the same
reasoning, therefore, there is no guarantee that even where radioisotope concordances do occur the resultant dates are
somehow objectively correct. In any case, the ages derived from the radioisotope systems should only be regarded as
maximum ages. The prolific evidence of open-system behavior and mixing and/or inheritance suggests that the true ages of
the strata may be considerably, or even drastically, younger. This is intolerable for uniformitarians, as their evolutionary
timescale is crucial to their paradigm. These endemic problems with the radioisotope dating methods demonstrate that the
conventional interpretation of radioisotope dating is not secure, and also provide evidence that indicates a much younger
earth.A very relevant example is the stark contrast between the U-Pb radioisotope age of 1500 Ma for the zircon grains in
the Jemez granodiorite of New Mexico and the He (derived from U decay) diffusion age of the same zircon grains of only
about 6,000 years.8, 9, 10, 11 This huge discrepancy can be explained if the rate of 238U decay was grossly accelerated at
some time(s) in the past. A proposed test of this explanation is to document whether there is a systematic pattern in the
discordances between the different radioisotope systems. If there is a systematic pattern, it may reflect differing amounts of
such accelerated nuclear decay in the different radioisotope systems over the same real time interval, due to their different
modes of decay and parent half-lives.12 The amphibolites in the Precambrian basement of the Grand Canyon were chosen
for this study for three reasons. First, the Grand Canyon is a well known and well studied area that contains a good, clear
strata cross-section representative of much of earth history. Second, as metamorphosed basalts, amphibolites consist of a
very simple two-component system, essentially just the minerals plagioclase and hornblende, which simplifies the
geochemistry of radioisotope systematics. And third, being Precambrian, these rocks should have accumulated large
enough amounts of the radioisotope decay products to produce isochrons with good statistics.
Geology of the Precambrian Basement Metamorphics in Grand Canyon
The east-west trending Grand Canyon presents spectacular exposures of the Lower Proterozoic (Paleoproterozoic) rocks of
the crystalline basement beneath the Colorado Plateau.13 In the Upper Granite Gorge, these rocks are continuously
exposed from river mile 78 to 120 (downstream from Lees Ferry), while there are discontinuous exposures in the Middle
Granite Gorge from mile 127 to mile 137 (fig. 1).14, 15 Powell was the first to identify the Precambrian granite and Grand
Canyon schist.16 Walcott identified Significance of Highly Discordant Radioisotope Dates for Precambrian Amphibolites in
Grand Canyon, USA the Vishnu terrane as a complex of schist and gneiss.17Subsequently, Campbell and Maxson
identified different mappable units called the Vishnu series18 and Brahma series.19 However, Campbell and
Maxson20 underestimated the structural complexities and probably overestimated the stratigraphic thickness when they
proposed that the combined stratigraphic sequence of these metasedimentary and metavolcanic rocks was 816 km thick.
This stratigraphic approach was called into question by Ragan and Sheridan,23 and subsequently Brown, Babcock, Clark,
and Livingston24 also emphasized the complex deformational features, so they lumped all of the metasedimentary and
metavolcanic rocks together under the name Vishnu Complex, the approach continued by Babcock,25 who used the term
Vishnu Metamorphic Complex.

Fig. 1. Simplified geologic map of Paleoproterozoic (lower Proterozoic) rocks in the Upper and Middle Granite Gorges,
Grand Canyon, northern Arizona (after Ilg et al.21, Karlstrom et al.22). Form lines outside the Paleoproterozoic exposures
show their interpretation of the trace of the regional foliation on the map surface. The transect is divided into metamorphic
domains that are generally separated by shear zones. Sample locations are marked.More recent detailed field mapping,
based on the approach that recognizes the need to pursue simultaneously both tectonic and stratigraphic subdivisions of
these Lower Proterozoic rocks, has resulted in a new geologic map (fig. 1). Thus, Ilg et al.26 and Karlstrom et al.27 have
proposed the new name of Granite Gorge Metamorphic Suite for the entire sequence of metamorphosed volcanic and
sedimentary rocks in the Grand Canyon. Furthermore, the new names assigned to the mappable rock units in the Upper
and Middle Granite Gorges (fig. 1), as well as the Lower Granite Gorge, are Brahma Schist for the mafic metavolcanic rocks
(after the Brahma series of Campbell and Maxson28), the Rama Schist for the felsic metavolcanic rocks, and Vishnu
Schist for the metamorphosed sedimentary rocks. This designation of only the metamorphosed sedimentary rocks as the
Vishnu Schist is probably that which was originally intended by Walcott,29 recommended by Noble and Hunter30 (their
Vishnu schist), and proposed by Campbell and Maxson31 (their Vishnu series). These metasedimentary and metavolcanic
rocks of the Granite Gorge Metamorphic Suite make up about half of the exposed rocks in the Granite Gorges of Grand
Canyon, the rest being intrusive rocks (granite, granodiorite, tonalites, and gabbros). Descriptive metamorphic rock names
are used for the rocks seen in outcrop and in thin section, and the original sedimentary or volcanic protoliths are inferred
from rock compositions and a limited number of primary structures that have survived the deposition and metamorphism.
Primary structures such as relict pillows and graded bedding show that the original sedimentary rocks were locally deposited
on a volcanic sequence, and that the mafic and felsic metavolcanic rocks are commonly interlayered. However, because
similar volcanogenic sequences could have been deposited at different times or in separate basins, and such differences
would be difficult to unravel due to the subsequent tectonism, this terminology can be considered mainly as lithologic, rather
than necessarily stratigraphic.The Rama Schist consists of quartzofeldspathic schist and gneiss with locally preserved
phenocrysts of quartz and feldspar, and possible relict lapilli, that suggest a felsic to intermediate volcanic origin.32, 33 It is
dominated by massive fine-grained quartzofeldspathic rocks, but also contains metarhyolites and interlayered micaceous
quartzofeldspathic schists and gneisses. The Rama Schist is commonly complexly injected with pegmatite and contains
leucocratic layers that may in part reflect preferential partial melting of these rocks due to the peak metamorphic conditions
of about 720 C and 6 kbar.34, 35 It is also locally interlayered with the mafic Brahma Schist.The Brahma Schist consists of
amphibolite, hornblende-biotite-plagioclase schist, biotite plagioclase schist, orthoamphibole-bearing schist and gneiss, and
metamorphosed sulfide deposits.36, 37 The petrology and geochemistry of Brahma Schist amphibolites were studied by
Clark,38, 39 who divided the amphibolites and mafic schists into five groups based on field occurrence and mineral
assemblage: (1) anthophyllite-bearing and cordierite-anthophyllite-bearing rocks (orthoamphibole schist), (2) early
amphibolites, (3) the Granite Park mafic body (Lower Granite Gorge area), (4) hornblende-bearing dikes, and (5) tremolitebearing dikes. Ilg et al.40 agreed with Clarks interpretation that the orthoamphibole-bearing (group 1) rocks are
metamorphosed, hydrothermally altered, mafic marine volcanic rocks, and that the early amphibolites (group 2) are
metamorphosed basalts and basaltic tuffs. Clarks groups 1 and 2 compose the supracrustal Brahma Schist, following
Campbell and Maxsons original usage of the term.41Massive amphibolites (part of Clarks group 2) make up 3040% of the
Brahma Schist. This unit does not typically preserve primary igneous features, but relict pillow structures are present at a
number of localities. Massive amphibolites occur in units several meters to tens of meters thick, and are composed of
plagioclase
and
hornblende,
plus
subordinate
quartz,
biotite,
clinopyroxene,
and
epidote
(plus

accessories).42, 43 Furthermore, these massive amphibolites have a tholeiitic character and trace element compositions
consistent with an island-arc environment. The biotite-plagioclase and hornblende-biotite-plagioclase schists (the remainder
of Clarks group 2) make up approximately 50% of the Brahma Schist in the Upper Granite Gorge. Although strong tectonic
layering has mostly obscured primary igneous textures, in several locations original textures are preserved, such as
subangular quartz + plagioclase + biotite fragments entrained in an amphibolitic matrix, which suggests that some of these
rocks may have been volcaniclastic breccias. Interlayered with the biotite schists are discontinuous meter-scale lenses of
garnet + diopside + epidote + calcite rocks, the protoliths of these lenses possibly being relatively thin layers of calcareous
shale or algal mats interbedded with submarine sediments.44 The Brahma Schist also contains exposures of
orthoamphibole-bearing rocks (Clarks group 1) (fig. 1). They are interpreted to be hydrothermally altered, mafic marine
volcanic rocks.45The presence of relict pillow basalt, orthoamphibolite rocks, and associated sulfide mineralization indicates
that the Brahma Schist was a product of dominantly mafic submarine volcanism. The Rama and Brahma metavolcanic
schists can be complexly interlayered so that contact relationships support variable relative ages between mafic and
intermediate metavolcanic rocks. However, in the Upper Granite Gorge, the Rama Schist is underneath and older than the
Brahma Schist.The Vishnu Schist consists of pelitic schist and quartz-biotite-muscovite schists that are interpreted as
metamorphosed sandstones and mudstones, with numerous calc-silicate lenses and pods that are interpreted to be
concretions.46, 47 Several-kilometer thick sections of meta-lithic-arenite and metagreywacke sequences exhibit rhythmic
banded (centimeter-to meter-scale) coarser and finer layers, with locally well-preserved bedding and graded
bedding48, 49 suggesting deposition as submarine turbidites. The original grain sizes in the Vishnu Schist metasedimentary
rocks probably ranged from medium-grained sand to silt and clay, while conglomerates are conspicuously absent, all of
which suggests a lack of high-energy proximal facies. The preserved relict graded bedding, association with metavolcanic
rocks containing pillow structures, lack of coarse sediments, and geochemical data50indicate that these Vishnu
metasedimentary units were deposited in submarine conditions on the flanks of eroding oceanic islands (an oceanic islandarc environment). The preserved graded bedding indicates that the Vishnu Schist was deposited stratigraphically above the
Brahma Schist, and the accessible exposures indicate that the contact between them is generally concordant, although
there is some interlayering of the contact in some places.These Brahma amphibolites have not previously been dated using
radioisotope methods. Metamorphic rocks in general are not easy to date with radioisotopes, because often it is not readily
apparent whether the results represent the date of the metamorphic event or the date of the cooling and crystallization of the
original rock. However, by U-Pb dating what is believed to be original zircon grains in metamorphosed felsic volcanic units
within the Brahma and Rama Schists, it has been proposed that the original basalt lavas were erupted between 1741 and
1750 Ma.51, 52 The subsequent metamorphism of these basalt lavas to form these Brahma amphibolites is believed to have
occurred between 1690 and 1710 Ma, based on U-Pb dating of monazite, xenotime, and titanite (sphene) in the overlying
Vishnu Schist and underlying Rama Schist, assuming that these minerals formed, or were radioisotopically reset, during the
metamorphism.53, 54
Fig. 2. Outcropping amphibolites below the Tapeats
Sandstone and the Great Unconformity beside the
Colorado River, just upstream of Blacktail Canyon (see fig.
1). Seven samples were collected along the river
here.Twenty-seven Brahma Schist amphibolite samples
were collected in the Upper and Middle Granite Gorges
(with a Scientific Research and Collecting Permit issued by
the Grand Canyon National Park): (1) three samples from
the Cottonwood Canyon area, (2) nine samples from the
Clear Creek area, including seven samples from a single 50
m long and 2 m wide amphibolite body just upstream from
the mouth of Clear Creek, (3) one sample from the
Cremation Creek area, (4) one sample from near the mouth
of Pipe Creek, (5) seven samples from outcrops just
upstream of Blacktail Canyon (fig. 2), and (6) six samples
from outcrops along the Colorado River between miles
126.5 and 129. All these locations are marked on Fig. 1. The small tabular
body of amphibolite near the mouth of Clear Creek was intensively sampled
because it appeared to show mineralogical variation through its width,
perhaps suggesting that it may have been a thin sill rather than a lava flow
(fig. 3). Otherwise, all the other samples were of massive amphibolite. In the
area just upstream of Blacktail Canyon, there was clear field evidence that
the amphibolites represented a series of basaltic lava flows, there being
well-defined competent layers 310 meters thick in succession along the
outcrop separated by structural breaks accompanied by leaching of the rock
(possibly paleoweathering), or in one instance by what appeared to be a thin
inter-flow sandstone layer (fig. 4). This sequence of metamorphosed basalt
flows was thus systematically sampled.
Chemical and Radioisotope Analyses
Fig. 3. A single amphibolite body (dark rock through the center of the
photograph in the erosion gully) that likely represents an original basalt lava
flow, just upstream of Clear Creek (see fig. 1). Seven samples were
collected from this outcrop and yielded widely divergent K-Ar model
ages.Split by hammering or cutting with a rock saw, a portion of each sample
was thin sectioned for subsequent petrographic analysis. Approximately 100
gram splits of each sample were then dispatched to the Amdel laboratory in
Adelaide, South Australia, where each sample was crushed and pulverized.
Whole-rock analyses were undertaken by total fusion and digestion of each
powdered sample followed by ICPOES (inductively coupled plasma-optical
emission spectrometry) for major and minor elements, and ICP-MS
(inductively coupled plasma-mass spectrometry) for trace and rare earth
elements. Separate analyses for Fe as FeO were also undertaken by wet
chemistry methods that were able to indicate any sample loss on ignition, primarily as H 2O or carbonate (given off as CO2).A

second representative set of 100 gram pieces of each sample was sent to the K-Ar dating laboratory at Activation
Laboratories in Ancaster, Ontario, Canada, for whole-rock K-Ar dating under the direction of the laboratory manager, Dr.
Yakov Kapusta. After crushing of the whole-rock samples and pulverizing them, the concentrations of K (weight %) were
measured by the ICP technique. The 40K concentrations (ppm) were then calculated from the terrestrial isotopic abundance
using these measured concentrations of K. The concentrations in ppm of 40Ar*, the supposed radiogenic 40Ar, were derived
from isotope dilution measurements with a noble gas mass spectrometer.55, 56
Fig. 4. Outcrop of amphibolite just
upstream of Blacktail Canyon (see
fig. 1). The vertical layering in the
amphibolite appears to parallel the
original layering of successive
horizontal basalt flows that were
subsequently tilted. The rock
hammer and brightly colored roll of
tape mark the top and bottom of
what was an original basalt
flow.Finally, a third representative
set of 100 gram pieces of each
sample was sent to the PRISE
laboratory in the Research School
of Earth Sciences at the Australian
National University in Canberra,
Australia, where under the direction
of Dr. Richard Armstrong, wholerock Rb-Sr, Sm-Nd, and Pb-Pb
isotopic analyses were undertaken.
After the sample pieces were
crushed
and
pulverized,
the
powders
were
dissolved
in
concentrated hydrofluoric acid, followed by standard chemical separation procedures for each of these radioisotope
systems. Once separated, the elements in each radioisotope system were loaded by standard procedures onto metal
filaments to be used in the solid source thermal ionization mass spectrometer (TIMS), the stateof- the-art technology in use
in this laboratory. Sr isotopes were measured using the mass fractionation correction 86Sr/88Sr = 0.1194, and the 87Sr/86Sr
ratios were reported normalized to the NBS standard SRM 987 value of 0.710207. Nd isotopes were corrected for mass
fractionation using146Nd/144Nd = 0.7219 and were normalized to the present-day143Nd/144Nd value of 0.51268 for standard
BCR-1 (a Columbia River basalt, Washington, sample). Pb isotope ratios were normalized to NBS standard SRM 981 for
mass fractionation.
Results
The whole-rock major and selected trace element data for all 27 amphibolite samples are listed in Table 1. These analytical
results are consistent with these amphibolites being tholeiitic basalts that were metamorphosed without any significant
chemical alteration (losses or additions). This can especially be seen in the high MgO, Cr, and Ni contents of the amphibolite
in the small body near Clear Creek, which is also high in rare earth elements. Otherwise the range of 43.8%51.2% SiO 2 is
typical of the variations found in basalt lavas. Especially noteworthy are the low K 2O contents (0.11%1.96%) of these
amphibolites, which are consistent with the absence of any pervasive alteration of the plagioclase in them to sericite,
confirmed by petrographic examination (fig. 5). This is particularly relevant to the suitability of these amphibolites for K-Ar
radioisotope dating. Any sericitic alteration of the plagioclase would represent post-formation hydrothermal alteration with
introduced K2O, so the K2O in these amphibolites is, thus, of the primary in situ origin required for valid K-Ar radioisotope
dating. Equally significant are the wide variations in the contents of Rb, Sr, Sm, Nd, U, and Pb in these amphibolites, which
make them extremely suitable for radioisotope dating using the isochron method, in spite of the simplicity of their dominant
mineralogy (plagioclase + hornblende biotite).The results of the K-Ar radioisotope analyses on these Brahma amphibolite
samples are summarized in Table 2. The K-Ar model ages were calculated for each sample analyzed using the standard
model-age equation, which assumes that 10.5% of the 40K atoms in each sample decay to 40Ar atoms. Furthermore,
because 40Ar is a common atmospheric gas which can leak into rocks and minerals making them appear older than their
actual ages, in conventional K-Ar model age determinations it is assumed that a certain proportion of the 40Ar in each rock
sample is contamination, and therefore, a certain proportion of the total 40Ar determined in the laboratory on each sample, in
accordance with the40Ar to 36Ar ratio of the present atmosphere, is subtracted so that only what is thus assumed to be the
radiogenic 40Ar in each sample is used in the model-age calculations.57 Furthermore, it is convention to assume that no
radiogenic 40Ar (written as 40Ar*) was present in the rock when it initially formed, so that all the 40Ar* now measured in the
rock has been derived from in situ radioactive decay of 40K. The reported error listed with each model age in Table 2
represents the estimated 1s (sigma) uncertainty due to the analytical equipment and procedure.
Table 1. Major oxide and selected trace element analyses.
Oxide/
Element

Clear Creek small body

Blacktail Canyon area sequence

BA-2 BA-7

BA-8

UM-1 UM-2

UM-3

UM-4

BA-5

BA-9

BA-10 BA-11 BA-12 BA-13 BA-14

SiO2 (%)

44.4

46.1

45.4

47.3

47.1

48.7

46.8

47.9

50.8

44.6

43.8

45.9

49.1

48.6

TiO2 (%)

1.05

1.00

1.09

0.68

0.78

0.62

0.47

1.03

1.13

2.30

1.21

0.70

0.80

1.06

Al2O3 (%)

14.4

12.9

13.9

9.53

11.0

8.29

6.00

14.5

13.4

13.2

14.4

14.8

14.6

14.9

Fe2O3 (%)

11.2

11.1

12.5

10.0

10.2

9.98

8.55

14.6

13.1

20.2

16.3

10.8

12.3

12.5

MgO (%)

10.6

11.5

10.6

15.6

13.7

15.5

19.6

7.09

7.14

4.86

9.21

5.05

8.20

5.40

MnO (%)

0.19

0.19

0.20

0.18

0.17

0.19

0.16

0.22

0.16

0.26

0.22

0.18

0.19

0.21

CaO (%)

12.1

12.4

10.7

12.0

12.2

11.9

12.2

11.2

9.22

10.7

9.62

14.2

11.3

11.1

Na2O (%)

0.82

0.80

0.82

0.66

0.82

0.63

0.20

2.10

3.50

2.14

1.79

2.33

2.38

2.71

K2O (%)

1.95

1.35

1.96

1.00

0.91

1.02

0.11

0.47

0.76

0.67

1.12

1.58

0.25

1.26

P2O5 (%)

1.02

0.70

0.81

0.54

0.57

0.54

0.35

0.12

0.12

0.53

0.15

0.07

0.07

0.11

LOI (%)

2.21

2.67

2.52

1.97

2.57

2.96

6.04

1.01

1.28

1.18

2.43

5.15

1.32

2.74

Total

99.94 100.71 100.50 99.46 100.02 100.33 100.48 100.24 100.61 100.64 100.25 100.76 100.51 100.59

Cr (ppm)

320

400

280

900

900

750

850

170

96

78

110

185

210

200

V (ppm)

260

240

250

200

210

180

125

290

310

320

320

200

230

310

Ni (ppm)

99

210

160

440

300

450

700

89

64

62

74

115

120

82

Co (ppm)

57

54

56

60

50

50

60

59

54

54

60

49

49

47.5

Cu (ppm)

49

29.5

440

13.5

36

115

66

73

84

32

16

96

84

78

Zn (ppm)

67

84

86

78

80

110

60

105

105

175

150

60

86

100

Rb (ppm)

91

70

120

58

41

62

1.0

8.0

15.5

11

33

50

2.4

35

Sr (ppm)

430

480

600

200

390

125

140

230

125

100

72

290

200

220

Pb (ppm)

11.5

13.5

12.0

4.5

13.0

6.5

5.5

4.0

7.0

6.5

6.5

3.0

5.0

2.0

Th (ppm)

10.5

6.5

7.0

5.5

6.0

4.4

4.3

0.25

0.29

0.65

0.24

0.16

0.17

0.34

U (ppm)

4.6

2.3

2.8

2.1

2.1

1.6

1.45

0.09

0.11

0.28

0.09

0.06

0.05

0.14

Ce (ppm)

145

98

105

91

84

66

54

5.5

6.0

15.0

5.0

3.5

3.5

6.0

La (ppm)

73

52

56

47.0

45.5

34

29.5

2.5

2.5

7.5

2.0

1.5

1.5

3.0

Nd (ppm)

91

100

115

57

88

64

56

5.5

8.5

21.0

8.5

5.5

6.0

8.5

Sm (ppm)

15.0

16.5

17.5

10.0

14.0

10.5

9.0

<0.02

2.5

5.5

2.4

1.65

1.70

2.3

RM
79.85

Clear
area

Table 1b. Major oxide and selected trace element analyses.


Oxide/
Element

River Miles (RM) 126.5-129 sequence

Cottonwood
Canyon

Creek Cremation
Creek

Pipe
Creek

VS13

BA-6

BA-15 BA-16 BA-17 BA-18 VS-8

VS-10 BA-1

VS-11

BA-3 VS-12

BA-4

SiO2 (%)

50.2

49.0

51.2

51.1

47.1

49.3

50.5

49.1

50.4

47.7

50.1

50.1

47.3

TiO2 (%)

1.31

1.95

2.33

1.12

1.14

1.75

0.99

1.11

1.20

0.79

0.92

0.76

0.49

Al2O3 (%)

13.2

12.7

12.3

13.1

13.7

12.2

13.8

13.0

13.6

14.7

12.6

14.7

14.0

Fe2O3 (%)

16.8

18.5

16.9

13.8

14.5

17.1

14.1

15.3

14.8

13.0

11.0

12.2

10.8

MgO (%)

5.12

5.13

4.62

6.77

6.74

4.89

6.16

6.11

6.14

6.29

11.4

7.39

11.0

MnO (%)

0.24

0.25

0.18

0.20

0.21

0.24

0.30

0.23

0.28

0.21

0.19

0.21

0.24

CaO (%)

8.36

8.89

7.93

8.78

10.9

8.19

9.34

9.59

8.41

13.8

9.32

9.57

11.8

Na2O (%)

3.17

2.19

2.66

3.08

2.36

2.63

2.15

1.63

1.93

0.69

0.77

2.71

2.17

K2O (%)

0.51

0.61

0.44

0.50

0.63

0.74

1.04

1.41

1.66

0.82

1.62

1.40

1.10

P2O5 (%)

0.16

0.18

0.37

0.16

0.16

0.31

0.15

0.16

0.15

0.24

0.17

0.12

0.06

LOI (%)

0.86

0.69

1.44

1.47

2.83

2.77

1.08

1.97

1.49

1.77

1.61

1.18

0.90

Total

99.93 100.09 100.37 100.08 100.27 100.12 99.61

99.61

100.06 100.01 99.70 100.34

99.86

Cr (ppm)

<2

<2

33

165

170

26

63

31

130

550

110

480

V (ppm)

340

650

390

320

350

380

270

300

330

180

195

230

240

Ni (ppm)

25

24

28

78

86

28

52

30

37

73

300

69

160

Co (ppm)

49.5

66

32

45.5

49.5

42

46

51

63

39

54

47.5

59

Cu (ppm)

39.5

120

150

23

23.5

45.5

130

43

96

60

6.5

48

2.0

Zn (ppm)

145

140

130

115

105

145

250

150

140

87

88

98

175

Rb (ppm)

11.0

8.5

4.7

9.5

16

10.5

17.5

29

30

22.5

82

44.5

9.0

Sr (ppm)

220

160

150

165

145

115

360

240

180

320

65

230

210

Pb (ppm)

5.0

5.0

8.0

5.5

3.5

6.5

34.5

11.5

8.0

10.0

6.5

6.0

6.0

Th (ppm)

0.81

0.92

1.10

0.29

0.37

0.63

1.0

0.63

4.1

0.24

4.9

0.87

1.95

U (ppm)

0.35

0.39

0.51

0.13

0.17

0.24

0.64

0.45

0.40

0.29

1.35

0.36

0.26

Ce (ppm)

13.5

15.5

18.5

5.5

8.0

14.0

11.0

16.5

18.0

11.5

43.0

9.5

5.5

La (ppm)

7.5

8.0

9.0

2.5

4.0

7.0

6.0

9.0

9.5

8.0

23.5

5.0

3.1

Nd (ppm)

11.5

11.5

23.5

9.0

11.0

18.5

9.5

13.5

13.0

9.5

21.5

8.0

3.8

Sm (ppm)

3.3

2.4

6.0

2.6

2.8

4.8

2.6

3.3

2.9

2.4

4.6

2.2

<0.02

All the results of the Rb-Sr, Sm-Nd, and Pb isotope analyses are listed in Table 3. As expected from the trace element
analyses, there is a relatively good spread in the isotope ratios for each of the radioisotope pairs. This is ideal for the plotting
of isochrons with good statistics for the fits to the data, and for significant age calculations.The Isoplot computer
program,58 which is now commonly utilized by the geochronology community, was used to process the analytical data for
each radioisotope system. This program utilizes the least-squares linear regression method59 to plot the isochron as the
best-fit regression line through the data. The slope of the isochron is then used by the program to calculate the isochron age
using the standard isochron-age equation. When plotted, each data point has assigned to it error bars that represent the
estimated 2s uncertainties due to the analytical equipment and procedure. The program also evaluates the uncertainties
associated with the calculated isochron age using a statistic known as the mean square of weighted deviates (MSWD),
which is roughly a measure of the ratio of the observed scatter of the data points from the best-fit line or isochron to the
expected scatter from the assigned errors and error correlations (including, but not limited to, the analytical equipment). If
the assigned errors are the only cause of scatter, so that the observed scatter approximates the expected scatter, then the
value of the MSWD will tend to be near unity. MSWD values much greater than unity generally indicate either
underestimated analytical errors or the presence of non-analytical scatter, while MSWD values much less than unity
generally indicate either overestimated analytical errors or unrecognized error-correlations. Thus it is crucial to estimate
adequately the analytical errors so that the observed scatter of the data points from the isochron line yields an MSWD near
unity. This was the procedure adopted here, so that the isochrons plotted in Figs. 68 have MSWDs near unity. The errors
for the isochron ages calculated from the isochrons represent the estimated 2s uncertainties. This does not mean that the
true age of the samples has a 95% probability of falling within the stated age interval, but rather only signifies that the mean
of the infinitely-replicated regressions would yield an isochron age within this interval.
Table 2. K-Ar isotope analyses and model ages.
Location

40
Sample K2O 40K
Ar* 40Ar*
Code (wt%) (ppm) (nl/g) (ppm)

40

Ar* 40Ar*/
Total40Ar
40
(%) Total Ar (ppm)

40

36

Ar/ Ar

36

UncerModel
tainty
Ar (ppm) K/ Ar Age
(Ma) (1
(Ma)
sigma)
40

36

BA-2

1.997 1.977 93.72 0.1671 94.03 0.9403

0.1777

4960.3

0.0000358 55223.5 1082.5 29

BA-7

1.738 1.721 79.36 0.1415 73.3 0.733

0.1930

1108.3

0.0001741 9885.1 1060.4 28

BA-8

1.752 1.734 95.08 0.1695 82.1 0.821

0.2065

1650.4

0.0001251 13860.9 1205.3 31

UM-1

1.015 1.005 87.97 0.1569 85.85 0.8585

0.1828

2092.5

0.0000873 11512.0 1666.5 63

UM-2

1.011 1.001 73.58 0.1312 77.9 0.779

0.1684

1339.0

0.0001257 7963.4 1482.8 39

Clear
UM-3
Creek
small body UM-4

0.776 0.768 53.92 0.0961 75.3 0.753

0.1276

1197.7

0.0001065 7211.3 1436.5 38

0.118 0.117 21.34 0.0380 49.5 0.495

0.0768

585.8

0.0001310 893.1

BA-5

1.474 1.459 21.95 0.0391 77.73 0.7773

0.0503

1389.8

0.0000361 40415.5 418.1 12

BA-9

0.837 0.829 42.34 0.0755 84.7 0.847

0.0891

1933.1

0.0000460 18021.7 1144.3 30

BA-10 0.687 0.680 45.98 0.0820 81.5 0.815

0.1006

1598.0

0.0000629 10810.8 1399.7 37

BA-11 0.904 0.895 49.63 0.0885 72.2 0.722

0.1226

1065.6

0.0001150 7782.6 1215.8 32

BA-12 1.304 1.291 41.19 0.0734 78.4 0.784

0.0936

1372.9

0.0000681 18957.4 794.6 21

BA-13 0.672 0.665 20.06 0.0358 58.7 0.587

0.0610

716.8

0.0000850 7823.5 758.6 21

BA-14 1.352 1.338 38.51 0.0687 84.6 0.846

0.0812

1925.3

0.0000421 31781.5 730.3 20

VS-13

0.528 0.523 ---

0.1029

870.5

0.0001182 4424.70 1259

BA-6

0.607 0.601 32.80 0.0585 79.41 0.7941

0.0737

1437.8

0.0000512 11738.3 1201.1 39

0.01263 1159.3

0.0000108 33703.7 405.1 10

BA-16 0.653 0.646 28.42 0.0507 62.6 0.626

0.0810

791.1

0.0001023 6314.8 1022.2 28

River Miles BA-17 0.613 0.607 24.76 0.04414 61.1 0.61


126.5129
sequence BA-18 1.323 1.310 71.82 0.1280 87.4 0.874

0.0722

760.9

0.0000948 6403.0 964.9 26

0.1465

2345.8

0.0000624 20989.9 1205.9 31

VS-8
1.138 1.126 --Cottonwoo
d Canyon VS-10 0.827 0.818 ---

0.1440 92.3 0.923

0.1560

4571.5

0.0000341 33020.5 1439

37

0.1147 89.5 0.895

0.1282

3507.5

0.0000365 22411.0 1532

39

Blacktail
Canyon
area
sequence

BA-15 0.368 0.364 5.28

0.05528 53.7 0.537

0.00941 74.5 0.745

2574.2 73

38

RM 79.85

BA-1

2.234 2.212 109.66 0.1955 83.4 0.834

0.2344

Clear
Creek

VS-11

0.272 0.269 ---

0.07028 693.5

0.0001013 2655.48 1318

0.1361

0.0000831 18194.9 975.4 24

0.03036 43.2 0.432

near Clear
Creek
BA-3

1.527 1.512 62.51 0.1115

Cremation
Creek
VS-12

0.571 0.565 ---

Pipe Creek BA-4

1.138 1.126 77.98 0.1390 92.08 0.9208

81.92 0.8192

0.07295 73.0 0.730

1783.2

1637.5

0.0001314 16834.1 1119.4 34


43

0.09993 1116.5

0.0000894 6319.91 1447

0.1510

0.0000403 27940.4 1372.9 26

3739.7

43

The K-Ar model ages in Table 2 range from 405.1 10 Ma to 2574.2 73 Ma. The other radioisotope pairs yielded isochron
ages of 1240 84 Ma (Rb-Sr), 1655 40 Ma (Sm-Nd), and 1883 53 Ma (Pb-Pb) (figs. 68). Unfortunately, the K-Ar
isotope data contained too much scatter to yield a statistically viable K-Ar isochron and isochron age. This suggests that a
significant component of the data has been perturbed by factors other than simple analytical error. Factors likely to be
responsible for the wide variations in the K-Ar model ages include contamination by open-system behavior such as
additions from the host wallrocks, and/or perturbing of the K-Ar radioisotope system.
Fig. 5. Representative photo-micrographs of the amphibolite samples collected for this study from outcrops along the
Colorado River corridor (see fig. 1). All photo-micrographs are at the same scale (20 or 1 mm = 40mm). The amphibolites
are here viewed under crossed polars and consist of hornblende (bright colors, two cleavages) and plagioclase (grey,
multiple twinning). There is only occasional minor sericite alteration (speckled, shimmering) of the plagioclase.
BA-1 (River Mile 78.5)
VS-11 (Clear Creek area)
BA-10 (Blacktail Canyon area)
BA-13 (Blacktail Canyon area)
BA-6 (River Miles 126.5129)
BA-15 (River Miles 126.5129)
Some of the 27 data points do not plot on the isochrons in Figs. 68. The Rb-Sr isochron is defined by only 19 samples, the
Sm-Nd isochron by 21 samples, and the Pb-Pb isochron by 20 samples. For each isochron the assigned analytical errors
are low and the statistics are excellent.
Fig. 6. 87Rb/86Sr versus 87Sr/86Sr isochron diagram for the Brahma amphibolites in Grand Canyon. Nineteen of the 27 wholerock samples were used in the isochron and age calculations. The bars represent the 2s uncertainties.The observed scatter
matches low assigned analytical errors whenever the MSWD equals unity. The excellent statistics for these isochrons,
coupled with the wide spread of the data points, yield isochron ages with low 2s uncertainties. If more data points were
included in the Isoplot analyses, both the Sm-Nd and Pb-Pb isochron ages would essentially be the same, but the 2s
uncertainties would be higher, whereas the Rb-Sr isochron age would be significantly lower. Only 610 samples are
routinely used in most radioisotope dating studies for plotting isochrons and calculating isochron ages. The use of 1921
samples in this study is exceptionally generous. In isochron regression analysis and isochron age calculations outlying data
points are frequently excluded in order to improve the isochron fits and the resultant statistics, even when the excluded data
plot as close to the isochrons as they do here.
Discussion
Fig. 7. 147Sm/144Nd versus 143Nd/144Nd isochron diagram
for the Brahma amphibolites in Grand Canyon. Twentyone of the 27 whole-rock samples were used in the
isochron and age calculations. The bars represent 2s
uncertainties.These 27 samples of the Brahma
amphibolites in Grand Canyon yielded an enormously
wide range of K-Ar model ages, from 405.1 10 Ma
to 2574.2 73 Ma, for a rock unit that is supposed to
be 17401750 Ma. Even samples only 0.84 meters
apart in the same outcrop of the small amphibolite
body near Clear Creek (table 2, lower right corner of
fig. 1, fig. 3) yielded K-Ar model ages of 1205.3 31
Ma and 2574.2 73 Ma. Petrographically these
amphibolites are dominated by hornblende which
shows no signs of any alteration to chlorite, while the
minor plagioclase is not replaced by sericite alteration,
which is confirmed by the low whole-rock K2O contents
of 0.111.96 wt% (table 1). Thus the wide range of KAr model ages for these amphibolites cannot be
explained by any significant variable alteration. These
differences instead could be explained easily by 40Ar*
loss from one part of the outcrop and accumulation of
excess 40Ar* in the other part of the same outcrop. This could
also account for why there is too much scatter in the K-Ar isotope
data for this rock unit to produce a viable isochron and a
statistically valid isochron age for it. Such an excessively wide
range of K-Ar model ages and the corresponding inability to
obtain a valid K-Ar isochron age, plus the uncertainty over 40Ar*
mobility and the role of excess 40Ar*, must cast some doubt over
the reliability of K-Ar radioisotope dating. While the Ar-Ar method
is now often preferred, due to its ability often to decipher loss or
accumulation of Ar, in this study the K-Ar method was used

because the Ar-Ar method is usually performed on individual minerals, whereas here the focus was on comparison of the
whole-rock analyses of the four major radioisotope systems.
Fig. 8. 206Pb/204Pb versus 207Pb/204Pb isochron diagram
for the Brahma amphibolites in Grand Canyon. Twenty of
the 27 whole-rock samples were used in the isochron
and age calculations. The ellipses represent the 2s
uncertainties.Snelling60, 61 reported much evidence of
the mobility of 40Ar* in crustal rocks. This can result
in 40Ar* loss from some rock units and some minerals
within rock units and 40Ar* excess in other rock units and
minerals. As an inert gas which does not chemically
bond with the crystal lattices of minerals, 40Ar* can
migrate from and through lava flows. Subsequent to their
extrusion these lava flows were buried under a
considerable thickness of sedimentary and other
volcanic strata, and subjected to metamorphism. Thus,
there were ample opportunities for40Ar* to be lost from
these rock units to surrounding rock units or the
atmosphere. In other rock units excess 40Ar* could
accumulate. Even within single rock units there could be
regions and minerals whose 40Ar* content has been
depleted, and nearby areas and minerals in which
excess 40Ar* has accumulated. This explanation might
well account for the wide variations in the individual sample K-Ar model ages for the metamorphosed basalts reported in
Table 2. Different samples of amphibolite units may contain different quantities of K, and therefore 40K (sometimes vastly
different quantities). Nevertheless, all samples from these amphibolites are supposed to be the same age. No matter what
their 40K concentrations are, a constant rate of 40K decay should have yielded the same proportional quantities of 40Ar*, so
that all samples yielded the same model age. Thus the wide range of K-Ar model ages recorded for these amphibolites must
be due to some cause, other than decay rate variability. The mobility of 40Ar* would seem to provide the most likely
explanation.Alternately, it might be postulated that substantial 40K decay has not really occurred but that the 40Ar in the lavas
was acquired instead from their mantle source and conduit contacts. However, it is certain that much radioactive decay has
occurred throughout the earths history, physical evidence for which is provided by radiohalos62, 63, 64, 65 and fission
tracks.66 This would thus imply that the K-Ar model ages and Rb-Sr, Sm-Nd, and Pb-Pb isochron ages yielded by these
ancient volcanic rocks should be due primarily to radioactive decay of the parent radioisotopes.Austin67 has discussed the
nature of the linear isotope plots for two other rock units. That discussion is equally relevant to the linear isotope plots
obtained for the Brahma amphibolites in Figs. 68. These isochron plots reveal an extraordinary linearity within the 87Rb87
Sr, 147Sm-143Nd, and 207Pb/206Pb-204Pb radioisotope systems. The extraordinary linearity implies a high degree of statistical
consistency undergirding these isochrons and isochron ages, with the isochrons passing within the relatively small
estimated range of analytical errors for every one of the data points. As already indicated, the observed scatter in these plots
has been fully accounted for by the assigned errors, as measured by the MSWD being at or near unity in each case. By
contrast, the attempted isochron plot for 40K-40Ar was not successful because the observed scatter was much too vast to be
accounted for, even if the analytical errors were assumed to be much larger than the laboratory estimates.Snelling68 has
reviewed the many problems that beset the major parent-daughter radioisotope pairs. Whole-rock Rb-Sr systems can be
disturbed and reset to give good-fit secondary isochrons even by relatively low-grade metamorphism when there may be
little field evidence and only relatively minor mineralogical alteration.69 Rb and Sr elemental abundances are known to have
been redistributed during metamorphism,70 but such redistribution would not necessarily be systematic, so the Rb-Sr data
would not then plot on an isochron as they do for these Grand Canyon Brahma amphibolites. 87Sr loss from hornblende at
high temperatures has been documented,71 but in that study the loss was due to contact metamorphism, whereas
systematic loss would not be likely during regional metamorphism where the whole region is at a similar temperature, as is
evident with the Grand Canyon Brahma amphibolites. Thus, it is not legitimate to discount the reliability of this whole-rock
Rb-Sr isochron age of 1240 84 Ma for these amphibolites because it is discordant with the Sm-Nd and Pb-Pb isochron
ages for the same rocks, and with the expected age of 17411750 Ma for the original basalt lavas based on a zircon U-Pb
date for associated felsic lavas.72,73 After all, the published 10-point whole-rock Rb-Sr isochron age for the Cardenas
Basalt in Grand Canyon74 is conventionally regarded as the best constrained radioisotope date for a Grand Canyon rock
unit where the Rb-Sr radioisotope system was closed, in contrast to the K-Ar radioisotope system.On the other hand, great
confidence has been placed in Sm-Nd isochron dating of Precambrian igneous and metamorphic rocks because of the
belief that the Sm-Nd radioisotope system is not reset, even during highgrade regional metamorphism.75, 76 Although some
studies have shown that the Sm-Nd radioisotope system is not remobilized during hydrothermal alteration of felsic rocks,
Poitrasson, Pin, and Duthou77 maintain that rare earth element mobility caused by certain hydrothermal conditions does
perturb the Sm-Nd system. However, in a study of high-grade metamorphic tonalitic gneisses and mafic rocks,
Whitehouse78 showed that the mafic rocks had retained the older true Sm-Nd age, whereas the Sm-Nd whole-rock
radioisotope systems in the non-mafic rocks had been reset to the same age as the U-Pb zircon and other whole-rock
radioisotope systems. It has already been noted that there are neither petrographic or geochemical indications in these
Grand Canyon Brahma amphibolites of hydrothermal alteration, and being mafic rocks it is thus most likely the Sm-Nd
radioisotope system in them has not been reset. Thus this 1655 40 Ma whole-rock Sm-Nd isochron age should be
considered as both valid and reliable (in conventional terms). Yet it is discordant with the published accepted ages for the
eruption of the original basalt lavas and their subsequent metamorphism, except perhaps that its upper limit does fall within
the age range for the metamorphism.The whole-rock Pb-Pb isochron method is still utilized because it is regarded as
reliable, even where U has been mobilized.79, 80 It is argued that if a group of rock samples all have the same age and
initial whole-rock Pb isotopic composition, they will then have developed with time different present-day, whole-rock Pbisotopic compositions according to their respective present-day U-Pb ratios. If the present-day, whole-rock Pb isotopic
compositions of a suite of rock samples is then plotted, as has been done here with these Grand Canyon Brahma
amphibolites, they are expected to form a straight-line array provided they have remained closed systems, the slope of the
array being dependent on the age of the rocks, in this case 1883 53 Ma. However, Pb-Pb isotopic linear arrays are also
known to be mixing lines, particularly for mafic volcanic rocks,81 as these amphibolites originally were. Faure admits the
problem with a word of caution: not all linear arrays on the Pb-Pb diagram are isochrons.82 This Pb-Pb isochron age for
these Braham amphibolites is also discordant with the accepted published age based on zircon U-Pb dating of associated

felsic volcanic rocks; yet there is no good reason to discard it as unreliable or invalid, because there is no way of
determining if this isochron represents a linear array inherited from the original mafic magma source. However, it is more
likely that high-grade metamorphism homogenized the whole-rock Pb isotopic compositions but did not reset the Pb-Pb
radioisotope system, as with the zircon U-Pb system in the associated felsic volcanic rocks.83, 84 It is often argued that
usually zircon U-Pb ages are the best, but they do suffer from the possibilities of the zircons being inherited grains so their
U-Pb ages are source ages, and of thermal resetting of the radioisotope systems within them. On the other hand, the wholerock Pb-Pb system is more likely to have been homogenized but not reset, so it could be argued the whole-rock Pb-Pb
isochron age is more reliable.Therefore, assuming that each of the Rb-Sr, Sm-Nd, and Pb-Pb isochron ages are valid and
reliable, it is clearly evident that there is gross disagreement between these radioisotope systems as to the age of these
amphibolites. Indeed, the discordances between the isochron ages are pronounced. Snelling et al.85 and Austin86 have
discussed explanations for these discordances and have concluded that, since the different radioisotope systems have to be
dating the same geologic event, the only explanation that can account for them is that the radioisotope decay rates have not
always been constant. Several examples were cited from the geologic literature that similarly report discordant isochron
ages for other rock units, most notably for the Great Dyke in Zimbabwe,87, 88 for the Stuart dyke swarm of central
Australia,89 and the Uruguayan dike swarm in South America.90 These examples, together with the large number of other
examples recorded by Snelling strongly suggest that, where two or more of the commonly-used radioisotope pairs are
applied to date rock units, discordances are the norm and not the exception.91Explanations for these discordances have
been attempted.92, 93, 94 For example, different isotope pairs have different closure temperatures that can therefore result
in different ages. There may also have been an open-system, magma mixing, inheritance, and/or paleoweathering. It is
sometimes suggested that there may have been differing chemical activity of the daughter elements in comparison with the
parent elements due to having different chemical properties, which is supposedly why the Pb-Pb age is often the highest.
Furthermore, submarine volcanic rocks have been known to retain Ar due to the hydrostatic pressure. Yet while those
geologic or geochemical explanations can often seem reasonable, they are not relevant to this study. In nearly every other
study reported in the literature, the discordant ages have been obtained by using the different radioisotope methods
on different whole-rock samples and/or different minerals. Similarly, in the many studies where concordant ages have been
obtained, such as for the Shap Granite in England as reported by Snelling, the different radioisotope methods have been
used again both on different samples and different minerals.95 However, in this study all four radioisotope methods were
used on the same whole-rock samples. This rules out the possibility of any of these postulated processes having any
significant effect on the resultant radioisotope ages. The high-grade metamorphism would have affected all four radioisotope
systems similarly. Using whole-rocks for the analyses homogenizes any different chemical or system behaviors in the
different minerals. In any case, these amphibolites are essentially just simple, two-mineral systems (fig. 5). Thus, this
somewhat uniform approach makes it more likely that the discordances are due to the radioisotope decay rates having not
always been constant.

Fig. 9. The isochron ages yielded by three radioisotope systems for the Brahma amphibolites, Grand Canyon, plotted
against the present half-lives of the parent radioisotopes according to their mode of decay.There is a pattern to the isochron
discordances. The isochron ages consistently indicate that the a-emitters ( 238U, 235U, 147Sm) yield older ages than the emitters (87Rb and 40K) when used to date the same geologic eventthat is, the formation of a specific rock unit. In the case
of the Brahma amphibolites, none of the isochron ages date the eruption of the original basalt lavas at 17401750 Ma, the
value based on U-Pb concordia dating of zircon grains in metamorphosed felsic volcanic lavas within the associated Brahma
and Rama Schists that are believed to have survived the metamorphism without the U-Pb radioisotope system being
reset.96, 97 Nor do the isochron ages yield the accepted date for the metamorphism of the original basalt lavas to form the
amphibolites. That event has been placed at 16901710 Ma, on the basis of U-Pb concordia dating of metamorphic
monazite, xenotime and titanite in the overlying Vishnu Schist and underlying Rama Schist.98, 99 Nevertheless, assuming
the isochron ages for the Brahma amphibolites are dating the same geologic event, the formation of the amphibolites, a
logical explanation of these data is that the radioisotope decay of the various parent isotopes has not always proceeded at
the rates described by the present-day decay constants. Thus, the discordances would instead be due to the parent
radioisotopes decaying at different rates from present rates, with the decay rates for the different parent isotopes having
been accelerated by different amounts. The data are consistent with the possibility that at some time or times in the past adecay was accelerated more than -decay.The correlation between the present radioactive decay constants for these a- and
-emitters and the ages they have yielded for the Brahma amphibolites is illustrated in Fig. 9. Austin 100 and
Snelling101 provide similar isochron age versus half-life and mode of decay diagrams for the radioisotope systems within
other rock units. In each of these examples the a-decaying isotopes ( 238U and 147Sm) yielded older isochron ages than the decaying isotopes (40K and87Rb). Among the -decaying isotopes, 87Rb has the smaller decay constant and thus the longer
half-life, yet in the cited
rock units it consistently
yields the older ages, for
example, double the K-Ar
isochron age of the
Cardenas Basalt.102 On
the other hand, even
though 147Sm
has
the
smaller decay constant
(and thus the longer halflife) of the a-decaying
isotopes, it does not
always yield the older
isochron age. It does for
other
rock
units,103, 104, 105 but not
for
the
Brahma
amphibolites
(fig.
9).
Perhaps
the
metamorphism of the original basalt lavas may have reset the Sm-Nd radioisotope system in the resulting Brahma
amphibolites but not perturbed the U-Pb radioisotope pairs. If this is the case, then it may also be possible that at some time
or times in the past the longer half-life a-emitter 147Sm had its decay accelerated more than the other a-emitters 238U
and 235U.
Fig. 10. The isochron ages yielded by three radioisotope systems for the Brahma amphibolites, Grand Canyon, plotted
against the atomic weights of the parent radioisotopes according to their mode of decay.There is another possibility
originally suggested by Austin and displayed by Vardiman et al. in their Fig. 5. 106 If these isochron ages are plotted against
the atomic weights of the parent radioisotopes (fig. 10), there is a clear relationship of increasing isochron age with atomic
weight. In this instance, U has a heavier atomic weight than Sm, and the Pb-Pb isochron age is older than the Sm-Nd
isochron age. And in this relationship the a-decay ages are older than the one -decay (Rb-Sr) age. Whatever mechanism
or mechanisms were involved in causing accelerated radioisotope decay, there thus were likely at least three factors
involved to produce differing amounts of acceleration for the different parent isotopesthe mode of decay, the present
halflife, and the atomic weight. However, with a combination of these three factors influencing the amount of acceleration of
the decay rates of the different parent radioisotopes, it might seem that these four methodsK-Ar, Rb-Sr (-decay), and
Sm-Nd, U-Pb (a-decay)may not be really enough to resolve the mystery of the mechanism for accelerated decay. Thus, in
future studies it is planned to utilize all four -decay radioisotope systemsK-Ar, Rb-Sr, Lu-Hf, and Re-Os.The critical
problems with accelerated radioisotope decay have been discussed by Vardiman et al.107 The question of how the heat
would have been dissipated is particularly relevant. Volumetric cooling due to concurrent stretching of the fabric of space
has been suggested as a viable mechanism.108 Kinnaird has demonstrated that a small change in the binding energies
within the nuclei of the parent radioisotopes not only causes the decay rates of the heavier elements to be greatly increased,
but drastically decreases the output of heat energy.109 Furthermore, if the accelerated nuclear decay is due to small
changes in the forces holding together the nuclei of the parent radioisotopes,110 then the accelerated nuclear decay would
not alter the electronic aspects of matter or affect the chemistry of the elements involved. On the other hand, there is clear
evidence that there may not have really been a heat problem. Because 238U radiohalos in biotite flakes are annealed at and
above 150 C, if there had been too much heat during accelerated 238U decay, no 238U radiohalos would have formed and
still be visible today.111In any case, it is clear that the different radioisotope systems produced discordant isochron ages for
the same geologic event (the eruption of the original basalt lavas or their subsequent metamorphism). For this discordance
to be due to changes in radioisotope decay rates, the a-decay rates must have been accelerated more than the -decay
rates. It is also possible that the longer the half-life of a radioisotope, whether a or -emitter, the more its decay has been
accelerated at some time or times in the pastthat is, the slower the decay rate, the more it is has been temporarily
increased. For these reasons, the isochron dating method cannot be relied on to give true real time ages. As with the K-Ar
radioisotope system, there are good reasons, many perhaps as yet unknown, for rejecting each of these radioisotope
systems as being capable of yielding reliable absolute ages.
Conclusions
The Brahma amphibolites of the Granite Gorge Metamorphic Suite of the Grand Canyon were originally erupted as basalt
lavas in a thick sequence of lavas and sediments that suffered high-grade regional metamorphism. This sequence was
intruded by granitic plutons to form the crystalline basement of the region. These amphibolites are a very simple whole-rock
system, primarily composed of hornblende, with minor subordinate plagioclase. Both petrographic examination and

geochemical analyses of the 27 samples used in this study do not suggest these rocks have suffered any significant
hydrothermal alteration or regressive metamorphism. The hornblende shows no sign of chloritisation, or the plagioclase of
sericitisation, and the K2O contents of the rocks are low.The samples yielded a wide range of K-Ar model ages, even for
adjacent samples from a single outcrop of the same original lava flow. No statistically viable K-Ar isochron age could be
calculated because of this scatter in the data. The best and most likely explanation is 40Ar* mobility within these rocks.The
Rb-Sr, Sm-Nd, and Pb-Pb radioisotope systems, on the other hand, yielded good, statistically valid, isochron ages of 1240
84 Ma, 1655 40 Ma, and 1883 53 Ma, respectively. These are clearly discordant with one another, with the published
17411750 Ma age for the original basalt lavas (based on zircon U-Pb dating of related felsic volcanics), and with the
published 16901710 Ma age for the subsequent metamorphism (based on monazite and titanite U-Pb dating of the
overlying Vishnu Schist). There are no clear reasons to reject any of these discordant isochron ages as invalid, or not dating
the same event. An explanation for this discordancy may be made on the basis of parent radioisotope decay at rates
different than presently measured during the interval since these rocks formed. This change in decay rate cannot have been
by a single common proportionality factor, applicable to all radioisotope species. Instead, the change in rate must have been
different for the different isotope pairs such that the U-Pb system decayed through 1,883 million years (in terms of todays
rate), while the Rb-Sr system decayed through 1,240 million years (likewise in terms of todays rate) over the identical actual
time interval.The pattern of isochron ages suggests several factors that may have been involved in the acceleration
mechanism(s)mode of decay (a or ), decay rate, and atomic weight. The a-decaying U and Sm yielded older ages than
the -decaying Rb, while the heavier atomic weight U yielded an older age than the Sm. In any case, if decay rates have not
been constant, radioisotope dating is unreliable and cannot provide valid absolute ages for rocks.
Acknowledgments
Numerous people made this research possible, and all are acknowledged and thanked. The samples were collected with the
permission of the National Park Service and the Research Office of the Grand Canyon National Park. Funding for the
sampling trips and analytical work was provided by the Institute for Creation Research from donations received towards the
RATE project. Help with sampling was provided by Tom Vail, Kurt Wise, and ICR raft tour participants. The reviewers and
the editor of the manuscript are thanked for their helpful input that greatly improved the readability of this paper. My wife
Kym and my family have tolerated my many long absences on field and other trips, and have supported my research
endeavors. Finally, the Lord has faithfully helped and guided to bring research projects like this to fruition.
Whole-Rock K-Ar Model and Isochron, and Rb-Sr, Sm-Nd, and Pb-Pb Isochron, Dating of the Somerset Dam
Layered Mafic Intrusion, Australia
by Dr. Andrew A. Snelling on March 24, 2010
Abstract
The Somerset Dam layered mafic intrusion in southeast Queensland, Australia, has been conventionally dated as Late
Triassic by the apparently successful application of radioisotopic dating techniques. Mineralogical, geochemical, and
isotopic evidence indicates that all of this gabbro intrusions cyclic units were derived coevally from the same parental
basaltic magma, with an initial homogeneous isotopic mixture ideal for yielding concordant isochron ages. However, newly
obtained K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic data from 15 whole-rock samples (representing all gabbro
macrolayers in four of the intrusions cyclic units) yield discordant isochron ages, although the excellent-fitting 15-point KAr isochron suggests the resultant 1748 Ma age (Middle Jurassic) should be regarded as the revised conventional age of
the layered intrusion. Nevertheless, it is concluded that these discordances between the radioisotope systems are likely due
to changes in their decay rates in the past, with the longer half-life b-emitter 87Rb being accelerated more and thus yielding
an older age. Furthermore, the Sr, Nd, and Pb isotopes indicate the parental basaltic magma was derived from a depleted
mantle source, while the large spread of Nd T DM ages suggests accelerated radioisotopic decay rates during the partial
melting and magma ascent. It is concluded that the Somerset Dam layered mafic intrusion has inherited the radioisotopic
signature of its mantle source, and so the conventional radioisotopic dating techniques do not provide its true age.
Keywords: gabbro, layered intrusion, Australia, potassium-argon, rubidium-strontium, samarium-neodymium, leadlead,
radioisotopic dating, whole-rock model ages, whole-rock isochron ages, discordances, decay constants, accelerated
decay, mantle source inheritance
This paper was originally published in the Proceedings of the Fifth International Conference on Creationism, pp. 305324
(2003) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh.
Introduction
The Somerset Dam layered mafic intrusion is situated immediately west of the village of Somerset Dam at 15232E and
277S, some 65 km northwest of the city of Brisbane in south-east Queensland on Australias east coast (Figure 1). The
outcrop is somewhat oval shaped, covering an area of about 4 km 2 with a diameter of about 1.5 km. It is a small layered
gabbro intrusion with an exposed stratigraphic thickness of 500 m on a steep hillside. It is a well-preserved, well-exposed,
steep-sided, discordant intrusion which is undeformed and unmetamorphosed. The roof and floor of the intrusion are not
exposed, and unknown thicknesses of the layered gabbros of the intrusion have been eroded from the top, and are
concealed below the exposed sequence. The location and nature of the feeder zone are unknown, yet it probably represents
a small, relatively shallow (35 km depth), subvolcanic magma chamber.1, 2

Due to its relatively small size, excellent preservation and good exposure this layered gabbro intrusion is ideal for the study
of the physical and chemical processes occurring in magma chambers emplaced at shallow crustal depths below volcanos
and feeding them.
Indeed,
layered
mafic
intrusions
provide
valuable
information on the
processes effecting
magmas while they
are
transported
from their mantle
sources until they
are emplaced. The
Somerset
Dam
intrusion though is
fundamentally
different from the
classic Skaergaard
intrusion
in
Greenland3, 4 and
the giant Bushveld
Complex intrusion
in
South
Africa,5, 6 since it is
characterized by several repetitions of similar patterns, and progressive differentiation is absent, suggesting venting of the
magma chamber through the overlying volcano and recharge from the magma source below several times during its
emplacement and cooling. The age of this intrusion and the timing of these processes have to be younger than the andesitic
lavas that the magma body has intruded, so radioisotope dating has been the only means for determining some time
constraints.
Figure 1. Location and general geology maps for the Somerset Dam layered mafic intrusion (after Mathison).7
Geologic Setting
The Somerset Dam layered mafic complex intrudes the andesitic lavas and tuffs of the Neara Volcanics, which have been
dated as early Middle Triassic based on plant spores in the underlying and overlying sedimentary formations,8, 9,10 and on
radioisotope dates of 24011 Ma (Ar-Ar), 232 Ma and 2428 Ma (K-Ar).11, 12, 13 Together with other acid to intermediate
volcanics and volcaniclastic and clastic sediments, the Neara Volcanic were deposited in the downfaulted Esk Trough, a
Mesozoic basin within the Paleozoic New England Orogen, elongated blocks of which are exposed to the west and east
paralleling the sides of the Esk
Trough.14, 15 To
the
south,
the
sediments and volcanics of the Esk
Trough plunge beneath the clastic
sediments and coal seams of the Upper
TriassicJurassic Moreton Basin.
The region is characterized by younger
basaltic volcanism, radioisotope dated
as
Late
Oligocene
to
Early
Miocene.16, 17These volcanic rocks
form a continuous belt extending northnorthwest from near the coast south of
Brisbane for over 200 km parallel and
close to the mountains west of the
Brisbane River Valley, which has been
eroded into the sediments and
volcanics of the Esk Trough. This
volcanism produced large, now eroded,
shield volcanoes, complex lava fields,
ring complexes and localized cones,
plugs,
laccoliths,
sills
and
dikes.18 Other
Tertiary
volcanic
remnants are found outcropping to the
east and northeast of the Somerset
Dam intrusion, perhaps suggesting a
wider distribution of these lava flows
across the region prior to the erosion.
The geological development of the
region has been carefully documented
by Cranfield et al.19 Based on
stratigraphic
relationships,
the
Somerset
Dam
layered
gabbro
intrusion is clearly younger than the
Neara Volcanics it intrudes, but just
how much younger has only been
determined by radioisotopic dating
which has yielded Late Triassic ages
for the intrusion.20, 21 Nevertheless, if
it
were
not
for
conventional
biostratigraphic and radioisotopic dating

a genetic relationship between the Somerset Dam gabbro and the basaltic volcanism which produced the Tertiary shield
volcanoes and lava fields may not be excluded. According to stratigraphic relationships, the sediments of the Moreton Basin
overly the Neara Volcanics and other sediments of the Esk Trough, with the erosion of both these basins apparently
commencing prior to the extrusion of the basalt and related lavas of the shield volcanos and lava fields.22,23 However, an
interpretation of the regional geologic record within a creation framework and timescale would place the deposition of the
volcanics and sediments in these Mesozoic basins as late in the Flood year, with the subsequent erosion at the end of the
Flood and extending into the post-Flood era to the present day. Thus if the Tertiary basaltic volcanic activity occurred early in
the post-Flood era, then it would not be far removed in time, or spatially, from the Somerset Dam gabbro intrusion that has
been interpreted as the remains of the magma chamber feeding a basaltic volcano.
Geology of the Intrusion
Figure 2. Geological map (top) and cross-sections (bottom) of the Somerset Dam layered mafic intrusion (after
Mathison24).
The geology of the Somerset Dam layered mafic intrusion and its immediate area has been thoroughly mapped and
investigated.25, 26, 27, 28, 29, 30, 31 The general geology of the immediate area is shown in Figure 1, while Figure 2 provides
a detailed geologic map and appropriate cross-sections of the intrusion, based on the definitive work of Mathison.32, 33, 34
Within this gabbro intrusion there is an exposed sequence of 22 saucer-shaped macrolayers, 350 m thick generally dipping
inwards at 1020. The contacts between these prominent layers are sharply defined, generally to within 10 cm, and are
phase, modal mineralogy and textural contacts. These macrolayers appear to be stratigraphically and laterally
homogeneous. The repetition of these macrolayers has allowed the recognition of at least six well-developed cyclic units,
ranging from 30 to 150 m thick (average about 80 m thick). The macrolayers are limited to only four main rock types, which
are defined in terms of their essential cumulus mineral phases (distinguished texturally from the intercumulus mineral
phases).35, 36 These four rock types constituting the macrolayers are leucogabbro or anorthosite (plagioclase cumulate),
troctolite (plagioclase + olivine cumulate), olivine gabbro (plagioclase + augite + olivine cumulate), and oxide (or ferri-)
gabbro (plagioclase + augite olivine + magnetite + ilmenite cumulate).The definition of these cyclic units, selecting which
of these macrolayers commences each cycle, is somewhat interpretive. The choice is strongly influenced by what is
expected to be the order of crystallization and the magma fractionation pattern, because there is commonly no clear field
evidence to identify the base and top of a cyclic unit. In the Somerset Dam gabbro intrusion, Mathison37, 38, 39, 40 chose to
define each cyclic unit to be the sequence troctoliteolivine gabbrooxide gabbroleucogabbro, because troctolite was
considered the least fractionated rock type, and cryptic trends generally suggested a reversal at the bases of the troctolites.
However, Mathison41 revised this choice of sequence in each cyclic unit so that anorthosite was defined as the basal layer
and oxide gabbro as the top layer in each cyclic unit, the choice subsequently followed by Walker.42 This interpretation
better fits the inferred order of crystallization, the oxide gabbro being the most fractionated rock type. Mineral compositions
in a typical cyclic unit therefore show a reversed fractionation trend in the sequence from anorthosite to troctolite, and a
normal fractionation trend from troctolite through olivine gabbro to oxide gabbro.43, 44 Whole-rock compositions also show
marked changes between these rock types in this cyclical sequence.45, 46, 47
Figure 3. Stratigraphic column for the
Somerset Dam Layered Mafic Intrusion
(after Mathison53) showing inferred cyclic
units,
rock
densities,
and
modal
compositions (blank = leucogabbro; dotted
= troctolite; dashed = olivine gabbro; black
= oxide gabbro; R.L. = best developed zone
of rhythmic layering)
Despite the remarkable similarity of
successive
cyclic
units,
significant
differences exist between them in the
sequences of macrolayers, thicknesses of
individual macrolayers and of the cyclic
units, mineral compositions and cryptic
patterns, average level of fractionation, and
the
sizes
of
the
reversals.48, 49, 50, 51 Figure 3 shows the
stratigraphic sequence in macrolayers
within the Somerset Dam intrusion, with the
cyclic units inferred by Mathison,52 the rock
densities, and the modal compositions. Unit
3, the thickest unit in the sequence, is
noteworthy by being different from the other
cyclic units. It appears to compromise two
incomplete cyclic units in which the
sequences are anorthositetroctolite (unit
3A) and leucotroctolitetroctolite olivine
gabbro (unit 3B). In the leucotroctolite is the
best developed zone of small-scale
rhythmic banding within the macrolayers of
the intrusion. Unit 4 is also very different in
that there is a 5-mthick zone with about 7%
olivine occurring near the middle of the
anorthosite at the base of the units sequence, so that the lower 30 m of unit 4 could be an incomplete or interrupted cyclic
unit, with the sequence anorthosite troctoliticanorthosite followed by the typical cyclic sequence from anorthosite through
to oxide gabbro.
Previous Radioisotopic Dating
The currently accepted conventional age of the Somerset Dam layered gabbro intrusion is based on two sets of K-Ar
determinations reported by Webb and McDougall.54 They made two measurements on hornblende from a hornblende
gabbro near the top of the intrusion, and two measurements on plagioclase from an olivine gabbro near the bottom of the

intrusion. Murphy, Trezise, Hutton, Cranfield, and Whittaker55 and Murphy, Trezise, Hutton, and Cranfield56 recalculated
the resultant K-Ar dates using the IUGS constants of Dalrymple.57 Thus the hornblende yielded revised ages of 218 and
220 Ma, and the plagioclase yielded revised K-Ar ages of 212 and 213 Ma. Of course, it would be expected on the
assumption of the sequential crystallization of the layering within the intrusion that the dates from the olivine gabbro near the
bottom of the intrusion should be older than the dates from the hornblende gabbro near the top, but it is reasonable to
attribute these differences in the age determinations to the within error of the technique. Taking these results as a group
yields an averaged age of 2164 Ma.Walker58 separated hornblende grains from the troctolite in cyclic unit 2 and the
olivine gabbro in cyclic unit 4, and used the argon laser incremental heating technique to make 40Ar/39Ar age determinations.
Only the hornblende grains from the olivine gabbro yielded a calculated apparent age of 221.50.9 Ma. Determinations
were also made on K-feldspar, muscovite, and biotite grains of alkali feldspar granite (granophyre) and quartz monzodiorite
units which he interpreted as being related to the layered intrusion (Figure 2), and together with the hornblende result, a
pooled apparent age of 224.24.8 Ma was calculated for the entire Somerset Dam igneous complex. Similarly,
Walker59undertook Rb-Sr age determinations on whole-rock samples of all four rock types in the layered intrusion plus the
nearby peripherally-related igneous units, as well as mineral determinations for an olivine gabbro and the peripheral granite
(granophyre). An isochron fitted to all these data using all the whole-rock samples and all the mineral separates yielded an
acceptable Rb-Sr isochron age for the whole Somerset Dam igneous complex of 225.32.3 Ma, which thus agreed
remarkably well with the pooled apparent Ar-Ar age.
Sample Collection
The field work and collection of samples was undertaken in June 1998 and November 1999. Access to suitable outcrops
was provided by a farm road that starts at the village of Somerset Dam, from the highway into the village and the dam itself,
which is a spur from the Brisbane Valley Highway just north of the town of Esk. The farm access road climbs the hillside just
to the west of the village and the dam, which is the hill containing the layered gabbro intrusion. Thus the road progressively
climbs up through the different macrolayers of the intrusion, with suitable outcrops in roadcuts and on the hillside either side
of the road. During the June 1998 field trip, twelve 23 kg samples were collected along this road traverse through the
intrusion, commencing with a sample of the oxide gabbro at the top of cyclic unit 2,60 with samples of all the successive
macrolayers for units 3 and 4, to two samples of the anorthosite macrolayer at the base of unit 5. The locations of these
samples are marked on Figure 2. During the November 1999 field trip a series of six 34 kg closely-spaced samples were
collected from the oxide gabbro macrolayer at the top of cyclic unit 2. The purpose of the close-spaced sampling was to
investigate mineralogical and geochemical variations within a macrolayer, and particularly to choose one suitable sample for
separation into its mineral constituents for radioisotopic dating of them in order to produce mineral isochrons, work still in
progress. The locations of these samples are also marked on Figure 2.
Laboratory Work
All samples were sent first for sectioningone thin section from each sample for petrographic analysis. Four of the six
closely-spaced samples from the oxide gabbro macrolayer at the top of cyclic unit 2 were then selected for analysis, along
with the 11 samples from the overlying macrolayers of units 3 and 4, and the anorthosite at the base of unit 5. Approximately
100 g of each of these samples was dispatched to the AMDEL Laboratory in Adelaide, South Australia, for whole-rock major,
trace and rare earth element analyses. A second representative set of pieces (50100 g) of each of these samples was then
sent to the K-Ar dating laboratory at Activation Laboratories in Ancaster, Ontario, Canada, for whole-rock K-Ar dating. A third
representative set of pieces (50100 g) of each of these samples was also sent to the PRISE Laboratory in the Research
School of Earth Sciences at the Australian National University in Canberra, Australia, for whole-rock Rb-Sr, Sm-Nd, and PbPb radioisotopic dating.
At the AMDEL Laboratory each sample was crushed and pulverized. Whole-rock analyses were undertaken by total fusion
of each powdered sample and then digesting them before ICP-OES (inductively coupled plasma optical emission
spectrometry) for major and minor elements, and ICP-MS (inductively coupled plasmamass spectrometry) for trace and
rare earth elements. Fe was analyzed for among the major elements by ICP-OES as Fe 2O3 and reported accordingly, but
separate analyses for Fe as FeO were also undertaken via wet chemistry methods that were also able to record the loss on
ignition, primarily representing H2O or any carbonate (given off as CO 2) in the samples. The detection limit for all major
element oxides was 0.01%. For minor and trace elements the detection limits varied between 0.5 and 20 ppm, and for rare
earth elements between 0.5 and 1 ppm.At Activation Laboratories the K-Ar analyses were performed under the direction of
the K-Ar Laboratory manager, Dr. Yakov Kapusta. No specific location or expected age information was supplied to the
laboratory, but a description of the samples was given so that the laboratory staff knew the types of samples they were
dealing with and could ensure their procedures were conducted appropriately. After crushing of the whole-rock samples and
pulverizing them, the concentrations of K (weight %) were measured by the ICP technique. The 40K concentrations (ppm)
were then calculated from the terrestrial isotopic abundance using these measured concentrations of K. The concentrations
in ppm of 40Ar*, the supposed radiogenic 40Ar, were derived using the conventional formula from isotope dilution
measurements on a noble gas mass spectrometer by correcting for the presence of atmospheric argon whose isotopic
composition is known.61 The ratios 40Ar*/total Ar and 40Ar/36Ar were also derived from the mass spectrometer
measurements.The Rb-Sr, Sm-Nd, and Pb-Pb isotopic analyses were undertaken at the PRISE Laboratory under the
direction of Dr. Richard Armstrong. No specific location or expected age information was supplied to the laboratory, although
samples were described as young gabbros so that the laboratory staff would optimize the sample preparation procedure in
order to obtain the best analytical results. At the laboratory the sample pieces were crushed and pulverized before being
dissolved in concentrated hydrofluoric acid, followed by the standard chemical separation procedures for each of these
radioisotope systems. Once separated, the elements in each radioisotope system were loaded by standard procedures onto
metal filaments to be used in the solid source thermal ionization mass spectrometer (TIMS), the state-of-the-art technology
in use in this laboratory. Sr isotopes were measured using the mass fractionation correction 86Sr/88Sr = 0.1194, and
the 87Sr/86Sr ratios reported are normalized to the NBS standard SRM 987 value of 0.710207. Nd isotopes were corrected
for mass fractionation using 146Nd/144Nd = 0.7219 and are normalized to the present-day 143Nd/144Nd value of 0.51268 for
BCR-1. Pb isotope ratios are normalized to NBS standard SRM 981 for mass fractionation.
Petrography and Chemistry
Each cyclic unit commences at its base with a macrolayer of leucogabbro (plagioclase cumulate or anorthosite). Plagioclase
(labradorite, 84% average modal composition62) is the main cumulus phase in the leucogabbros, with very minor cumulus
olivine (trace) and sometimes traces of cumulus magnetite. Intercumulus minerals (which crystallized subsequent to, and
therefore in the remaining space between, the cumulus phases63) in order of decreasing abundance include augite (3%),
brown hornblende (3%), ilmenite (2%), magnetite (1%), apatite (0.4%), and orthopyroxene (0.1%). The crystallization order
is thus inferred to have been plagioclase first, then minor olivine (if present), augite, magnetite and ilmenite, rare
orthopyroxene, brown hornblende, and finally apatite. The leucogabbros are consistently more altered than the other rock
types, and the augite (invariably ophitic or poikilitic) is usually highly altered to felted aggregates of a low-Al actinolitic pale

green amphibole (uralite, 6% average modal composition), which is thus commonly the next most abundant mineral to
plagioclase. The leucogabbros are mainly orthocumulates64 as indicated by their overall texture, the proportion of
intercumulus minerals, the relatively abundant brown hornblende, the extensive normal zoning in plagioclase (1520 mole
% An [anorthite]), the generally higher content of incompatible elements such as K and P, and probably also by the
extensive alteration that probably reflects a relatively high trapped fluid content upon crystallization.65Cumulus phases in
the troctolites (plagioclase-olivine cumulates) include plagioclase (labradorite, 75% average modal composition,66 olivine
(18%) and traces of Cr-magnetite, while the intercumulus minerals include mainly augite (3%), brown hornblende (0.5%),
magnetite (0.4%) and ilmenite (0.6%), with trace amounts of orthopyroxene and apatite. The crystallization order is inferred
to have been the same as for the leucogabbros, except that minor Cr-magnetite crystallized early, possibly after plagioclase.
At least some of the plagioclase had to have crystallized before the olivine because olivine grains commonly contain small
laths of plagioclase. The plagioclase laths are generally smaller and thinner than in the other rock types of the intrusion, and
they are strongly laminated. Olivine grains are subhedral when enclosed in augite, but are partly interstitial with respect to
plagioclase. Augite grains are always poikilitic or ophitic, and enclose numerous laths of plagioclase and a few grains of
olivine. However, the modal proportion and texture of augite also varies, and the troctolite in unit 4 particularly has
conspicuous augite oikocrysts (510 mm diameter). Nevertheless, the main variation in the troctolites is in the relative modal
proportions of plagioclase and olivine (see Figure 3), those troctolites with >80% plagioclase being leucotroctolites. Some
alteration of the augite has again resulted in aggregates of uralite, which can reach an average modal composition of 2%.67
In the olivine gabbros (plagioclase-augite-olivine cumulates) cumulus augite (20% average modal composition68 joins the
cumulus assemblage of plagioclase (labradorite, average 68%) and olivine (average 6%) of the troctolites, and the
intercumulus minerals are mainly brown hornblende (2%), magnetite (1%), ilmenite (1%), very minor orthopyroxene
(average 0.2%) occurring exclusively as partial reaction rings on some olivine grains, and trace apatite. The crystallization
order was probably similar to that for the leucogabbros. The augite is subhedral and twinned, and never surrounds olivine.
Again, it can be altered to uralite, which sometimes averages 1.5%. The brown hornblende is often ophitic and may be
present in amounts up to 5% by volume. Transitions between olivine gabbros and oxide gabbros are observed in the upper
cyclic units.Cumulus phases in the oxide gabbros (plagioclase-augite-magnetite-ilmenite cumulates) are mainly plagioclase
(labradorite, 55% average modal composition69) and augite (average 28%) with 03% olivine, some magnetite (average
7%) and ilmenite (average 6%). At least some of the magnetite is cumulus (subhedral octahedral), but extensive
intercumulus overgrowth has occurred so that most Fe-Ti oxide grain boundaries are interstitial to plagioclase and augite.
The augite is frequently twinned, and occurs as well-defined subhedral prisms. Although alteration of the oxide gabbros is
minimal, there are still some traces of uralite. Some brown hornblende is again present (up to an average of 1.5%), and
other intercumulus minerals are rare, while orthopyroxene is completely absent. The oxide gabbros have the most
adcumulate character, that is, during magmatic crystallization the interstitial liquid had been displaced by outgrowth from the
cumulus crystals so that the latter are interlocking.70 The oxide gabbros are also the most fractionated compositions of the
four main rock types in a typical cyclic unit. Though originally termed ferrigabbros by Mathison,71 because the total iron
content is similar to that of some of the ferrogabbros in the Skaergaard layered mafic intrusion in east Greenland, in these
Somerset Dam gabbros the Fe3+/Fe2+ is high (>0.7) due to the abundant magnetite, so Mathison72 referred to them as oxide
gabbros.The whole-rock major element oxide and selected trace and rare earth element analytical results from the 15
samples selected for this study are listed in Table 1. The distribution of the major element oxides between the four different
rock types in the macrolayers of each cyclic unit follows the pattern in the average whole-rock major element oxide
compositions for these rocks already reported by Mathison.73, 74 The oxide gabbros have the lowest SiO 2 and highest
Fe2O3 contents, and the anorthosites (leucogabbros) have the highest SiO 2 and lowest Fe2O3 contents. The Al2O3content is
highest in anorthosite at the base of each cyclic unit and decreases upwards through troctolite and olivine gabbro to the
lowest content in the oxide gabbro, which parallels the decreasing plagioclase content75 (see Figure 3). Mathison76 found
that the compositional changes with increasing height and increasing fractionation in a typical cyclic unit were clearly evident
from the major element oxide data, such as that in Table 1, and generally involve increases in total iron (both Fe 2+ and Fe3+),
the Fe3+/Fe2+ratio, TiO2 and MnO, and decreases in Al2O3, SiO2, Na2O, and K2O.The selected trace and rare earth element
data in Table 1 are also similar to the comparable data obtained by Mathison77 and Walker.78 In summary, Cr, Ni, V, and Cu
show the greatest variation both within a cyclic unit and throughout the intrusions stratigraphic succession. Furthermore, as
expected, the pattern of variations in each cyclic unit of the major and trace elements parallels the changes in mineralogy.
Thus the P2O5 content parallels the presence of accessory apatite, Sr parallels the abundance of plagioclase, and rare earth
elements such as La, Ce, and Nd are often slightly more concentrated in the oxide gabbros. Mathison79, 80, 81 also
analyzed the compositions of the major minerals (plagioclase, olivine, augite, orthopyroxene, brown hornblende, magnetite,
ilmenite, and biotite) and plotted the compositional variations through each cyclic unit for the entire exposed stratigraphic
sequence of the intrusion. He found that the most sensitive parameters for defining the cryptic compositional trends in the
minerals are An in plagioclase, Fo and Ni in olivine, and Cr in magnetite and augite.
Radioisotopic Dating Results
The K-Ar analytical data and K-Ar model ages for all 15 samples in this study are listed in Table 2. These model ages are
calculated using the standard equation of Dalrymple and Lanphere82 utilizing the ratio of the abundances of 40Ar* (the
radiogenic 40Ar) to 40K listed in Table 2. The model age method assumes no radiogenic 40Ar was present when the basaltic
magma was intruded to form each cyclic unit, and as it differentiated to form the layered gabbros while cooling. The model
ages range from 181.38 Ma (one-sigma error) to 252.89 Ma, with the mean age being 206.6 Ma (+46.2 Ma, 25.3 Ma, n
= 15). The wide variation in these model ages is not able to be explained readily, because they do not form a predictable
sequence of decreasing age with the formation of each cyclic unit in an ascending order from bottom to top of the intrusion.
Indeed, there is no recognizable pattern, except that many of the model ages are discordant from one another. The mean
model ages for the cyclic units sampled are 203.2 Ma (n = 4) for cyclic unit 2, 227.2 Ma (n = 5) for cyclic unit 3, 192.6 Ma (n
= 4) for cyclic unit 4, and 189.7 Ma (n = 2) for cyclic unit 5. Whereas there is an apparent younging trend upwards from
cyclic unit 3 to cyclic units 4 and 5, cyclic unit 2, which should be the oldest of those sampled yields a mean model age
younger than the overlying cyclic unit 3. Furthermore, where in cyclic units 3 and 4 the sequence of macrolayers has been
sampled and analyzed, these macrolayers should all be the same age, because they differentiated and fractionated from the
same pulse of intruded basaltic magma in each case, but the model ages obtained are discordant. Even where three
troctolite samples were collected in ascending order within that macrolayer within cyclic unit 3, each of those troctolite
samples yielded strongly discordant model ages. Similarly, samples SDI-2A4 and SDI-2A6, stratigraphically only several
meters apart in the same oxide gabbro macrolayer in cyclic unit 2, yield strongly discordant model ages of 194.78 Ma and
239.49 Ma respectively.
Table 1. Whole-rock major element oxide and selected trace element analyses of 15 samples from the Somerset Dam
layered mafic intrusion, near Brisbane, Australia. (Analyst: AMDEL, Australia; April, 2000).
Sample SDI-

SDI-

SDI-

SDI-

SDI-

SDI- SDI-

SDI- SDI- 3DSDI-

SDI-

SDI- 4BSDI- 4CSDI-

SDI-

2A1

2A3

2A4

2A6

2B

3A

3B

3C

3E

4A

4D

4E

Rock
Type

Oxide Oxide Oxide Oxide Anor- Troc- TrocGabbro Gabbro Gabbro Gabbro thosite tolite tolite

Troc- Olivine Anor- Troctolite Gabbro thosite tolite

Olivine Oxide Anor- AnorGabbro Gabbro thosite thosite

Cyclic
Unit

SiO2(%) 43.6

45.3

44.2

39.9

47.5

48.1 45.9

TiO2(%) 4.75

3.97

3.91

5.21

0.30

Al2O3(%) 14.6

17.1

17.2

17.2

Fe2O3(%
)
15.4

12.6

12.8

MgO
(%)

5.98

5.57

MnO
(%)

0.19

46.8 50.9

49.1

48.4

47.2

43.8

50.7

50.9

0.58 0.41

0.33 0.49

0.50

0.52

1.27

4.85

1.13

0.88

25.5

21.3 18.1

24.6 20.1

24.1

20.7

20.2

15.9

25.5

24.8

19.2

4.70

6.76 9.00

5.08 4.59

4.77

7.50

10.4

15.1

5.63

5.49

5.85

7.77

6.54

8.65 14.0

6.97 6.46

4.67

8.72

6.19

5.56

2.07

2.62

0.14

0.13

0.18

0.07

0.10 0.14

0.08 0.08

0.08

0.12

0.11

0.15

0.06

0.08

CaO (%) 11.5

10.5

10.9

6.64

12.3

10.3 9.84

11.9 13.2

12.4

10.4

10.9

11.8

10.3

10.2

Na2O
(%)

2.55

3.17

3.19

2.79

2.30

2.85 1.91

2.27 3.04

2.99

2.77

3.03

2.39

3.92

4.02

K2O (%) 0.15

0.24

0.23

0.12

0.12

0.20 0.09

0.14 0.18

0.27

0.22

0.19

0.11

0.27

0.49

P2O5(%) 0.23

0.30

1.20

0.04

0.06

0.11 0.04

0.06 0.04

0.15

0.07

0.05

0.01

0.06

0.08

Lol (%)

0.57

0.09

0.17

0.77

0.89

0.99 0.91

1.36 0.55

0.94

0.86

0.22

0.43

0.54

1.07

TOTAL

99.52

98.98

99.78

99.52

100.28 99.94 100.34 99.59 99.63

99.97 100.28 99.76

100.1

100.18 100.63

Cr (ppm) 70

70

60

20

90

100

200

120

550

200

380

140

40

40

70

V (ppm) 470

340

340

440

30

60

60

40

90

60

70

350

550

140

110

Ni (ppm) 26

92

115

230

91

36

39

81

49

Co
(ppm)

40

40

40

60

30

40

60

30

20

20

40

60

50

20

20

Cu
(ppm)

90

84

250

<2

12

32

33

13

38

20

45

210

60

13

Zn
(ppm)

83

59

59

94

34

46

63

55

33

45

55

60

79

42

74

Rb
(ppm)

4.5

4.5

3.5

1.5

2.5

3.5

1.5

2.5

3.5

5.0

4.0

3.5

1.5

5.0

15.5

Sr (ppm) 450

550

500

550

600

450

400

550

470

550

490

500

410

700

700

Y (ppm) 16

11

16

10

10

Zr (ppm) 90

60

50

40

30

40

30

30

30

40

50

40

40

40

50

Ba
(ppm)

50

60

60

40

45

45

30

40

50

65

60

55

35

75

100

Pb
(ppm)

25

20

15

30

<5

<5

<5

<5

<5

10

<5

<5

25

Th
(ppm)

33.5

20.5

18.5

17.5

9.5

11.0 12.0

9.5

10.0

9.5

11.0

11.5

14.0

8.0

8.5

U (ppm) 1.5

0.5

<0.5

<0.5

<0.5

<0.5 <0.5

<0.5 <0.5

<0.5

<0.5

<0.5

<0.5

<0.5

<0.5

La
(ppm)

11

13

10

Ce
(ppm)

12

10

16

12

10

10

Nd
(ppm)

13.5

9.0

16.0

2.0

4.5

6.5

4.0

4.0

5.5

7.5

7.0

4.5

4.5

4.5

6.0

Sm
(ppm)

3.5

2.5

4.0

<0.5

1.0

1.5

1.0

1.0

1.5

2.0

2.0

1.0

1.5

1.0

1.5

Table 2. Whole-rock K-Ar data for the Somerset Dam layered mafic intrusion, near Brisbane, Australia. (Analyst: Dr. Y.
Kapusta, Activation Laboratories, Ancaster, Canada; April, 2001). These data yield 15 point isochron ages of 1748 Ma (KAr) and 18123 Ma (40K/36Ar40Ar/36Ar).
Sample
Code

Rock Type

K2O (wt40K
%)
(ppm)

40

Ar*
(ppm)

Total 40Ar
(ppm)

40
36

Ar
Ar

/36

Ar (ppm)

40

K /36Ar

Model
Uncertainty (Ma)
Age (Ma) (1 sigma)

SDI-4E

Leuco
Gabbro

0.498

0.493

0.005445 0.007467

405.96 0.0000183 26939.9 184.2

+7

SDI-4D

Leuco
Gabbro

0.302

0.299

0.003516 0.004254

359.54 0.0000118 25338.9 195.1

+8

SDI-4C

Oxide
Gabbro

0.166

0.165

0.001804 0.002040

334.72 0.0000060 27500.0 182.7

+9

SDI-4B

Olivine
Gabbro

0.223

0.221

0.002676 0.002945

325.73 0.0000090 24555.6 201.1

+8

SDI-4A

Troctolite

0.223

0.221

0.002398 0.002657

328.03 0.0000080 27625.0 181.3

+8

SDI-3E

Leuco
Gabbro

0.224

0.222

0.002749 0.003588

386.34 0.0000092 24130.4 205.3

+8

SDI-3D

Olivine
Gabbro

0.212

0.210

0.002692 0.003818

419.73 0.0000090 23333.3 211.9

+8

SDI-3C

Troctolite

0.140

0.138

0.001870 0.002397

379.43 0.0000063 21904.8 222.7

+7

SDI-3B

Troctolite

0.100

0.099

0.001532 0.001932

373.25 0.0000051 19411.8 252.8

+9

SDI-3A

Troctolite

0.198

0.231

0.002459 0.003134

377.34 0.0000083 27831.3 208.2

+9

SDI-2B

Leuco
Gabbro

0.110

0.109

0.001590 0.001925

358.46 0.0000053 20566.0 240.3

+8

Oxide
SDI-2A6 Gabbro

0.100

0.009

0.001444 0.001851

379.55 0.0000048 20625.0 239.4

+9

Oxide
SDI-2A4 Gabbro

0.513

0.508

0.005953 0.009920

492.15 0.0000201 25273.6 194.7

+8

Oxide
SDI-2A3 Gabbro

0.251

0.248

0.002896 0.004060

415.04 0.0000097 25567.0 193.9

+7

Oxide
SDI-2A1 Gabbro

0.205

0.203

0.002250 0.003125

411.10 0.0000076 26710.5 184.9

+9

Figure 4. 40K versus 40Ar* in the Somerset Dam layered mafic


intrusion, all 15 samples being used in the isochron and age
calculations. The bars represent the two-sigma uncertainties.
Figure 4 is the 40K versus 40Ar* diagram in which all 15 samples
from the Somerset Dam layered mafic intrusion have been
plotted. The plotted error bars for each data point are the twosigma uncertainties, but there is still a very strong linear trend
apparent in the data, and this has been plotted as an isochron
using the Isoplot program of Ludwig83 that utilizes the leastsquares linear regression method of York.84 All 15 samples were
included in the regression calculation and produced an excellent
fit. The isochron age calculated from the slope of the line is
1748 Ma (two-sigma error). The two-sigma uncertainty in this
isochron age is very small because the two-sigma uncertainties
assigned to each of the data points are also very small. The
excellent fit of this isochron line to the data is highlighted by the
very low MSWD (mean square weighted deviatesa measure of
the ratio of the observed scatter of the data points from the bestfit line to the expected scatter from the assigned errors and error
correlations) and the very high probability (0.98). The initial 40Ar* is 0.00029 ppm, which is a very small quantity of excess
radiogenic 40Ar in violation of the assumption of zero 40Ar* in the model age technique. This K-Ar isochron age is
significantly discordant with the three sets of published K-Ar determinations on hornblende and plagioclase of Webb and
McDougall,85 recalculated by Murphy et al86,87 using the IUGS constants of Dalrymple,88 of 218 and 220 Ma, and 212 and
213 Ma respectively, which yield an averaged age of 2164 Ma. Similarly, this 15-point K-Ar isochron age of 1748 Ma is
strongly discordant with the pooled apparent age of 224.24.8 Ma based on 40Ar/39Ar determinations on three K-feldspar,
four muscovite, one hornblende and four biotite grains by Walker.89The whole-rock Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic
data for all 15 samples from the intrusion are listed in Table 3. As anticipated, the radioisotopic ratios in these 15 gabbro
samples show little variation in all three radioisotope systems. This is because of the similar mineralogy in these gabbros,
and in spite of the modal variations in these minerals between the different macrolayers. Even the significantly lower
plagioclase and higher magnetite and ilmenite contents in the oxide gabbros compared to the leucogabbros (anorthosites)
makes little difference in the variations in the respective radioisotopic ratios.

Figure 5. 87Rb/86Sr versus 87Sr/86Sr diagram for the Somerset Dam layered mafic intrusion. Only 14 of the 15 samples were
used in the isochron and age calculations. The bars
represent the two-sigma uncertainties.
Figure 5 shows 87Rb/86Sr plotted against 87Sr/86Sr for
the layered mafic intrusion, based on the data in Table
3. The error bars again represent the two-sigma
uncertainties in the data points, which were small in
absolute terms but appear quite large with respect to
the measured radioisotopic ratios, as can be seen in
Figure 5. The regression analysis using the Isoplot
program of Ludwig90 yielded a reasonable isochron
fit to 14 of the 15 samples, with a reasonably low
MSWD and a moderate probability. The troctolite
sample SDI-3A is quite clearly an outlier. The
resultant isochron age of 393170 Ma has a large
two-sigma error, because of both the two-sigma
uncertainties of the data points and the very narrow
spread in the 87Rb/86Sr ratios of the data points. This
14-point whole-rock Rb-Sr isochron age appears to
be significantly higher than the 12-point composite
whole-rock (eight samples) and mineral (four
minerals) Rb-Sr isochron age of 225.32.3 Ma
obtained by Walker.91 Even though the two-sigma error envelope of the higher Rb-Sr isochron age obtained here includes
the lower Rb-Sr isochron age of Walker and its two-sigma error envelope, the higher age does appear discordant in
absolute terms with the younger age. The initial 87Sr/86Sr ratio obtained here is much higher than that obtained from the
isochron fit of Walker, so if the anorthosite sample SDI-4E is also excluded from the regression analysis here, the resultant
isochron line is significantly steepened with a slightly lower initial 87Sr/86Sr ratio, and thus yields an even higher isochron
age of 572310 Ma (Figure 5) that is clearly discordant with the Walker Rb-Sr isochron age.
Table 3. Whole-rock Rb-Sr, Sm-Nd, Pb-Pb radioisotope data for the Somerset Dam layered mafic intrusion, near Brisbane,
Australia. (Analyst: Dr R.A. Armstrong, PRISE, Australian National University, Canberra; March-July 2002).
Cyc- Rb
Sm
87
147
Samp Rock
Sr
Rb/86S 87 86
Nd
Sm/144N 143Nd/144N
T (Ma 206Pb/204P 207Pb/204P 208Pb/204P
lical (ppm
Sr/ Sr (ppm
eNd(to) DM
le
Type
(ppm) r
(ppm) d
d
)
b
b
b
Unit )
)
SDI2A1

Oxide
gabbr
o
2

503.8
0.70292
10.15
1.92 9
0.0110 9
2.985 0
0.17782

0.512989 +6.03 686.5 18.660

15.589

38.529

SDI2A3

Oxide
gabbr
o
2

560.9
0.70294
2.71 4
0.0140 4
2.632 7.753 0.20531

0.512990 +6.05 2923.0 18.721

15.702

38.748

SDI2A4

Oxide
gabbr
o
2

596.4
0.70310
18.67
7.83 0
0.0380 3
4.878 3
0.15797

0.512956 +5.38 532.3 18.518

15.600

38.396

SDI2A6

Oxide
gabbr
o
2

622.8
0.70291
1.38 2
0.0064 6
0.366 1.679 0.13178

0.512913 +4.54 442.3 18.458

15.572

38.310

SDI2B

Anorthosit
e
3

710.0
0.70282
2.05 8
0.0083 7
0.821 3.571 0.13908

0.512939 +5.05 432.3 18.520

15.578

38.382

SDI3A

Troctolite 3

534.2
0.70274
3.77 0
0.0204 0
1.301 5.400 0.14568

0.512923 +4.74 510.2 18.538

15.578

38.422

SDI3B

Troctolite 3

498.6
0.70277
1.26 3
0.0073 7
0.724 3.265 0.13411

0.512922 +4.72 437.9 18.510

15.590

38.389

SDI3C

Troctolite 3

604.8
2.19 8
0.015

0.70285
2
0.823 3.718 0.13394

0.512906 +4.41 467.6 18.532

15.610

38.443

SDI3D

Olivin
e
gabbr
o
3

644.6
0.70291
2.85 8
0.0128 9
1.643 5.584 0.17789

0.512986 +5.97 700.6 18.632

15.598

38.550

SDI3E

Anorthosit
e
4

631.0
0.70290
3.23 3
0.0148 9
1.374 5.579 0.14893

0.512931 +4.90 516.9 18.393

15.570

38.252

SDI4A

Troctolite 4

615.3
0.70295
4.38 4
0.0206 7
1.170 5.266 0.13438

0.512902 +4.33 477.9 18.739

15.625

38.690

SDI4B

Olivin 4
e
gabbr

3.59 625.0 0.0166 0.70292 1.130 4.148 0.16474


0
5

0.512951 +5.29 657.2 18.551

15.601

38.455

o
SDI4C

Oxide
gabbr
o
4

450.6
0.70297
1.33 8
0.0085 9
1.427 4.394 0.19644

0.512986 +5.97 1454.4 18.613

15.605

38.513

SDI4D

Anorthosit
e
5

820.2
0.70304
5.06 5
0.0178 1
1.144 4.867 0.14215

0.512906 +4.41 521.3 18.861

15.604

38.611

SDI4E

Anorthosit
e
5

786.6
0.70313
16.40 8
0.0603 8
1.414 6.242 0.13694

0.512909 +4.47 479.9 18.547

15.562

38.395

Notes:
Measured, present-day 87Sr/ 86Sr ratios (2s) are normalized to 86Sr/88Sr = 0.1194
The 87Sr/86Sr ratios are reported relative to a value of 0.71020726 (2s) for the NBS standard SRM987 run with these
samples.
Measured, present-day 143Nd/144Nd ratios (2s) are normalized to 146Nd/144Nd = 0.7219
eNd (t0) refers to the present-day, calculated value relative to a depleted mantle value of 0.51268 for BCR-1
Depleted mantle model ages (TDM) calculated using present-day depleted mantle values of 143Nd/144Nd = 0.51315
and147Sm/144Nd = 0.2136
206
Pb/204Pb ratios are normalized to NBS standard SRM981 for mass fractionation, while all other ratios are calculated by a
double spiking routine.
Figure 6. 147Sm/144Nd versus 143Nd/144Nd diagram for the
Somerset Dam layered mafic intrusion, all 15 samples
being used in the isochron and age calculations. The
bars represent the two-sigma uncertainties.
Figure 6 is the 147Sm/144Nd versus 143Nd/144Nd diagram
on which all 15 data points for the Somerset Dam
layered mafic intrusion in Table 3 have been plotted.
The regression analysis on all 15 data points using the
Isoplot program of Ludwig92 yielded a goodfitting
isochron with a moderate probability and a low MSWD.
The resultant isochron age of 19948 Ma has a
moderate two-sigma error due to the narrow spread in
the radioisotopic ratios, even though the two-sigma
uncertainties for the individual data points are quite
small in absolute terms. This all suggests that this is an
acceptable Sm-Nd isochron age for this intrusion,
which because of its moderate two-sigma error
envelope would appear to be concordant with both the
excellent 15-point whole-rock K-Ar isochron age of
1748 Ma obtained in this study, and the reasonable 12point composite whole-rock and mineral Rb-Sr isochron
age of 225.32.3 Ma obtained by Walker.93 Yet if the two oxide gabbro samples SDI-2A3 and SDI-4C are excluded from
the regression analysis here, the slope of the resultant isochron line is significantly steeper, resulting in a much older 13point Sm-Nd isochron age of 25976 Ma (Figure 6).
Figure 7. 206Pb/204Pb versus 207Pb/204Pb diagram for
the Somerset Dam layered mafic intrusion. Only 13 of
the 15 samples were used in the isochron and age
calculations. The error ellipses represent the twosigma uncertainties.
Figure 7 shows206Pb/204Pb plotted against 207Pb/204Pb
for the 15 whole-rock samples from the Somerset
Dam layered mafic intrusion listed in Table 3. Only 13
of the 15 samples were used in the regression
analysis (the oxide gabbro sample SDI-2A3 obviously
being an outlier) and yielded an isochron fit with an
age of 14251000 Ma. The probability of this fit is
moderate and the MSWD is reasonably low, while the
two-sigma uncertainties in the data points (plotted as
error ellipses in Figure 7) are reasonably small in
absolute terms, yet the two-sigma error in the
isochron age is unacceptably large because of the
extremely narrow spread in the data points.
Otherwise, it would be considered discordant with all
other isochron ages obtained on this intrusion, apart
perhaps from the Rb-Sr isochron age obtained in
this study.
Discussion
The Somerset Dam layered gabbro intrusion has been interpreted as a small sub-volcanic magma chamber located at some
35 km depth, where each cyclic unit resulted from the entry of new magma into the chamber.94, 95 The question therefore
arises as to whether it is legitimate to use these radioisotopic isochron techniques to date the intrusion, when each of these
cyclic units in effect represents a distinct event in a time sequence of infillings of this sub-volcanic magma chamber.
However, even in conventional timeframe these processes can be extremely rapid, conceivably occurring over hundreds of

years or less (in conventional terms). Such time differences cannot be detected by these radioisotopic dating techniques, so
the cyclic units in this intrusion are not only comagmatic and cogenetic, but are essentially coeval, which are necessary
conditions for the application of these radioisotopic isochron techniques. In any case, this small sub-volcanic magma
chamber has been interpreted as an open system with a connection to the surface, because a closed system would have
resulted in a gradual increase in the degree of fractionation in successive cyclic units, which is observed in many large
layered mafic intrusions such as the Skaergaard and Bushveld,96, 97, 98 but which is not observed in this Somerset Dam
layered mafic intrusion. As an open system this magma chamber would have had a connection to the surface through the
overlying volcano, with space for successive magma renewals being made available by simultaneous venting periodically to
the surface of a considerable volume of volcanic lavas. Unfortunately, the postulated vent system from this magma chamber
is not now exposed, due to the top of the intrusion having been eroded away. Walker99 invoked a cyclic model involving a
gradual increase and sudden release of pressure in the magma chamber that then explains a number of features of this
layered gabbro intrusion, including the order of its layers, their inhomogeneity and lateral extent, the variations in grain sizes,
the nature of the contact between each cyclic unit, and the similarity of the rock types in the macrolayers within the cyclic
units. A gradual increase in pressure would have occurred after a blockage of the outlet conduit, followed by a sudden
decrease when the blockage cleared with the next eruption. It is therefore obvious that this cyclic sequence responsible for
macrolayering in this intrusion would not have required large amounts of time. Thus including samples of macrolayers from
several cyclic units in regression analyses to obtain isochron ages, as has been done in this study, is totally reasonable
and permissible.The isotope plots reveal extraordinary linearity within the 40K-40Ar, 87Rb-87Sr, 147Sm-143Nd, and 207Pb-206Pb204
Pb radioisotope systems. Each of the four radioisotope pairs produces a 15-point, whole-rock plot. The four data plots
(Figures 4 through 7) contain 60 data points, with 57 points following the linear trends. It is both remarkable and significant
that only three of the whole-rock data points plot slightly off the linear trends. There is no readily apparent explanation for
this, except perhaps difficulties in the analytical procedures in obtaining precision and reproducibility because of the low
levels in the samples of elements such as Pb, a comment made by Armstrong (personal communication) with respect to the
Pb radioisotopic results plotted in Figure 7. The stray data point in Figure 5 is more problematic, the only other possible
explanation being contamination from xenoliths known to occur within the intrusion.100
Giem101 has suggested the possibility that the remarkably linear radioisotope ratios in isochron plots were derived, not by
radioactive decay, but by mixing of two compositionally-different magmas. Such a model has been discussed by Snelling,
Austin, and Hoesch102 in the context of a granophyre magma (high K, Ar, Rb, Sm, Nd, and U) supposedly having been
mixed with a diabase magma (lower K, Ar, Rb, Sm, Nd, and U) in different proportions to produce the Bass Rapids sill in
Grand Canyon. Application of a similar mixing model to the Somerset Dam layered mafic intrusion would have to postulate
that its macrolayering and cyclic units were produced by the mixing of two basaltic magmas with different compositions.
These two magmas, and the various mixtures of them, would never have been in an isotopically homogeneous condition, so
that the radioisotopes in them may then have formed the mixing lines in Figures 4 through 7 without radioisotope decay
within the rocks.However, mineral isochron plots provide the data critical for testing this magma mixing model.103 Mineral
phases within any single rock should be homogeneous, because the mixing model supposes rocks crystallized from large,
locally mixed, batches of melt, and since crystallization radioisotope decay has been minor. Yet in the Bass Rapids sill
Snelling et al.104 found that the radioisotopes differed significantly between mineral phases within a diabase sample.
Similarly, preliminary data obtained from the mineral phases of a gabbro sample from the Somerset Dam layered mafic
intrusion show significant differences in their radioisotopic content. Thus, significant radioisotope decay, not mixing, is
favored as the explanation for the extraordinary linearity in the isochron plots. Furthermore, the petrographic and
geochemical data from the Somerset Dam intrusion, as from other layered mafic intrusions, argues strongly that instead
unmixing has occurred as crystal fractionation and magmatic differentiation resulted in chemical and gravitational
segregation from the initially homogeneous, molten condition of a single basaltic magma. Thus the best explanation is the
exact opposite of the mixing model.Thus the 40K-40Ar, 87Rb-87Sr, 147Sm-143Nd, and 207Pb-206Pb-204Pb radioisotopic data
obtained in this study provide strong evidence that the cyclic units in the Somerset Dam layered gabbro intrusion were all
intruded while in an isotopically mixed condition from the same homogeneous source. Indeed, primary evidence of the
original homogeneous mixture of the Ar, Sr, Nd, and Pb isotopes both in the magma source and throughout all the sampled
cyclic units of the layered intrusion is provided by the extraordinary linearity within the plots of the whole-rock radioisotopic
data (Figures 4 through 7). This evidence is also consistent with other inferences about the rapid sequential intrusion and
differentiation of the cyclic units of the intrusion derived from a chemically and isotopically homogeneous basaltic magma
source. Even though some 30% of the parental basaltic magma may have crystallized in transit before entering the magma
chamber,105 the subsequent mineralogical segregation to form the macrolayers within each cyclic unit was produced by
crystal accumulation, fractional crystallization, and crystal differentiation as the residual magmatic fluid compositions
changed.106 Therefore, at the time of intrusion each of the cyclic units and the macrolayers within them had the same Ar,
Sr, Nd, and Pb isotopic ratios that were inherited from the homogeneous parental basaltic magma source. This has to be
implicit in, and is conventionally by definition, the agreed initial condition of these coeval macrolayers and cyclic units in
order for radioisotopic dating of the intrusion to be achievable.However, the present parent-daughter radioisotopic ratios
within the whole-rock samples from this intrusion do not produce a consistent picture of the age of the formation of this
layered intrusion. Indeed, the four radioisotope systems yield whole-rock isochron ages that are discordant, so that it
appears difficult to constrain the age of the intrusion. However, the K-Ar whole-rock isochron is the most tightly constrained
with the best regression statistics, which thus suggests that 1748 Ma (two-sigma error) may be the true conventional age of
the intrusion. On the other hand, this 15-point whole-rock K-Ar isochron age is strongly discordant with the 12-point
combined whole-rock and mineral Rb-Sr isochron age of 225.32.3 Ma (two-sigma error) obtained by Walker,107 which in
turn is discordant with the 14-point whole-rock Rb-Sr isochron age of 393170 Ma (twosigma error) obtained in this study.
However, Walker used four data points (spiked and unspiked whole-rock determinations, and spiked K-feldspar and
muscovite determinations) from the one sample of alkali feldspar granite (granophyre) to dramatically improve the spread of
the data when plotted on the 87Rr/86Sr versus 87Sr/86Sr diagram, based on the assumption that the granophyre is related to
the layered gabbro intrusion as a late-stage differentiate from the same magma source. Yet the field relationships constrain
the relative ages so that the granophyre was emplaced after the quartz diorite, which in turn was emplaced after the layered
mafic intrusion.108, 109 Thus if the four granophyre data points are excluded from Walkers regression analysis because it is
unclear exactly how the granophyre is related to the layered intrusion, except that it has a younger relative age, then his
remaining eight data points do not define an Rb-Sr isochron and yield an age for the layered mafic intrusion. The scatter in
his data is large and the spread of data is so small.But what about the pooled apparent age of 224.24.8 Ma (two-sigma
error), based on 40Ar/39Ar determinations on three K-feldspar, four muscovite, one hornblende, and four biotite grains, that is
supposed to be confirmation of this 12-point Rb-Sr isochron age of 225.32.3 Ma?110 Is this remarkably close agreement
real or contrived? The reality is that the three K-feldspar and four muscovite grains analyzed came from the alkali feldspar
granite (granophyre), while the four biotite grains came from a quartz monzodiorite that like the granophyre is younger

relatively than the layered mafic intrusion, to which their relationship has not been clearly demonstrated. Thus only one
hornblende grain from an olivine gabbro within the mafic layered intrusion was used in the calculation of this pooled
apparent age, and that was only because it gave a similar age to the K-feldspar, muscovite and biotite grains from the
granophyre and quartz monzodiorite.111 Indeed, a total of seven integrated ages were obtained on hornblende grains from
just two samples from the layered mafic intrusion (the olivine gabbro and a troctolite), but these ages ranged from 55.26.7
Ma to 221.50.9 Ma (two-sigma error), which was clearly an unacceptable result. Thus neither the pooled Ar-Ar results nor
the Rb-Sr isochron of Walker yield an acceptable concordant conventional age for only the Somerset Dam layered mafic
intrusion. Therefore, it must be concluded that the 15-point whole-rock K-Ar isochron age of 1748 Ma (two-sigma error)
obtained in this study is the most reliable conventional age determination for the Somerset Dam layered mafic intrusion,
which implies that the conventional age reported in the literature should be revised upward from Late Triassic to Middle
Jurassic.This conventional age of 1748 Ma for the layered mafic intrusion therefore defines when the intrusion was
isotopically homogeneous with respect to Ar. However, the 15-point Sm-Nd isochron plot (Figure 6) is strongly linear, giving
the age for initial homogeneous Nd as 19948 Ma (two-sigma error), while the 14-point Rb-Sr isochron plot (Figure 5) is
also strongly linear, yielding the age from initial homogeneous Sr as 393170 Ma (two-sigma error). Although the
uncertainties associated with these Rb-Sr and Sm-Nd whole-rock isochrons are larger, their ages are somewhat discordant
with the K-Ar isochron age, particularly the Rb-Sr. How then could the same suite of whole rocks have Sr isotopes mix at
393170 Ma (Figure 5), but not have the Sr remixed within the whole rocks by the event that thoroughly mixed the Nd
isotopes within the same whole rocks at 19940 Ma (Figure 6), while the production of 40Ar* from radioisotopic decay of 40K
only commenced at 1748 Ma (Figure 4), assuming no initial 40A an event which presumably should have remixed both the
Sr and Nd isotopes in the same whole rocks? While the excellent 15-point K-Ar whole-rock isochron age of 1748 Ma
would seem most likely to be the true age of the initial isotopic mixing, no internally consistent age emerges from these
radioisotopic data.Austin112 has already documented four categories of discordance found in cogenetic suites of rocks, that
thus test the assumptions of radioisotopic dating. The first of these categories is when two or more wholerock isochron
ages are discordant. The whole-rock radioisotopic data reported here from the Somerset Dam layered mafic intrusion thus
call into question the assumptions of radioisotopic dating. However, as already argued, there is corroborative evidence that
the layered gabbro intrusion initially had a homogeneous mixture of the same Ar, Sr, Nd, and Pb isotopic ratios as in its
parental basaltic magma. Thus the assumption about initial conditions for the intrusion and these radioisotope systems must
be valid, in spite of the isochron age discordances. Furthermore, the evidence for open-system behavior is limited to
perhaps contamination of Sr in the Rb-Sr radioisotope system in one troctolite sample, so the closed-system assumption is
not unreasonable. Therefore, as already suggested by Snelling et al.,113 the differences in these isochron ages could be
due to changes in the decay rates of the different radioisotope systems at some time or times in the past, rather than being
caused by errors in the determination of the radioisotopic decay constants. According to Steiger and Jger,114 the U
decay constants have been measured to four significant figures, so no significant errors would be expected to occur in the
Pb-Pb systems. Steiger and Jger also recommend the decay constant for 87Rb of 1.42 10-11 that is in wide use, but
Begemann et al.115 have recommended a figure of 1.4060.008 10 -11. However, this small change in the decay constant
would not close the discordance between the Rb-Sr and either the Pb-Pb or Sm-Nd systems. Furthermore, Begemann et al.
reported that the decay constant for 147Sm has generally been agreed to have been determined to three significant figures
(6.54 10-12).Snelling et al.116 present data that indicate that the a-emitting radioisotopes (238U, 235U, and 147Sm) yield
older ages than the b-emitting radioisotopes (87Rb and 40K) when used to date the same geologic event. In other words, the
discordances in their calculated isochron ages are due to the different parent radioisotopes decaying at different rates over
the same time period since that geologic event. They found that their data are consistent with the possibility that a-decay
was accelerated more than b-decay at some time or times in the past. While the same pattern of discordances between the
a-emitting radioisotopes and the b-emitting radioisotopes is not as clearly evident in the radioisotopic whole-rock isochron
dating of the formation of the Somerset Dam layered mafic intrusion, there is comparative consistency. Perhaps the
inconsistencies are due to the youthfulness of the conventional age of the Somerset Dam layered mafic intrusion (Middle
Jurassic) compared to the conventional age of the Bass Rapids diabase sill in Grand Canyon (Upper Precambrian).
Additionally, the similarity in mineralogy of all the gabbros in the layered intrusion does not provide a large spread in the
concentrations of the parent radioisotopes, so a shorter real-time period of accelerated decay would yield an even narrower
spread in the values of the resultant daughter radioisotopes.Furthermore, Snelling et al.117 report that their data are also
consistent with the possibility that the longer the half-life of the a- or b-emitting radioisotope the more its decay has been
accelerated, relative to the other a- or bemitting radioisotopes, at some time or times in the past. This same pattern can be
seen between the b-emitting radioisotopes in the Somerset Dam layered mafic intrusion radioisotopic data. The Rb-Sr
whole-rock isochron age of 393170 Ma (Figure 5) is older than the K-Ar whole-rock isochron age of 1748 Ma (Figure
4), which is consistent with 87Rb having a longer half-life than 40K. However, the same pattern of isochron ages is not seen
for the a-emitting radioisotopes in the Somerset Dam layered mafic intrusion. Further similar studies of suitable rock units
are in progress in order to continue testing this suggested pattern in the effect of accelerated decay in the past on these
radioisotope systems.Finally, Snelling118 has extensively documented the role of geochemical processes in the earths
mantle and crust. These have resulted in the inheritance of radioisotopic ratios from the mantle and crustal magma sources
of the intrusive and volcanic rocks now found at the earths surface. From studies of present oceanic basalts it has been
determined that Sr, Nd, and Pb isotopes can be used very effectively to identify the distinctive geochemistry of mantle
reservoirs that were the sources of the magmas found today as basaltic oceanic crust and lavas on oceanic
islands.119Because mafic rocks intruded into the earths crust were also derived by partial melting of upper mantle sources,
Sr, Nd, and Pb isotopes can similarly provide valuable information on the geochemical history of the mantle during the
operation of past global tectonic processes.
Table 4. Whole-rock Rb-Sr, Sm-Nd, Pb-Pb radioisotope data for the Somerset Dam layered mafic intrusion, near Brisbane,
Australia. (Analyst: Dr R.A. Armstrong, PRISE, Australian National University, Canberra; March-July 2002).
Cyc- Rb
Sm
87
147
Samp Rock
Sr
Rb/86S 87 86
Nd
Sm/144N 143Nd/144N
T (Ma 206Pb/204P 207Pb/204P 208Pb/204P
lical (ppm
Sr/ Sr (ppm
eNd(to) DM
le
Type
(ppm) r
(ppm) d
d
)
b
b
b
Unit )
)
SDI2A1

Oxide
gabbr
o
2

503.8
0.70292
10.15
1.92 9
0.0110 9
2.985 0
0.17782

0.512989 +6.03 686.5 18.660

15.589

38.529

SDI2A3

Oxide
gabbr
o
2

560.9
0.70294
2.71 4
0.0140 4
2.632 7.753 0.20531

0.512990 +6.05 2923.0 18.721

15.702

38.748

SDI2A4

Oxide
gabbr
o
2

596.4
0.70310
18.67
7.83 0
0.0380 3
4.878 3
0.15797

0.512956 +5.38 532.3 18.518

15.600

38.396

SDI2A6

Oxide
gabbr
o
2

622.8
0.70291
1.38 2
0.0064 6
0.366 1.679 0.13178

0.512913 +4.54 442.3 18.458

15.572

38.310

SDI2B

Anorthosite 3

710.0
0.70282
2.05 8
0.0083 7
0.821 3.571 0.13908

0.512939 +5.05 432.3 18.520

15.578

38.382

SDI3A

Troctolite 3

534.2
0.70274
3.77 0
0.0204 0
1.301 5.400 0.14568

0.512923 +4.74 510.2 18.538

15.578

38.422

SDI3B

Troctolite 3

498.6
0.70277
1.26 3
0.0073 7
0.724 3.265 0.13411

0.512922 +4.72 437.9 18.510

15.590

38.389

SDI3C

Troctolite 3

604.8
2.19 8
0.015

0.70285
2
0.823 3.718 0.13394

0.512906 +4.41 467.6 18.532

15.610

38.443

SDI3D

Olivin
e
gabbr
o
3

644.6
0.70291
2.85 8
0.0128 9
1.643 5.584 0.17789

0.512986 +5.97 700.6 18.632

15.598

38.550

SDI3E

Anorthosite 4

631.0
0.70290
3.23 3
0.0148 9
1.374 5.579 0.14893

0.512931 +4.90 516.9 18.393

15.570

38.252

SDI4A

Troctolite 4

615.3
0.70295
4.38 4
0.0206 7
1.170 5.266 0.13438

0.512902 +4.33 477.9 18.739

15.625

38.690

SDI4B

Olivin
e
gabbr
o
4

625.0
0.70292
3.59 0
0.0166 5
1.130 4.148 0.16474

0.512951 +5.29 657.2 18.551

15.601

38.455

SDI4C

Oxide
gabbr
o
4

450.6
0.70297
1.33 8
0.0085 9
1.427 4.394 0.19644

0.512986 +5.97 1454.4 18.613

15.605

38.513

SDI4D

Anorthosit
e
5

820.2
0.70304
5.06 5
0.0178 1
1.144 4.867 0.14215

0.512906 +4.41 521.3 18.861

15.604

38.611

SDI4E

Anorthosit
e
5

786.6
0.70313
16.40 8
0.0603 8
1.414 6.242 0.13694

0.512909 +4.47 479.9 18.547

15.562

38.395

Notes:
Measured, present-day 87Sr/86Sr ratios (2s) are normalized to 86Sr/88Sr = 0.1194
The 87Sr/86Sr ratios are reported relative to a value of 0.71020726 (2s) for the NBS standard SRM987 run with these
samples.
Measured, present-day 143Nd/144Nd ratios (2s) are normalized to 146Nd/144Nd = 0.7219
eNd (t0) refers to the present-day, calculated value relative to a depleted mantle value of 0.51268 for BCR-1
Depleted mantle model ages (TDM) calculated using present-day depleted mantle values of 143Nd/144Nd = 0.51315
and 147Sm/144Nd = 0.2136
206
Pb/204Pb ratios are normalized to NBS standard SRM981 for mass fractionation, while all other ratios are calculated by a
double spiking routine.
Table 3 lists the relevant Sr, Nd, and Pb isotopic ratios for all 15 whole-rock samples of the macrolayers in the Somerset
Dam layered gabbro intrusion. These are plotted on isotope correlation diagrams in Figure 8, which also show the isotopic
characteristics of the defined mantle reservoirs.120 Also provided in Table 3 are the calculated epsilon Nd values, eNd(to), for
the individual samples at the present day (t o) compared to the presentday value of CHUR, the CHondrite Uniform Reservoir
in the mantle.121, 122 The magnitudes of the eNd (to) values reflect the degree of the time integrated depletion in Nd relative
to CHUR. Additionally, Table 3 lists the calculated depleted mantle Nd model ages (T DM) for each of the samples. These
are calculated with reference to the present-day isotopic values for the depleted mantle (DM) reservoir from which the
continental crust is regarded as having been largely extracted over time.123, 124,125

Figure 8. Isotope correlation diagrams on which are plotted all


15 samples from the Somerset Dam layered mafic intrusion
(after Rollinson129). The positions of the main oceanic mantle
reservoirs identified by Zindler and Hart130 are shown: DM =
depleted mantle, BSE = bulk silicate earth, EMI and EMII =
enriched mantle I and II, HIMU = high mantle U/Pb ratio,
PREMA = frequently observed PREvalent MAntle composition.
(a) 143Nd/144Nd versus87Sr/86Sr. The mantle array is defined by
many oceanic basalts and a bulk earth value for 87Sr/86Sr can
be obtained from this trend. (b) 87Sr/86Sr versus 206Pb/204Pb.
(c) 143Nd/144Nd versus 206Pb/204Pb. The yellow field is the midocean ridge basalts (MORB). The 206Pb/204Pb value of the bulk
earth differs from that of Allegre, Lewin, and Dupre.131
The sample Nd and Sr isotopic ratios plotted in Figure 8(a) are
identical to those obtained from the Somerset Dam layered
mafic intrusion by Walker.126 The depleted mantle (DM)
region where most mid-ocean ridge basalts (MORBs) plot is
connected to the bulk silicate earth (BSE) region by the mantle
array along which most basalts from ocean islands, intraoceanic island arcs and continents plot, including mantleplume-related continental flood basalts and hot-spot
continental basalts.127The gabbros from the Somerset Dam
layered intrusion plot just outside and on the edge of the
mantle array midway between the depleted mantle reservoir
and the bulk silicate earth, within the mantle reservoir known
as mantle with high U/Pb ratio (HIMU) and adjacent to the
frequently observed prevalent mantle reservoir (PREMA). This
indicates that the mantle source material from which partial
melting produced the parental basaltic magma for the layered
gabbro intrusion was already depleted by previous extraction
of crustal magmatic materials. That is, the mantle source had
already had a history of earlier extraction of basaltic magma
by partial melting. Furthermore, there is no evidence of any
significant crustal contamination of the parental basaltic
magma to the Somerset Dam layered gabbros during its
passage from its mantle source up into the upper continental
crust.
In Figures 8(b) and (c) the Somerset Dam layered gabbros
also plot in the general area of mantle depletion, again near the PREMA mantle reservoir, but this time in proximity to the
MORB field rather than within the HIMU mantle reservoir. This is consistent with the conclusions drawn from Figure 8(a),
namely, that the mantle source of the parental basaltic magma to the layered gabbros had already had a previous history of
depletion via partial melting, and that the parental basaltic magma suffered negligible crustal contamination as it ascended
from its mantle source up into the sub-volcanic magma chamber. Furthermore, the tight clustering of the data points
throughout Figure 8 confirms that the parental basaltic magma for all the cyclic units in the Somerset Dam layered gabbro
intrusion came from the same mantle source region.Consistent with this evidence of a depleted mantle source for the
parental basaltic magma are the depleted mantle Nd model ages (T DM) of the gabbro macrolayers listed in Table 3. They
generally range between 432.3 Ma and 700.5 Ma, but two oxide gabbro samples yield much greater ages of 1454.4 Ma
and 2923.0 Ma. The depleted mantle Nd model ages are interpreted as the time in the past when the magma, that
eventually formed the gabbros represented by the samples, was separated from the mantle reservoir.128 However, all the
mineralogical, geochemical and other isotopic evidence points to the parental basaltic magma for all the cyclic units of the
layered gabbro intrusion having been extracted from the same mantle source at the same time. If the layered gabbros
conventional intrusion age is 1748 Ma, but the parental basaltic magma separated from its mantle source at, for example,
432.3 Ma, then does this imply that the magma required more than 250 million years to ascend from the mantle to the upper
crust? Certainly not, if todays magma ascension rates of only years are any guide. Perhaps this vast spread of depleted
mantle Nd model ages for these layered gabbros is further evidence of accelerated radioisotopic decay rates during the
time the parental basaltic magma was being partially melted from its mantle source, and then ascending up into the upper
crust and into the sub-volcanic magma chamber to form the layered intrusion.However, if the postulated greater acceleration
of 87Rb decay compared to 40K decay was proportional to the half-lives, then a corresponding acceleration of 147Sm decay
would also be expected to have occurred. The 147Sm/40K half-lives ratio is 83, while the 87Rb/40K half-lives ratio is 37.
However, the Sm-Nd and K-Ar isochron ages are nearly equal (19948 Ma compared with 1748 Ma). The 147Sm/87Rb halflives ratio is 2.2, yet the Sm-Nd isochron age (19948 Ma) is only about half the Rb-Sr isochron age (393170 Ma). On
the other hand, the Pb-Pb isochron age of 14251000 Ma yields a similar range of ages (4252425 Ma) to the depleted
mantle Nd model ages (432.32923.0 Ma). Thus the Pb-Pb isochron age may be a radioisotopic/geochemical signature
of the mantle source of the basaltic magma, that ascended to form the layered gabbro intrusion, similar to the depleted
mantle Nd model ages.On this basis an alternative model for the radioisotopic ages of the Somerset Dam layered gabbro
intrusion would be that the mantle source of the basaltic magma originally had K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic
ratios corresponding to ages within the 4252425 Ma range. Such radioisotopic ratios could have been geochemical
design features characteristic of the original primordial creation of the earth and its mantle, or of modifications made after
initial creation of the earth as the mantle formed via differentiation.132 Subsequently, during partial melting of the mantle
source, and ascent and intrusion of the resultant basaltic magma, these radioisotopic ratios were perturbed, the K-Ar
radioisotope system being affected the most because of Ar being an inert gas, while Rb-Sr ratios were affected less than
Sm-Nd ratios.Nevertheless, it is not inconsistent for accelerated nuclear decay to have provided the heat that powered the
differentiation of the initially created earth into its core and mantle divisions and the subsequent development of the crust out
of the mantle in the early part of the Creation Week.133 Such accelerated nuclear decay could have thus produced these
geochemical design features of vast depleted mantle Nd model ages and the age spread in the Pb radioisotopes in the
mantle source area from which the parental basaltic magma was later extracted by partial melting during the Flood to

produce the Somerset Dam gabbro intrusion with inherited radioisotopic arrays. Some additional minor accelerated nuclear
decay during the Flood may have also contributed to the radioisotopic ratios now measured in the different gabbro
macrolayers of this intrusion, particularly producing the discordances between the different radioisotope systems in the
observed pattern as a result of the acceleration factor being proportional to the half-life of each radioisotopic system. Thus
accelerated nuclear decay is the favored model for explaining the radioisotopic systematics found in this layered gabbro
intrusion.In any case, at the very least, it can be concluded that the Somerset Dam layered mafic intrusion has inherited the
radioisotopic signature of its mantle source, and thus its present radioisotopic ratios do not provide the true age of the
intrusion by the conventional radioisotopic dating techniques.
Conclusions
Mineralogical, geochemical, and isotopic evidences all indicate that the cyclic units of gabbro macrolayers in the Somerset
Dam layered mafic intrusion are coeval and were derived from the same parental basaltic magma. The original endowment
of Ar, Sr, Nd, and Pb isotopes was homogeneously mixed, a necessary condition for successful conventional radioisotopic
dating, particularly by the isochron method. However, the four analyzed radioisotope systems yield discordant whole-rock
isochron ages. The K-Ar radioisotope system produces an excellent 15-point isochron age of 1748 Ma (Middle
Jurassic), which should be regarded as the revised conventional age of this layered intrusion because the previously
published ages that are discordant with it all rely on radioisotopic determinations on granitic rocks whose relationship to the
layered gabbro intrusion is unclear. It is thus concluded that the discordances between these radioisotope systems in dating
this same geologic event are likely due to changes in the decay rates of the different radioisotopic systems at some time or
times in the past. It would also appear that of the b-emitting radioisotopes, 87Rb with its longer half-life yields an older age
than 40K, suggesting 87Rb decay was accelerated more than 40K decay.The Sr, Nd, and Pb isotopes also reveal that the
parental basaltic magma of the Somerset Dam layered gabbro intrusion was partially melted from a depleted mantle source
in a single episode, and that it suffered negligible crustal contamination as it ascended up and into the sub-volcanic magma
chamber. However, the gabbros yield depleted mantle Nd model ages that supposedly indicate this separation from the
mantle source by partial melting occurred hundreds of millions of years before the resultant basaltic magma formed the
layered intrusion. This could be better explained by accelerated radioisotopic decay rates during partial melting and magma
ascension. Nevertheless, it is concluded that the Somerset Dam layered mafic intrusion has inherited the radioisotopic
signature of its mantle source, and thus its present radioisotopic ratios do not provide its true age by the conventional
radioisotopic dating techniques.
Acknowledgments
Initial work on this research was made possible by the logistical support of Answers in Genesis (Australia) for the first
sample-collecting field trip, but the project has subsequently been fully supported by the Institute for Creation Research.
Early help and discussions with Tas Walker are acknowledged, but all the work reported here is my own. All funding of the
considerable cost of the whole-rock geochemical and radioisotopic analyses was provided by donors to the RATE
(Radioisotopes and the Age of The Earth) project, whose contributions are gratefully acknowledged.
Radioisotopes in the Diabase Sill (Upper Precambrian) at Bass Rapids, Grand Canyon, Arizona
An Application and Test of the Isochron Dating Method
by Dr. Andrew A. Snelling, Dr. Steve Austin, and Bill Hoesch on March 10, 2010
Abstract
The five-point Rb-Sr whole-rock isochron age of 1.07 Ga for the diabase sill at Bass Rapids, Grand Canyon, has been
regarded for 20 years as an excellent example of the application of conventional radioisotopic dating. Initial thorough
isotopic mixing within the sill is ideal for yielding concordant whole-rock isochron and mineral isochron ages. However, our
new K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotope data from 11 whole-rock samples (eight diabase, three granophyre) and
six mineral phases separated from one of the whole-rock diabase samples yield discordant whole-rock and mineral isochron
ages. These isochron ages range from 841.5164 Ma (whole-rock K-Ar) to 1375170 Ma (mineral Sm-Nd). Although
significant discordance exists between the K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotope methods, each method appears to
yield concordant ages internally between whole rocks and minerals. Internal concordance is best illustrated by the Rb-Sr
whole rock and mineral isochron ages of 105546 Ma and 105948 Ma, respectively. It is therefore argued that only
changing radioisotope decay rates in the past could account for these discordant isochron ages for the same geologic
event. Furthermore, these data are consistent with alpha decay having been accelerated more than beta decay, and with
the longer the present half-life the greater being the acceleration factor.
Keywords: diabase, sill, Grand Canyon, potassium-argon, rubidium-strontium, samarium-neodymium, lead-lead,
radioisotopic dating, model ages, whole-rock isochron ages, mineral isochron ages, discordance, decay constants,
accelerated decay
This paper was originally published in the Proceedings of the Fifth International Conference on Creationism, pp. 269284
(2003) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh.
Introduction
The 1.1-billion-year rubidium-strontium isochron date for the Cardenas Basalt is widely regarded as perhaps the best age
obtained for Grand Canyon strata.1, 2 Similarly, the diabase sill at Bass Rapids (fig. 1) has yielded a 1.07-billion-year
rubidium-strontium isochron date,3 providing apparent confirmation of the relationship between the diabase sills and the
Cardenas Basalt flows. The Bass Rapids diabase sill provides every indication that it was well mixed isotopically when it
was intruded, even though during cooling the sill segregated mineralogically and chemically by crystal settling to produce a
granophyre on top of the diabase. Such a condition of initial thorough isotopic mixing of the original magma body followed
by rapid chemical segregation is suited to the assumptions of whole-rock and mineral isochron dating. How then are the
radioisotope daughters distributed through the granophyre and diabase, and through the mineral phases of the latter? The
various radioisotope pairs would be expected to give concordant whole-rock isochron and mineral isochron ages. However,
published potassium-argon model ages for the diabase sills (and the Cardenas Basalt) are significantly younger than their
associated rubidium-strontium isochron ages.4, 5
Geologic Setting
Mafic igneous rocks occur as sills, dikes, and flows in the thick succession of strata making up the middle Proterozoic Unkar
Group of the Grand Canyon, Arizona (fig. 1). The Unkar Group sedimentary sequence is comprised of four formationsin
ascending order, the Bass Limestone, Hakatai Shale, Shinumo Quartzite, and the Dox Sandstonewhich are overlain by
the 300 meter plus thick flow sequence of lava flows of the Cardenas Basalt.6, 7 The younger Precambrian sediments of the
Nankoweap Formation and the Chuar Group overlie this Unkar Group succession, which unconformably rests on the early
Proterozoic metamorphic and igneous crystalline basement.8, 9, 10

Fig. 1. Location of the Bass Rapids diabase sill in Grand Canyon, northern Arizona.The diabase sills and dikes are believed
to be the intrusive equivalents of the
Cardenas lava flows, but they are not
found in direct association with the
Cardenas Basalt.11, 12, 13 Thus the
relationship between them is obscure
because the direct feeders to the
flows have never been recognized
among
the
available
diabase
outcrops. The diabase sills are, in
fact, confined to the lower part of the
Unkar Group, particularly intruding
near the boundary between the Bass
Limestone and Hakatai Shale, while
the related dikes are intruded into all
the formations above the sills along
faults
that
predate,
or
are
contemporaneous with, the sills.
These mafic sills crop out in seven
locations along a 7080 km length of
the Grand Canyon (fig. 1), whereas
the Cardenas Basalt flows are
restricted to the area around Basalt
Canyon in the eastern Grand Canyon. The sills range in thickness from 20 meters (about 65 feet) near Hance Rapids in the
east to more than 200 meters (655 feet) near Tapeats Creek in the west. They are composed chiefly of medium-grained
ophitic olivine-rich diabase that is uniform in texture and mineralogy from sill to sill in the Canyon. The dikes have a similar
composition but are finer grained, as are the chilled margins of the sills. Early in-place differentiation and crystal settling in
the sills is evidenced by granophyre layers up to 10 meters thick and felsite dikes, and by layers which are richer in
olivine.The thick sill near Bass Rapids was chosen for this study because not only are there good outcrop exposures of it,
but because there is a well-defined 6 meter thick granophyre layer on top of the 85
meter thick diabase. Furthermore, more geochemical and radioisotopic analyses
have been undertaken previously on this sill than any of the other sills.
Previous Work
Noble was the first to describe the diabase sills in the Bass Canyon-Shinumo Creek
area.14, 15 Maxson16, 17 mapped the intrusive rocks of the Grand Canyon but did
not describe the diabase sills and dikes. Detailed mapping and sampling of the sills
and dikes, and the Cardenas Basalt flows, followed by petrographic examination
and chemical analysis of the samples collected, were reported by
Hendricks18, 19 and Hendricks and Lucchitta.20 They found that chemical variation
diagrams indicated a potential common parentage for the diabase in the sills and
the lower third of the basalt flows. However, the flows in the upper two-thirds of the
Cardenas Basalt sequence were found to be much more silicic than the diabase
sills, and, therefore, it was concluded that they probably were not emplaced during
the same phase of igneous activity. Nevertheless, the mineral composition of the
unaltered basalt flows in the bottom third of the sequence is similar to that of the
diabase sills, which suggested that those lavas and the diabase sills were comagmatic and probably coeval. Thus, they concluded that the basalt lavas in the top
two-thirds of the sequence were extruded after differentiation of the parent magma.
Fig. 2. Diagrammatic section through the Bass Rapids sill showing the granophyre
capping on the diabase, the contact hornfels, and the location of
samples.Paleomagnetic observations and radioisotopic age determinations by
Elston and Gromm,21 Elston and McKee,22 and Elston23 suggest that the
diabase sills may be slightly older than the Cardenas Basalt flows, perhaps as much
as 4050 Ma. The isotopic (Rb- Sr) determinations yielded an age of 109070 Ma
for the flows of the Cardenas Basalt,24, 25whereas the sill in the Shinumo CreekBass Canyon area had a five-point isochron age of 107030 Ma.26 Although these
ages are obviously identical, the flows and sills were found to have different
initial 87Sr/86Sr ratios (0.706500.0015 and 0.704200.0007 respectively) and
apparently distinctly different paleomagnetic pole positions. A K-Ar model age of
944 Ma was obtained on pyroxene extracted from a sample of the diabase sill,
presumably at Hance Rapids, by Ford, Breed, and Mitchell,27 while Elston and
McKee obtained two K-Ar model ages of 91340 Ma (for pyroxene in the diabase
sill at Hance Rapids), and 95430 Ma (for plagioclase from the diabase sill at
Tapeats Creek). Additionally, Elston and McKee reported a total fusion 40Ar/39Ar age
of 90735 Ma for pyroxene from the diabase sill near Shinumo Creek and a40Ar/39Ar
isochron age of 904100 Ma from seven-step incremental heating of a whole-rock
sample from the diabase sill at Tapeats Creek.Hendricks and Stevenson28, 29 have
summarized most of the details of the Unkar Group, including the diabase sills and
dikes, and the Cardenas Basalt flows. Subsequently, while focusing on the
Cardenas Basalt, Larson, Patterson, and Mutschler30 found that, whereas the
major-element chemistry of the diabase sills exhibited similarities and dissimilarities
with the lower-member flows of the Cardenas Basalt,31the trace and rare earth
element data from a sample of the sill at Hance Rapids show very similar variation
patterns to those in the lower-member flows of the Cardenas Basalt. Only Ti and P
contents were markedly higher in the sill, and the negative Eu anomaly for the sill
was smaller than that for the lower-member Cardenas Basalt flows. Thus Larson,
Patterson, and Mutschler suggested a common origin for the diabase of the sills

and the basalt of the lower-member flows similar to continental flood basalts, except that the higher Ti and P contents of the
diabase may indicate that the magma that fed the intrusions did not also directly feed the flows of the lower member.
Alternately, they suggested that the higher silica, Ti and P contents of the basalt flows were due either to greater crustal
contamination of the basalt magma on its passage to the earths surface, or heterogeneity in the mantle source.Finally,
Austin and Snelling32 obtained K-Ar data on five further samples, one from the diabase dike in Red Canyon adjacent to
Hance Rapids, one from the diabase sill near Hance Rapids, and three from the sill near Bass Rapids (one diabase and two
granophyre). The model ages ranged from 70315 Ma to 89520 Ma. When combined with the two samples analysed by
Elston and McKee,33 the K-Ar data yielded a K-Ar seven-point isochron age of 83752 Ma, significantly discordant with the
Rb-Sr five-point isochron age of 107030 Ma also obtained by Elston and McKee, and the Rb-Sr ten-point isochron age of
110366 Ma obtained by Larson, Patterson, and Mutschler34 for the Cardenas Basalt flows and a sample of the Hance
Rapids diabase sill. It was concluded that this discordance could not be explained by argon loss due to either resetting or
leakage. Austin and Snelling offered three alternative explanations(a) argon inheritance, (b) argon mixing, or (c) change in
the radioisotopic decay rates that affected 87Rb and 40K decay by different factors.
Sample Collection, Preparation and Analysis
Eleven whole rocks were collected from a composite section through the sill at Bass Rapids (north bank of the Colorado
River at mile 107.6108.0), the same section sampled by Hendricks and Lucchitta35 some 800 meters east of Shinumo
Creek. The samples were chosen to represent the overall petrographic variability within the complete thickness of the sill, as
depicted in Fig. 2. These whole-rock samples were prepared by clean laboratory techniques as -200-mesh powders for
chemical and isotopic analyses. Thin-sawed slices of the whole rocks were prepared as thin sections for petrographic
analysis. One of the eight diabase samples representative of the sill was crushed to -140 to +270 mesh grains, and the
various minerals within the powder were progressively concentrated by centrifugation in different heavy liquids, followed by
further cleaning using a strong magnet. Six mineral phases were thus separated from whole-rock diabase sample DI-13
(biotite, clinopyroxene, normal plagioclase, high-density plagioclase, olivine, and magnetite). X-ray diffraction analysis and
optical microscopy were used to confirm the identity and purity of the minerals concentrated.XRAL Laboratories of Don Mills,
Ontario, using XRF (x-ray fluorescence), ICP (inductively coupled plasma), and ICP-MS (mass spectrometer) methods,
performed chemical analyses of the whole-rock powders for 67 elements. Whole rocks were subjected to standard K-Ar
analysis by Geochron Laboratories, Cambridge, Massachusetts (R. Reesman, analyst) and Activation Laboratories,
Ancaster, Canada (Y. Kapusta, analyst). Whole rocks and mineral separates were analyzed by mass spectrometry for Rb-Sr,
Sm-Nd, and Pb-Pb isotopes by the University of Colorado (G. L. Farmer, analyst). The University of Colorado
measurements of Sr, Nd, and Pb isotopes were carefully calibrated by internationally recognized standards. All the resultant
isotopic data were then analyzed and isochrons plotted using the computer program called Isoplot.36
Petrography and Chemistry
The sill at Bass Rapids just east of Shinumo Creek is similar to other sills within the Unkar Group being composed of olivine
diabase, but it is capped by granophyre (fig. 2), making this sill a classic example of inplace differentiation of a basaltic
magma. The 6-m-thick granophyre consists predominantly of K feldspar (5560%) and quartz (1225%), with biotite,
plagioclase, some clinopyroxene, and titanomagnetite making up the remaining 2028%. The rock is holocrystalline, coarsegrained, and has a well-developed granophyric texture in which quartz, plagioclase, biotite, clinopyroxene, and
titanomagnetite fill interstices between the orthoclase crystals. The transition between the granophyre and diabase below
occurs over a vertical distance of less than 1 m and is a zone rich in biotite and accessory minerals.37 Apatite makes up as
much as 510% of the rock. ilmenite and sphene are prominent, and zircon with reaction halos occurs within the biotite
grains.The olivine diabase interior of the sill is medium- to coarse-grained, containing plagioclase (3045%), olivine (20
35%), clinopyroxene (1530%), titanomagnetite and ilmenite (5%), and biotite (1%), with accessory apatite and sphene. The
texture is diabasic to subophitic, although a crude alignment of feldspar laths can be seen in many places. The plagioclase
laths (composition An45-60 [4560% anorthite]) average 1.5 mm in length and are partially to completely altered to sericite.
Both normal and reverse zoning of crystals are common. Anhedral to subhedral olivine crystals up to 1 mm in diameter are
often partially altered along borders and fractures to chlorite, talc, magnetite, iddingsite, and serpentine. Fresh grain interiors
have compositions of approximately Fo80 (80% forsterite) and interference colors that suggest normal zoning. Plagioclase
laths and olivine grains are often enclosed by large, optically continuous, poikilitic clinopyroxene grains, giving the rock its
subophitic texture. The clinopyroxene is brownish-pink, non-pleochroic augite which is usually fresh. Large irregular grains
of titanomagnetite partly altered to hematite and biotite, as well as primary pleochroic brown biotite partially altered to
chlorite, occupy interstices between the plagioclase and olivine grains.The olivine concentration tends to increase towards
the center of the sill, whereas the clinopyroxene decreases. Immediately below the granophyre the diabase contains about
5% modal olivine, which increases rapidly to 2030% through the central part of the sill.38 About 15 m above the base of
the sill is an olivine-rich layer that contains about 50% modal olivine, and then the olivine content of the diabase decreases
to about 10% near the base. Hendricks39 and Hendricks and Lucchitta have suggested that this distribution of the olivine in
the sill can be explained by the process of flow differentiation, which involves the movement of early-formed olivine grains
away from the margins of the sill during flow of the intruding magma.40, 41, 42 It is envisaged that, as the magma intrudes up
through the conduit and then outward to form the sill, hydrodynamic forces concentrate toward the center of the moving
mass the olivine crystals that have formed early in the cooling history of the magma, even before the emplacement of the
sill. As the magma also moves laterally, gravity acting on the olivine crystals could have produced a gradational change in
the olivine content from the lower contact upward, while causing an abrupt change in olivine from the upper contact
downward. Once emplaced, crystallization of the remaining liquid magma within the sill would then have yielded the
remaining minerals in relatively constant proportions.43
Table 1. Whole-rock, major-element oxide and selected trace element analyses of 11 samples from the Bass Rapids sill,
Grand Canyon, northern Arizona. Sample locations are shown in Fig. 1. (Analyst: XRAL Laboratories of SGS Canada, Don
Mills, Ontario; January 1997 and February 2002).
Sample

DI-10

DI-11

DI-16

DI-17

DI-15

DI-18

DI-14

DI-13

DI-19

DI-7

DI-22

2m

3.8m

5.5m

7.5m

21m

29m

49m

59m

71m

73m

86m

SiO2 (%)

60.4

60.9

57.8

46.5

45.4

46.0

45.2

44.7

44.5

45.2

46.2

TiO2 (%)

0.903

1.18

0.03

0.16

0.25

0.16

0.17

0.18

0.17

0.16

0.21

Al2O3 (%)

14.8

15.4

14.2

11.6

13.8

15.8

14.0

14.6

12.8

15.7

14.9

Fe2O3 (%)

5.96

4.79

8.24

16.2

14.1

12.5

13.3

12.4

12.6

10.3

13.5

Position
top)

(from

MgO (%)

5.87

5.57

6.25

8.40

10.9

11.4

13.5

15.3

16.5

13.2

9.06

MnO (%)

<0.01

0.01

0.03

0.16

0.25

0.16

0.17

0.18

0.17

0.16

0.21

CaO (%)

0.27

0.09

0.65

4.16

6.89

8.13

7.68

7.80

4.16

8.48

7.15

Na2O (%)

0.57

1.47

1.63

3.15

2.06

2.18

1.86

1.87

1.71

1.93

2.23

K2O (%)

8.05

7.75

6.64

1.32

1.82

0.97

0.81

0.68

0.64

0.62

1.90

P2O5 (%)

0.03

0.02

0.32

1.23

0.51

0.26

0.28

0.09

0.27

0.19

0.43

S (%)

0.007

0.004

0.02

0.03

0.12

0.15

0.12

0.03

0.09

0.052

0.10

Lol (%)

2.85

2.90

2.95

3.15

2.2

1.75

1.85

1.2

2.7

2.4

2.45

TOTAL

99.8

100.2

100.0

99.0

99.9

100.4

100.3

100.1

100.4

99.4

100.30

Cr (ppm)

55

14

69

52

237

326

317

395

460

400

165

V (ppm)

108

116

72

200

195

142

140

88

131

110

231

Ni (ppm)

31

11

30

25

214

244

317

455

558

478

191

Co (ppm)

15

12

21

38

62

68

64

67

83

69

56

Cu (ppm)

3.4

7.9

15.6

51.4

98.3

129

63.9

30.3

55.0

31.0

101

Zn (ppm)

23.1

33.8

48.4

79.0

203

88.3

78.6

81.3

116

82.5

125

Rb (ppm)

104

106

87

23

39

23

18

16

23

14

51

Sr (ppm)

36

34

113

168

342

470

363

441

379

441

395

Zr (ppm)

228

313

349

342

160

92

116

82

116

68

161

Nb (ppm)

16

18

10

11

21

Ba (ppm)

567

633

613

327

322

184

134

167

161

239

352

Pb (ppm)

<2

<2

<2

<2

<2

<2

<2

Th (ppm)

9.4

12.0

12.2

6.4

<0.5

2.3

2.3

0.7

2.4

0.6

1.2

U (ppm)

1.9

2.5

4.0

4.8

1.7

<0.5

1.3

0.9

1.9

<0.5

0.9

La (ppm)

36.6

31.9

29.4

39.5

15.5

7.1

8.2

3.6

7.5

7.3

11.9

Ce (ppm)

80.8

68.9

65.5

94.0

38.2

17.8

20.5

8.2

18.8

16.9

29.7

Nd (ppm)

38.0

31.6

34.4

59.4

24.2

11.7

13.1

6.0

11.9

10.6

19.4

Sm (ppm)

8.1

6.2

7.9

13.3

6.3

3.1

3.3

1.5

2.9

3.2

4.6

Cl (ppm)

430

600

1050

580

1620

835

1070

363

842

954

2690

Although there is a general uniformity of the diabase throughout the sill, there are two types of textural variations, first
described by Noble.44 First, there are lumps or balls similar in mineralogy to the surrounding diabase, that is, olivine and
plagioclase with augite filling interstices. The plagioclase laths in the lumps are up to 7.5 mm in length, filling embayments in
large olivine crystals. The augite occurs as ophitic intergrowths with the plagioclase. Second, pegmatite veins consisting of
plagioclase and augite with a very similar texture are found in the upper part of the sill. These textural variations undoubtedly
represent segregation features produced during crystallization of the sill.The lower chilled margin and contact of the sill with
the underlying Hakatai Shale is covered, but is probably similar to the fine-grained chilled margins found in most of the other
sills intruding the Unkar Group in Grand Canyon. The upper contact of the sill is marked by the 6-m thick capping of
granophyre, the contact with the overlying Hakatai Shale is sharp (fig. 2), and no xenoliths of Hakatai Shale are found in the
granophyre, suggesting that it was not produced by assimilation of the shale. Instead, the transition zone between the
granophyre and the diabase beneath it in the sill suggests that the granophyre was a residual magma that floated to the top
of the sill as the diabase crystallized, so that there was little late-stage mixing of it with the diabase part of the sill.45Contact
metamorphism of the Hakatai Shale has occurred above and below the sill, the shale being altered to a knotted hornfels (fig.
2). This contact metamorphism is greater below the sill than above it. The hornfels below the sill extends for 5 m below the
contact and forms a prominent outcrop. Biotite porphyroblasts as much as 0.25 mm in size occur within 5 cm of the contact
with the sill, and at 10 cm the shale is a knotted hornfels containing porphyroblasts of andalusite and cordierite(?) that have
been replaced pseudomorphically by muscovite and green chlorite respectively. These porphyroblasts become larger and
less numerous away from the sill, reflecting a slower rate and lower density of nucleation. No recrystallization of the shale
has occurred beyond 5 m below the sill contact, while the mineralogy of the metamorphism suggests that it was of low
medium grade.The whole-rock, major element oxide and selected trace element analyses of the samples, whose locations
are shown in Fig. 2, are listed in Table 1. The major element oxide percentages are very similar to those reported by
Hendricks.46 The granophyre as expected has a much higher SiO2 content than the diabase making up the main portion of
the sill, because the granophyre contains free quartz. Similarly, the diabase has a higher Fe 2O3 and MgO content than the
granophyre because of its olivine and augite content, the MgO concentration increasing towards the central part of the sill
due to the higher olivine content there. Similarly, the higher Al 2O3 and CaO values in the center of the sill would result from
the concentration there of more calcic plagioclase. The high P 2O5 content of sample DI-17 at the top of the diabase in close
proximity to the granophyre is consistent with the apatite that is abundant in the transition zone. On a total alkalis-silica
(TAS) diagram the diabase plots in the alkali olivine basalt field,47, 48 using their own data and that of Hendricks and

Lucchitta,49 have suggested that chemically the diabase sills in the Unkar Group exhibit similarities and dissimilarities with
the lower-member flows of the Cardenas Basalt, which are commonly regarded as the extrusive equivalents of these
intrusive diabase sills.Based on selected trace element data for one sample of the Hance Rapids sill, Larson, Patterson, and
Mutschler50concluded that the variation in those data was very similar to trace element patterns for the lowermember flows
of the Cardenas Basalt. The selected trace element data in Table 1 for the Bass Rapid diabase sill are also similar.
Furthermore, the differences in trace element contents of the granophyre compared to the diabase are very obvious, and
reflect the mineralogical differences. For example, Cr, Ni, Co, Cu, and Zn are much higher in the diabase than the
granophyre, because of the olivine and trace sulfides found in the diabase that are not in the granophyre. In contrast, Ba,
Rb, La, Ce, and Nd are much higher in the granophyre than in the diabase, reflecting differences in the feldspar contents of
the two rock types, orthoclase being dominant in the granophyre, whereas plagioclase is dominant in the diabase and
contains higher Sr. The higher content of Cl in the diabase parallels the higher content of P 2O5 due to the presence of trace
apatite. Zr is as expected higher in the granophyre where zircon is more likely to be in trace amounts.
Radioisotope Results
The K-Ar analytical data and K-Ar model ages for all 11 samples are listed in Table 2. These model ages are calculated by
the standard equation of Dalrymple and Lanphere51 using the ratio of the abundances of 40Ar* (the radiogenic40Ar) to 40K
listed in Table 2. The model age method assumes no radiogenic 40Ar was present when the basaltic magma cooled to form
the diabase sill. The model ages range from 65615 Ma (million years with one-sigma error) to 105324 Ma, with the mean
age being 816 Ma (n = 11). The wide variation in these model ages is not able to be explained readily, because they are not
easily predicted by any possible sequence in the formation of the sill, such as the bottom and top of the sill cooling before
the center of the sill, or the granophyre cooling before the diabase below. Indeed, there is no recognizable pattern, except
that the model ages are discordant from one another. The mean model age for the granophyre is 863.3 Ma (n = 3),
whereas the mean model age for the diabase is 798.3 Ma (n = 8), but the model age for sample DI-14 in the center of the
diabase sill is much older at 91422 Ma. Furthermore, pairs of samples very close to one another give highly discordant
model ages, such as granophyre samples DI-10 and DI-11 which are only 1.8 m apart and yet yield model ages of 89520
Ma and 72114 Ma, and diabase samples DI-19 and DI-7 which are only 2 m apart and yet yield model ages of 86624 Ma
and 72820 Ma respectively.
Table 2. K-Ar data for the Bass Rapids diabase sill, Grand Canyon, northern Arizona. (Analysis: Dr. R. Reesman, Geochron
Laboratories, Cambridge, Massachusetts, and Dr. Y. Kapusta, Activation Laboratories, Ancaster, Canada).
Model
Age
(Ma)

Uncertainty
(Ma)
(1
sigma)

131184.6

895

20

0.000078

104692.3

721

14

4875.0

0.000095

60063.2

974

20

0.1345

2388.0

0.000056

24982.1

1053

24

86.15

0.1406

2613.0

0.000054

48777.8

656

15

0.06572

76.45

0.08596

1370.0

0.000063

21301.6

692

14

0.950

0.06567

76.1

0.08629

1486.0

0.000058

16379.3

914

22

0.958

0.948

0.05014

83.45

0.06008

1896.0

0.000032

29625.0

737

18

71 m

0.778

0.770

0.04973

75.8

0.06561

1293.0

0.000051

15098.0

866

24

DI-7

73 m

0.754

0.747

0.03893

80.0

0.04866

1623.3

0.000030

24900.0

728

20

DI-22

86 m

2.157

2.135

0.11107

8.79

1.26359

3367.0

0.000375

5693.3

740

22.4

Sampl
e

Position K2O (wt40K


(from top) %)
(ppm)

40

Ar*
(ppm)

40

Ar*
(%)

Total40Ar
(ppm)

40

DI-10

2m

8.61

8.527

0.5737

96.4

0.5951

9155.4

0.000065

DI-11

3.8 m

8.245

8.166

0.4206

94.3

0.4460

5717.9

DI-16

5.5 m

5.764

5.706

0.4281

92.85

0.4611

DI-17

7.5 m

1.413

1.399

0.1162

86.4

DI-15

21 m

2.661

2.634

0.1211

DI-18

29 m

1.356

1.342

DI-14

49 m

0.959

DI-13

59 m

DI-19

36

Ar/ Ar

36

Ar (ppm)

40

36

K/ Ar

Fig. 3. 40K versus 40Ar in the Bass Rapids diabase sill, all 11
samples being used in the isochron and age calculations. The
bars represent the two-sigma uncertainties.
Fig. 3 is the 40K versus 40Ar diagram for the Bass Rapids diabase
sill. The error bars plotted with the data are the estimated twosigma uncertainties, and the strong linear trend that is apparent
is plotted as an isochron using the Isoplot program of
Ludwig52 that utilizes the least-squares linear regression
method of York.53 All 11 samples were included in the
regression calculation, although the assigned two-sigma errors
were large. The isochron age calculated from the slope of the
line is 841.5164 Ma (two-sigma error). The initial 40Ar is zero, so
this is consistent with the assumption of zero 40Ar* in the model
age technique. This K-Ar isochron age is discordant with the
published five-point Rb-Sr whole-rock isochron age for the sill of
107030 Ma.54 Note also that the slope of the line is heavily
influenced by the three granophyre data points with their high K contents due to their contained orthoclase. Because all the
samples are cogenetic it was important that all are included in the calculation, even though it leads to a large two-sigma
uncertainty in the isochron age. This large uncertainty must be due to more than analytical errors, and is thus indicative of
the minor hydrothermal alteration present (plagioclase altered to sericite) and perhaps some contamination of the
granophyre from the hornfels wall-rock during contact metamorphism.

Fig. 4. 40K/36Ar versus 40Ar/36Ar in the Bass Rapids sill, all 11


samples being used in the isochron and age calculations. The
bars show the two-sigma uncertainties.
Fig. 4 shows 40K/36Ar plotted against 40Ar/36Ar for the sill, based
on the data in Table 2. The error bars again represent the twosigma uncertainties in the data points, which again were large,
with ten of the 11 samples included in the regression analysis.
The calculated isochron age is therefore 844.7179 Ma with an
initial 40Ar/36Ar value of 249. This is much less than the present
atmospheric 40Ar/36Ar value of 295.5, and suggests the possibility
of a small Ar loss or that the regression line needs to be
appropriately adjusted. This would reduce the isochron age, and
make it even more discordant with the published Rb-Sr isochron
age, even though it would still be concordant with the K-Ar
isochron age determined here and shown in Fig. 3. Alternately,
the low 40Ar/36Ar value could indicate incorporation into the
basaltic magma of primitive argon thus inherited from its mantle
source.55The whole-rock Rb-Sr, Sm-Nd, and Pb-Pb
radioisotopic data for all 11 samples from the sill are
listed in Table 3. As anticipated, the radioisotopic
ratios in the three granophyre samples are distinctly
different to those obtained from the eight diabase
samples. This reflects the major and trace element
differences between these two rock types and their
different mineralogies, the granophyre having a much
higher K2O content (table 2) than the diabase
because of the abundant orthoclase in it. Thus, the
Rb content of the granophyre is higher than that of the
diabase, whereas the Sr content is higher in the
diabase because it partitions with the Ca in
plagioclase. The generally higher rare earth element
and Pb contents of the granophyre likewise leads to
significantly different radioisotopic ratios in the
granophyre compared to the diabase. These
differences are ideal for plotting of isochrons because
of the larger spreads in the radioisotopic ratios.
Fig. 5. 87Sr/86Sr versus 87Sr/86Sr diagram for the Bass
Rapids diabase sill, all 11 whole-rock samples being used in the isochron and age calculations. The bars represent the twosigma uncertainties.
Fig. 5 shows 87Rb/86Sr plotted against 87Sr/86Sr for the sill, based on the data in Table 3. The error bars again represent the
two-sigma uncertainties in the data points, which were small. The regression analysis using the Isoplot program of
Ludwig56 yielded an excellent-fitting isochron with a high probability and low MSWD (mean square of weighted deviatesa
measure of the ratio of the observed scatter of the data points from the best-fit line to the expected scatter from the assigned
errors and error correlations). The resultant isochron age of 105546 Ma is only marginally less than the five-point Rb-Sr
isochron age of 107030 Ma obtained by Elston and McKee.57 At 0.7043, the initial 87Sr/86Sr for this isochron is virtually
identical to the value of 0.704200.0007 obtained by Elston and McKee. Significantly, when we added the Rb-Sr data for the
five Elston and McKee samples to that of our 11 samples the resulting regression analysis yielded an even better 16-point
isochron with a higher probability (0.86) from the same two-sigma uncertainties for each of the data points. The isochron
age of 105544 Ma is identical, as is the initial 87Sr/86Sr. Nevertheless, the uncertainty of 44 Ma is higher than the 30 Ma
obtained by Elston and McKee, but a lot of this uncertainty is due to the poorer fit of the two high Rb granophyre samples
DI-10 and DI-11.
Table 3. Whole-rock Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic data for the Bass Rapids diabase sill, Grand Canyon, northern
Arizona. (Analyst: Associate Professor G. L. Farmer, University of Colorado) at Boulder).
Positio
147
206
Sam
Rb
Sr
Sm
Nd
Sm144N 143Nd144N
Pb204P 207Pb204P 208Pb204P
87
n (from
Rb86Sr 87Sr 86Sr
eNd(tO) TDM(Ga)
ple
(ppm) (ppm)
(ppm) (ppm) d
d
b
b
b
top)
DI10

2m

101.5 35

8.4302 0.83703 8.25

37.8

0.132

0.512070 -11.08

22.948

15.933

42.233

DI11

3.8m

95.8

33

8.439

32.59 0.150

0.511992 -12.60

25.432

16.135

45.201

DI16

5.5m

82.8

106

0.74129
2.2643 7
7.98

36.3

0.1329

0.512084 -10.81 1.80

21.923

15.878

41.482

DI17

7.5m

24.2

154

0.71332
0.4539 9
14.55 62.4

0.1411

0.512391 -4.82 1.36

19.368

15.702

38.574

DI15

21m

38.5

312

0.70913
0.3568 9
6.0

25.0

0.1456

0.512441 -3.84 1.34

17.255

15.475

36.981

DI18

29m

21.8

422

0.70635
0.1492 9
3.08

12.5

0.1495

0.512458 -3.51 1.38

17.358

15.494

37.003

DI-

49m

11.7

328

0.1026 0.70546 3.37

13.8

0.1473

0.512443 -3.80 1.37

18.355

15.561

38.154

0.82481 8.11

14

DI13

59m

15.4

383

0.70481
0.1165 8
2.65

10.8

0.1480

0.512438 -3.90 1.40

17.699

15.510

37.353

DI19

71m

16.9

329

0.70501
0.1487 9
3.60

15.0

0.1452

0.512466 -3.36 1.28

17.260

15.452

36.854

DI-7

73m

11.5

347

0.70450
0.0959 2
1.64

6.04

0.164

0.512554 -1.64

17.407

15.480

37.005

DI22

86m

51.3

371

0.71130
0.4005 6
5.13

21.1

0.1471

0.512446 -3.75 1.36

19.429

15.687

38.711

Fig. 6. 147Sm/144Nd versus 143Nd/144Nd diagram for all 11 wholerock samples of the Bass Rapids diabase sill. The bars
represent the two-sigma uncertainties. The Sm- Nd mineral
isochron of Fig. 9 is shown for comparison. The eight diabase
samples plot on the mineral isochron whereas the three
granophyre samples (DI-10, DI-11, DI-16) do not, suggesting
they have been contaminated from the hornfels wall-rock.
Fig. 6 is the 147Sm/144Nd versus 143Nd/144Nd diagram for the Bass
Rapids diabase sill using the data in Table 3. These whole-rock
samples are tightly grouped showing that little variation within
the Sm-Nd system exists within the whole rocks. No Sm-Nd age
information can be derived from the 11 whole-rock samples. All
eight diabase samples do suggest a line in Fig. 6, but
Isoplot58 attaches little age significance to it. The Sm-Nd
mineral isochron (see below) is plotted in Fig. 6 and appears to
pass through the eight whole-rock diabases. The three
granophyre samples (DI-10, DI-11, and DI-16) plot on the
diagram in a random scatter widely separated from any
apparent relationship with the eight diabase samples, which is suggestive of contamination from the overlying hornfels wallrock, perhaps by some assimilation of Nd.
Fig. 7 shows 206Pb/204Pb plotted against 207Pb/204Pb for the whole-rock samples from the Bass Rapids diabase sill. All 11
samples were used in the regression analysis and yielded an isochron fit with an age of 1249140 Ma, with a high
probability and a low MSWD. The relatively large two-sigma uncertainty in the resultant age is largely due to the size of the
two-sigma errors in the data points represented by the ellipses on the diagram. On the other hand, the three granophyre
samples give a greater spread to the data which otherwise yields good regression statistics for the isochron.
Fig. 7. 206Pb/204Pb versus 207Pb/204Pb diagram for the Bass
Rapids diabase sill, using all 11 whole-rock samples in the
isochron and age calculations. The error ellipses represent
the two-sigma uncertainties
Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic data for the six
minerals separated from the whole-rock diabase sample
DI-13 are listed in Table 4. The strong partitioning of the
relevant trace elements into the different mineral phases is
evident as expected. For example, Rb is high in the biotite,
whereas the Sr is high in the plagioclase. This was
expected to provide a good spread in the radioisotopic
data, and improve the statistics of the isochron fits.
Fig. 8 is the 87Rb/86Sr versus 87Sr/86Sr diagram for the six
mineral fractions from sample DI-13, plus the whole rock.
The regression analysis using the Isoplot program of
Ludwig59 produced an excellent isochron fit, with a very
high probability and low MSWD. The resultant isochron
age is 105948 Ma, the two-sigma uncertainty being
moderate because the two-sigma error bars on the data
points were also moderate. This mineral isochron age is, of course, totally concordant with the whole-rock Rb-Sr isochron
age (105546 Ma, Fig. 5), but at 0.70302 the initial87Sr/86Sr is marginally lower than that for the whole-rock isochron.
Table 4. Mineral Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic data for diabase samples DI-13 from the Bass Rapids sill, Grand
Canyon, northern Arizona. (Analyst: Associate Professor G. L. Farmer, University of Colorado at Boulder.)
147
206
Sm
Nd
Sm/144N 143Nd/144N
Pb/204P 207Pb/204P 208Pb/204P
eNd(tO) TDM(Ga)
(ppm) (ppm) d
d
b
b
b

Rb
Sr
(ppm) (ppm)

87

Wholerock

15.4

383

0.1165

0.70481
8
2.65

10.8

0.1480

0.512438 -3.90 1.40

17.699

15.510

37.353

Biotite

92.2

187

0.72474
1.4294 6
2.37

11.3

0.1266

0.512225 -8.06 1.43

17.457

15.486

37.150

Clinopyr
oxene

0.74

32.8

0.70416
0.0651 6
5.19

14.3

0.2190

0.512922 +5.54

17.463

15.477

37.191

Plagiocla
se
28.5

784

0.70476
0.1050 9
0.63

9.4

0.0408

0.511448 -23.21 1.40

17.194

15.471

36.913

Fraction

Rb/86Sr 87Sr/ 86Sr

0.36

7.62

0.70475
0.1348 2
1.15

5.9

0.1189

0.512097 -10.55 1.52

17.384

15.536

37.171

Highdensity
Plagiocla
se
27.3

634

0.70496
0.1244 7
8.96

39.1

0.1385

0.512403 -4.58 1.29

17.085

15.461

36.791

Magnetit
e
1.01

7.98

0.70840
0.3669 8
0.88

27.1

0.0196

0.511161 -28.81 1.48

18.423

15.640

38.035

Olivine

Fig. 8. 87Rb/86Sr versus 87Sr/86Sr diagram for six mineral fractions


from diabase samples DI-13 (plus the wholerock) from the Bass
Rapids diabase sill. All seven data points were used in the
isochron and age calculations, and the bars represent the twosigma uncertainties.
Fig. 9 shows the 147Sm/144Nd versus 143Nd/144Nd diagram for the
six mineral fractions, plus the whole rock, of sample DI-13 from
the Bass Rapids diabase sill. The diagram shows an excellent
large spread amongst the seven data points, from magnetite
with
the
lowest147Sm/144Nd
ratio
through
to
the
highest 147Sm/144Nd ratio in the clinopyroxene. The regression
analysis again produced an excellent isochron fit with a good
probability and a low MSWD. The resultant mineral isochron age
is 1375170 Ma. The relatively large two-sigma uncertainty in
the isochron age is of course due to the relatively large twosigma error bars for each of the data points, and to the scatter of
some of the data points (for example, the whole rock and
clinopyroxene) either side of the isochron (the line of best
fit).
Finally,
Fig.
10
shows 206Pb/204Pb
plotted
against 207Pb/204Pb for the six mineral fractions from
sample DI-13, plus the whole rock, from the Bass Rapids
diabase sill. These data are tightly grouped, showing that
uranium is not strongly partitioned within the mineral
phases within a single rock. Isoplot60 shows no significant
Pb- Pb age information can be derived from these seven
data points. However, the Pb-Pb whole-rock isochron of
Fig. 7 when plotted on Fig. 10 passes through the
twosigma error ellipses of all seven data points. Thus, the
Pb-Pb mineral data appear to be concordant with the
1249140 Ma whole-rock Pb-Pb isochron.
Discussion
Fig. 9. 147Sm/144Nd versus 143Nd/144Nd diagram for six
mineral fractions from diabase sample DI-13 (plus the
whole-rock) from the Bass Rapids diabase sill. All seven
points were used in the isochron and age calculations, and
the bars represent the two-sigma uncertainties.Isotope
plots reveal extraordinary linearity within the 40K-40Ar, 87Rb-87Sr, 147Sm-143Nd, and 207Pb-206Pb-204Pb radioisotope systems.
Each of the four radioisotope pairs produces an elevenpoint, whole-rock plot. The five data plots (Figs. 3 through
7) contain 55 data points with 51 points following linear
trends. Remarkably, only four of the whole-rock data
points plot significantly off the linear trends. Three of the
data points plotting significantly off the line in Fig. 6 are
easily explained by the granophyres assimilation of
neodymium due to contamination from the adjoining
hornfels just above the sill (Fig. 2). Certain hydrothermal
conditions have been shown to cause rare earth element
mobility in rhyolite and granite,61 the Nd isotopes being
perturbed during hydrothermal alteration. That such
hydrothermal alteration of the granophyre in the sill has
occurred during contact metamorphism with the overlying
shale is evidenced by plagioclase altered to sericite and
biotite altered to chlorite.Some creationists may want to
consider the possibility that the remarkably linear isotope
ratios within the diabase and granophyre were derived, not
by radioisotope decay, but by mixing of two different
magmas. Such a model has been proposed by
Giem62 and discussed by Austin and Snelling.63 We might suppose the sill at Bass Rapids was formed from a granophyre
magma (higher K, Ar, Rb, Sm, Nd, and U) combined with different proportions of a diabase magma (lower K, Ar, Rb, Sm, Nd,
and U). The mixing model supposes these two magmas, and their various magma mixtures, were never in an isotopically
homogeneous condition. Isotopes from these two magma types may then have formed the mixing lines in Figs. 3 through 7
without radioisotope decay within the rocks. As pointed out by Austin,64 mineral isochron plots provide the data critical for
testing the magma mixing model. Mineral phases within any single rock should be homogeneous, because the mixing model
supposes rocks crystallized from large, locally mixed, batches of melt, and since crystallization radioisotope decay has been

minor. For the Bass Rapids diabase sill, however, radioisotopes differ significantly between mineral phases within diabase
sample DI-13. Thus, significant radioisotope decay, not mixing, is the favored explanation of the extraordinary linearity.
Furthermore, the petrographic and geochemical data from the Bass Rapids sill argues that unmixing has occurred
(chemical and gravitational segregation from an initially homogeneous, molten condition). The best explanation is exactly
opposite the mixing model.
Fig. 10. 206Pb/204Pb versus 207Pb/204Pb diagram for six mineral fractions from diabase sample DI-13 (plus the whole-rock)
from the Bass Rapids diabase sill. The error ellipses represent the two sigma uncertainties, and the Pb-Pb whole-rock
isochron of Fig. 7 is shown for comparison.
40
K-40Ar, 87Rb-87Sr, 147Sm-143Nd, and 207Pb-206Pb-204Pb radioisotope data provide strong evidence that the Bass Rapids
diabase sill was intruded while in an isotopically mixed, homogeneous condition. Initially, the sill was chemically and
isotopically homogeneous when the basaltic magma was intruded rapidly into the Hakatai Shale. The subsequent
mineralogical segregation within the sill was produced by flow differentiation and gravitational settling, resulting in olivine
diabase overlain by granophyre. At that time of intrusion the different parts of the newly formed sill had the same Ar, Sr, Nd,
and Pb isotopic ratios. This must be, and is conventionally by definition, the agreed initial condition in order for radioisotopic
dating of the diabase to be achievable.What then can be said about the present isotopic ratios within whole rocks and
minerals of the sill? Do parentdaughter radioisotope ratios produce a consistent picture of the age of the sill? Two
significant Rb-Sr wholerock and mineral isochron plots (Figs. 5 and 8) appear to constrain the age of the diabase sill to
105546 Ma (two-sigma error). That is the currently accepted age of this Grand Canyon diabase according to Elston and
McKee65 and Larson, Patterson, and Mutschler66 when the diabase sill was isotopically homogeneous with respect to Sr.
However, the Sm-Nd mineral isochron plot (Fig. 9) is strongly linear giving the age from initial homogeneous Nd as
1375170 Ma (two-sigma error). Although the uncertainty associated with this Sm-Nd mineral isochron is larger, its age is
clearly discordant with Rb-Sr. How could the suite of minerals in sample DI-13 have Nd isotopes mixed at 1375170 Ma
(Fig. 9) but not have Nd remixed within the minerals by the event that thoroughly mixed the Sr isotopes within the minerals
at 105948 Ma (Fig. 8)?The Pb-Pb whole-rock isochron plot (Fig. 7) gives the age from initial homogeneous Pb as
1249140 Ma (two sigma-error), again discordant with Rb-Sr. How could the suite of whole rocks within the sill have Pb
isotopes mixed at 1249140 Ma (Fig. 7) but not have Pb remixed within the rocks by the event that thoroughly mixed the Sr
isotopes at 105546 Ma (Fig. 5)? Both of these ages are discordant with the K-Ar whole-rock isochron age (Fig. 3) of
841.5164 Ma assuming no initial 40Ar. Which of these is the true age of the initial isotopic mixing? No internally consistent
age emerges from these data.Indeed, Austin67 has already documented that, when the mineral isochron method is applied
as a test of the assumptions of radioisotopic dating, discordances inevitably result. According to Austin four categories of
discordance are found in cogenetic suites of rocks(a) two or more discordant whole-rock isochron ages, (b) a whole-rock
isochron age older than the associated mineral isochron ages, (c) two or more discordant mineral isochrons from the same
rock, and (d) a whole-rock isochron age younger than the associated mineral isochron ages. Our radioisotope data from the
Bass Rapids diabase sill exhibit all four categories of isochron discordance. Thus the assumptions of radioisotopic dating
must be questioned.However, as already argued, there is corroborative evidence that the sill initially had a homogeneous
mixture of the same Ar, Sr, Nd, and Pb isotopic ratios. Thus the assumption about the initial conditions for the sill and these
radioisotope systems must be valid. Furthermore, the evidence for open-system behavior is limited to the neodymium in the
Sm-Nd radioisotope system in the granophyre whole-rock samples (perturbed by contamination from the overlying hornfels
wall-rock), so the closed-system assumption in this instance is not unreasonable. Therefore, could the differences in the
calculated ages above be caused by errors in determining the constants of radioisotope decay? According to Steiger and
Jger,68 uranium decay constants are measured reproducibly to four significant figures. Therefore, no significant error
occurs with Pb-Pb isochrons. Steiger and Jger recommended the decay constant for87Rb of 1.42 10-11 that is in wide use,
but Begemann et al.69 recommend 1.4060.008 10-11. However, this small change would not close the discordance
between the Rb-Sr and either the Pb-Pb or Sm-Nd systems. According to Begemann et al.,70 researchers generally agree
that the decay constant for 147Sm has been determined to three significant figures (6.54 10 -12).Our data indicate that the
alpha emitters (238U, 235U, and 147Sm) have yielded older ages than the beta emitters ( 87Rb and 40K) when used to date the
same geologic event, that is, the intrusion of the Bass Rapids diabase sill. A logical explanation of these data is that the
radioisotope decay of the various parent isotopes has not always proceeded at the rates described by modern decay
constants, the discordances being due to the different parent radioisotopes decaying at different rates over the same time
period since the formation of the sill. In other words, the decay of these parent radioisotopes was accelerated by different
amounts. Thus our data are consistent with the possibility that alpha decay was accelerated more than beta decay at some
time or times in the past.Furthermore, our data also show that there is a correlation between the present radioactive decay
constants for these alpha and beta emitters and the ages they have yielded for this same geologic event. Of the alpha
emitters, 147Sm has the smallest decay constant (and thus the longest half-life) and it yielded the oldest age, a mineral
isochron age of 1375170 Ma (Fig. 9), compared to the Pb-Pb whole-rock isochron age of 1249140 Ma (Fig. 7). Similarly,
of the beta emitters, 87Rb has the smaller decay constant (and thus the longer half-life) and it yielded the older ages, a
whole-rock isochron age of 105546 Ma (Fig. 5) and a mineral isochron age of 105948 Ma (Fig. 8), compared to the K-Ar
whole-rock isochron age of 841.5164 Ma (Fig. 3). Thus our data are also consistent with the possibility that the longer the
half-life of the alpha or beta emitter the more its decay has been accelerated, relative to the other alpha or beta emitters, at
some time or times in the past.It is recommended that further similar studies of suitable rock units be undertaken to confirm
these findings.
Conclusion
The distributions of radioisotopes and their daughter isotopes in the middle Proterozoic Bass Rapids diabase sill in Grand
Canyon reveal a glaring problem with the assumptions of conventional radioisotopic dating. Even though several lines of
evidence confirm that the daughter isotopes were homogeneously mixed when the basaltic magma was intruded initially to
form the sill, the four analyzed radioisotope systems yield discordant whole-rock isochron ages for this geologic event.
Although significant discordance exists between the K-Ar, Rb-Sr, Sm-Nd, and Pb-Pb radioisotope methods, each method
appears to yield concordant ages internally between whole rocks and minerals. Internal concordance is best illustrated by
the Rb-Sr whole-rock and mineral isochron ages of 105546 Ma and 105948 Ma, respectively. Furthermore, the only
evidence of open-system behavior is contamination of the whole-rock Sm-Nd radioisotope system in the granophyre
immediately adjacent to the overlying hornfels wall-rock. Therefore, it is concluded that it is the constant decay rates
assumption of conventional radioisotopic dating that is potentially invalid, and thus changing decay rates in the past could
account for the demonstrated discordances between the resultant isochron ages. Furthermore, our data are consistent with
the possibilities that, at some time or times in the past, decay of the alpha emitters ( 238U, 235U, and 147Sm) was accelerated
more than decay of the beta emitters ( 87Rb and 40K), and the longer the present half-life of the alpha or beta emitter the more
its decay was accelerated relative to the other alpha or beta emitters.

Acknowledgments
Grand Canyon National Park provided special use permits allowing access to the remote site and granting permission to
collect rock samples. Private donors provided financial support through the RATE project, and the project preceding RATE,
both administrated at the Institute for Creation Research.

Discordant Potassium-Argon Model and Isochron Ages for Cardenas Basalt (Middle Proterozoic) and Associated
Diabase of Eastern Grand Canyon, Arizona
by Dr. Andrew A. Snelling and Dr. Steve Austin on February 3, 2010
Abstract
For more than twenty years it has been
known that the Rb-Sr and K-Ar systems
give discordant ages for Cardenas
Basalt
and
associated
Proterozoic
diabase sills and dikes of Grand Canyon.
Thirteen new K-Ar analyses of Proterozoic
mafic rocks of Grand Canyon are added to
nine published K-Ar analyses. We report a
new fourteen-point K-Ar isochron age of 516 30 Ma which is strongly discordant with the published Rb-Sr isochron age
of 1.07 0.07 Ga for Cardenas Basalt. By more than doubling the K-Ar data set we can test explanations for why the
discordance exists. Advocates of the Rb-Sr isochron, recognizing the strong geochemical similarity of rubidium and
potassium, have not argued for significant potassium addition to these rocks. Addition of potassium during alteration of
these rocks would explain the anomously young K-Ar age, but it would also add rubidium and invalidate the Rb-Sr isochron
age. Instead, advocates of the Rb-Sr isochron have argued only for significant argon loss. Two argon loss models (episodic
loss and continuous loss) are tested in an attempt to explain why these altered rocks have about half the 40Ar required by the
conventional Rb-Sr interpretation. Both argon loss models, although attempting to maintain the assumptions of conventional
geochronology, fail to explain the data, especially the new data we offer. Three models are proposed as alternatives to
argon loss models, but these invalidate using the K-Ar system as conventional geochronology would assume.
Shop Now
Keywords: basalt, diabase, Grand Canyon, potassium-argon dating, model ages, isochron ages, discordance, argon
leakage models, inheritance, mixing, change of decay
This paper was originally published in the Proceedings of the Fourth International Conference on Creationism, pp. 35-51
(1998) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh.
IntroductionThe Cardenas Basalt (fig. 1) is a succession of lava flows over 300-m-thick occurring deep within a thick
succession of strata in the eastern Grand Canyon, Arizona. The 1.1-billion-year Rb-Sr isochron date for the Cardenas Basalt
is widely regarded by geologists as the best age yet obtained for Grand Canyon strata.1, 2 That age agrees with what
most geologists have come to believe about other Precambrian strata and the pervasive mafic intrusive and extrusive
magmatism of the Middle Proterozoic of the southwestern United States.3, 4, 5 Although we might suppose that
radioisotopes could definitely date Cardenas Basalt, a significant problem has arisen. Published potassium-argon (K-Ar)
ages for Cardenas Basalt and Precambrian
diabase are significantly younger than their
associated
rubidium-strontium
(Rb-Sr)
ages.6 Why are K-Ar dates significantly younger
than the accepted Rb-Sr age? We seek to
examine the potassium and argon systematics of
Cardenas
Basalt
along
with
associated
Precambrian diabase dikes and sills. We seek to
test theories which explain why K-Ar ages are
significantly younger than the widely accepted RbSr isochron age.
Geology of Sills, Dikes, and Flows
Mafic igneous rocks occur as sills, dikes, and flows
in the Upper Precambrian (Proterozoic) Unkar
Group of the Grand Canyon, Arizona (figs. 1 and 2).
The extrusive rocks form a 300-m-thick flow
sequence called the Cardenas Basalt which rests
conformably on the Dox Formation in the middle of
the 4000-m-thick Proterozoic Grand Canyon
Supergroup
sedimentary
sequence.7 This
sedimentary sequence rests nonconformably on
the older metamorphic and igneous rocks, including
the Vishnu Schist and Zoroaster Granite.
Fig. 1. Stylized block diagram showing the
stratigraphic position of Cardenas Basalt and
Precambrian Diabase in the eastern Grand
Canyon.
In the type section at Basalt Canyon8, 9 the
Cardenas Basalt consists of a 100-m-thick lower
member composed of about six, coarsely ophitic
flows of olivine basalt which vary in thickness from
about 3 to 25 m. Beds of siltstone and sandstone 1.5 to 3 m thick occur between some of the flows. All flows possess

vesicular tops and bottoms and massive to columnar-jointed central portions. Typically this lower member is poorly exposed,
weathering to spheroidal masses surrounded by granular debris. Before alteration, this mediumgrained basalt was similar in
texture and mineralogy to the sills and dikes, while petrologic and chemical data suggest to some that this lower member is
a spilitic hyaloclastite which is the altered effusive equivalent of those mafic intrusives.10 However, the spheroidal masses
may simply have resulted from onion-skin weathering rather than being suggestive of pillow structures, and therefore, the
field evidence could indicate subaerial extrusion of these lower member basalts in a series of low-viscosity pahoehoe flows.
In contrast, the 200-m-thick upper member of the Cardenas Basalt comprises four to six, aphyric, intersertal to intergranular
flows that change sequentially from basaltic andesite to basalt and then back to basaltic andesite upward through the
section. The dramatic change to the resistant, finer-grained flows of the upper member coincides with the abrupt change in
magma chemistry. Individual flows vary in thickness from about 20 to 50 m, each flow being separated from adjacent flows
by laterally persistent siltstone and sandstone beds which generally range in thickness from 0.3 to 3 m. Amygdaloidal zones,
from 2 to 5 m thick, are common near the bases and tops of the flows; scoriaceous flow breccia is conspicuous at the tops
of several of them. Some of the flows exhibit crude columnar jointing in their middle to upper portions, whereas none of the
flows exhibits any evidence of interaction with water.

Fig. 2. Location of Cardenas Basalt and Precambrian diabase in eastern Grand Canyon.
The sills are not found in direct association with the Cardenas Basalt, but are confined to the lower part of the Unkar Group,
particularly near the boundary between the Bass Formation and the Hakatai Shale, while the related dikes are intruded
above the sills along faults that predate or are contemporaneous with the sills. These mafic sills crop out in seven locations
along a 7080 km length of the Grand Canyon (fig. 2) and range in thickness from 23 to 300 m. They are composed chiefly
of medium-grained ophitic diabase. The dikes have a similar composition but are finer grained, as are the chilled margins of
the sills. Early differentiation and crystal settling in the sills is evidenced by granophyre layers up to 10-m thick and felsite
dikes, and by olivinerich layers. The relationship of these intrusives to the Cardenas flows remains obscure because direct
feeders to the flows and sills have never been recognized.
Sample Selection for K-Ar Analysis
The published geologic literature contains seven K-Ar analyses of Cardenas Basalt and two K-Ar analyses of Precambrian
diabase.11, 12, 13, 14, 15 Descriptions of these rocks are found in the appendix of this paper and the K-Ar data have been
collected into Table 1. General locations are shown in Fig. 2. Our goal was to more than double the K-Ar data for these
Precambrian rocks of Grand Canyon, allowing a better and more comprehensive interpretation of K-Ar ages. Thus,
collection of new samples of Cardenas Basalt and Precambrian diabase had three purposes:
to complement and enlarge the published isotopic data set by sampling basalt and diabase in areas or stratigraphic levels
not already described,
to select a variety of rock types, especially those with different potassium concentrations, in order to illustrate the variation of
argon as a function of potassium,
to maintain precise geographic and stratigraphic data on samples so that double-checking and further sampling could be
conducted.
Complete location data for the nine published K-Ar analyses and the 13 new analyses are found in the appendix. The new
K-Ar data have been combined with the published data into table 1. Of the 13 new analyses, eight are from Cardenas Basalt
and five are from Precambrian diabase.A total of 19 new samples of Cardenas Basalt were collected. All new samples,
except two, are from Basalt Canyon, one of the best and most studied outcrops. The Basalt Canyon section has poor
exposure of the lower third of the Cardenas Basalt, so we collected two samples from the superior exposure of the lower 50
m at Lava Chuar Canyon. A split from each sample was submitted for 66 element geochemical analysis, and a slab was
retained for thin-section petrographic analysis. The geochemical and petrographic data were used to evaluate each sample
for its suitability for K-Ar analysis. We found an extreme variation in potassium in our new samples of Cardenas Basalt from
2.1% K2O to 10.4% K2O by weight. There was also a significant variation in silica (46.6 to 57.8% by weight SiO 2). Eight
samples were selected for K-Ar analyses, including seven from Basalt Canyon (samples A:C-1, A:C-2, A:C-7, A:C-10, A:C13, A:C-16, and A:C-19) and one from Lava Chuar Canyon (sample A:C-14).Eleven new samples of Precambrian diabase
were collected. Each sample was submitted for 66 element geochemical analysis, and, after thin-section petrographic
analysis, each was evaluated for its suitability for K-Ar analysis. Five diabase samples were selected for K-Ar analysis.
Three samples (A:DI-7, A:DI-10, and A:DI-11) come from the very thick sill complex within the Hakatai Shale at Bass
Canyon, including two samples of the high-potassium granophyre zone near the top and one sample of very low-potassium

diabase near the base of the sill. One sample (A:DI-9) is from the sill within the Bass Formation at Hance Rapids, and
represents a whole-rock analysis of rock from which a pyroxene concentrate analysis has been published for the sample we
call EM:(Han). One new sample (A:DI-5) is from a dike within the Hakatai Shale at Red Canyon and probably represents the
same intrusive event that formed the diabase sill at Hance Rapids.
Sample Preparation and Analysis
Two kilogram samples of rock were returned to the petrographic laboratory and sawed to remove exterior surfaces which
could be contaminated. Thin-sawed samples of each rock were retained for thin-section petrographic analysis. Interior
blocks each representative of the whole rock and each weighing about 1 kg were washed, dried, and then crushed in an iron
mortar. After milling and grinding, the residue was sieved. The 80200 mesh (0.180.075 mm) particles were retained for KAr analysis. The approximate 20-gram split of particles finer that 200 mesh was submitted for bulk geochemical analysis (the
66 element MER package by XRAL Laboratories of Don Mills, Ontario) and the remainder of the processed rock powder
was archived with the intention of performing further isotopic analyses.
Table 1. K-Ar data for Cardenas Basalt and Precambrian diabase of Grand Canyon. Thirteen samples coded with A: are
new data collected and analyzed for this study. The other samples are from earlier publication. All samples are described in
the Appendix.
Unit/
K2O
Sample
%)
Code

(wt40

40
40
K
Ar*
Total40Ar*
K (ppm) (mol/g) x40Ar* (ppm) (mol/g) x40Ar* (%)
(ppm)
10-8
10-9

Total40Ar 36Ar
(mol/g) x(ppm)
10-8
10-5

36
Ar
x(mol/g)
10-13

UncerModel
tainty in
xAge
Ma
on
(Ma)
age (1s)

Carden
as FBM:
(Bas)
2.40

2.377

5.95

0.1507

3.77

96.6

0.1559

3.90

1.6

4.50

853

15

Carden
as
MN:b-8

3.10

3.070

7.68

0.1850

4.63

98

0.1886

4.72

1.2

3.20

820

20

Carden
as
MN:Tb8

3.33

3.298

8.25

0.1930

4.83

95

0.2050

5.08

3.1

8.60

800

20

Carden
as
MN:b-4

4.185

4.1447

10.37

0.2390

5.98

99

0.2414

6.04

0.73

2.04

791

20

Carden
as EM:
(Pal)
2.089

2.0689

5.18

0.1291

3.23

98

0.1319

3.30

0.80

2.23

843

34

Carden
as
LPM:4D 1.231

1.469

3.05

0.1077

2.70

85.2

0.1263

3.16

5.7

15.7

957

35

Carden
as
LPM:4E 1.282

1.269

3.18

0.10018

2.51

94.1

0.1063

2.66

1.9

5.18

1013

37

Carden
as A:C1
4.005

3.965

9.923

0.2822

7.06

97.4

0.2897

7.25

2.3

6.37

935

19

Carden
as A:C2
3.97

3.934

9.845

0.2448

6.13

97.1

0.2518

6.30

2.1

5.90

841

17

Carden
as A:C7
0.782

0.775

1.94

0.05887

1.47

82.8

0.0711

1.78

3.7

10.4

984

24

Carden
as A:C10
2.32

2.295

5.744

0.1049

2.62

94.3

0.1111

2.78

1.9

5.25

653

15

Carden
as A:C13
7.43

7.355

18.41

0.2907

7.27

93.4

0.3109

7.78

6.2

17.1

577

12

Carden
as A:C14
4.17

4.129

10.33

0.2023

5.06

95.2

0.2122

5.31

3.0

8.38

962

14

Carden
as A:C16
10.5

10.398

26.02

0.4133

10.3

98.7

0.4196

10.5

1.9

5.34

580

12

Carden
as A:C19
6.64

6.573

16.45

0.2878

7.20

97.7

0.2945

7.37

2.0

5.69

630

13

Diabase
EM:
(Han)
0.303

0.300

0.751

0.207

0.519

64.4

0.0322

0.806

3.5

9.71

914

40

1.307

3.27

0.0955

2.39

92.5

1.1031

2.58

2.4

6.56

954

30

Diabase 1.32
EM:

(Tap)
Diabase
A:DI-5
2.002

1.982

4.96

0.1171

2.93

89.35

0.1311

3.281

4.3

11.88

806

18

Diabase
A:DI-7
0.754

0.747

1.87

0.03893

0.974

80.0

0.0487

1.22

3.0

8.24

728

20

Diabase
A:DI-9
1.51

1.496

3.74

0.07468

1.87

89.8

0.0832

2.08

2.6

7.18

703

15

Diabase
A:DI-10 8.61

8.527

21.34

0.5737

14.36

96.4

0.5951

14.89

6.5

17.94

895

20

Diabase
A:DI-11 8.245

8.166

20.43

0.4206

10.52

94.3

0.4460

11.16

7.8

21.66

721

14

Potassium and argon were measured in the 13 new whole-rock samples by Geochron Laboratories of Cambridge,
Massachusetts, under the direction of Richard Reesman, the K-Ar laboratory manager. The 13 preparations were submitted
to Geochron Laboratories with the statement that each sample was a whole rock preparation from igneous rock of general
basaltic composition and that the lab should expect high argon concentration. The lab was not given any location or age
information.
The analytical methods used for the nine published determinations for K-Ar appear to be comparable to our 13 new
determinations. Comparability includes sample preparation, analytical technique and reporting. Two published whole-rock
Cardenas analyses (LPM:4D and LPM:4E by Larsen, Patterson, and Mutschler16) were also performed by Geochron
Laboratories using the same equipment. Three Cardenas analyses (MN:b-8, MN:Tb-6, and MN:b-4) and two diabase
analyses (EM:(Han) and EM:(Tap)) performed at the USGS lab also used flame photometry and the standard isotopedilution technique with mass spectrometer. Also, samples FBM:(Bas) and EM:(Pal) had the same analytical technique.
Therefore, we can say that the published K-Ar analyses, although performed at three or four different labs, show evidence of
significant standardization in sample preparation, analytical method, and reporting.The new analytical data as well as the
published analytical data are reported in Table 1. For comparability we have converted the data to the same format. The
concentration of K2O (weight%) was measured by the flame photometry method, the reported value being the average of
two readings from each sample. The 40K concentration (ppm) was calculated from the terrestrial isotopic abundance using
the concentration of K. The concentration in ppm of 40Ar*, the supposed radiogenic argon, was derived by the conventional
formula from isotope-dilution measurements on a mass spectrometer by correcting for the presence of atmospheric argon
whose isotopic composition is known. The concentration of 36Ar is also calculated from the 40Ar and total Ar measurements.
Petrography and Chemistry
The original mineralogy of the Cardenas lavas was difficult to decipher upon petrographic examination of the collected
samples due to their extensive alteration and some weathering effects. Nevertheless, estimates were made of each
samples original mineral composition, and these data are found in the appendix. Some data on whole-rock chemistry are
also included.According to our estimates, our samples of the lower member flows appear to have originally consisted of 45
50 vol.% plagioclase, 2025 vol.% augite, 1518 vol.% olivine, 57 vol.% titanomagnetite and ilmenite, and 510 vol.%
glass (mesostatis)/groundmass. Our apparent overestimates of the augite and olivine contents compared to those reported
by Larsen, Patterson, and Mutschler17 are due to the difficulty of distinguishing the alteration products of augite and olivine
from those of the originally glassy groundmass. Nevertheless, this mineral association and the subequant ophitic augite
grains enclosing the plagioclase laths mean the rock is appropriately classified as an olivine basalt.The post-extrusion
alteration of all lower member flows is similar. The late-stage glassy groundmass has been altered to chlorite, epidote and/or
red-brown clay, olivine has been completely replaced by chlorite, clay, and hematite, plagioclase has been largely altered to
saussurite and/or replaced by fine-grained sericite, chlorite, and clay, and augite has generally only sustained minimal
replacement by chlorite, clay, and hematite. Chlorite, calcite, and quartz (mostly chalcedony) commonly form amygdules and
vein fillings.Whereas it could be expected that this alteration may have changed the original rock chemistry, Larsen,
Patterson, and Mutschler18 used an isochron plot to show that these rocks have probably undergone minimal enrichment or
depletion of individual major-element oxides during hydration. However, because of the abundant clay alteration, the loss on
ignition (loss at 1000 C after samples initially dried to constant weight) varies between 4 and 10 wt%, which is almost
exclusively attributable to water loss. Otherwise, the variable contents of CaO (4.06.5 wt%), Na 2O (0.54.5 wt%) and K2O
(2.07.5 wt%) are due to the alteration mineralogy, while the relatively high total iron-oxide content (FeO plus
Fe2O3 analyses recalculated to give a total expressed as *FeO) averaging 10 wt% suggests a tholeiite, and the average 16.5
wt% Al2O3 is compatible with a high-alumina basalt of shoshonitic affinities.The pronounced rock-type variations of the upper
member flows are reflected in the wider ranges in the estimated original mineral compositions4560 vol.% plagioclase,
2030 vol.% augite, 518 vol.% olivine, 25 vol.% titanomagnetite and ilmenite, and 715 vol.% glass
(mesostasis)/groundmass. Larsen, Patterson, and Mutschler19reported the sporadic occurrence of pigeonite, and because
of the aphyric texture and lower olivine content suggested all the flows can be classified as tholeiitic olivine basalts or
basaltic andesites.The variable chemical alteration of the upper member flows is generally similar to that of the lower
member flows, but is often less extensive. Again the style and mineralogy of the alteration is reflected in the major-element
oxide contents in whole-rock analyses, but the loss on ignition attributable to water loss at 2.03.5 wt% is consistently much
lower than that of the lower member flows, indicative of the alteration being less extensive. Both Al 2O3 and total iron-oxide
content (as *FeO) are relatively uniform in all samples, Al 2O3 averaging about 14.4 wt% and *FeO about 12.9 wt%, with the
exception of A:C-19 with 20.1 wt% *FeO. Samples A:C-16 and A:C-19 with 10.4 wt% and 8.1 wt% K 2O respectively,
compared to the 3 wt% K2O average of the other samples, are heavily altered, but they also only contain trace CaO and
Na2O compared with the 4.3 wt% CaO and 3.5 wt% Na2O averages in the other samples.Larsen, Patterson, and
Mutschler20 concluded that some of these chemical variations between the upper member flows appear to be related to
initial differences in magma composition. They found that K 2O increased abruptly in the top two flows, while CaO decreased
abruptly (A:C-19 is our sample from the top flow). However, our samples do not follow all the chemical trends they observed
in their samples, which again emphasizes the complexity of the interplay between original magma compositions and
subsequent alteration during and after lava cooling. Nevertheless, it is possible to conclude from the major-element oxide
analyses of these rocks that they are tholeiitic, varying from quartz tholeiites to tholeiitic andesites. 21The intrusive rocks of
the sills and dikes are all diabases, except for the granophyres found segregated by differentiation at the tops of some of the
thicker sills (for example, the sill at Bass Canyon). Mineralogically, with 45 vol.% plagioclase, the diabases are similar to the
lower member flows, and they are also ophitic. However, the main observable difference between the lower member flows
and these mafic intrusives is in the color of the augite in thin section, which is in turn due to differences in the TiO 2 content of

the augite Larsen, Patterson, and Mutschler.22Chemically, the mafic intrusives exhibit similarities with the lower member
flows, but there are also dissimilarities Larsen, Patterson, and Mutschler. The diabases have not suffered alteration as
intense as that in the lower member flows (nor of that in the upper member flows), and what minor alteration there is, is
primarily sericite and chlorite. The granophyre is distinctive in outcrop due to its felsic character, marked by 2025 vol.%
quartz, 810 vol.% orthoclase, and 510 vol.% biotite.Previous workers appear to have created a conundrum in their
interpretation of K-Ar dating of these basalts and diabases. By accepting the Rb-Sr system as an isochron, they have
argued that the original rubidium in these rocks has been a closed system, a major assumption for a valid Rb-Sr isochron.
The observation of the strong linearity of Rb-Sr plots, indeed, is an argument for a closed system for rubidium in these rocks
as pointed out by Larsen, Patterson, and Mutschler.23 However, potassium is an alkali element very similar geochemically
to rubidium, and the argument, by its association, makes potassium in these basalts and diabases a closed system as well.
Workers24, 25 who have accepted the Rb-Sr isochron have not argued, for example, that potassium has been added to
these basalts and diabases during the hydration process which altered mineral phases within these rocks. Metasomatic
addition of potassium could explain younger-than-expected ages. That has not been argued, and apparently for good
reason. Instead, previous workers have questioned only the retention of argon by these rocks. Accepting an argon loss
model for these rocks, however, as we point out below, raises many inconsistencies and questions. Hence, the creation of a
conundrum. Although we agree that a strong case can be made against using the K-Ar system to date these rocks, the
concentration and isotopic abundances of argon in these rocks needs to be explained by an adequate model.
Results
Fig. 3. 40Ar versus 40K for Cardenas Basalt.
Isochron and age calculated from 14 of 15
samples. One sample data point (A:C-1) was not
included in the isochron calculation. Bars
represent 2s uncertainties.
K-Ar model ages for each of the 22 analyses
are listed in table 1. These model ages are
calculated by the standard equation of Dalrymple
and Lanphere26 using the mole ratio of 40Ar*
to 40K, both abundances being listed in table 1.
The model age method assumes no
radiogenic 40Ar was present when diabase and
lavas cooled to form rocks. For the Cardenas
Basalt the K-Ar model ages range from 57712
Ma (million years with 1s error) to 101337 Ma,
with the mean age of 816 Ma (n = 15). For the
Precambrian diabase the model ages range
from 70315 Ma to 95430 Ma, with the mean
age of 817 Ma (n = 7). Although the mean ages
for Cardenas Basalt and Precambrian diabase
are essentially the same (816 Ma), the wide
variation in model ages remains unexplained. Ages for Cardenas Basalt are not as predicted by superposition (that is,
relative positions in the succession of lava flows). For example, sample A:C-13 from the base of the Basalt Canyon section
is, therefore, the oldest of the Cardenas samples according to superposition of the lava flows. Yet, A:C-13 gives
the youngest age of all of the 15 Cardenas samples (577 Ma). Sample LPM:4E comes from near the top of the Cardenas
Basalt at about 220 m above the base, and, therefore, is one of the youngest flows according to superposition. Yet, LPM:4E
gives theoldest age of 101337 Ma. Also, discordant model ages were obtained for the granophyre zone near the top of the
sill at Bass Rapids: 89520 Ma and 72114 Ma (diabase samples A:DI-10 and A:DI-11 respectively). The geologic
observations at Bass Rapids require that the granophyre
zone represents a single cooling unit at the top of the sill
and the two samples are separated by only a few meters.
Furthermore, for the Hance Rapids diabase sill, pyroxene
mineral concentrate sample EM:(Han) gave a model age
of 91440 Ma, whereas the whole-rock sample A:DI-9
from the same sill gave 70315 Ma.
Fig. 4. 40Ar versus 40K for Precambrian diabase.
Isochron and age calculated from all seven data
points. Bars represent 2s uncertainties.
Fig. 3 is the 40Ar versus 40K graph for Cardenas Basalt.
We plot the data with 2s estimated error bars. A strong
linear trend is apparent which appears to represent an
isochron. The twoerror regression method of York27 was
used to fit the line to the data. Of the 15 points, only one
(A:C-1) was recognized by the computer operation to lie
significantly off the line, and that sample was not
included in the regression calculation. The equation of
the line is shown in Fig. 3 and is plotted. There is a
significantly non-zero 40Ar value for zero 40K, which would
appear to refute the zero 40Ar* assumption of the model
age technique. The computer calculated 40Ar initial is
1.760.27 10-9 mole/gram (2s error). The slope of that
line was used to calculate the isochron age of 51630
Ma (2s error).
Fig. 3 also includes the 1100 Ma reference isochron
which is widely assumed to be the true age of Cardenas
Basalt based on the Rb-Sr isochron.28 The 40Ar-40K data

plot significantly beneath what would be expected for 1100 Ma age and have significantly lower slope than what would be
expected for 1100 Ma age.
Fig. 5. 40Ar/36Ar versus 40K/36Ar for Cardenas Basalt. Isochron and age calculated from fourteen of fifteen data points.
Data point A:C-13 was not included in the isochron calculation. Bars represent 2s uncertainties.
Fig. 4 is the 40Ar versus 40K graph for Precambrian diabase. Two-sigma error bars and the best fit two-error regression
line29 are plotted. All seven data points are accepted by our computer model in plotting the line, the slope of which is most
strongly affected by the two highpotassium granophyre samples (A:DI-10 and A: DI-11). A very small, essentially zero, value
for initial 40Ar is calculated and the isochron age is 83752 Ma (2s error). The 1100 Ma reference isochron shows the
discordance with the 1100 Ma Rb-Sr isochron age.
Figs. 5 and 6 are the 40Ar/36Ar versus 40K/36Ar graphs for Cardenas Basalt and Precambrian diabase. Significant analytical
error builds up in the measurement of the abundance of 36Ar as the ratio of 40Ar to 36Ar increases. That is why data points in
the upper right of Figs. 5 and 6 have large error bars. Even considering these larger errors, linear trends are apparent. The
two-error regression method is used to plot lines in Figs. 5 and 6. For the 40Ar/36Ar versus 40K/36Ar plot in Fig. 5 for Cardenas
Basalt the isochron age is 75651 Ma. Most important is the y-intercept indicating significant initial 40Ar/36Ar = 787118 (2s)
which is 2.7 times higher than 40Ar/36Ar = 295.5, the present atmospheric value. This argues that a major part of the 36Ar is
not atmospheric contamination but was included in the rocks when they formed. The associated diabase of Fig. 6 gives a
linear array with an isochron age of 67635 Ma and initial 40Ar/36Ar about 1.5 times the modern atmosphere. In each case
the calculated K-Ar isochron age is significantly less than the generally accepted Rb-Sr isochron age of 1100 Ma.
Discussion of Discordant Ages
Fig. 6. 40Ar/36Ar versus 40K/36Ar for Precambrian diabase.
Isochron and age calculated from six of seven data points.
Data point EM:(Tap) was not included in the isochron
calculation. Bars represent 2s uncertainties.
Why does the K-Ar data yield significantly younger ages for
Cardenas Basalt and Precambrian diabase than the Rb-Sr
data? Five explanations should be considered.
1. Argon reset model

When McKee and Noble30 published their Rb- Sr isochron


age of 1070 Ma (recalculated with new Rb decay constant)
for Cardenas Basalt, they explained the four K-Ar model
ages (81920, 80920, 79020, and 84334 Ma) available
to them as caused by resetting during an episode of
heating about 800 Ma ago. They assumed that a significant
quantity of40Ar was released by metamorphism. When
another Cardenas KAr model age became available
(85515 Ma), Elston and McKee31reiterated this
interpretation specifying the Cardenas heating event at
82326 Ma. Elston and McKee argued from three new
diabase model ages (91340, 95430, and 904100 Ma)
that another earlier heating event reset diabase ages at
93025 Ma when the diabase was more deeply buried than
Cardenas Basalt.
Fig. 7. Diagrams illustrating the argon reset model for
Cardenas Basalt. The two diagrams follow sample A:C-16
through the postulated metamorphic event which allows
removal of about half of the samples argon. Other samples
on the diagrams would follow similar pathways.
Fig. 7 shows how the reset model would interpret the linear
arrays of 40Ar versus 40K and 40Ar/36Ar versus 40K/36Ar for
our larger data set of Cardenas Basalt. Because the reset
model assumes the Cardenas Basalt has an age of 1100
Ma, a large quantity of 40Ar must be removed from these
rocks to explain why they have only half the argon
expected. Fig. 7 shows how a heating event at 516 Ma
(584 million years after the eruption of the Cardenas
Basalt) removed almost all of the early-formed argon
essentially resetting the isochron to zero age making, then,
a horizontal linear array. The passage of 516 Ma since the
heating event allowed the present buildup of 40Ar by decay
of 40K.
Fig. 7 shows sample A:C-16 on todays linear array at point
C'. However, 1100 Ma ago (slightly less than one half-life
of 40K) sample A:C-16 had nearly twice its present quantity
of 40K (approximately 480 10-9mole/gram) and a very low quantity of argon (most of the argon is assumed to have vented
as the lava flows cooled). Argon buildup should have moved the sample to point C on the 1100 Ma isochron. The postulated
heating event deflected sample A:C-16 when it had an age of 584 Ma (516 Ma ago) from point B (on the 584 Ma isochron)
to point B' (on the 0 Ma isochron) as 14 10 -9 mole/gram of 40Ar and a significant quantity of 36Ar were lost. During the 516
Ma since the heating event, sample A:C-16 moved from point B' (on the 0 Ma reset isochron) to point C' (on the 516 Ma

isochron). Other samples of Cardenas Basalt followed similar trajectories on the 40Ar versus 40K and 40Ar/36Ar versus 40K/36Ar
graphs leaving the linear array suggesting an isochron age of 51630 Ma.Our larger data set shows that the heating event
postulated by McKee and Noble32 and Elston and McKee33 could not have reset the Cardenas Basalt K-Ar system 82326
Ma ago. Our new data set which forms the 14-point linear array does two things: 1. it invalidates the zero initial argon
assumption of the K-Ar model ages; and 2. it specifies the heating event at 51630 Ma ago if a reset model is to be imposed
on the data.However, an age of 51630 Ma is Upper Cambrian, not Proterozoic. The conventional interpretation of Grand
Canyon Upper Cambrian stratigraphy at Basalt Canyon is that the Tapeats Sandstone (Upper Cambrian) was sitting level on
the floor of an ocean just above the Great Unconformity and just a few hundred meters above 12 dipping Cardenas Basalt.
No trace of the assumed Upper Cambrian heating event is seen in the Tapeats Sandstone at Basalt Canyon. Therefore, the
reset model has incorrect timing and fails to explain the
K-Ar data within conventional geology.
2. Argon leakage model
Fig. 8. K-Ar model age versus K2O for Cardenas Basalt
and Precambrian diabase. Bars represent 2s
uncertainties.Could continuous argon leakage during the
history of Cardenas Basalt and Precambrian diabase
explain why these rocks have lost almost half of their 40Ar
which would be expected to build up in 1100 Ma? Larson,
Patterson, and Mutschler34 defended prolonged burial
metamorphism to explain loss of argon in Cardenas
Basalt. For seven K-Ar dates from Cardenas Basalt, they
observed a negative correlation between the abundance
of potassium and the K-Ar model age. All 22 samples are
plotted in the K-Ar model age versus K 2O graph in Fig. 8.
Negative correlation within Cardenas Basalt is apparent,
a trend that is not as strong within the diabase. Our
increased data set shows that the three highestpotassium samples of Cardenas Basalt (A:C-13, A:C16, and A:C-19) do, indeed, have the lowest K-Ar
model ages. Larson, Patterson, and Mutschler
interpreted this trend as indicative of higher burial
alteration of glassy and aphanitic (mesostasis) material
in the more-felsic, higher-viscosity, higher-potassium
flows.
Fig. 9 depicts the continuous argon leakage model for
Cardenas Basalt. Samples should be positioned on the
1100 Ma isochron, but, according to the leakage
model, because of diffusion of argon out of glassy and
crystalline material, are positioned on the 516 Ma
isochron. The model leads to a rather remarkable
explanationthe idea of a leakage isochron. The
explanation appears remarkable because leakage
would be thought of as a random process, not as a
process so strongly correlated with the abundance of
potassium. Sample A:C-16 in Fig. 9 illustrates the
argon leakage model. That sample sits at point C'
today, but, according to the leakage model, was at
point A about 1100 Ma ago. Originally, sample A:C-16,
and all the other Cardenas samples, had almost twice
the quantity of 40K than today. According to the leakage
model, continuous argon loss occurred as sample A:C16 moved away from its original position at or near
point A. After 584 Ma passed (at 516 Ma), the sample
should have been at point B. Instead of moving
through point B to point C, argon diffusion deflected the
trajectory through point B' to C'. Other samples should
have followed similar trajectories, but with significantly
less argon loss. To make the 516 Ma isochron by such
a process requires an extraordinary correlation
between the quantity of argon leakage and the quantity
of potassium. Because Cardenas Basalt differs widely
in its abundance of potassium, Elston and
McKee35rejected the continuous leakage model:
Slow leakage would have had to occur in the exact proportions
in both the minerals and the whole rock to produce the similar
agesan unlikely situation.
Fig. 9. Diagrams illustrating the argon leakage model for the
Cardenas Basalt. The two diagrams follow sample A:C-16
through continuous argon leakage which is postulated to
remove about half of the samples argon. Other samples follow
similar pathways.
Because 36Ar is nonradiogenic, we can use it as a tracer to test
any model which requires significant leakage of 40Ar. The two
isotopes have similar mass and nearly identical chemical
properties. Diffusion should not select one isotope over the
other. Two requirements of the leakage model are:the rocks

must have had significant 36Ar originally, andthere must have been significant 36Ar loss by leakage, especially high leakage
in the rocks with high potassium.
Fig. 9b illustrates the need for significant initial 36Ar and the requirement for significant 36Ar leakage. In order for highpotassium sample A:C-16 to acquire its present 40K/36Ar of 5 105 according to the leakage model, it has to begin near point
A in Fig, 9b with 40K/36Ar of about 4 10 5. This is a significant quantity of initial 36Ar, and about 45% of it appears to have
been lost by leakage during 1100 Ma of 40K decay to move the sample to its present position at point C' in Fig, 9b. If
significantly less 36Ar was present initially and sample A:C- 16 began near point D in Fig, 9b, leakage would deflect that
samples argon and potassium composition toward point F', very different from its proper position at C'. Other samples of
basalt and diabase could be used to illustrate the above conclusion about high initial 36Ar and significant36Ar loss that the
leakage model requires. Rocks with the highest 40K should have experienced the highest 36Ar loss.
Fig. 10. 36Ar versus 40K for Cardenas Basalt and Precambrian diabase. Note that non-negative correlation lines argue that
the high-potassium samples (also high40Ar) have not lost significantly more 36Ar. Positive correlation appears to fail a
requirement of argon loss models.Is there a strong negative correlation between the abundances of 36Ar and40K in Cardenas
Basalt and Precambrian diabase as the leakage model would require? Fig, 10 is the plot of the present abundances of 36Ar
and40K in the 22 Grand Canyon samples. Linear regression indicates lines in Fig. 10 possessing positive slopes,
not negative slopes. The two highpotassium Cardenas Basalt samples (A:C-16 and A:C-13) have more, not less, 36Ar than
the average of the low-potassium Cardenas samples. The two high-potassium granophyres (A: DI-10 and A: DI-11) have the
highest 36Ar (and highest 40Ar) of all 22 samples. Why would high-potassium granophyre, which supposedly experienced
very significant argon leakage, retain so much 36Ar? These data do not fit well with the continuous argon leakage model.
3. Argon inheritance model
An alternate model to the above two argon-loss models can be called the inheritance model. We could suppose that
Precambrian magmas of Grand Canyon had differing potassium concentrations and that those magmas with higher
potassium had higher partial pressures of argon. Furthermore, we might suppose that these Precambrian magmas did not
vent completely their argon. By this mechanism high-potassium magmas would create lava flows and sills possessing higher
argon, and a linear array could characterize the 40Ar versus 40K plot. By this mechanism the argon of Grand Canyon igneous
rocks would have been largely inherited from its pre-existent source. The model may explain the linear-array plots of data
such as Figs. 3, 4, 5, 6, and 8.
Although a strong case can be made that some of the argon is inherited, the extreme argon variation within a single sill is
not well explained. Why would the two high-potassium granophyre samples (A:DI-10 and A:DI-11) have about ten times
more 40Ar than the single sample of normal diabase (A:DI-7)? All three samples come from what seems to be the same
diabase sill at Bass Rapids, and would seem to require the same magma source. Therefore, the argument for inheritance of
major argon might appear weak.Nevertheless, there are numerous examples in the relevant literature of excess argon that
has thus been inherited,36including the 1986 lava dome at Mount St. Helens (Washington, USA)37 and recent lava flows at
Mt. Ngauruhoe (North Island, New Zealand).38 Admittedly, these known examples of inheritance are usually modern or
recent volcanic rocks and usually only involve a couple of hundred thousand years of radiometric age, but this is thus
suggestive of the possibility of inheritance in ancient volcanic rocks, such as these in Grand Canyon.In spite of the doubts
already expressed, it is still possible that excess argon may have been inherited by the Cardenas Basalt flows and the mafic
intrusives. Thus some of the variation in K-Ar model ages between the different flows, sills, and dikes could be attributable to
both differing amounts of excess argon in the magma over the duration of this igneous activity, and to how much excess
argon was degassed from the cooling rocks and how much retained. In the case of the diabase sill at Bass Rapids, this
might also explain why there is ten times more 40Ar in the granophyre than in the normal diabasejust as differentiation
separated the lighter felsic minerals to the top of the sill to form the granophyre, so too the inherited 40Ar diffused upwards
during cooling of the sill to concentrate also in the granophyre.
4. Argon mixing model
An alternate model similar to the inheritance model involves mixing of two different kinds of magmas. Rather than
inheritance of potassium and argon from numerous different magma types, we could suppose only two original types of
magma. The first magma we could suppose was like sample A:DI-7 with low potassium and low argon, whereas the second
magma we could suppose was like sample A:C-16 with very high potassium and very high argon. Because the two magma
end-members display a ten-fold difference in potassium and argon, the combination of different proportions of these two
magmas could produce the various concentrations of potassium and argon we have today in the basalt and diabase.
Furthermore, the mixing process of various proportions of the two magmas would produce a linear pattern or mixing line on
a graph of 40Ar versus 40K (that is, Fig. 3). For example, a magma mixed with a significant quantity of A:C-16 would lie on the
line between A:DI-7 and A:C-16, but nearer A:C-16. Another magma formed mostly of A:DI-7 would also lie on the line
between the two end-members, but closer to A:DI-7. That mixing line would be identical to an isochron but have
absolutely no age significance.Two observations appear to be fatal to the argon mixing model. First, the single diabase sill
at Bass Rapids would have to reflect the character of essentially two unmixed magmas. The high-argon granophyre at the
top of the sill could not be differentiated by gravitational removal of mafic minerals (the conventional interpretation of
granophyre), but would have to be a completely different type of magma within the sill. How could two types of magma be
intruded into a narrow fracture, presumably during a single catastrophic event, but retain their original chemical identities by
not mixing? This is a significant problem for both the inheritance and mixing models.The second problem for the mixing
model involves other major and trace elements. If the concentrations of potassium and argon in diabase and basalt are
controlled by a binary mixing process, then the other elements should reflect the mixing process as well. Samples A:DI-7
and A:C-16 should be end-members of other mixing lines besides potassium and argon. However, analysis of the
geochemical data does not produce other-element mixing lines with A:DI-7 and A:C-16 as obvious end-members. The twocomponent mixing model does not fit the geochemical data in a convincing way.
5. Change of decay model
Discordance between Rb-Sr and K-Ar dating methods for Grand Canyon Precambrian rocks has led many geologists to
propose some very speculative models suggesting significant argon loss. These argon loss models, either by reset event or
continuous leakage, have significant problems explaining the data. These difficulties arise because the basalt and diabase
are interpreted within an evolutionary context with the assumption of constancy of radioisotope decay. Could the discordant
ages be better interpreted if a singularity disrupted normal radioactive decay sometime in the past? Could 87Rb decay
and/or 40K decay constants be altered to make the data be interpreted in a concordant way? The mathematics simply
requires a change to one or both decay constants and they could be concordant. The important question remains as to how
the physics of such nonconstant decay could occur. One intriguing explanation is that both 87Rb and 40K decay were
accelerated in the past, and that 87Rb decay was accelerated more than 40K decay. We might imagine the physics of the
acceleration of the decay process is mass dependent, with b-decay of more-massive 87Rb having occurred faster than bdecay of less-massive 40K. This explanation, with appropriate increases in the decay constants of 87Rb and 40K, could bring

the two radioisotope clocks into concordance. This model would have application in explaining other discordances between
Rb-Sr and K-Ar data where the K-Ar age is much younger than the Rb-Sr age.
Other Considerations
The negative correlation between the abundance of K 2O and the K-Ar model ages was initially observed by Larson,
Patterson, and Mutschler.39 The seven samples they found were strongly correlated, and all came from the Cardenas
Basalt upper member flows. If, therefore, the samples from the lower member flows (A:C-10, A:C-13, and A:C-14) are
ignored in the graph in Fig. 8, then the negative correlation of K-Ar model age with K 2O abundance is quite pronounced
(A:C-10 and A:C-14 are clearly outriders). This relationship thus needs to be explained, as does the poor correlation in the
Cardenas Basalt lower member flows and the mafic intrusives.The abundance of K 2O is, of course, related to the mineral
compositions of these mafic rocks. In the granophyre, the major K 2O-bearing minerals are orthoclase and biotite, whereas in
the basalts and diabases neither augite nor olivine has any K2O in it, the only major mineral constituent with more than a
trace of K2O being plagioclase. This only leaves the glassy mesostasis/groundmass as potentially the primary bearer of the
K2O contents of these mafic rocks. It represents the last-crystallized material or the dregs of cooling/crystallization,
probably being quite felsic and thus high in silica and alkalis. However, the highest K 2O contents in our samples do not
correlate with the highest volume percentage of glassy mesostasis/groundmass (see appendix).The other possibility is that
the K2O contents reside in the alteration mineralogy, but sericite is the only likely candidate, being consistently present as
one of the alteration products of plagioclase. On the other hand, the glassy mesostasis/groundmass is often altered to
chlorite, epidote, and clay, none of which contain more than trace amounts of K 2O, unless the clay were illite. As already
noted, the highest-K2O samples in our Cardenas Basalt data set (A:C-13, A:C-16 and A:C-19) have the lowest K-Ar model
ages, but these samples are not necessarily all the most intensely altered. Yet the two granophyre samples (A:DI-10 and
A:DI-11) have equally high K2O contents and yield higher K-Ar model ages, though one of them has a virtually identical K-Ar
model age to the diabase (A:DI-7) beneath it in the same sill, even though the diabase only has a K 2O content less than a
third of that of the granophyre.Another puzzle is why recent lavas retain excess argon and thus yield anomalously high K-Ar
model ages, whereas these ancient lavas of the Cardenas Basalt and the diabase sills yield anomalously low K-Ar model
ages (according to the accepted age derived in this case from a Rb-Sr isochron). If indeed these mafic rocks were 1100 Ma
old, then, there has to have been significant Ar loss, yet neither the reset model for episodic Ar loss nor the leakage model
for continuous Ar loss correlates with the data, especially the 36Ar data. Thus it is highly questionable whether the K-Ar
system in these rocks is a valid indicator of their age.
Conclusion
Isotopic studies of the Cardenas Basalt and associated Proterozoic diabase sills and dikes have produced a geologic
mystery. Using the conventional assumptions of radioisotope dating, the Rb-Sr and K-Ar systems should give concordant
ages. However, it has been known for over 20 years that the two systems give discordant ages, the K-Ar age being
significantly younger than the Rb-Sr age. Previous workers have argued that a linear array of the Rb-Sr in these basalts
and diabases represents a closed system. Therefore potassium, rubidiums sister alkali element, would also appear to
represent a closed system. The published explanations for the discordant ages impugn the 40Ar of the K-Ar system, and, at
the same time, defend the Rb-Sr system. Our new data, when combined with the published data, allow models for
explaining the K-Ar system to be tested.The argon reset model was the first explanation proposed for the discordance. A
metamorphic event is supposed to have expelled significant argon from these rocks. The reset model is unable to reconcile
the new data, leading to a metamorphic event which is excessively young and inconsistent with the conventional
stratigraphic interpretation of the Grand Canyon. The argon leakage model also attempts to explain why these rocks have
about half the argon which seems to be required by the Rb-Sr system. The leakage model supposes an incredible
improbability. Both the old and new data imply that the rocks leaked argon in nearly exact proportion to the abundance of
potassium producing a leakage isochron, an explanation not supported by a quantity of an appropriate mineral or
mesostasis phase. Strong negative correlation between K-Ar model age and K 2O in the upper portion of the Cardenas
Basalt does not find an obvious petrographic interpretation consistent with mineral phases and their degrees of alteration.
Furthermore, reset and leakage models have difficulty explaining the abundance of initial 36Ar in the rocks, especially the
abundance of36Ar in those rocks which supposedly leaked the most 40Ar.Three alternatives are suggested to the two argon
loss models. The argon inheritance model and argon mixing model simply propose that argon is positively correlated with
potassium from its magma source or produced by a mixing process, and that the linear relationship on a plot of 40Ar
versus 40K is an artifact of the magma, not produced by radioisotope decay within these rocks. Inheritance of argon is a
better model for Cardenas Basalt and diabase than mixing. The change of decay model goes to the physics of
radioisotope decay and proposes a fundamental change in87Rb and/or 40K decay. All three explanations offered as
alternatives to the argon loss models invalidate using the K-Ar system as conventional geochronology would assume.
Appendix: Sample name, location and cescription
Samples whose numbers begin with A: are new rocks collected for K-Ar analyses by the authors. Samples whose numbers
do not begin with A: are those with published K-Ar analyses from the geologic literature. Stratigraphic positions use the
informal numbering terminology of Hendricks and Lucchitta.40 Oxide percentages, if available, are reported on volatile free
and reduced iron basis. Volcanic rock classification is according to the alkali-silica diagram after Le Bas et al.41
FBM:(Bas)
Whole rock preparation of Cardenas Basalt from a random sample from the strata of Basalt Canyon collected and analyzed
by Ford, Breed, and Mitchell.42 Level within Basalt Canyon section was not specified. The rock is simply called basalt and
is the rock on which the first published K-Ar analysis was performed.
MN:b-8
Whole rock preparation of Cardenas Basalt from Basalt Canyon collected and analyzed by McKee and Noble.43 Near top of
formation at about 270 m above base; equivalent to Hendricks and Lucchitta unit 30.
MN:Tb-6
Whole rock preparation of Cardenas Basalt from Basalt Canyon collected and analyzed by McKee and Noble.44 Near top of
formation at 252 m above base; equivalent to Hendricks and Lucchitta45 unit 28. Rock is basaltic trachyandesite with 55.3
wt% SiO2 and 5.4 wt% total alkalis, recalculated to volatile free).
MN:b-4
Whole rock preparation of Cardenas Basalt from small canyon west of Tanner Canyon (southeast of Basalt Canyon about 2
km and on the south side of the Colorado River). Collected and analyzed by McKee and Noble.46 Equivalent to the middle
of the Basalt Canyon section about 140 m above base; probably equivalent to Hendricks and Lucchitta unit 11.
EM:(Pal)
Whole rock preparation of Cardenas Basalt from Palisades Creek. Analysis for K-Ar by M. L. Silberman, as reported by
Elston and McKee.47 Stratigraphic level was not specified.
LPM:4D

Whole rock preparation of Cardenas Basalt from Basalt Canyon flow 4, upper member collected and analyzed by Larson,
Patterson, and Mutschler.48 Probably equivalent to unit 24 of Hendricks and Lucchitta at about 220 m above base. Called a
tholeiitic basalt with 52.0 wt% SiO2 and 3.7 wt% total alkalis.
LPM:4E
Whole rock preparation from a second sample of Cardenas Basalt from Basalt Canyon flow 4, upper member (the same
flow as sample LPM:4D) collected and analyzed by Larson, Patterson, and Mutschler.49 Probably equivalent to unit 24 of
Hendricks and Lucchitta at about 220 m above base. Called a tholeiitic basalt with 51.8 wt% SiO2 and 3.5 wt% total alkalis.
A:C-1
Whole rock preparation of Cardenas Basalt from middle third of Basalt Canyon section. Equivalent to Hendricks and
Lucchitta unit 9 at 103 m above base. Rock is trachyandesite with 56.6 wt% SiO2 and 6.1 wt% total alkalis. Estimated
original mineral composition is plagioclase 55 vol.%, augite and olivine 30 vol.%, glass (mesostasis) 13 vol.%, and
titanomagnetite and ilmenite 2 vol.%.
A:C-2
Whole rock preparation of Cardenas Basalt from middle third of Basalt Canyon section. Equivalent to Hendricks and
Lucchitta unit 9 at 108 m above base. A sample of the same lava flow unit as sample A:C-1. Rock is trachyandesite with
57.1 wt% SiO2 and 7.5 wt% total alkalis. Estimated original mineral composition is plagioclase 48 vol.%, augite and olivine
30 vol.%, glass (mesostasis)/groundmass 15 vol.%, titanomagnetite, and ilmenite 4 vol.%, orthoclase 3 vol.%, and pyrite
trace.
A:C-7
Whole rock preparation of Cardenas Basalt from upper third of Basalt Canyon section. Equivalent to Hendricks and
Lucchitta unit 24 at 213 m above base. Rock is basaltic trachyandesite with 53.9 wt% SiO 2 and 5.8 wt% total alkalis.
Estimated original mineral composition is plagioclase 60 vol.%, augite 20 vol.%, olivine 5 vol.%, glass
(mesostasis)/groundmass 10 vol.%, and titanomagnetite and ilmenite 5 vol.%.
A:C-10
Whole rock preparation of Cardenas Basalt from lower third of Basalt Canyon section. Equivalent to Hendricks and Lucchitta
unit 7 at 83 m above base. Sample comes from center of spherical mass within more weathered and altered flow. Rock is
basaltic trachyandesite with 53.1 wt% SiO2 and 6.7 wt% total alkalis. Estimated original mineral composition is plagioclase
45 vol.%, augite 25 vol.%, olivine 18 vol.%, titanomagnetite and ilmenite 7 vol.%, and glass (mesostasis)/groundmass 5 vol.
%.
A:C-13
Whole rock preparation of Cardenas Basalt from lower third of Basalt Canyon section. Equivalent to Hendricks and Lucchitta
unit 1 at less than 1 m above the first Dox intertongue at the base of the type section. Sample is lowest flow and may be
somewhat contaminated with material included from underlying Dox Formation. Rock is tephrite basanite with 46.6 wt%
SiO2 and 8.0 wt% total alkalis. Estimated original mineral composition is plagioclase 50 vol.%, augite 20 vol.%, olivine 15
vol.%, glass (mesostasis)/groundmass 10 vol.%, and titanomagnetite and ilmenite 5 vol.%.
A:C-14
Whole rock preparation of Cardenas Basalt from exposure on north side of Lava Chuar Canyon. Representative of the lower
50 m of the formation, the sample is probably equivalent to Hendricks and Lucchitta unit 5 about 45 m above the base of the
formation. Rock is trachyandesite with 58.3 wt% SiO 2 and 6.8 wt% total alkalis. Estimated original mineral composition is
plagioclase 45 vol.%, augite 23 vol.%, olivine 17 vol.%, glass (mesostatis)/groundmass 10 vol.%, and titanomagnetite and
ilmenite 5 vol.%.
A:C-16
Whole rock preparation of Cardenas Basalt from upper third of Basalt Canyon section. Equivalent to Hendricks and
Lucchitta unit 26 at 230 m above base. Sample comes from the flow just beneath the thick lapillite unit. Rock is
trachyandesite with 57.8 wt% SiO2 and 10.5 wt% total alkalis. Estimated original mineral composition is plagioclase 50 vol.
%, augite 30 vol.%, olivine 8 vol.%, groundmass/glass (mesostasis) 7 vol.%, and titanomagnetite and ilmenite 5 vol.%.
A:C-19
Whole rock preparation of Cardenas Basalt from upper third of Basalt Canyon section. Equivalent to Hendricks and
Lucchitta unit 31 at 290 m above the base. Rock is tephrite basanite with 47.5 wt% SiO 2 and 8.1 wt% total alkalis. Strong
resemblance to sample A:C-13 which also has low silica. Estimated original mineral composition is plagioclase 45 vol.%,
augite 25 vol.%, olivine 18 vol.%, groundmass/glass (mesostasis) 7 vol.%, and titanomagnetite and ilmenite 5 vol.%.
EM:(Han)
Pyroxene concentrate from the diabase sill at Hance Rapids. Collected and analyzed by Elston and McKee.50
EM:(Tap)
Plagioclase concentrate from the diabase sill at Tapeats Creek. Collected and analyzed by Elston and McKee.51
A:DI-5
Whole rock preparation from the diabase dike on the east side of Red Canyon south of Hance Rapids. Rock has 50.2 wt%
SiO2 and 5.0 wt% total alkalis. Estimated mineral composition is plagioclase 45 vol.%, olivine 30 vol.%, augite 25 vol.%, and
titanomagnetite and ilmenite 5 vol.%.
A:DI-7
Whole rock preparation from the diabase sill within the Hakatai Shale at Bass Canyon on the north side of the Colorado
River at mile 107.6. Rock has 47.2 wt% SiO2 and 2.7 wt% total alkalis. Estimated mineral composition is plagioclase 45 vol.
%, olivine 34 vol.%, augite 15 vol.%, titanomagnetite and ilmenite 5 vol.%, and biotite 1 vol.%.
A:DI-9
Whole rock preparation from middle of diabase sill on north side of Colorado River below Hance Rapids at mile 76.8. The sill
intrudes the Bass Formation. Rock has 49.5 wt% SiO 2 and 4.6 wt% total alkalis. Estimated mineral composition is
plagioclase 45 vol.%, augite 30 vol.%, olivine 20 vol.%, and titanomagnetite and ilmenite 5 vol.%.
A:DI-10
Whole rock preparation from fine-grained granophyre zone 4 m from the top of the diabase sill at Bass Canyon at mile 108
about 250 m height above the Colorado River on the north side. The sill intrudes the Hakatai Shale. Rock has 62.8 wt%
SiO2 and 9.0 wt% total alkalis. Estimated mineral composition is plagioclase 48 vol.%, quartz 25 vol.%, orthoclase 10 vol.%,
augite 10 vol.%, biotite 5 vol.%, and titanomagnetite and ilmenite 2 vol.%.
A:DI-11
Whole rock preparation from same location as A:D-10 except that the rock is coarse-grained granophyre 7 m from the top of
the diabase sill and about 3 to 4 m from the base of the granophyre zone. The sill intrudes the Hakatai Shale. Rock has 63.0
wt% SiO2 and 9.5 wt% total alkalis. Estimated mineral composition is plagioclase 40 vol.%, quartz 20 vol.%, augite 15 vol.%,
biotite 10 vol.%, orthoclase 8 vol.%, and titanomagnetite and ilmenite 2 vol.%

The Relevance of Rb-Sr, Sm-Nd, and Pb-Pb Isotope Systematics to Elucidation of the Genesis and History of
Recent Andesite Flows at Mt. Ngauruhoe, New Zealand, and the Implications for Radioisotopic Da
by Dr. Andrew A. Snelling on January 6, 2010

Abstract
Mt. Ngauruhoe in the Taupo Volcanic Zone of New Zealand erupted andesite lava flows in 1949 and 1954, and avalanche
deposits in 1975. Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic analyses of samples of these andesites, as anticipated, did not
yield any age information, although the Pb isotopic data are strongly linear. When compared with recent andesite flows
from the related adjacent Ruapehu volcano, the Sr-Nd-Pb radioisotopic systems plotted on correlation diagrams provide
information about the depleted mantle source for the parental basalt magmas and the source of the crustal contamination
that produced the andesite lavas from them. The variations in both the depleted mantle Nd model ages and the Pb
isotopes also suggest radioisotopic heterogeneity in the mantle wedge 80 km below the volcano where partial melting has
occurred, contaminated by mixing with trench sediments scraped off the interface with the subducting slab. Thus the
radioisotopic ratios in these recent Ngauruhoe andesite flows were inherited, and reflect the origin and history of the mantle
and crustal sources from which the magma was generated. By implication, the radioisotopic ratios in ancient lavas
throughout the geologic record are likely fundamental to their geochemistry, characteristic of their origin and history rather
than necessarily providing valid conventional ages.
Shop Now
Keywords: andesite, 19491975 flows, Mt. Ngauruhoe, New Zealand, Rb-Sr, Sm-Nd, Pb-Pb, radioisotopes, petrogenesis,
depleted mantle, magma genesis, crustal contamination, subduction, mixing, inherited radioisotopic ratios, invalid
conventional ages
This paper was originally published in the Proceedings of the Fifth International Conference on Creationism, pp. 285303
(2003) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh.
Introduction
Fig. 1. The location of Mt. Ngauruhoe in the Taupo Volcanic Zone
(TVZ), New Zealand, showing the main structural features. The
hatched area is the andesite arc, and the solid triangles are
basalt-andesite volcanoes. The inset shows the major
components of the boundary between the Australian and Pacific
Plates in the New Zealand region (arrows indicate relative
motions).12With the development of isotope geochemistry in the
last 3540 years has come the realization that radioisotopes may
not always provide reliable age measurements. It has been
discovered that recent and historic lavas, particularly on oceanic
islands, yield incredibly old radioisotopic ages.1, 2, 3 This has
led to the recognition that the radioisotopes in these lavas reflect
the isotopic compositions of the mantle sources of these lavas,
and of any crustal contamination the magmas may have
incorporated during ascent and extrusion.4, 5, 6, 7 The present
burgeoning isotope geochemistry literature, reporting increasing
numbers of ever more accurate and sophisticated radioisotopic
determinations, has only refined the modeling of mantle sources
and discussion of their origin, while ignoring the obvious
implications for the radioisotopic age determinations of
ancient lavas being published in the same literature. It
was deemed timely, therefore, to undertake an isotope
geochemical study of some recent lavas not previously
investigated. There were two objectivesto explore the
meaning of the radioisotopic ratios in terms of the
petrogenesis of the lavas; and thus to recognize the
implications for radioisotopic age determinations on
both recent and ancient lavas.Chosen for this study was
Mt. Ngauruhoe, an andesite stratovolcano rising 2291
m above sea level within the Tongariro Volcanic Center
of the Taupo Volcanic Zone, North Island, New Zealand
(fig. 1).8, 9 Though not as well known as its active
neighbor, Mt. Ruapehu (about 12 km to the south),
Ngauruhoe is an imposing, almost perfect cone that
rises more than 1000 m above the surrounding
landscape. Eruptions from a central 400 m diameter
crater have constructed the steep (33), outer slopes of
the cone.10, 11
Geologic Setting
The Taupo Volcanic Zone is a volcanic arc and marginal
basin of the Taupo-Hikurangi arc-trench (subduction)

system,13, 14 and is a southward extension on the Tonga-Kermadec arc into the continental crustal environment of New
Zealands North Island. It has been interpreted as oblique subduction of the Pacific Plate beneath the Australian Plate (fig.
1). The zone extends approximately 300 km north-northeast across the North Island from Ohakune to White Island and is up
to 50 km wide in the central part, narrowing northwards and southwards. This volcano-tectonic depression (the TaupoRotorua Depression15) comprises rhyolitic volcanic centers (Rotorua, Okataina, Maroa, and Taupo), plus the calc-alkaline
Tongariro Volcanic Center. The latter is part of a young (<0.25 Ma) andesite-dacite volcanic arc with no associated rhyolitic
volcanism extending along the eastern side of the zone.16The Tongariro Volcanic Center extends for 65 km southsouthwest from Lake Taupo at the southern end of the Taupo Volcanic Zone (fig. 1), and consists of four large predominantly
andesite volcanoesKakaramea, Pihanga, Tongariro, and Ruapehu (fig. 2); two smaller eroded centers at Maungakatote
and Hauhungatahi; a satellite cone and associated flows at Pukeonake; and four craters at Ohakune (fig.
2).17, 18, 19, 20Most vents lie close to the axis of a large graben in which Quaternary volcanic rocks overlie a basement of
Mesozoic greywacke and Tertiary sediments.21, 22 North-northeast-trending normal faults with throws up to 30 m cut the
volcanoes within the graben. Nearly all vents active within the last 10,000 years lie on a gentle arc which extends 25 km
north-northeast from the Rangataua vent on the southern slopes of Ruapehu through Ruapehu summit and north flank
vents, Tama Lakes, Ngauruhoe, Red Crater, Blue Lake, and Te Mari craters. None of the young vents lie on the mapped
faults, which mostly downthrow towards the axis of the graben. The vent lineation lies along this axis, which is considered to
mark a major basement fracture23, 24, 25 that allows the intrusion of andesite dikes.The Tongariro volcanics unconformably
overlie late Miocene marine siltstones beneath Hauhungatahi. A minimum age for the onset of volcanism is indicated by the
influx of andesite pebbles in early Pleistocene conglomerates of the Wanganui Basin to the south. 26, 27 Wilson et
al.28 suggest a maximum late Pliocene age of 2.0 Ma, based on K-Ar dating of the earliest lavas of the Hauhungaroa
cone.29 The oldest dated lavas from the Tongariro Volcanic Center are hornblende andesites: exposed at Tama Lakes
between Ngauruhoe and Ruapehu, dated at 0.26 0.003 Ma; from Ruapehu, dated at 0.23 0.006 Ma; and from
Kakaramea, dated at 0.22 0.001 Ma (all potassium-argon dates).30The Tongariro volcano itself is a large volcanic massif
that consists of at least twelve composite cones, the youngest and most active of which is Ngauruhoe. A broad division has
been made into older (>20 ka) and younger (<20 ka) lavas.31,32 There is a north-northeast alignment of the younger vents
of Tongariro, particularly evident between Te Mari and Ngauruhoe.
Ngauruhoe
Fig. 2. Location and deposits of the Tongariro Volcanic Center.45, 46, 47
Ngauruhoe is the newest cone of the Tongariro Volcanic Center, and has been active for at least 2.5 ka.33, 34, 35 It has been
one of the most active volcanoes in New Zealand, with more than 70 eruptive episodes since 1839, when the first steam
eruption was recorded by European settlers.36, 37, 38 Prior to European colonization the Maoris witnessed many eruptions
from the mountain.39 The first lava eruptions seen by European settlers occurred between April and August 1870, with two
or three flows witnessed spilling down the north-western flanks of the volcano on July 7. 40, 41 Following that event there
were pyroclastic (ash) eruptions every few years, with major explosive activity in AprilMay 1948.42The next lava extrusion
was in February 1949, beginning suddenly with ejection of incandescent blocks, and a series of hot block and ash flows
down the north-western slopes on February 9.43, 44 The southern sub-crater filled with lava, which by late on February 10
had flowed over the lowest part of the rim and down the northwest slopes of the cone. By February 12 the flow had ceased.
Subsequent mapping placed its volume at about 575,000 m3 (fig. 3).48, 49 Further explosive pyroclastic (ash) eruptions
followed, reaching a maximum about February 1921. The eruptions ended on March 3.The eruption from May 13, 1954 to
March 10, 1955 began with explosive ejection of ash and blocks. Red-hot lava had been seen in the crater five months
previously.50, 51 The eruption was remarkable for the estimated large volume of almost 8 million m 3 of lava that then flowed
from the crater from June through September 1954, and was claimed to be the largest flow of lava observed in New Zealand
(that is, by European settlers).52,53 The lava was actually expelled from the crater in a series of seventeen distinct flows on
the following dates:54, 55
Fig. 3. Map of the north-western slopes of Mt. Ngauruhoe showing
the lava flows of 1949 and 1954, and the 1975 avalanche
deposits.56, 57, 58, 59, 60 The location of samples collected for this
study are marked.
June 4, 30;
July 8, 9, 10, 11, 13, 14, 23, 28, 29, 30;
August 15(?), 18; and
September 16, 18, 26.
Fig. 3 shows the distribution of those 1954 lava flows that are still
able to be distinguished on the northwestern and western slopes of
Ngauruhoe. All flows were of aa lava (as was the February 11, 1949
flow), typified by rough, jagged, clinkery surfaces made up of blocks
of congealed lava. The lava flows were relatively viscous, some being
observed at close quarters slowly advancing at a rate of about 20 cm
per minute.61, 62, 63 The August 18 flow was more than 18 m thick
and still warm almost a year after being erupted. Intermittent
explosive eruptions and spectacular lava fountaining during June and
July 1954 built a spatter-and-cinder cone around the south subcrater,
modifying the western summit of the mountain. Activity decreased for
two months after the last of the lava flows on September 26, but
increased again during December 1954 and January 1955, with lava
fountaining and many highly explosive pyroclastic (ash) eruptions. The last ash explosion was reported on March 10, 1955,
but red-hot lava remained in the crater until June 1955.64, 65
After the 19541955 eruption, Ngauruhoe steamed semi-continuously, with numerous small eruptions of ash derived from
comminuted vent debris. Incandescent ejecta were seen in January 1973, and ash containing juvenile glassy andesite
shards erupted in December 1973.66 Canon-like highly explosive eruptions in January and March 1974, the largest since
19541955, threw out large quantities of ash and incandescent blocks, one of which was reported as weighing 3,000 tonnes
and thrown 100 m.67, 68 Pyroclastic avalanches flowed from the base of large convecting eruption columns, down the west
and north slopes of the cone; and the crater became considerably shallower.69, 70A series of similar but more violent
explosions occurred on February 19, 1975, accompanied by clearly visible atmospheric shock waves and condensation
clouds.71, 72, 73, 74 Ash and blocks up to 30 m across were ejected and scattered within a radius of 3 km from the summit.
The series of nine, canon-like, individual eruptions followed a 1.5 hour period of voluminous gas-streaming emission, which

formed a convecting eruption plume between 11 km and 13 km high.75, 76, 77 The explosions took place at 2060 minute
intervals for more than five hours. Numerous pyroclastic avalanches were also generated by fallback from the continuous
eruption column. They were a turbulent mixture of ash, bombs, and larger blocks which rolled swiftly down Ngauruhoes
sides at about 60 km per hour.78, 79 The deposits from these avalanches and the later explosions accumulated as sheets of
debris in the valley at the base of the cone, but did not extend beyond 2 km from the summit. It is estimated that a minimum
bulk volume of 3.4 million m3 of pyroclastic material was erupted in the seven-hour eruption sequence on that day.80 Fig. 3
shows the location of these avalanche deposits.There have been no eruptions since February 1975. A plume of steam or
gas was often seen above the summit of the volcano for many years, as powerful fumaroles in the bottom of the crater
discharged hot gases. The temperature of these fumaroles in the crater floor has steadily cooled since 1979, suggesting that
the main vent has become blocked.
Sample Collection
Field work and collection of samples was undertaken in January 1996. The Ngauruhoe area was accessed from State
Highway 47 via Mangateopopo Road. From the parking area at the end of the road the Mangateopopo Valley walking trail
was followed to the base of the Ngauruhoe cone. From there the darker-colored recent lava flows were clearly visible and
easily identifiable against the lighter-colored older portions of the north-western slopes (fig. 3).
Eleven 23 kg samples were collectedtwo each from the February 11, 1949, June 4, 1954, and July 14, 1954 lava flows,
and from the February 19, 1975 avalanche deposits; and three from the June 30, 1954 lava flows. The sample locations are
marked on Fig. 3. Care was taken to ensure correct identification of each lava flow, and that the samples collected were
representative of each flow and of any variations in textures and phenocrysts in the lavas.
Laboratory Work
All samples were sent first for sectioningone thin section from each sample. These were subsequently all carefully
examined under a petrographic microscope and their mineralogy and textures recorded. A set of representative pieces from
each sample (approximately 100 g) was then dispatched to the AMDEL Laboratory in Adelaide, South Australia, for wholerock major, trace and rare earth element analyses. A representative set (50100 g from each sample) was also sent to
Geochron Laboratories in Cambridge (Boston), Massachusetts, for whole-rock potassium-argon (K-Ar) dating.81 A third
representative set (50100 g from ten of the samplestwo samples from each flow) was sent to the PRISE Laboratory in
the Research School of Earth Sciences at the Australian National University in Canberra, Australia, for Rb-Sr, Sm-Nd, and
Pb-Pb isotopic determinations.
At the AMDEL Laboratory each sample was crushed and pulverized. Whole-rock analyses were undertaken by total fusion
of each powdered sample and then digesting them before ICP-OES (inductively coupled plasmaoptical emission
spectrometry) for major and minor elements, and ICP-MS (inductively coupled plasmamass spectrometry) for trace and
rare earth elements. Fe was analyzed for amongst the major elements by ICP-OES as Fe 2O3 and reported accordingly.
Additionally, separate analyses for Fe as FeO were undertaken via wet chemistry methods, which were also used to
measure the loss on ignition (primarily H2O content). The detection limit for all major element oxides was 0.01%. For minor
and trace elements the detection limits varied between 0.5 and 20 ppm, and for rare earth elements between 0.5 and 1
ppm.
The Rb-Sr, Sm-Nd, and Pb-Pb isotopic analyses were undertaken at the PRISE Laboratory under the direction of Dr.
Richard Armstrong. No specific sample location or expected age information was supplied to the laboratory, although
samples were described as young andesites so that the laboratory staff would optimize the sample preparation procedure in
order to obtain the best analytical results. At the laboratory the sample pieces were crushed and pulverized, and then
dissolved in concentrated hydrofluoric acid, followed by the standard chemical separation procedures for each of these
isotopic systems. Once separated, the elements in each isotopic system were loaded by standard procedures onto metal
filaments to be used in the solid source thermal ionization mass spectrometer (TIMS), the state-of-the-art technology in use
in this laboratory. Sr isotopes were measured using the mass fractionation correction 86Sr/88Sr = 0.1194. The 87Sr/86Sr ratios
reported were normalized to the NBS standard SRM 987 value of 0.710207. Nd isotopes were corrected for mass
fractionation using 146Nd/144Nd = 0.7219, and were normalized to the present-day 143Nd/144Nd value of 0.51268 for BCR-1. Pb
isotope ratios were normalized to NBS standard SRM 981 for mass fractionation.
Petrography and Chemistry
Clark82 reported that most of the flows from Ngauruhoe are labradorite-pyroxene andesite with phenocrysts of plagioclase
(labradorite), hypersthene, and rare augite in a hyalopilitic (needle-like microlites set in a glassy mesostasis) groundmass
containing abundant magnetite. However, all lavas, lapilli, and incandescent blocks that have been analyzed from eruptions
in the twentieth century also contain olivine; so that chemically they may be classed as low-silica (or basaltic) andesites
(using the classification scheme of Gill83). Previously published analyses84, 85, 86, 87, 88, 89,90, 91 show only trivial changes
in composition of the lavas and pyroclastics between 1928 and 1975. In fact, the 1954 and 1974 andesites are so similar
that Nairn et al.92 suggested that a solid plug of 1954-andesite was heated to incandescence and partially remobilized on
top of a rising magma column in 1974. This plug was disrupted and blown from the vent as ejecta ranging in texture from
solid blocks, through expanded scoria to spatter bombs.
Table 1 lists the whole-rock major, trace, and rare earth element analyses of the 11 samples collected in this study.
Comparison
of
these
data
for
each
flow
with
the
corresponding
published
Ngauruhoe
data93, 94, 95, 96, 97, 98, 99, 100indicates that in their bulk chemistries all the samples analyzed (and thus all the flows) are
virtually identical to one another, the trivial differences, even among the trace and rare earth elements, being attributable to
the statistics of analytical errors, sampling and natural variations. Thus it is not unreasonable to conclude that these basaltic
andesites are cogenetic, coming from the same magma and magma chamber, even as they have been observed to flow
from the same volcano.
Table 1. Whole-rock, major-element oxide, trace element, and rare earth element analyses of five recent (1949, 1954, 1975)
lava flows at Mt. Ngauruhoe, New Zealand (Analyst: AMDEL, Adelaide; April 1996).
Sample

1A

1B

2A

2B

3A

3B

Flow Date February 11, 1949

June 4, 1954

June 30, 1954

SiO2 (%)

56.7

56.2

55.3

55.8

56.3

55.9

TiO2 (%)

0.79

0.85

0.74

0.77

0.76

Al2O3 (%)

17.2

17.3

16.5

17.3

Fe2O3 (%)

9.10

9.63

9.26

MgO (%)

4.28

3.84

5.21

3C

4A

4B

5A

5B

July 14, 1954

February 19, 1975

55.6

56.1

55.6

56.0

55.4

0.75

0.74

0.75

0.84

0.79

0.78

17.0

16.9

16.7

16.9

17.5

17.0

16.5

9.23

9.11

9.17

9.59

9.29

9.61

9.25

9.43

4.71

4.75

5.00

5.09

4.71

3.84

4.31

5.27

MnO (%)

0.15

0.16

0.15

0.15

0.15

0.15

0.16

0.15

0.16

0.15

0.16

CaO (%)

7.61

7.93

8.22

8.29

7.95

8.16

8.17

8.00

8.17

7.83

8.56

Na2O (%)

3.08

3.19

2.91

3.03

3.06

2.98

2.95

3.02

3.11

3.08

2.86

K2O (%)

1.15

1.01

1.05

1.00

1.10

1.08

1.06

1.12

1.04

1.10

1.09

P2O5 (%)

0.13

0.13

0.12

0.12

0.13

0.13

0.13

0.13

0.13

0.14

0.14

L0l (%)

0.42

0.48

0.51

0.37

0.38

0.50

0.53

0.62

0.42

0.70

0.41

TOTAL

100.61

100.72

99.97

100.77

100.69

100.72

100.72

100.79

100.42

100.31

100.60

Cr (ppm)

60

60

120

80

80

100

100

80

40

60

100

Sc (ppm)

25

30

30

30

25

30

30

30

30

30

30

V (ppm)

220

240

220

220

220

220

220

220

260

240

240

Ni (ppm)

15

10

31

21

24

29

29

25

18

29

Co (ppm)

20

20

20

20

20

20

20

20

20

20

20

Cu (ppm)

180

38

63

57

69

51

69

67

22

51

78

Zn (ppm)

87

93

87

84

86

89

85

84

91

88

87

Ga (ppm)

24

22

23

28

27

26

26

27

29

27

28

Rb (ppm)

36.0

26.5

29.0

32.0

32.5

30.0

29.5

35.5

32.0

31.5

31.0

Sr (ppm)

220

165

200

240

240

240

220

260

260

240

320

Y (ppm)

19

18

15

19

18

19

18

18

20

19

17

Zr (ppm)

100

80

80

80

80

80

80

100

80

100

80

Hf (ppm)

10

Nb (ppm)

<10

<10

<10

<10

<10

<10

<10

<10

<10

<10

<10

Ba (ppm)

200

165

175

185

220

200

200

220

190

200

220

Ta (ppm)

<2

<2

<2

<2

<2

<2

<2

Pb (ppm)

20

15

10

10

20

10

10

15

10

15

20

Th (ppm)

5.0

3.0

2.5

3.0

3.0

3.0

3.0

3.5

3.0

3.0

3.5

U (ppm)

2.0

1.5

1.0

1.0

1.0

1.0

1.0

1.5

1.5

1.0

1.0

La (ppm)

12

11

11

11

11

11

11

11

12

Ce (ppm)

25

16

19

22

22

23

22

25

21

22

24

Pr (ppm)

Nd (ppm)

18.5

14.0

10.5

12.5

12.0

13.0

13.5

14.0

14.0

11.5

14.0

Sm (ppm)

3.5

3.0

1.5

2.5

2.5

2.5

2.5

2.5

3.0

2.5

3.0

Eu (ppm)

1.0

1.0

0.5

1.0

1.0

1.0

1.0

1.0

1.0

1.0

1.0

Gd (ppm)

Tb (ppm)

0.5

<0.5

<0.5

0.5

<0.5

<0.5

0.5

0.5

0.5

<0.5

0.5

Dy (ppm)

4.0

3.5

2.5

3.5

3.0

3.5

3.0

3.5

4.0

3.0

3.5

Ho (ppm)

1.0

1.0

0.5

1.0

1.0

1.0

1.0

1.0

1.0

1.0

1.0

Er (ppm)

Tm (ppm)

<1

<1

<1

<1

<1

<1

<1

<1

<1

<1

<1

Yb (ppm)

Lu (ppm)

<0.5

<0.5

<0.5

<0.5

<0.5

<0.5

<0.5

<0.5

<0.5

<0.5

<0.5

K/Rb

319

381

362

313

338

360

359

315

325

349

351

Ca/Sr

346

481

411

345

331

340

371

308

314

325

268

Nevertheless, Nairn et al.101 suggested that even though the 1949 and 1954 lavas were both olivine-bearing andesites,
analyses of them showed the 1954 lavas to be slightly more basic than the 1949 lava, with slightly higher MgO, CaO, and
total iron oxides, but lower SiO2 and alkalis. However, these trends are not duplicated with any statistical significance in the
analytical results of this study (table 1). At least they found that their analyses of the 1974 lava blocks and bombs were
identical within the limits of error with the 1954 lavas, which was also substantiated in this study with respect to the 1975
avalanche material and the 1954 lavas (table 1).Clark102 and Cole103 recognized five lava types in the Tongariro Volcanic
Center based on the modal proportions of the phenocryst minerals. Graham104 modified this scheme to six types based on
a combination of mineralogy and chemistry, but given their uniform bulk chemistry and petrology, these Ngauruhoe lava

flows group together as plagioclase-pyroxene andesite within Grahams Type 1. Cole et al.105 have described Type 1
lavas as volumetrically dominant within the Tongariro Volcanic Center, and as exhibiting coherent chemical trends with
increasing silica content. They are relatively Fe-rich and follow a typical calc-alkaline trend on the AFM diagram.Adapting the
terminology of Gill,106 the Ngauruhoe lavas are described as basic andesites (5358 wt% SiO 2).107 Their designation as
plagioclase-pyroxene andesites is based on the predominant phenocrysts present, with plagioclase greater than or equal to
pyroxene. The proportion of phenocrysts to groundmass, and the different mineral proportions among the phenocrysts, in
the samples collected for this study very closely resemble those in samples reported from the same flows. 108, 109All
samples of the five lava flows examined in this study exhibit a porphyritic texture, with phenocrysts (up to 3 mm across)
consistently amounting to 3540% by volume. The phenocryst assemblage is dominated (2:1) by plagioclase, but
orthopyroxene and augite (clinopyroxene) are always major components, while olivine and magnetite are only present in
trace amounts. This POAM (plagioclase-olivine/orthopyroxene-augite-magnetite) phenocryst assemblage is a typical
anhydrous mineralogy.110 The groundmass consists of microlites of plagioclase, orthopyroxene, and clinopyroxene, and is
crowded with minute granules of magnetite and/or Fe-Ti oxides. Small amounts (910%) of brown transparent (acidresiduum) glass are also present, and the overall texture is generally pilotaxitic (the lath-shaped microlites being generally
interwoven in an irregular unoriented fashion).Steiner111 stressed that xenoliths are a common constituent of the 1954
Ngauruhoe lavas, but also noted that Battey112 reported the 1949 Ngauruhoe lava was rich in xenoliths. All samples in this
study contained xenoliths, including those from the 1975 avalanche material. However, many of these aggregates are more
accurately described as glomerocrysts and mafic (gabbro, websterite) nodules Graham, Cole, Briggs, Gamble, and
Smith.113 They are 35 mm across, generally have hypidiomorphic-granular textures, and consist of plagioclase,
orthopyroxene, and clinopyroxene in varying proportions, and very occasionally olivine. The true xenoliths are often rounded
and invariably consist of fine quartzose material. Steiner114 also described much larger xenoliths of quartzo-feldspathic
composition and relic gneissic structure.The plagioclase phenocrysts have been recorded as ranging in composition from
An89 to An40 (andesine to bytownite), but in Ngauruhoe lavas are usually labradorite (An68-55). They are subhedral and
commonly exhibit complex oscillatory zoning with an overall trend from calcic cores to sodic rims.115, 116, 117 Thin outer
rims are usually compositionally similar to groundmass microlites. Twinning and hourglass structures are
common.Orthopyroxene predominates (>2:1) over clinopyroxene. Subhedral-euhedral orthopyroxene is typically pleochroic
and sometimes zoned. Compositions range from Ca4Mg74Fe22 to Ca3Mg47Fe50,118, 119 but representative bulk and partial
analyses of Ngauruhoe orthopyroxenes120, 121, 122 indicate a hypersthene composition predominates, which is confirmed
by optical determinations.123, 124 Euhedral-subhedral clinopyroxene is typically twinned and zoned, but compositions show
a restricted range of Ca43Mg47Fe10 to about Ca35Mg40Fe25, all of which is augite.125, 126The olivine present is strongly
magnesian, analyses indicating some compositional zoning from Fo88 to Fo78. The magnetite present in the groundmass is
titanomagnetite, judging from the amount of TiO2 present in the whole-rock analyses (table 1), but some ilmenite is likely to
occur sporadically in association with it.127, 128As already noted, the trace and rare earth element data in Table 1 indicate
that all these samples (and thus all the flows) are virtually identical to one another, the minor differences being attributable to
the statistics of analytical errors, sampling and natural variations, and thus are trivial. These data are also comparable to
published data.129, 130, 131, 132As noted by Graham and Hackett, in these Type 1 or plagioclase-pyroxene andesites Cr
and Ni follow a trend similar to that of MgO, so that in samples where the MgO concentration is lower, the Cr and Ni
concentrations are also lower, while at the same time the Al 2O3 concentration is higher, reflecting less olivine and pyroxene
in those samples. The K/Rb ratios vary from 313 to 381 (average 343) and are within the range, and right on the average, of
those ratios for plagioclase-pyroxene andesites reported by Cole.133 Both Rb and Ba are, as expected, positively related to
K. Similarly, Sr is positively correlated with Ca, and the Ca/Sr ratios vary from 268 to 481 (average 349), which matches
published data.134 On the other hand, the Zr contents of these lava flows range from 80 to 100 ppm (average 85 ppm),
which is less than the published data for plagioclase-pyroxene andesites; whereas the Cu ranges from 22 to 180 ppm, which
is higher than the published data.The normalized geochemical distribution patterns of the trace elements in these lava flows
show the characteristic enrichment of large ion lithophile elements (LILE) such as Sr, K, Rb, Ba, and Th, and depletion of
high field strength elements (HFSE) such as Nb, P, Zr, Ti, and Y, in calc-alkaline lavas compared to normal (N-type) midocean ridge basalts (MORB).135 The samples from these lava flows also show enrichment of the rare earth elements when
compared to chondrite values, and show greater enrichment of the light rare earth elements (LREE) than the heavy rare
earth elements (HREE), identical to published data.136 These plagioclasepyroxene andesites also all have negative Eu
anomalies.
K-Ar Isotope Systematics
Snelling137 reported having obtained K-Ar model ages for these same samples of recent Mt. Ngauruhoe andesite flows of
<0.27 to 3.5 Ma. Such results were expected, as meaningful dates from historic lava flows are not usually obtained, which is
recognized in the standard scientific literature.138, 139 These dates could not be reproduced, even from splits of the same
samples from the same flow. This apparent inconsistency merely indicates variation in the excess 40Ar* (radiogenic 40Ar)
content. Indeed, Ar contamination at such low concentration levels is often expected, but this problem of excess 40Ar* in
historic lava flows is still well documented in the literature. It was concluded that this excess 40Ar* had been inherited by
these magmas during their genesis in the upper mantle, and therefore has no age significance.
Sr-Nd-Pb Isotope Geochemistry
The results of the Rb-Sr, Sm-Nd, and Pb-Pb isotopic analyses are listed in table 2. The epsilon () values are a measure of
the deviations of the isotopic ratios in the samples from the expected value in a uniform reservoir.140 In Table 2 the epsilon
values for both Sr and Nd isotopes have been calculated, comparing the present-day isotopic ratios as measured in these
samples with the present-day isotopic ratios of the depleted mantle, the postulated source of the magmas. The positive
epsilon values imply that these rocks have greater 87Sr/86Sr and 143Nd/144Nd ratios than those in the depleted mantle source
region(s) of the magmas from which they cooled.141 Calculating the epsilon values for individual samples can be a test of
cogenicity or of contamination in the samples, or can be a measure of the extent to which the magmas, when they cooled to
form the rocks, had fractionated relative to their postulated depleted mantle source(s). The T DM notation refers to the
depleted mantle model ages calculated from the Nd isotope ratios, and is a measure of the length of time each sample has
been separated from the mantle from which it was originally derived.142It needs to be remembered that the basis of these
model age calculations is an assumption about the isotopic composition of the mantle source region from which the
magmas that cooled to form these samples were originally derived. Note that the calculated depleted mantle model ages for
these recent lava flows range from 724.5 Ma to 1453.3 Ma.
Table 2. Rb-Sr, Sm-Nd, and Pb-Pb isotopic analyses of five recent (1949, 1954, 1975) lava flows at Mt. Ngauruhoe, New
Zealand (Analyst: Dr. R. A. Armstrong, PRISE, Australian National University, Canberra; February 1997 March 2000).
Sample

1A

1B

Flow Date

February 11, 1949

2A

2B

June 4, 1954

3A

3C

June 30, 1954

4A

4B

July 14, 1954

5A

5B

February 19, 1975

Rb (ppm)

36.19

45.73

36.85

104.92

37.80

24.78

69.85

50.33

55.16

36.13

87

117.94

149.01

120.08

341.88

123.18

80.74

227.62

163.99

179.73

117.74

Sr (ppm)

237.27

203.03

249.44

228.36

238.86

236.23

239.50

269.70

232.20

304.81

86

267.17

228.62

280.87

257.13

268.95

266.00

269.68

303.70

261.45

343.22

Rb (nm/g)

Sr (nm/g)

87

86

Rb/ Sr

0.4414

0.65177

0.4275

1.32962

0.4580

0.30355

0.84405

0.5400

0.68745

0.3431

87

Sr/86Sr

0.70558

0.705153

0.70552

0.705461

0.70566

0.705401

0.705539

0.70516

0.705599

0.70529

Sr(t0)

+42.41

+36.34

+41.56

+40.72

+43.55

+39.87

+41.83

+36.47

+42.68

+38.29

Sm (ppm)

5.319

3.499

3.225

4.057

3.667

3.361

3.121

3.584

3.816

4.671

Sm(nm/g) 5.303

3.490

3.215

4.035

3.656

3.352

3.113

3.573

3.805

4.657

20.886

15.640

13.660

14.256

14.912

13.968

13.258

14.515

15.742

19.350

34.447

25.795

22.529

23.513

24.594

23.038

21.867

23.940

25.965

31.914

147

Nd (ppm)
144

Nd
(nm/g)
147

Sm/144Nd

0.15401

0.13528

0.14276

0.17161

0.14869

0.14550

0.14236

0.14933

0.14655

0.14595

143

Nd/144Nd

0.512751

0.512778

0.512731

0.512749

0.512704

0.512720

0.512740

0.512739

0.512742

0.512775

Nd(t0)

+1.38

+1.91

+0.99

+1.35

+0.47

+0.78

+1.17

+1.15

+1.21

+1.85

TDM (Ma)

1020.4

724.5

901.7

1453.3

1047.0

962.4

877.5

974.7

927.6

845.2

206

Pb/204Pb

18.842

18.802

18.821

18.838

18.832

18.828

18.797

18.841

18.833

18.805

207

204

Pb/ Pb

15.635

15.631

15.629

15.632

15.632

15.636

15.629

15.637

15.630

15.620

208

204

38.749

38.701

38.716

38.740

38.735

38.735

38.694

38.748

38.740

38.687

Pb/ Pb

Notes:
Measured, present-day 87Sr/86Sr ratios ( 2s) are normalized to 86Sr/88Sr = 0.1194
The 87Sr/86Sr ratios are reported relative to a value of 0.710207 26 ( 2s) for the NBS standard SRM 987 run with these
samples
Sr(t0) refers to the present-day, calculated value relative to a depleted mantle 87Sr/86Sr ratio today of 0.7026143, 144
Measured, present-day 143Nd/144Nd ratios ( 2s) are normalized to 146Nd/144Nd = 0.7219
Nd(t0) refers to the present-day, calculated value relative to a depleted mantle value of 0.51268 for BCR-1
206
Pb/204Pb ratios are normalized to NBS standard SRM 981 for mass fractionation, while all other ratios are calculated by a
double spiking routine.
Because these samples are from recent lava flows (only 2854 years old), the isotope ratios of these samples were not
expected to yield any age information. Nevertheless, a thorough analysis of the data was still undertaken to test for any
age information they might still yield. Selective plotting of the data does yield some seemingly valid isochrons, such as a 5point Rb-Sr isochron yielding an apparent age of 133 87 Ma and a 5-point Sm-Nd isochron yielding an apparent age of
197 160 Ma. The goodness of fit statistics for these two isochrons yield low MSWD values of 0.97 and 0.82 respectively
(mean square of weighted deviatesa measure of the ratio of observed scatter from the best-fit line to the expected scatter
from the assigned errors and error correlations). However, the probabilities of these fits being meaningful are only moderate
and the assigned error margins on each of the determined isotopic ratios required to constrain the fits are intolerably large,
resulting in final error margins that are more than 50% of the apparent isochron ages. Such selective manipulation of the
data is thus not only misleading, but completely meaningless. However, better apparent results are obtainable with the Pb
isotopic data, a 7-point isochron yielding a 207Pb-206Pb age of 3908 390 Ma. The statistics of this fit are much better, with
small error margins for each data point and a reasonable MSWD value of 1.07, but the probability of the fit is only moderate
and this apparent isochron has intercepts with the Pb isotope growth curve at -92 Ma and 3921 Ma. For comparison, the Pb
isotopic data also yield a 9-point 208Pb-206Pb line of best fit with a low MSWD value of 0.45 and a high probability of 0.87.
These outcomes would thus seem to have some validity and meaning to them, implying some significance to these trends in
the Pb isotopic data.
Discussion
The andesites of the Taupo Volcanic Zone exhibit compositional characteristics of continental arc lavas worldwide,145and it
is therefore probable that they were generated by similar processes, namely differentiation of mantle-derived basaltic
magmas.146 Petrographic studies of the basalts of the Taupo Volcanic Zone suggested to Gamble et al.147 that crystal
fractionation and accumulation processes could account for much of the diversity observed in their major-element
compositions. They concluded that the parental basaltic magmas must be broadly similar to each other, and thus much of
the diversity observed in the andesites can be attributed to secondary processes as the magmas ascended through
complex plumbing systems in the mantle and crust beneath this volcanic arc. Although primary andesitic magmas have
been postulated for some volcanic arcs,148 and some experimental studies indicate that andesitic magmas may be
generated from hydrated peridotite in the mantle above subduction zones,149, 150 it is now generally agreed that this
mechanism does not explain the origin and formation of the majority of the andesites in the Taupo Volcanic Zone, including
these recent andesite flows at Mt. Ngauruhoe. Few if any of these andesites have the required attributes (olivine with a
composition greater than Fo90, high Cr, Ni, and Mg, plus low phenocryst contents) to have been derived directly from an
andesitic magma generated in the mantle. Cole151, 152 favored an origin for these Taupo Volcanic Zone andesites from a
primary andesitic magma produced by the melting of subducting oceanic crust and assimilated greywacke and other
sediments dragged into the mantle with the subducting slab of oceanic crust, but melts from a subducted slab or the lower
crust are also unlikely to be of andesitic composition.153In one of the earliest investigations of the petrogenesis of these
volcanic rocks of the Taupo Volcanic Zone, not only were the lavas and pyroclastics analyzed, but also the Permian to
Jurassic interbedded greywackes, siltstones, and shales that are spatially related to, and underlie, the volcanics and that
were therefore regarded as a potential source of crustal contamination.154, 155 Using the Sr, Rb, K, U, and Th abundances
in these rocks and their 87Sr/86Sr ratios, it was concluded that the data were most consistent with the production of the
andesites by partial assimilation of sedimentary material by a basaltic magma derived from the upper mantle; the adjacent

greywackes, siltstones, and shales being the most likely sedimentary material; and the unassimilated gneissic xenoliths
probably representing the basement rocks to those sediments. Cole et al.156 found that the rare earth element
geochemistry of the andesites suggested the andesite magmas were generated in the upper mantle wedge associated with
the subducting slab of the Pacific Plate beneath the Australian Plate (fig. 1, inset), but some crustal contamination also
occurred. Indeed, there are several indicators that suggest the andesites have undergone crustal contamination prior to
eruption. These include the frequent occurrence of crustal xenoliths,157, 158 elevated and correlated radiogenic and stable
isotope ratios,159,160, 161 and relatively enriched LILE contents, significantly higher than those predicted by crystal
fractionation processes alone.162, 163Of the possible assimilants, the most likely is the meta-greywacke basement of the
Torlesse and/or Waipapa Terranes. Strong correlations between combinations of radiogenic and stable isotopes trend
generally towards the compositional fields of these basement rocks, either partly or wholly overlapping with Waipapa
compositions.164 In published models of the postulated assimilation, the more siliceous and radiogenic Torlesse metagreywacke was preferred to the Waipapa, since a substantially smaller volume was required to satisfy the resultant
petrogenesis models.165, 166, 167, 168 However, although much of the isotopic data for the andesites of the Taupo Volcanic
Zone would seem to be explained by binary mixing between primitive basalt magmas and metasedimentary rocks,169 the
actual process is undoubtedly more complex. Bulk assimilation is unlikely on thermochemical grounds,170 given the
relatively low magmatic temperatures and highly porphyritic nature of the andesites, suggesting that combined assimilation
and fractional crystallization (AFC)171 would have been operating. Thus, as proposed by Graham and Hackett, the true
assimilant was probably a minimum partial melt of meta-greywacke with higher SiO 2 and LILE than its bulk parent,172 but
with similar isotopic composition (assuming equilibrium melting has taken place).173
Fig. 4. Plot of Sr and Nd isotopic compositions of samples of the modern (19451996) Ruapehu andesites (after Gamble et
al.174) and the modern (1949, 1954, 1975) Ngauruhoe andesites of this study, showing how similar in isotopic composition
these andesites are.Graham and Hackett175 and Graham and Cole176 suggested that Type 1 andesites had been
generated from low-alumina basalt, such as a Ruapehu basalt, by AFC (assimilation and fractional crystallization), involving
POAM (plagioclase, olivine/orthopyroxene, augite, and magnetite)177 fractionation and assimilation of Torlesse metagreywacke. Indeed, Graham and Hackett178 used least squares geochemical modeling to show how the andesite magma
from which a 1954 Ngauruhoe lava cooled179 could have been generated from a parental basalt magma with the
composition of a Ruapehu basalt by a process of combined assimilation of crustal material (addition of 6% assimilant) and
crystal fractionation (0% removal of crystals). Furthermore, the presence of xenoliths in the Ngauruhoe andesite flows,
particularly the vitrified meta-greywacke and gneissic xenoliths, confirm conclusively that the assimilant was most likely a
partial melt of gneiss composed of meta-greywacke that was originally a greywacke in the adjacent greywacke-siltstoneshale sediments of the Torlesse Terrane, which outcrops to the east of both Mt. Ngauruhoe and Mt. Ruapehu, and which is
intersected in boreholes under the Taupo Volcanic Zone.180,181, 182
Given the close proximity of Ngauruhoe and Ruapehu to one another as adjacent volcanoes developed from related vents in
the Tongariro Volcanic Center,183, 184, 185 it would be expected that recent andesite lava flows from these two volcanoes
could well have been generated from the same source in a similar manner. Gamble et al.186 have compiled and reviewed
the available geochemical and isotopic data for andesite lava
flows between 1945 and 1996 from the Ruapehu volcano, and
concluded that collectively the data show trends with time of
increasing SiO2abundance and rising 87Sr/86Sr ratios. This is
consistent with broad control of the magma chemistry, and thus
the resultant andesite lavas, by assimilation and crystal
fractionation (AFC) processes. Furthermore, they found that
the magmas emplaced at Ruapehu during that 50 year period
show geochemical variability that spans most of the range
shown by lavas erupted over the entire history of the volcano.
In comparing the geochemical data of the Ngauruhoe andesite
lavas analyzed for this study with the Ruapehu andesites, it is
evident that there are major differences in the major element
compositions (for example, the Ruapehu andesites all have
higher SiO2 contents). However, the87Sr/86Sr and 143Nd/144Nd
ratios of the recent andesite lava flows from these two adjacent
volcanoes are very similar, as can be readily seen in Fig. 4.
Because it is the Sr-Nd-Pb isotope geochemistry of recent
lavas
that
is
used
to
elucidate
their
petrogenesis,187, 188, 189 it is not unreasonable therefore to
interpret the Sr-Nd-Pb isotope geochemistry of these recent
Ngauruhoe andesite lavas in conjunction with the comparable
and almost identical isotope data of the closely related recent andesite flows at Ruapehu, and within the framework of the
interpreted petrogenesis of the andesites of these two volcanoes.
Fig. 5. Sr-Nd isotope correlation diagram showing the Sr and Nd isotopic compositions of the recent Ngauruhoe lavas,
which plot within the compositional field of the spatially and genetically related Ruapehu andesite lavas. The mantle array is
defined by MORB and OIB.190 The field of the Torlesse greywacke is defined by two metasediment samples and two
vitrified metasediment xenoliths.191When plotted on a broader-scale Sr-Nd isotope correlation diagram the recent
Ngauruhoe andesite lavas not only plot within the field defined by the recent Ruapehu andesite lavas, but they plot to the
right of the mantle array, the band of isotope compositions defined by mid-ocean ridge basalts (MORB) and ocean island
basalts (OIB)192 (fig. 5). Also shown on Fig. 5 is the field of Sr and Nd isotopic compositions for the Torlesse greywacke,
defined by two metasediment samples and two vitrified metasediment xenoliths.193 Because the magmas that were
extruded as these recent Ngauruhoe and Ruapehu lavas were originally sourced in the mantle, their compositions should
have originally been in the field of the mantle array. (These magmas had to be sourced in the mantle because the
temperatures in the lower crust are insufficient to generate magmas of a basaltic composition, and because the lower crust
is of granitic composition whereas the mantle is basaltic.) Thus whereas these Ngauruhoe and Ruapehu magmas were
originally basaltic, by the time of their extrusion they were andesitic in composition. The data plotted on Fig. 5 suggest that
crustal contamination was responsible for the andesitic composition, and that the Torlesse greywacke has the isotopic
composition consistent with the shift in the magmas compositions away from the mantle array due to that crustal
contamination. Graham and Hackett194 found that within the broad field in which the recent Ruapehua lavas plot, to the
right away from the mantle array, there is a trend in the Type 1 lavas from basalt through andesite to dacite of
increasing 87Sr/86Sr and decreasing 143Nd/144Nd. Similar trends in volcanic rocks in other locations have been similarly

explained in terms of crustal assimilation.195 These data, including the recent Ngauruhoe andesites, are consistent with the
classification by Nohda196 of the Taupo Volcanic Zone as a continental island arc.This contamination trend that explains the
andesite compositions is more pronounced, and therefore graphically illustrated, in the 87Sr/86Sr versus Nd diagram of Fig. 6.
The recent Ngauruhoe andesites plot within and beyond the field defined by other Taupo Volcanic Zone Type 13
andesites.197 The majority of these Ngauruhoe Type 1 andesite lavas plot outside this field and would seem comparable to
the Taupo Volcanic Zone rhyolites that plot in the same area,198 perhaps suggesting an even greater degree of fractional
crystallization than the other andesites. The representative AFC (assimilation and fractional crystallization) curves were
constructed
(after
DePaolo199)
for
various
combinations of a parental Ruapehu basalt, Torlesse
and Waipapa assimilants, and bulk distribution
coefficients (DSr, DNd). It is immediately evident that the
AFC mixing curve for a nominal average Waipapa
Terrane meta-greywacke is tangential well to the right
of, and for >90% crystallization falls well short of, the
Type 13 andesites field; and the Ngauruhoe andesites
in particular. In contrast, the lowalumina Ruapehu
basaltTorlesse Terrane meta-greywacke AFC curves lie
within and to the right of the Type 13 andesites field, so
any substantial Waipapa component would push them
further to the right. This would thus appear to rule out
any Waipapa component. The curve best matching the
trend in the recent Ngauruhoe andesites data does
closely match one of these Ruapehu basaltTorlesse
AFC curves, with the degree of crystallization required
to explain the isotope shifts being between 40 and 70%.
Although this87Sr/86Sr versus Nd trend could be
interpreted as due to source-related processes, as
suggested in the Aleutian Arc,200 Graham et
al.201 found that this could not be true for the 18O
versus 87Sr/86Sr trend for the Taupo Volcanic Zone Type
13 andesites, where the initially sharp increase in 18O
with respect to 87Sr/86Sr is unmistakably due to crustal assimilation.202, 203 A similar explanation is evident for the andesite
of the Lesser Antilles Arc.204
Fig. 6. An 87Sr/86Sr versus Nd diagram showing where the Ngauruhoe andesites in this study plot in relation to the field of
compositions for Taupo Volcanic Zone Type 13 andesites.208 The representative AFC curves are for various combinations
of the postulated Ruapehu basalt parent magma,209 assimilant and bulk distribution coefficients (DSr, DNd). Tick marks
indicate the degree of crystallization and are given in intervals of 0.1 to a maximum of 0.9. In the inset the T-K field
represents the lavas of the Tonga-Kermadec Arc.
The inset to Fig. 6 shows the broader isotopic context of these recent Ngauruhoe andesites. Whereas the volcanics of the
Tonga-Kermadec Arc205 were generated from a mantle source less depleted than mid-ocean ridge basalts (MORB),206 it
would appear that the parent basaltic magmas to the Taupo Volcanic Zone Type 13 andesites were derived from an even
less depleted mantle source. Furthermore, the increase in 87Sr/86Sr that shifts both the Tonga-Kermadec lavas and the Taupo
Volcanic Zone andesites out of the mantle array207 is indicative of crustal contamination/assimilation, as are the high
positive Sr values for the recent Ngauruhoe andesites (table 2), although a portion of these large positive values could also
indicate a high degree of fractionation of the magmas relative to their postulated depleted mantle source. Nevertheless, the
extreme Sr and Nd isotopic enrichment of the Torlesse assimilant, demonstrated to be the most likely crustal contaminant in
these andesites, underlines how only a small proportion of this contaminant (postulated at around 6%)210 is required.Plots
of the Pb isotope data also graphically illustrate these trends (fig. 7). Gamble et al.211 found no systematic variation with
time in the Pb isotopic data from the recent Ruapehu andesites, while samples from the 19951996 Ruapehu lavas
yielded 206Pb/204Pb ratios which spanned the entire range of compositions from the previous 50 years. Nevertheless, the
modern andesitic lavas were less radiogenic than the majority of prehistoric Ruapehu andesites and Taupo Volcanic Zone
basalts, with all samples defining a potential mixing array towards the Torlesse Terrane meta-greywacke basement, the
major upper crustal component in the region and thus the putative crustal contaminant for most Taupo Volcanic Zone
magmas.212, 213, 214 Furthermore, the Pb isotope data would suggest that the Taupo Volcanic Zone basalts were derived
from a depleted mantle source that was more radiogenic than that from which the Pacific MORB was generated. Because
the parent basalt magmas that were contaminated to form the andesite lavas were probably from identical depleted mantle
sources as the magmas that produced the Taupo Volcanic Zone basalts, the less radiogenic portion of the andesite fields in
Fig. 7 suggest that the sources for all the parental basalt magmas were much closer to the composition of the depleted
mantle sources for the Pacific MORB, while the parental magmas for the Taupo Volcanic Zone basalts must have thus
suffered from some crustal contamination similar to the andesites, as suggested by the Sr-Nd isotopic data.215

Fig. 7. 207Pb/204Pb and 208Pb/204Pb versus206Pb/204Pb plots for


modern
(19451966)
and
prehistoric
Ruapehu
andesites,224Taupo Volcanic Zone basalts225, 226 and recent
Ngauruhoe andesites (this study). The Pb isotopic data
suggest contamination of a Pacific MORB source227 with the
Torlesse basement metasediments.228, 229 NHRL is the
Northern Hemisphere Reference Line.230
As already noted, the Pb isotopic data for the recent
Ngauruhoe andesite lavas are strongly linear, particularly
the 208Pb-206Pb data. The207Pb-206Pb data yield an apparent 7point isochron with a 3908 390 Ma age. Not only is this
age somewhat meaningless based on the statistics of the fit;
these data lie to the right of the geochron, the Pb isochron
corresponding
to
the
presumed
age
of
the
earth.216, 217, 218This is the lead paradox, which implies
that the depleted mantle has an average composition more
radiogenic than the geochron. This is the opposite of the
postulated behavior expected in conventional radiogenic
isotope geology. Nevertheless, these linear arrays in the
recent Ngauruhoe andesite data not only imply Pb isotopic
mixing of the assimilated Torlesse contaminant with the
parental basalt magmas, but also mixing of leads of differing
isotopic compositions in the depleted mantle source of these
magmas.219, 220 It would also seem highly significant that
the linear arrays of the recent Ngauruhoe andesite Pbisotopic data parallel the Pb isotopic compositional field for
the Taupo Volcanic Zone basalts, as well as being parallel to
the Northern Hemisphere Reference Line (NHRL)221 and
sub-parallel to the compositional field of Pacific
MORB.222 Indeed, this would appear to imply that much of
the Pb isotopic variation in both the recent Ngauruhoe
andesites and the Taupo Volcanic Zone basalts is probably
due to Pb isotopic variations in the depleted mantle source
region for the parental basalt magmas. Superimposed on this
mantle source Pb-isotopic variation is the Torlesse basement
contamination. This would explain the shift in the Ngauruhoe
andesite linear arrays upwards toward the Pb isotopic
compositional field for the Torlesse basement while still
parallel to the Pb-isotopic compositional field of the Taupo
Volcanic Zone basalts that are supposed to be similar to the
compositions of the basalt magmas parental to these recent
Ngauruhoe andesites.Using trace elements and their ratios,
Gamble et al.223 reviewed the question of source
heterogeneity and fertility of the basalt magmas of the TaupoKermadec Volcanic Arc systems, and concluded that magma
sources were MORB-like, as is also suggested by the Sr-NdPb isotopic data. Thus, using combinations of a MORB-source with typical Kermadec Trench-Hikurangi Trough sediments
and with Torlesse metasediments, Gamble et al.233undertook mixing calculations based on isotope pairs such
as 206Pb/204Pb versus 143Nd/144Nd and 87Sr/86Sr (fig. 8). It was thus found that the incorporation of a relatively small amount
(up to ~5%) of New Zealand continental sediment ~20 ppm Pb, ~23 ppm U and ~14 ppm Ththe average New Zealand
Torlesse metasediment of Graham et al.234 to basalt magmas from a MORB-source could have brought about massive
shifts in the isotopic compositions of the resulting basalts of the Kermadec-Taupo Arc, the sediment-hosted Pb having a
swamping affect on the mantle Pb. Indeed, these bulk mixing calculations showed that mixing <5% of a trench sediment,
with a composition such as that of sample A305, with MORB mantle embraced the entire Sr, Nd, and Pb isotope data array
of the Kermadec-Taupo Arc magmas (Fig. 8). Similar calculations based on trace elements produced multi-element
diagrams that mimicked multi-element diagrams for the arc basalts, displaying many features typical of subduction zone
magmas. Fig. 8 also has plotted on it the Sr-Pb and Nd-Pb data for the recent Ngauruhoe andesites, which sit between the
calculated bulk mixing curves for MORBtrench sediment and MORBTorlesse basement, as do most of the basaltic
samples in the Gamble et al.235 study. Note that the overall range of isotopic compositions for the Torlesse
metasediments236 spans the range of isotopic compositions between these two bulk mixing curves, so it can be reasonably
concluded that the recent Ngauruhoe andesite lavas resulted from the contamination of a basaltic magma from a MORBsource with around 5% Torlesse basement metasediment. Thus the Gamble et al.237 study confirms the conclusions
reached previously by Graham and Hackett.238
Fig. 8. Plots of 87Sr/86Sr and 143Nd/144Nd versus 206Pb/204Pb for the recent Ngauruhoe andesites. Calculated bulk mixing
curves for MORB-sediment A305 (continuous line) and MORB-Torlesse basement (dashed line) are shown.231 T is the
average composition of the Torlesse metasediments and the horizontal bar delineates the range of their
compositions.232 Tick marks are percent of sediment added to the MORB-source end-member.However, whereas Graham
and Hackett239 regarded this crustal contamination as being the secondary process of crustal assimilation whereby magma
compositions are influenced en route to the surface through the crust, Gamble et al.240 concluded that the crustal
contamination was due to the source-modifying process of sediment subduction into the mantle. The two processes are
fundamentally different. However, in the continental New Zealand situation, as a result of the oblique plate convergence and
rapid uplift, the sediments which are being subducted offshore in the Hikurangi Trough are essentially the same as the crust
made up of Torlesse metasediment basement which is available for assimilation.241 From modeling based upon Sr and Nd
isotope data, Gamble et al.242 had concluded that small amounts of crustal assimilation (up to 10%) could account for the
compositions of the basalts in the Taupo Volcanic Zone. However, Gamble et al.243conclusively demonstrated, from isotope
and multi-element variations with latitude in the volcanics and ocean-floor sediments, that the amounts of sediments which

contributed to the sources of the arc magmas had a diminishing continental contribution from ~5% in the south (Hikurangi
Trough) to <1% in the north (Kermadec Trench). This is consistent also with currently available seismic and sediment
dispersion data along the Kermadec-Hikurangi Margin. These data indicate northward thinning of the trench fill
deposit.244 Furthermore, since there is no continental crust as basement to the Kermadec Island Arc, the crustal
contamination there of <1%, derived from the bulk mixing curves on the Sr-Pb and Nd-Pb plots (fig. 8), had to be contributed
by subduction of the trench sediments. Thus it can be concluded that the ~5% crustal contamination in the Taupo Volcanic
Zone lavas, including the recent Ngauruhoe andesites, was similarly contributed by subduction of the thicker sediments in
the Hikurangi Trough to the east of the Taupo Volcanic Zone.The petrogenetic model favored by Gamble et al., 245 which is
consistent with all the isotopic data discussed, is shown in Fig. 9, based on Tatsumi246 and Davies and Stevenson.247 This
model envisages a zone of melt formation approximately coincident to the volcanic front, which includes Ruapehu and
Ngauruhoe, and a melt generation region delimited by the interface of the subducting slab, the base of the arc lithosphere
(of continental New Zealand) and two vertical columns, one delineating the volcanic front, the other, the coupled back-arc
basin. Fluids liberated from the descending slab ascend into and enrich the overlying peridotitic mantle wedge. In the region
immediately adjacent to the slab mineral reactions would probably have stabilized amphibole. Sediment scraped from the
upper surface of the slab is incorporated into the mantle wedge along the slab-mantle interface. Mantle flow parallel to the
slab-wedge interface carries the amphibole peridotite down to higher pressures. There the amphibole breaks down, giving
rise to amphibole dehydration, while progressive dehydration reactions in the slab itself lead to fluid transfer from the slab
into the mantle wedge. Both processes produce partial melting as amphibole breaks down over the depth range 112 19
km.248, 249 The lower density melt then rises and pools in the upwelling melt column, eventually penetrating upwards into
the overlying arc lithosphere to fill magma chambers that then erupt when full.
Fig. 9. Dynamic petrogenetic model
for andesite magma genesis beneath
the Kermadec-Taupo Volcanic Arc
subduction system.250, 251, 252 Flow
lines (arrows with double lines) show
mantle flowing from the back-arc
region into the mantle wedge, where
the isotherms are inverted owing to
the cooling effect of the cold
subducting
Cretaceous
slab.253 Sediment that was deposited
on the oceanic crust and thus also
subducted is mixed into the wedge
assemblage along the interface.
Progressive dehydration reactions in
the slab lead to fluid transfer from the
slab into the mantle wedge. In the
juxtaposed wedge, amphibole (amph)
is stabilized, but then breaks down
over the depth range 112 19
km,254 inducing partial melting. In the
resulting melt column, the first formed
melts accumulating closest to the
slab-mantle interface will be most
susceptible to fluxing from the slab.
Above this zone, melting will continue.
The rising melts will eventually pool in
the melt column, and the resulting
magma finally ascends into the
overlying arc lithosphere along fracture conduits, filling magma chambers and triggering eruptions.It is abundantly clear,
therefore, from the foregoing discussion and the isotopic data yielded by these recent Ngauruhoe andesite flows, that no
age information is provided by the Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic systems. This is even true of the calculated
depleted-mantle Nd model ages, which range from 724.5 Ma to 1453.3 Ma. By definition these model ages are a measure
of the length of time each sample has been separated from the mantle from which it was originally derived.255 But all the
evidence is consistent with the magmas responsible for these recent lava flows having formed only recently from partial
melting of the mantle wedge beneath the volcano. The subduction of the Pacific Plate beneath New Zealand only began
recently, even in conventional terms, because the subducting slab is regarded in conventional terms as
Cretaceous.256 Even the Torlesse metasediments that underlie the Taupo Volcanic Zone and are adjacent to it yield a RbSr whole-rock isochron age of only 141 3 Ma.257 This is interpreted to represent the timing of lowgrade metamorphism
of these sediments which are regarded in conventional terms to be of Permian to Jurassic age (between 290 and 142 Ma).
However, these Nd model ages are a product of the Sm-Nd radioisotopic system within the mantle wedge that has been
partially melted near the interface with the subducting slab. Since the subduction is only recent, the melt that separated was
formed only recently. Thus these Nd isotopic signatures have been inherited from the partial melting of the mantle wedge,
which of course has not occurred over a period of 400500 million years, as these Nd model ages (table 2) would otherwise
imply. Therefore the Nd isotopes, as the Sr and Pb isotopes, have only been used to elucidate the genesis and history of
these recent Ngauruhoe andesite flows. The radioisotopic ratios are characteristic of the heterogeneity in the mantle source
of the magmas, and of the crustal contamination assimilated in the original basaltic magmas. These conclusions support the
contention of Snelling258 that the radioisotopes in these lava flows are an artifact of the mixing of mantle and crustal
sources; and that the radioisotopic ratios, which could be interpreted in terms of millions of years of radioactive decay, are
instead geochemical characteristics of the mantle and crustal rocks that were produced at their origin and during their
subsequent history. The obvious implication of this, ignored in the conventional scientific literature, is that if the radioisotopic
ratios do not provide age information on recent lava flows because they instead represent fundamental geochemical
signatures of the mantle and crustal sources, then the radioisotopic ratios in ancient lava flows also cannot be relied upon to
give valid age information, because they too are the products of the sources and history of the mantle and crustal rocks
that produced them.This conclusion does not deny the occurrence of radioisotopic decay in rocks and minerals within the
geologic record, because there is physical evidence of that decay occurring (such as mature 238U and 232Th radiohalos in

granitic rocks.259 It has been suggested that such radioisotopic decay must have occurred at an accelerated rate during
distinct events in the earths history, such as during the early part of the Creation week and during the Flood. 260 Such
accelerated decay would have left the mantle and crustal sources of magmas with radioisotopic signatures that now yield
apparent ages in conventional terms representing millions and billions of years of nuclear decay at todays rates, thus
endowing both recent and ancient lava flows with meaningless radioisotopic ages, except perhaps in a relative sense
within the creation framework for the earths true history.Furthermore, if radioisotopic decay was accelerated during the
Flood, the resultant accelerated generation of heat is likely to have powered the acceleration of other geologic processes,
such as the processes associated with plate tectonics. These would have been responsible during the Flood for
catastrophically subducting the pre-Flood ocean floor,261 and mixing mantle and crustal rocks and their radioisotopic
contents. At the close of the Flood these plate tectonic processes would have decelerated to their current almost
imperceptible rates, but the tectonic framework for the development of the Taupo Volcanic Zone of New Zealands North
Island would have been put in place. Geophysical investigations indicate that the Pacific Plate today continues to be
obliquely subducted beneath the Australian Plate, on which most of New Zealands North Island sits. The volcanoes of the
Taupo Volcanic Zone, including Ngauruhoe and Ruapehu in the Tongariro Volcanic Center, are about 80 km directly above
the subducting Pacific Plate. A zone of earthquakes reveals where the movement is still taking place.262 Partial melting in
the mantle wedge adjacent to the interface with the subducting slab provides a melt with inherited radioisotopic ratios that
rises through the mantle wedge from below 80 km depth in the melt column (Fig. 9). The melt then ascends through
fractures and conduits in the crust above, which is probably less than 20 km thick,263 to slowly replenish magma chambers
that have in recent decades occasionally erupted to extrude the andesite lava flows. This combination of processes confirms
that the radioisotopic ratios in recent lavas are fundamental geochemical characteristics of these rocks, and have no age
significance. It is a reasonable general implication that radioisotopic ratios must be among the fundamental geochemical
characteristics of ancient lava flows also, reflecting their origin and history rather than indicating valid ages.
Conclusions
The Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic ratios in these samples of the recent (19491975) andesite lava flows at Mt.
Ngauruhoe, New Zealand, as anticipated, do not yield any meaningful age information, even with selective manipulation of
the data. Instead, these data provide evidence of the mantle source of the lavas, of magma genesis, and of crustal
contamination of the parental basalt magmas. Subduction of the Pacific Plate beneath the Taupo Volcanic Arc has carried
trench sediments with itsediments identical in composition to the Torlesse metasediment basement underlying, and
outcropping adjacent to, these volcanoes. Scraped off the subducting slab, the sediments have contaminated the basalt
magmas generated by partial melting of the peridotitic mantle wedge at the mantle-slab interface. The resultant andesite
magmas rose in the melt column through the mantle wedge, and then ascended through fracture conduits in the overlying
crust into magma chambers below the volcanoes that erupted when full.The Sr-Nd-Pb radioisotopic systematics are thus
characteristic of the depleted mantle source, modified by mixing with the crustal contaminant. Variations in the depleted
mantle Nd model ages, which range from 724.5 to 1453.3 Ma, and which are meaningless in this recent (even in
conventional terms) tectonic and petrogenetic framework, and the Pb isotopic linear arrays, indicate geochemical
heterogeneity in the mantle wedge. Thus the radioisotopic ratios in these recent Ngauruhoe andesite lava flows were
inherited from both the peridotitic mantle wedge and the subducted trench sediments, and are fundamental characteristics
of their geochemistry. They therefore only reflect the origin and history of the mantle and crustal sources from which the
magma was generated, and therefore have no age significance.By implication, the radioisotopic ratios in ancient lavas found
throughout the geologic record are likely fundamental characteristics of their geochemistry. They therefore probably only
reflect the magmatic origin of the lavas from mantle and crustal sources, and any history of mixing or contamination in their
petrogenesis, rather than any valid age information. Even though radioisotopic decay has undoubtedly occurred during the
earths history, conventional radioisotopic dating of these rocks therefore does not necessarily provide valid absolute ages
for them. This is especially so if accelerated nuclear decay accompanied the catastrophic operation of those geologic and
tectonic processes responsible for the mixing of the radioisotopic decay products during magma genesis.
Acknowledgments
The initiation of this research at Mt. Ngauruhoe, New Zealand, was made possible by the logistical support of Answers in
Genesis (Australia), who funded both the sample collection and the whole-rock geochemical analyses. The subsequent
continuing support for this research by the Institute for Creation Research is gratefully acknowledgedparticularly its
funding of the significant cost of the Rb-Sr, Sm-Nd, and Pb-Pb radioisotopic analyses as part of the RATE (Radioisotopes
and the Age of The Earth) project.

The Cause of Anomalous Potassium-Argon Ages for Recent Andesite Flows at Mt. Ngauruhoe, New Zealand, and
the Implications for Potassium-Argon Dating
by Dr. Andrew A. Snelling on December 30, 2009
Abstract
New Zealands newest and most active volcano, Mt. Ngauruhoe in the Taupo Volcanic Zone, produced andesite flows in
1949 and 1954 and avalanche deposits in 1975. Potassium-argon dating of five of these flows and deposits yielded K-Ar
model ages from <0.27 Ma to 3.50.2 Ma. Dates could not be reproduced, even from splits of the same samples from
the same flow, the explanation being variations in excess 40Ar* content. A survey of anomalous K-Ar dates indicates they
are common, particularly in basalts, xenoliths, and xenocrysts such as diamonds that are regarded as coming from the
upper mantle. In fact, it is now well established that there are large quantities of excess 40Ar* in the mantle, which in part
represent primordial argon not produced byin situ radioactive decay of 40K and not yet outgassed. And there are mantle
crust domains between, and within which, argon circulates during global tectonic processes, magma genesis, and mixing of
crustal materials. This has significant implications for the validity of K-Ar and 40Ar/39Ar dating.
Shop Now
Keywords: andesite, 19491975 flows, Mt. Ngauruhoe, New Zealand, potassium-argon dating, anomalous model ages,
excess 40Ar*, excess40Ar* in rocks and minerals, upper mantle, geochemical reservoirs, mantle-crust domains, crustal
mixing, magma genesis
This paper was originally published in the Proceedings of the Fourth International Conference on Creationism, pp. 503525
(1998), and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh
(www.csfpittsburgh.org).
Introduction
Mt. Ngauruhoe is an andesite stratovolcano of 2291 m elevation, rising above the Tongariro volcanic massif within the
Tongariro Volcanic Center of the Taupo Volcanic Zone, North Island, New Zealand (fig. 1).1,2 Though not as well publicized

as its neighbor, Mt. Ruapehu (about 12 km to the south), Ngauruhoe is an imposing, almost perfect cone that rises more
than 1000 m above the surrounding landscape. Eruptions from a central 400 m diameter crater have constructed the steep
(33) outer slopes of the cone.3, 4
Geologic Setting
The Taupo Volcanic Zone, a volcanic arc and marginal basin of the Taupo-Hikurangi arc-trench (subduction) system, 5 is a
southward extension on the Tonga-Kermadec arc into the continental crustal environment of New Zealands North Island. It
has been interpreted as oblique subduction of the Pacific plate beneath the Australian plate. The zone extends
approximately 300 km north-northeast across the North Island from Ohakune to White Island (fig. 1) and is up to 50 km wide
in the central part, narrowing northward and southward. This volcanotectonic depression (Taupo-Rotorua depression6)
comprises four rhyolitic centers (Rotorua, Okataina, Maroa, and Taupo), plus the calc-alkaline Tongariro Volcanic Center,
part of a young (<0.25 Ma) andesite-dacite volcanic arc with no associated rhyolitic volcanism extending along the eastern
side of the zone.7The Tongariro Volcanic Center extends for 65 km south-southwest from Lake Taupo at the southern end of
the Taupo Volcanic Zone (fig. 1) and consists of four large predominantly andesite volcanoes (Kakaramea, Pihanga,
Tongariro, and Ruapehu; fig. 2); two smaller eroded centers at Maungakatote and Hauhungatahi; a satellite cone and
associated flows at Pukeonake; and four craters at Ohakune (fig. 2).8, 9
Fig. 1. The location of Mt. Ngauruhoe in the Taupo Volcanic Zone (TVZ), New Zealand, showing the main structural
features. The shaded area is the andesite arc, and the inset
shows the major components of the boundary between the
Australian and Pacific Plates in the New Zealand region (arrows
indicate relative motions). Solid triangles are basalt-andesite
volcanoes.15
Most vents lie close to the axis of a large graben in which
Quaternary volcanic rocks overlie a basement of Mesozoic
greywacke and Tertiary sediments.10, 11 North-northeasttrending normal faults with throws up to 30 m cut the volcanoes
within the graben. Nearly all vents active within the last 10 ka lie
on a gentle arc which extends 25 km north-northeast from the
Rangataua vent on the southern slopes of Ruapehu through
Ruapehu summit and north flank vents, Tama Lakes, Ngauruhoe,
Red Crater, Blue Lake, and Te Mari craters. None of the young
vents lie on the mapped faults, which mostly downthrow toward
the axis of the graben. The vent lineation lies above this axis,
which is considered to mark a major basement fracture that
allows the intrusion of andesite dikes.12, 13, 14The Tongariro
volcanics unconformably overlie late Miocene marine siltstones
beneath Hauhungatahi, and a minimum age for the onset of
volcanism is measured by the influx of andesite pebbles in early
Pleistocene conglomerates of the Wanganui Basin to the
south.16, 17The oldest dated lavas from the Tongariro massif are
hornblende andesites exposed at Tama Lakes between
Ngauruhoe and Ruapehu, at 0.260.003 Ma; from Ruapehu,
0.230.006 Ma; and from Kakoramea, 0.220.001 Ma
(potassium-argon dates).18Tongariro itself is a large volcanic
massif that consists of at least twelve composite cones, the
youngest and most active of which is Ngauruhoe. A broad
division has been made into older (>20 ka) and younger (<20 ka)
lavas.19, 20 There is a north-northeast alignment of the younger vents of Tongariro, particularly evident between Te Mari and
Ngauruhoe.
Ngauruhoe
Ngauruhoe is the newest cone of the Tongariro massif and has been active for at least 2.5 ka. 21, 22, 23 It has been one of
the most active volcanoes in New Zealand, with more than seventy eruptive episodes since 1839, when the first steam
eruption was recorded by European settlers.24, 25, 26 Prior to European colonization the Maoris witnessed many eruptions
from the mountain.27 The first lava eruption seen by European settlers occurred between April and August 1870, with two or
three flows witnessed spilling down the northwestern flanks of the volcano on July 7.28, 29 Following that event there have
been pyroclastic (ash) eruptions every few years,30 with
major explosive activity in AprilMay 1948.
Fig. 2. Location and deposits of the Tongariro Volcanic
Center.41, 42
The next lava extrusion was in February 1949, beginning
suddenly with ejection of incandescent blocks, and a series of
hot block and ash flows down the northwestern slopes on
February 9.31, 32 The southern sub-crater filled with lava,
which by late on February 10 had flowed over the lowest part
of the rim and down the northwest slopes of the cone. By
February 12 the flow had ceased moving, subsequent
mapping placing its volume at about 575,000 m3 (fig.
3).33, 34 Further explosive pyroclastic (ash) eruptions
followed, reaching a maximum about February 19-21 . The
eruptions ended on March 3.The eruption from May 13, 1954,
to March 10, 1955, began with explosive ejection of ash and
blocks, although red-hot lava had been seen in the crater five
months previous.35, 36 The eruption was remarkable for the
estimated large volume of almost 8 million m3 of lava that then
flowed from the crater from June through September 1954,
and was claimed to be the largest flow of lava observed in

New Zealand (that is, by the European settlers).37, 38 The lava was actually expelled from the crater in a series of
seventeen distinct flows on the following dates:39, 40
June 4, 30
July 8, 9, 10, 11, 13, 14, 23, 28, 29, 30
August 15 (?), 18
September 16, 18, 26
Fig. 3. Map of the northwestern slopes of Mt. Ngauruhoe
showing the lava flows of 1949 and 1954 and the 1975
avalanche deposits.48, 49, 50, 51, 52 The location of samples
collected for this study are marked.
Fig. 3 shows the distribution of those 1954 lava flows that are
still able to be distinguished on the northwestern and western
slopes of Ngauruhoe. All flows were of aa lava (as was the
February 1949 flow), typified by rough, jagged, clinkery
surfaces made up of blocks of congealed lava. The lava flows
were relatively viscous, some being observed at close quarters
slowly advancing at a rate of about 20 cm per
minute.43, 44, 45 The August 18 flow was more than 18 m thick
and still warm almost a year after being erupted. Intermittent
explosive eruptions and spectacular lava fountaining during
June and July 1954 built a spatter-and-cinder cone around the
south sub-crater, modifying the western summit of the
mountain. Activity decreased for two months after the last of
the lava flows on September 26, but increased again during
December 1954 and January 1955 with lava fountaining and
many highly explosive pyroclastic (ash) eruptions. The last ash
explosion was reported on March 10, 1955, but red-hot lava
remained in the crater until June 1955.46, 47After the 1954
1955 eruption, Ngauruhoe steamed semi-continuously, with numerous small eruptions of ash derived from comminuted vent
debris. Incandescent ejecta were seen in January 1973, and ash erupted in December 1973 contained juvenile glassy
andesite shards.53 Cannon-like, highly explosive eruptions in January and March 1974, the largest since 19541955, threw
out large quantities of ash and incandescent blocks, one of which was reported as weighing 3000 tonnes and thrown 100
m.54, 55 Pyroclastic avalanches flowed from the base of large convecting eruption columns, down the west and north slopes
of the cone, and the crater became considerably shallower.56, 57A series of similar but more violent explosions occurred on
February 19, 1975, accompanied by clearly visible atmospheric shock waves and condensation clouds.58, 59, 60, 61 Ash and
blocks up to 30 m across were ejected and scattered within a radius of 3 km from the summit. The series of nine cannonlike, individual eruptions followed a 1.5 hour period of voluminous gas-streaming emission, which formed a convecting
eruption plume between 11 km and 13 km high.62, 63, 64 The explosions took place at 2060 minute intervals for more than
five hours. Numerous pyroclastic avalanches were also generated by fallback from the continuous eruption column, the
avalanches consisting of a turbulent mixture of ash, bombs, and larger blocks which rolled swiftly down Ngauruhoes sides
at about 60 km per hour.65, 66 The deposits from these avalanches and the later explosions accumulated as sheets of
debris in the valley at the base of the cone, but did not extend beyond 2 km from the summit. It is estimated that a minimum
bulk volume of 3.4 million m3 of pyroclastic material was erupted in the seven-hour eruption sequence on that day.51 Fig. 3
shows the location of these avalanche deposits.There have been no eruptions since February 1975. A plume of steam or
gas is still often seen above the summit of the volcano, as powerful fumaroles in the bottom of the crater discharge hot
gases. However, the temperature of these fumaroles in the crater floor has steadily cooled significantly since 1979,
suggesting that the main vent is becoming blocked.
Sample Collection
Field work and collection of samples was undertaken in January 1996. The Ngauruhoe area was accessed from State
Highway 47 via Mangateopopo Road. From the parking area at the end of the road, the Mangateopopo Valley walking trail
was followed to the base of the Ngauruhoe cone, from where the darker-colored recent lava flows were clearly visible and
each one easily identified on the northwestern slopes against the lighter-colored older portions of the cone (fig. 3).Eleven 2
3 kg samples were collected: two each from the February 11, 1949, June 4, 1954, and July 14, 1954, lava flows and from
the February 19, 1975, avalanche deposits; and three from the June 30, 1954 lava flows. The sample locations are marked
on Fig. 3. Care was taken to ensure correct identification of each lava flow and that the samples collected were
representative of each flow and any variations in textures and phenocrysts in the lavas.
Laboratory Work
All samples were sent first for sectioning one thin section from each sample for petrographic analysis. A set of representative
pieces from each sample (approximately 100 g) was then dispatched to the AMDEL Laboratory in Adelaide, South Australia,
for whole-rock major, minor, and trace element analyses. A second representative set (50100 g from each sample) was
sent progressively to Geochron Laboratories in Cambridge (Boston), Massachusetts, for whole-rock potassium-argon (K-Ar)
datingfirst a split from one sample from each flow, then a split from the second sample from each flow after the first set of
results was received, and finally, the split from the third sample from the June 30, 1954, flow.At the AMDEL Laboratory each
sample was crushed and pulverized. Whole-rock analyses were undertaken by total fusion of each powdered sample and
then digesting them before ICP-OES for major and minor elements, and ICP-MS for trace and rare earth elements. Fe was
analyzed for amongst the major elements by ICP-OES as Fe 2O3 and reported accordingly, but separate analyses for Fe as
FeO were also undertaken via wet chemistry methods. The detection limit for all major element oxides was 0.01%. For
minor and trace elements the detection limits varied between 0.5 and 20 ppm, and for rare earth elements between 0.5 and
1 ppm.The potassium and argon analyses were undertaken at Geochron Laboratories under the direction of Richard
Reesman, the K-Ar laboratory manager. No specific location or expected age information was supplied to the laboratory.
However, the samples were described as andesites that probably contained low argon and therefore could be young, so as
to ensure the laboratory took extra care with the analytical work.Because the sample pieces were submitted as whole rocks,
the K-Ar laboratory undertook the crushing and pulverizing preparatory work. The concentrations of K 2O (weight %) were
then measured by the flame photometry method, the reported values being the averages of two readings for each
sample.67, 68 The 40K concentrations (ppm) were calculated from the terrestrial isotopic abundance using the measured
concentrations of K2O. The concentrations in ppm of 40Ar*, the supposed radiogenic 40Ar, were derived using the

conventional formula from isotope dilution measurements on a mass spectrometer by correcting for the presence of
atmospheric argon whose isotopic composition is known.69 The reported concentrations of 40Ar* are the averages of two
values for each sample. The ratios 40Ar*/Total Ar and 40Ar/36Ar are also derived from measurements on the mass
spectrometer and are also the averages of two values for each sample.
Table 1. Whole-rock, major-element oxide analyses of recent lava flows at Mt. Ngauruhoe, New Zealand, as reported in the
literature.
1

SiO2

56.63

57.24

55.90

56.22

56.2

55.83

55.73

56.05

TiO2

0.81

0.81

0.76

0.76

0.8

0.76

0.77

0.84

Al2O3

16.71

16.75

16.90

16.63

16.6

17.03

17.23

16.08

Fe2O3

1.16

1.54

2.10

2.37

1.4

2.14

2.13

3.24

FeO

7.00

6.44

6.30

6.14

7.0

6.36

6.47

5.57

MnO

0.16

0.12

0.15

0.15

0.1

0.16

0.16

0.15

MgO

4.85

4.58

5.20

5.24

5.2

4.79

4.89

5.03

CaO

8.16

7.95

8.40

8.31

8.3

8.37

8.52

7.93

Na2O

2.85

2.74

2.60

3.14

3.1

2.93

2.86

2.82

K2O

1.16

1.49

1.00

1.14

1.2

1.20

1.08

1.23

P2O5

0.16

0.17

0.10

0.17

0.2

0.15

0.15

0.15

H2O

0.31

0.24

0.06

0.19

N/A

0.14

0.10

0.38

TOTAL

99.95

100.07

99.47

100.46

100.1

99.86

100.09

99.47

Ejecta from March 1928 eruption73 74


Lava from February 1949 eruption75, 76, 77
Lava from June 30, 1954, flow78, 79, 80
1954 lava (VU 29250)81
1954 lava (VU 29250)82, 83
Average of four lava flows from 1954 eruptions84, 85
Average of five blocks and bombs from January and March 1974 eruptions86, 87
Lapilli from February 19, 1975, eruption88
Petrography and Chemistry
Clark reported that most of the flows from Ngauruhoe are labradorite-pyroxene andesite with phenocrysts of plagioclase
(labradorite), hypersthene, and rare augite in a hyalopilitic (needle-like microlites set in a glassy mesostasis) groundmass
containing abundant magnetite.70 However, all lava, lapilli, and incandescent blocks that have been analyzed from
eruptions this century also contain olivine; chemically they may be classed as low-silica (or basaltic) andesites (using the
classification scheme of Gill).71 The published analyses in Table 1 show only trivial changes in composition between 1928
and 1975. In fact, the 1954 and 1974 andesites are so similar that Nairn, et al., suggested that a solid plug of 1954 andesite
was heated to incandescence and partially remobilized on top of a rising magma column in 1974.72 This plug was disrupted
and blown from the vent as ejecta ranging in texture from solid blocks, through expanded scoria to spatter bombs.Table 2
lists the whole-rock major element analyses of the eleven samples collected in this study. Comparison of the data for each
flow with the corresponding data in Table 1 indicates that in their bulk chemistries all the samples analyzed (and thus all the
flows) are virtually identical to one another, the trivial differences being attributable to the statistics of analytical errors,
sampling, and natural variations. Thus it is not unreasonable to conclude that these basaltic andesites are cogenetic,
coming from the same magma and magma chamber, even as they have been observed to flow from the same volcano.
Nevertheless, Nairn, et al., suggested that even though the 1949 and 1954 lavas were both olivine-bearing andesite, the
chemical analyses (table 1) showed the 1954 lava to be slightly more basic than the 1949 lava, with slightly higher MgO,
CaO, and total iron oxides, but lower SiO 2 and alkalis.89 However, these trends are not duplicated with any statistical
significance by the analytical results of this study (table 2). At least, they found that their analyses of the 1974 lava blocks
and bombs were identical within the limits of error with the 1954 lava (table 1), which was also substantiated in this study
with respect to the 1975 avalanche material and the 1954 lava (table 2).
Table 2. Whole-rock, major-element oxide analyses of five recent lava flows at Mt. Ngauruhoe, New Zealand (Analyst:
AMDEL, Adelaide; April 1996.)
1A

1B

2A

2B

3A

3B

3C

4A

4B

5A

5B

SiO2

56.7

56.2

55.3

55.8

56.3

55.9

55.6

56.1

55.6

56.0

55.4

TiO2

0.79

0.85

0.74

0.77

0.76

0.75

0.74

0.75

0.84

0.79

0.78

Al2O3

17.2

17.3

16.5

17.3

17.0

16.9

16.7

16.9

17.5

17.0

16.5

Fe2O3*

9.10

9.63

9.26

9.23

9.11

9.17

9.59

9.29

9.61

9.25

9.43

MnO

0.15

0.16

0.15

0.15

0.15

0.15

0.16

0.15

0.16

0.15

0.16

MgO

4.28

3.84

5.21

4.71

4.75

5.00

5.09

4.71

3.84

4.31

5.27

CaO

7.61

7.93

8.22

8.29

7.95

8.16

8.17

8.00

8.17

7.83

8.56

Na2O

3.08

3.19

2.91

3.03

3.06

2.98

2.95

3.02

3.11

3.08

2.86

K2O

1.15

1.01

1.05

1.00

1.10

1.08

1.06

1.12

1.04

1.10

1.09

P2O5

0.13

0.13

0.12

0.12

0.13

0.13

0.13

0.13

0.13

0.14

0.14

L0I

0.42

TOTAL 100.51

0.48

0.51

0.37

0.38

0.50

0.53

0.62

0.42

0.70

0.41

100.72

99.97

100.77

100.69

100.72

100.72

100.79

100.42

100.35

100.60M

*Total Fe as Fe2O3
February 11, 1949, flow, samples A and B
June 4, 1954, flow, samples A and B
June 30, 1954, flow, samples A, B, and C
July 14, 1954, flow, samples A and B
February 19, 1975, flow, samples A and B
Clark and Cole recognized five lava types in the Tongariro Volcanic Center based on the modal proportions of the
phenocryst minerals.90,91 Graham modified this scheme to six types based on a combination of mineralogy and chemistry,
but given their uniform bulk chemistry and petrology, these Ngauruhoe lava flows group together as plagioclase-pyroxene
andesite within Grahams Type 1.92 Cole, et al., have described Type 1 lavas as volumetrically dominant within the
Tongariro Volcanic Center and as exhibiting coherent chemical trends with increasing silica content.93 They are relatively
Fe-rich and follow a typical calc-alkaline trend on the AFM diagram.Adapting the terminology of Gill, the Ngauruhoe lavas
are described as basic andesites (5358 wt% SiO2).94, 95 Their designation as plagioclase-pyroxene andesites is based on
the predominant phenocrysts present, with plagioclase greater than or equal to pyroxene. Two modal analyses are listed in
Table 3 which very closely resemble the samples collected for this study.All samples of the five lava flows examined in this
study exhibited a porphyritic texture, with phenocrysts (up to 3 mm across) consistently amounting to 3540% by volume.
The phenocryst assemblage is dominated (2:1) by plagioclase, but orthopyroxene and augite (clinopyroxene) are always
major components, while olivine and magnetite are only present in trace amounts. This POAM phenocryst assemblage is a
typical anhydrous mineralogy.98 The groundmass consists of microlites of plagioclase, orthopyroxene, and clinopyroxene,
and is crowded with minute granules of magnetite and/or Fe-Ti oxides. Small amounts (910%) of brown transparent (acidresiduum) glass are also present, and the overall texture is generally pilotaxitic.
Table 3. Modal analyses of two recent lava flows at Mt. Ngauruhoe, New Zealand, as reported in the literature.
g* in groundmass
Component
1
2
Ngauruhoe VU 29250, a 1954 flow.96
Plagioclase
22.6
21.6
Olivine-bearing low-Si andesite, June 30, 1954, Ngauruhoe flow.97
Steiner stressed that xenoliths are a common constituent of the 1954 Ngauruhoe
Augite
2.6
2.6
lava, but also noted that Battey reported the 1949 Ngauruhoe lava was rich in
xenoliths. All samples in this study contained xenoliths, including those from the
Orthopyroxene
6.0
5.8
1975 avalanche material.99, 100However, many of these aggregates are more
Olivine
0.2
0.2
accurately described as glomerocrysts and mafic (gabbro, websterite)
nodules.101 They are 35 mm across, generally have hypidiomorphic-granular
Iron Oxide

g*
textures, and consist of plagioclase, orthopyroxene, and clinopyroxene in varying
proportions, and very occasionally olivine. The true xenoliths are often rounded
Xenoliths
2.6
4.5
and invariably consist of fine quartzose material. Steiner also described much
Groundmass
66.0
65.3
larger xenoliths of quartzo-feldspathic composition and relic gneissic structure.
The plagioclase phenocrysts have been reported as ranging in composition from
TOTAL
100.0
100.0
An89 to An40 (andesine to bytownite), but in Ngauruhoe lavas are usually
labradorite (An68-55). They are subhedral and commonly exhibit complex oscillatory zoning with an overall trend from calcic
cores to sodic rims.102, 103, 104 Thin outer rims are usually compositionally similar to groundmass microlites. Twinning and
hourglass structures are common.Orthopyroxene predominates (>2:1) over clinopyroxene. Subhedral-euhedral
orthopyroxene is typically pleochroic and sometimes zoned. Compositions range from Ca 4 Mg74 Fe22 to Ca3 Mg47 Fe50, but
representative bulk and partial analyses of Ngauruhoe orthopyroxenes indicate a hypersthene composition predominates,
which is confirmed by optical determinations.105, 106, 107, 108, 109, 110, 111 Euhdral-subhedral clinopyroxene is typically
twinned and zoned, but compositions show a restricted range of Ca 43 Mg47 Fe10 to about Ca35 Mg40 Fe25, all of which is
augite.112, 113
The olivine present is strongly magnesian, analyses indicating some compositional zoning from Fo 88 to Fo78. The magnetite
present in the groundmass is titanomagnetite, judging from the amount of TiO 2 present in whole-rock analyses (tables 1 and
2), but some ilmenite is likely to occur sporadically in association with it.114, 115
K-Ar Results
All analytical results received from Geochron Laboratories are listed in table 4, grouped in chronological order according to
the historic date of each flow. The 40Ar* quantity refers to the amount of radiogenic 40Ar measured in each sample. All other
quantities are self-explanatory, some of them being calculated from the analytical results supplied by the laboratory.
The age of each sample is calculated from the analytical results using the general modelage equation:116, 117
(1)
where:
t

= the age

= the decay constant of the parent isotope

Dt = the number of daughter atoms in the rock presently


Do = the number of daughter atoms initially in the rock
Pt = the number of parent atoms presently in the rock
To date a rock, Dt and Pt are measured, and equation (1) can then be used if an assumption about the original quantity of
daughter atoms (Do) is made. Applied specifically to K-Ar dating, equation (1) thus becomes:

(2)
where:

= the age in Ma (millions of years)

5.543 10-10 = the current estimate for the decay constant of 40K
0.1048

= the estimated fraction of 40K decays producing 40Ar

40

= the calculated mole ratio of radiogenic 40Ar to 40K in the sample

Ar*/40K

It should be noted that to make equation (2) equivalent to equation (1), 40Ar* is assumed to be equal to (Dt - Do), which thus
means the 40Ar* measurement has included within it an assumption concerning the initial quantity of 40Ar in the rock, namely,
no radiogenic argon is supposed to have existed when the rock formed (that is, D o = 0). Thus equation (2) yields a model
age assuming zero radiogenic argon in the rock when it formed.The model ages listed in Table 4 range from <0.27 Ma to
3.50.2 Ma. However, it should be noted that the samples, one from each flow, that yielded model ages of <0.27 Ma and
<0.29 Ma (that is, below the detection limits of the equipment for 40Ar*) were all processed at the K-Ar laboratory in the same
batch, suggesting the possibility of a systematic problem with the analytical procedure and equipment (in particular, the gas
extraction line). When this question was raised with the laboratory manager, Richard Reesman, he kindly rechecked his
equipment and then re-ran several of the samples, producing similar results and thus ruling out a systematic laboratory
error.
However, an independent blind check was then made, by submitting to the K-Ar laboratory duplicate splits from two samples
already analyzed, to establish if results really were reproducible. The samples chosen were the A and B samples of the June
30, 1954, flow, because their first splits had produced the lowest and highest model ages, <0.27 Ma and 3.50.2 Ma,
respectively. The results of these additional analyses are shown in Table 4 as A#2 and B#2, and yielded model ages of
1.30.3 Ma and 0.80.2 Ma, respectively. Clearly, reproducibility was not obtained, but this is not surprising given the
analytical uncertainties at such low to negligible levels of 40Ar*, which are at the detection limits of the laboratorys equipment
(R. Reesman, pers. commun., February 19, 1997, and January 26, 1998).
Discussion
In spite of the wide variations in model ages obtained between and within these recent lava flows, and of the difficulties
obtaining analytical reproducibility, it is apparent that the cause of the anomalous K-Ar model ages is excess argon in the
lavas, that is, non-zero concentrations of radiogenic argon (40Ar*). This of course is contrary to the assumption of zero
radiogenic argon in equation (2) for calculating the model ages. When analyzed the oldest of the lavas was less than 50
years old, so there has been insufficient time since cooling for measurable quantities of 40Ar* to have accumulated within the
lavas due to the slow radioactive decay of 40K. Thus the measurable 40Ar* cant be from in situ radioactive decay since
cooling, and therefore must have been present in the molten lavas when extruded from Mt. Ngauruhoe.
No Radiogenic Argon Assumption Violated by Many Anomalous Ages
The assumption of no radiogenic argon ( 40Ar*) when the rocks formed is usually stated as self-evident. For example, Geyh
and Schleicher state:
What is special about the K-Ar method is that the daughter nuclide is a noble gas, which is not normally incorporated into
minerals and is not bound in the mineral in which it is found.118 (p. 56)
Similarly, Dalrymple and Lanphere state:
a silicate melt will not usually retain the 40Ar that is produced, and thus the potassium-argon clock is not set until the
mineral solidifies and cools sufficiently to allow the 40Ar to accumulate in the mineral lattice.119 (p. 46)
Dalrymple has recently put the argument more strongly:
The K-Ar method is the only decay scheme that can be used with little or no concern for the initial presence of the daughter
isotope. This is because 40Ar is an inert gas that does not combine chemically with any other element and so escapes easily
from rocks when they are heated. Thus, while a rock is molten the 40Ar formed by decay of 40K escapes from the
liquid.120 (p. 91)
Table 4. K-Ar analytical results and model ages for five recent lava flows at Mt. Ngauruhoe, New Zealand (Analyst:
Geochron Laboratories, Boston; July 1996, December 1996, September 1997, and January 1998.) Constants used: 40K/K =
1.193 10-4g/g; Fraction of 40K decays to40Ar* = 0.1048; Decay constant of 40K = 5.543 10-10yr-1; Atmospheric 40Ar/36Ar =
295.5
Flow
Date

Sample

K2O (wt40K
%)
(ppm)

40

Ar*
(ppm)
10-4

36

40

Ar*
(%)

40

Ar*
Total40Ar

Total40Ar
(ppm)

40

Ar
(ppm) 40Ar* 40K
10-5

Model Uncertainty
Age (Ma)(Ma) (1s)

Ar36Ar

Februar
y
11, A
1949
B

1.066
1.057

1.272
1.261

<0.2
0.75

3.75

0.0375

0.0020

296
310

0.645

<0.000016
0.000059

<0.27
1.0

0.2

June 4, A
1954
B

1.034
1.087

1.234
1.297

<0.2
1.10

3.00

0.030

0.00367

288
308

1.192

<0.000016
0.000085

<0.27
1.5

0.1

1.074
0.977

1.281
1.166

<0.2
0.87

2.75

0.0275

0.00316

285
306

1.033

<0.000016
0.000075

<0.27
1.3

0.3

B#1
B#2

0.944
0.995

1.126
1.186

2.28
0.56

4.05
1.70

0.405
0.017

0.00691
0.00329

311
304

2.222
1.082

0.000202
0.000047

3.5
0.8

0.2
0.2

1.097

1.308

0.93

3.75

0.0375

0.00248

311

0.798

0.000071

1.2

0.2

July 14, A
1954
B

1.033
1.004

1.232
1.198

0.69
<0.2

3.40

0.034

0.00203

301
283

0.675

0.000056
<0.000017

1.0
<0.29

0.2

Februar
y
19, A
1975
B

1.128
1.048

1.346
1.250

0.79
<0.2

2.55

0.0255

0.00310

307
291

1.010

0.000059
<0.000016

1.0
<0.27

0.2

A#1
A#2
June
30,
1954

However, these dogmatic statements by Dalrymple are inconsistent with even his own work on historic lava flows, some of
which he found had non-zero concentrations of 40Ar* in violation of this key assumption of the K-Ar dating method.121 He
does go on to admit, Some cases of initial 40Ar remaining in rocks have been documented but they are uncommon, but
then refers to his study of 26 historic, subaerial lava flows.122, 123 Five (almost 20%) of those flows contained excess
argon, but Dalrymple still then says that excess argon is rare in these rocks! The flows and their ages were:124

1.60.16 Ma
Hualalai basalt, Hawaii (AD 18001801)

1.410.08 Ma

Mt. Etna basalt, Sicily (122 BC)

0.250.08 Ma

Mt. Etna basalt, Sicily (AD 1792)

0.350.14 Ma

Mt. Lassen plagioclase, California (AD 1915)

0.110.03 Ma
0.270.09 Ma

Sunset Crater basalt, Arizona (AD 10641065)

0.250.15 Ma

Far from being rare, there are numerous examples reported in the literature of excess 40Ar* in recent or young volcanic rocks
producing excessively old whole-rock K-Ar ages:
Akka Water Fall flow, Hawaii (Pleistocene)

32.37.2 Ma125

Kilauea Iki basalt, Hawaii (AD 1959)

8.56.8 Ma126

Mt. Stromboli, Italy, volcanic bomb (September 23, 1963)

2.42 Ma127

Mt. Etna basalt, Sicily (May 1964)

0.70.01 Ma128

Medicine Lake Highlands obsidian, Glass Mountains, California (<500 years


old)
12.64.5 Ma129
Hualalai basalt, Hawaii (AD 18001801)

22.816.5 Ma130

Rangitoto basalt, Auckland, New Zealand (<800 years old)

0.150.47 Ma131

Alkali basalt plug, Benue, Nigeria (<30 Ma)

95 Ma132

Olivine basalt, Nathan Hills, Victoria Land, Antarctica (<0.3 Ma)

18.00.7 Ma133

Anorthoclase in volcanic bomb, Mt. Erebus, Antarctica (1984)

0.640.03 Ma134

Kilauea basalt, Hawaii (<200 years old)

218 Ma135
42.94.2 Ma136

Kilauea basalt, Hawaii (<1000 years old)

30.33.3 Ma137

East Pacific Rise basalt (<1 Ma)

6907 Ma138
58010 Ma139

Seamount basalt, near East Pacific Rise (<2.5 Ma)

700150 Ma140

East Pacific Rise basalt (<0.6 Ma)

24.21.0 Ma141
40

Other studies have also reported measurements of excess Ar* in lavas. Fisher investigated submarine basalt from a
Pacific seamount and found the largest amounts of excess 4He and 40Ar ever recorded (at that time).142 McDougall not
only found extraneous radiogenic argon present in three of the groups of basalt flows on the young volcanic island of
Runion in the Indian Ocean, but extraneous argon was also detected in alkali feldspar and amphibole in hyperbyssal
drusy syenites that are exposed in the eroded core of Piton des Neiges volcano.143 Significant quantities of excess 40Ar*
have also been recorded in submarine basalts, basaltic glasses and olivine phenocrysts from the currently active Hawaiian
volcanoes, Loihi Seamount and Kilauea, as well as on the flanks of Mauna Loa and Hualalai volcanoes, also part of the
main island of Hawaii, and in samples from the Mid-Atlantic Ridge, East Pacific Rise, Red Sea, Galapagos Islands,
McDonald Seamount and Manus Basin.144, 145, 146, 147 Patterson, Honda, and McDougall claimed that some of the initial
Loihi analytical results were due to atmospheric contamination of the magma either during intrusion or eruption, but
subsequent work has confirmed that the excess 40Ar* is not from atmospheric contamination at all.148, 149, 150
Excess 40Ar* Occluded in Minerals
Austin has investigated the 1986 dacite lava flow from the postOctober 26, 1980, lava dome within the Mount St. Helens
crater and has established that the 10-year-old dacite yields a whole-rock K-Ar model age of 0.350.05 Ma due to
excess 40Ar* in the rock.151 He then produced concentrates of the constituent minerals, which yielded anomalous K-Ar
model ages of 0.340.06 Ma (plagioclase), 0.90.2 Ma (hornblende), 1.70.3 Ma (pyroxene), and 2.80.6 Ma (pyroxene
ultra-concentrate). While these mineral concentrates were not ultra-pure, given the fine-grained glass in the groundmass
and some Fe-Ti oxides, it is nonetheless evident that the excess 40Ar* responsible for the anomalous K-Ar ages is retained
within the different constituent minerals in different amounts. Furthermore, the whole-rock age is very similar to the age of
the plagioclase concentrate because plagioclase is the dominant constituent of the dacite.That the excess 40Ar* can be
occluded in the minerals within lava flows, rather than between the mineral grains, has been established by others also.
Laughlin, et al., found that the olivine, pyroxene, and plagioclase in Quaternary basalts of the Zuni-Bandera volcanic field of
New Mexico contained very significant quantities of excess 40Ar*, as did the olivine and clinopyroxene phenocrysts in
Quaternary flows from New Zealand volcanoes.152, 153 Similarly, Poths, Healey, and Laughlin separated olivine and
clinopyroxene phenocrysts from young basalts from New Mexico and Nevada and then measured ubiquitous excess argon
in them.154 Damon, Laughlin, and Precious have reported several instances of phenocrysts with K/Ar ages 17 million
years greater than that of the whole rocks, and one K/Ar date on olivine phenocrysts of greater than 110 Ma in a recent
(<13,000 year old) basalt.155 Damon, et al., thus suggested that large phenocrysts in volcanic rocks contain the
excess 40Ar* because their size prevents them from completely degassing before the flows cool, but Dalrymple concluded
that there does not appear to be any correlation of excess 40Ar* with large phenocrysts or with any other petrological or
petrographic parameter.156Most investigators have come to the obvious conclusion that the excess 40Ar* had to have been
present in the molten lavas when extruded, which then did not completely degas as they cooled, the excess 40Ar* becoming
trapped in the constituent minerals, and in some instances, the rock fabrics themselves. Laboratory experiments have
tested the solubility of argon in synthetic basalt melts and their constituent minerals near 1300C at one atmosphere
pressure in a gas stream containing argon.157, 158 When quenched, synthetic olivine in the resultant material was found to

contain 0.34 ppm 40Ar*. Broadhurst, et al., commented that The solubility of Ar in the minerals is surprisingly high, and
concluded that the argon is held primarily in lattice vacancy defects within the minerals.159In a different experiment,
Karpinskaya, Ostrovskiy, and Shanin heated muscovite to 740860C under high argon pressures (28005000
atmospheres) for periods of 3 to 10.5 hours.160 The muscovite absorbed significant quantities of argon, producing K/Ar
ages of up to 5 billion years, and the absorbed argon appeared like ordinary radiogenic argon ( 40Ar*). Karpinskaya
subsequently synthesized muscovite from a colloidal gel under similar argon pressures and temperatures, the resultant
muscovite retaining up to 0.5 wt% argon at 640C and a vapor pressure of 4000 atmospheres.161 This is approximately
2,500 times as much argon as is found in natural muscovite. These experiments show that under certain conditions argon
can be incorporated into minerals and rocks that are supposed to exclude argon when they crystallize.
Applications to the Mt. Ngauruhoe Andesite Flows
Therefore, the analytical results from the very recent (19491975) andesite flows at Mt. Ngauruhoe, New Zealand, that yield
anomalous K-Ar model ages because of excess 40Ar* are neither unique nor an artifact of poor analytical equipment or
technique. This realization that the presence of the excess 40Ar* in these rocks is both real and measurable, and has not
been derived from radioactive decay of 40K in situ, leads to the obvious questions as to whether there is any pattern in the
occurrences of excess 40Ar*, and from whence came this excess 40Ar*?It is clear that the excess 40Ar* was in the lavas when
they flowed from the Mt. Ngauruhoe volcano and were trapped in the andesite as it cooled. That there were gases in the
lavas is readily evident from the copious frozen bubble holes now in the rock, implying that much of the gas content
escaped as the lavas flowed and cooled. When choosing samples, care was taken to select pieces from each flow that were
different from one another (for example, copious frozen gas bubble holes compared with virtually no such holes). It is
hardly surprising, therefore, that the 40Ar* measurements on four of the five flows were consistent with such differences the
samples from each flow which had very few or virtually no frozen gas bubble holes yielded excess 40Ar* and thus
anomalous K-Ar model ages, whereas the other samples from each of these flows that contained copious frozen gas
bubble holes failed to yield detectable 40Ar* (<0.27 Ma and <0.29 Ma in Table 4).
The exception was the June 30, 1954, flownot only was this expected relationship between excess 40Ar* and lack of
frozen gas bubble holes not duplicated, but analyses on duplicate splits off the same samples yielded widely divergent
results (<0.27 Ma versus 1.30.3 Ma and 3.50.2 Ma versus 0.80.2 Ma; see Table 4). Thus the presence (or absence) of
excess 40Ar* must also depend on which portion of a rock sample is being analyzed, which in turn implies dependence on
the mineral constituents present, including the glass in the groundmass. As already noted, Austin found widely different
amounts of excess 40Ar* in the mineral separates concentrated from Mount St. Helens 1986 dacite, while numerous other
studies have located excess 40Ar* in phenocrysts.162, 163, 164, 165, 166
Cooling Rates, Pressures, and Potassium Alteration
Another factor is the rate of cooling of lavas. Dalrymple and Moore found that the 1 cm thick glassy rim of a pillow in a
Kilauea submarine basalt had greater than forty times more excess 40Ar* than the basalt interior just 10 cm below.167The
glassy pillow rim is, of course, produced by rapid quenching of the hot basalt lava immediately as it contacts the cold ocean
water, so the excess 40Ar* in the lava is rapidly trapped and retained. Dymond obtained similar results on four deep-sea
basalt pillows from near the axis of the East Pacific Rise.168 Dalrymple and Moore also found that the excess 40Ar* contents
of the glassy rims of basalt pillows increased systematically with water depth, leading them to conclude that the amount of
excess 40Ar* is a direct function of both the hydrostatic pressure and the rate of cooling.169 In a parallel study, Noble and
Naughton reported K-Ar ages from zero to 22 Ma with increasing sample depth for submarine basalts probably less than
200 years old, also from the active Kilauea volcano.170Seidemann has reported yet another intriguing relationship.171 He
analyzed deep-sea basalt samples obtained from DSDP drillholes in the floor of the Pacific Ocean basin and found that K-Ar
ages increased with increasing K contents of the basalts, a relationship he noted also appeared in similar data published
by DSDP staff (Seidemann172, fig. 1). In basalt pillows the K content increases from the margin to a maximum at an
intermediate distance into the pillows, whereas holocrystalline basalts show a decrease of K inward from the
margin.173 Seidemann concluded, as had others before him, that submarine weathering adds K to the basalts, as does
alteration at the time of formation, whereas the glassy pillow margins are largely impervious to seawater. The net result,
however, is unreliable K-Ar dates, because the measured 40Ar* was probably not derived by radioactive decay of the
measured 40K contents. Seidemann also determined that sediment cover is not a significant barrier to the diffusion of K into
basalt.It is possible that some of these factors are relevant to the pattern of excess 40Ar* measured in the samples from the
recent Mt. Ngauruhoe andesite flows. For example, the surfaces of the flows would have cooled more rapidly to crusts on
top of the still molten flow interiors, which would certainly have been the case with the August 18, 1954, flow that was
reported as being 18 m thick. Furthermore, the overburden pressure within the deep interiors of such thick flows would
likewise inhibit degassing of the lava as it cooled. However, so long as the crusts on the tops of the flows remained
unbroken and intact they would have sealed in the molten lava and its contained gases, including excess 40Ar*. But this
sealing was probably short-lived, because today the flows mostly outcrop as piles of pieces of andesite that look like rubble
(typical aa lavas). The continued flow of the lavas down the sides of the volcano would have broken up the crusts as soon
as they congealed, as would contraction with cooling, thus enabling the molten interiors to degas as they cooled. So the
speed of cooling was likely the most relevant factor, and this would have varied laterally and vertically within the flows, even
at localised scales of a few centimeters.Any effects of weathering on the K contents of these flows can be discounted. On
the one hand these are subaerial flows that do not appear to have been subjected to leaching or addition of K, or to any Krich alteration for that matter, while on the other they have very uniform K contents (see tables 1, 2, and 4). This is not
unexpected, given the fact they flowed from the same magma source/chamber close together timewise. The K-Ar data in
table 4 do not reveal any discernible relationship between K contents and K-Ar model ages of these flows, unlike the
negative correlation found in the Middle Proterozoic Cardenas Basalt upper member flows in Grand Canyon, Arizona.174
Perhaps the key issues, though, are where this excess 40Ar* has come from, and whether it has been derived from
radioactive decay of 40K. One possibility is that the excess 40Ar* can be accounted for by radioactive decay during long-term
residence of magmas in chambers before eruption. Esser, McIntosh, Heizler, and Kyle discounted this option for the Mt.
Erebus anorthoclase phenocrysts.175 Dalrymple found that whereas the Mt. Lassen (1915) plagioclase phenocrysts yielded
excess 40Ar* and an anomalous K-Ar model age, a plagioclase from the 1964 eruption of Surtsey only had argon whose
isotopic composition matched that of air.176 Because phenocrysts usually crystallize from lavas after eruption, they may
arbitrarily trap excess 40Ar* during lava cooling, 40Ar* that will thus not be from in situ 40K radioactive decay.
Negative K-Ar Model Ages and Atmospheric Argon
Another relevant consideration bearing on these issues is the observation noted by Dalrymple that some modern lava
samples actually yield negative K-Ar model ages, apparently due to excess 36Ar177. Air has an 40Ar/36Ar ratio of 295.5, but
some of Dalrymples samples had ratios less than 295.5 (and hence negative ages). Some of the Mt. Ngauruhoe samples
in this study also yielded 40Ar/36Ar ratios less than 295.5 (see table 4). According to the straightforward interpretation of the

K-Ar dating methodology, this should be impossible. Dalrymple was not willing to attribute these anomalous ratios to
experimental error, and neither was Richard Reesman of Geochron Laboratories.
Dalrymple suggested three possible explanations that might account for the excess 36Ar178:
incorporation of primitive argon,
production of 36Ar by the radioactive decay of 36Cl, or
fractionation of atmospheric argon by diffusion.
He rejected the possibility of significant 36Ar formation in situ from nuclear reactions (option 2) because the Cl content of
basalts and the production rate of 36Cl by cosmic-ray neutrons both are too low to account for any significant amount of 36Ar.
Instead, Dalrymple seemed to favor option 3, that when atmospheric argon diffused back into lavas as they cooled, 36Ar
diffused in preferentially. However, he also recognized the weakness of this argumentit is difficult to explain why some
lavas are enriched in 36Ar while others are not (as at Mt. Ngauruhoe also). To be consistent, if fractionation of atmospheric
argon occurred during diffusion, then this would mean that even supposedly zero age lavas actually have an apparent age,
and that most lavas do not degas upon eruption. In fact, depending on how strong the fractionation of 36Ar was during
diffusion, it could even be that all lavas do not completely degas.This only leaves Dalrymples option 1, that the lavas with
the anomalously high 36Ar come from areas of the mantle (and perhaps also the crust) which have primordial argon that has
not been diluted with radiogenic 40Ar and have not completely degassed. However, this means that there is no reason to
assume that lavas whose argon matches that in the atmosphere have degassed either, because they may have simply
started with argon which matches atmospheric argon. Nevertheless, Dalrymple is convinced that much of the volatile
juvenile content may still be present in volcanic rocks quenched on the ocean floor.179 Indeed, Dalrymple has specifically
defined excess 40Ar* as 40Ar that is not attributed to atmospheric argon or in situ radioactive decay of 40K.180 Krummenacher
is more cautious, attributing anomalous 40Ar/36Ar ratios and excess 40Ar* to the mass fractionation effect on argon of
atmospheric isotopic composition trapped in the lavas, as well as to the presence of magmatic argon different in isotopic
composition.181
The Role of Xenoliths
Is the excess 40Ar* simply magmatic argon, that is, argon that collects in the magma and then is inherited by the lavas from
it? Funkhouser and Naughton found that the excess 40Ar* in the 18001801 Hualalai flow, Hawaii, resided in fluid and
gaseous inclusions in olivine, plagioclase, and pyroxene in ultramafic xenoliths in the basalt.182 The quantities of
excess 40Ar* were sufficient to yield K-Ar model ages from 2.6 Ma to 2960 Ma. However, Dalrymple subsequently only
used the presence of the ultramafic xenoliths and their excess 40Ar* contained in inclusions to explain partly the excess 40Ar*
and anomalous K-Ar model ages he obtained from the same 18001801 Hualalai flow, suggesting instead that the large
single inclusions are not directly responsible for the excess argon in the flows and that the 40Ar* is distributed more uniformly
throughout the rocks.183 Nevertheless, those K-Ar and Ar-Ar geochronologists who are concerned about the excess 40Ar* in
their samples undermining their dating are careful to check for xenoliths and xenocrysts. Esser, et al., did so and
discounted xenocrystic contamination.184Xenoliths are present in the Ngauruhoe andesite flows (table 3), but they are
minor and less significant as the location of the excess 40Ar* residing in these flows than the plagioclase and pyroxene
phenocrysts and the much larger glomerocrysts of plagioclase, pyroxene, or plagioclase and pyroxene that predominate.
The latter are probably the early-formed phenocrysts that accumulated together in the magma within its chamber prior to
eruption of the lava flows. Nevertheless, any excess 40Ar* they might contain had to have been supplied to the magma from
its source. The xenoliths that are in the andesite flows have been described by Steiner as gneissic and are therefore of
crustal origin, presumably from the basement rocks through which the magma passed on its way to eruption.185
Noble Gases from the Mantle
With the advent of the necessary technology, the isotopic concentrations and ratios of noble gases (including argon) in rock
and mineral samples are now obtainable. Honda, et al., have reported such analyses on submarine basalt glass samples
from Loihi Seamount and Kilauea, Hawaii, and concluded that helium and neon isotopic ratios in particular, being uniquely
different from atmospheric isotopic ratios, are indicative of the mantle source area of the plume responsible for the Hawaiian
volcanism rather than from atmospheric contamination of the magma.186, 187 The40Ar/36Ar ratios are consistent with
excess 40Ar* having also come with the magma from the mantle. A subsequent study, in which a larger suite of basalt glass
samples, and also samples of olivine phenocrysts, from the same and additional Hawaiian volcanoes were analyzed,
concluded that the isotopic systematics indicate that the helium and neon have been derived from the mantle and have not
been preferentially affected by secondary processes.188Consequently, the excess 40Ar* also in these samples would have
been also carried from the upper mantle source area of these basalts by the magma plume responsible for the volcanism.
Moreira, Kunz, and Allgre have suggested, based on new experimental data from single vesicles in mid-ocean ridge basalt
samples dredged from the North Atlantic, that the excess 40Ar* in the upper mantle may be almost double previous estimates
(that is, almost 150 times more than the atmospheric content relative to 36Ar), and represents a primordial mantle component
not yet outgassed.189, 190Burnard, Graham, and Turner obtained similar results on the same samples, but maintained that
because some of the36Ar is probably surface-adsorbed atmospheric argon, the upper mantle content of excess 40Ar* could
be even ten times higher.191Similar results have been obtained from ultramafic mantle xenoliths in basalts from the
Kerguelen Archipelago in the southern Indian Ocean, and the considerable excess 40Ar* measured concluded to be a part of
the mantle source signature of this hotspot volcanism.192 However, it has not only been the suboceanic mantle that has
thus been sampled for its excess 40Ar* via such magma plumes. Matsumoto, Honda, McDougall, and OReilly have reported
high40Ar/36Ar ratios in spinel-lherzolites from five eruption centers in the youthful (<7 Ma) Newer Volcanics of southeastern
Australia.193 These anhydrous lherzolites have compositions representative of the upper lithospheric mantle, and the
significant excess 40Ar* in them indicates the presence of a subcontinental mantle reservoir with a very high 40Ar/36Ar ratio,
and thus substantial excess 40Ar*, similar to that found in mid-ocean ridge and plume/hotspot basalts. Another example is
the Cardenas Basalt and associated diabase (Middle Proterozoic) of eastern Grand Canyon, regarded as part of the
pervasive mafic mid-continental magmatism of the southwestern United States and thus also sourced from the
subcontinental mantle. Austin and Snelling have found that the 40Ar/36Ar40K/36Ar isochrons for fourteen and six samples of
these rocks, respectively, yield initial 40Ar/36Ar ratios of 787118 and 45342, indicative of some initial excess 40Ar*.194
Sampling the Mantle with Diamonds and Their Inclusions
Another means of sampling the mantle is the study of diamonds and their micro-inclusions. It is now firmly established that
diamonds are thermodynamically stable in the pressure-temperature regime in the mantle at depths greater than 150 km,
and their origin is believed to extend back to the Archean and the early crust of the Earth. 195, 196Diamonds are formed in a
number of processes associated with two rock types, ecologite and peridotite, xenoliths of which are also brought up into the
upper crust with diamonds from the upper mantle below continental Precambrian shields (cratons) by kimberlite and
lamproite pipe eruptions.197, 198, 199 Even though the host kimberlite or lamproite may be relatively young (even in
conventional terms), many diamonds date back to the Archean and thus the early history of the Earth.200, 201 To account
for all this evidence, it is postulated that the formation of most diamonds was closely associated with subduction of the

Archean oceanic crust into the mantle, the required carbon (which was originally thought to be primordial carbon already in
the mantle) now believed to derive from sedimentary marine carbonates and biogenic carbon from bacteria/ algae in the
sediments subducted with the oceanic crust.202, 203, 204,205, 206The noble gas contents of diamonds are consistent with
their ancient and mantle origin, high helium isotopic ratios (290 times the atmospheric ratio) being regarded as primordial
and rivaling those measured for the sun today.207, 208Of significance here is the postulation that He, Ar, K, Pb, Th, and U
are added to the convecting upper mantle circulation, and the proportions and isotopic compositions are strongly determined
by entrainment from the lower mantle (below 670 km).209, 210 This is reflected in those Ar isotopic measurements that have
been made on diamonds and their micro-inclusions.Rather than focus on attempting to date only diamond micro-inclusions
as others had done, Zashu, Ozima, and Nitoh carefully selected ten Zaire diamonds and examined them for purity before
undertaking K-Ar dating analyses of the diamonds themselves.211 However, at the outset they noted that there had been
almost no direct radiometric dating of diamonds except for conventional K-Ar dating, and the results had been questionable
due to the possible presence of excess 40Ar*. To avoid this problem, they used the K-Ar isochron dating method. Their
experimental data showed good linear correlations, but these isochrons yielded an age of 6.00.3 Ga, which of course was
unacceptable because these diamonds would be older than the earth itself. Mistakes in the experimental procedure were
easily discounted, so they were forced to conclude that excess 40Ar* was responsible, and that it needed to be in a fluid state
to ensure the homogenization necessary to give such a constant 40Ar/K ratio. Alternately, they speculated that the diamonds
might differ in K isotopic composition from common potassium, but this was discounted in a follow-up study in which it was
found that 40K was present in these diamonds in normal abundance.212 Because 40Ar/39Ar analyses yielded the same
unacceptable age, it was concluded that the excess 40Ar* was not generated in situ, but was an inherited or trapped
component from the mantle reservoir when and where the diamonds formed.These Zaire diamonds are not the only ones
which have yielded excess 40Ar*. Phillips, Onstott, and Harris used a laser-probe to 40Ar/39Ar date eclogitic clinopyroxene
inclusions in diamonds from the Premier kimberlite, South Africa, and found moderate 40Ar/36Ar ratios indicative of much less
excess 40Ar* than in the Zaire diamonds.213 The age of these eclogitic diamonds was thus determined to be 1.1980.014
Ga, much younger than the 3.3 Ga peridotitic diamonds at Kimberley and Finsch, also in South Africa, so Phillips, et al.,
interpreted the moderate excess 40Ar* as characteristic of mantle conditions prevailing at the time and in the region of
Premier eclogitic diamond formation.214Zashu, et al., postulated that the excess 40Ar* in the Zaire diamonds needed to be in
a fluid state.215 Though Navon, Hulcheon, Rossman, and Wasserburg did not analyze for argon when they investigated
fluids in micro-inclusions in diamonds from Zaire and Botswana, they found a high content of volatiles and incompatible
elements in the uniform average composition of the micro-inclusions, with the amounts of water and CO 2 (in carbonates)
almost an order of magnitude higher than the volatile contents of kimberlites and lamproites (host rocks to diamonds).216 At
13 wt%, the chlorine levels were also much higher than those of kimberlites (<0.1%), although the bulk composition of the
micro-inclusions, including the high K2O content (up to 29.7 wt%), resembled that of such potassic magmas. They
concluded that these micro-inclusions represent the volatile-rich (~40% volatiles) fluid or melt from the upper mantle in
which the diamonds grew, and that because of the high volatile content in this hydrous mantle fluid, high levels of rare gases
may also be expected and explain the high 40Ar/K ratios (the excess 40Ar*) and anomalous ages.
As a result of continued investigation of the Zaire cubic diamonds, which produced 40Ar/39Ar age spectra yielding a ~5.7 Ga
isochron, Ozima, Zashu, Takigami, and Turner discovered that just as there was an excellent correlation between their
potassium contents and 40Ar/36Ar ratios, there is also a correlation between their chlorine contents and 40Ar.217 They
concluded from their data
that the 40Ar is an excess component which has no age significance, and that the 40Ar and its associated potassium are
contained in sub-micrometer inclusions of mantle-derived fluid.
Turner, Burgess, and Bannon also used the 40Ar/39Ar technique through correlations with K, Cl, and 36Ar to unscramble the
mixtures of radiogenic and parentless (excess) Ar components in fluid inclusions in coated Zaire diamonds and in olivine
from an East African mantle xenolith.218 Their results proved conclusively that 40Ar is present in a widespread chlorine-rich
component, which implies the existence of H2O/CO2-rich phases with 40Ar/Cl ratios that are remarkably uniform over large
distances, with enrichments of these two incompatible elements by almost four orders of magnitude relative to bulk uppermantle values. Clearly, excess 40Ar* is abundant in the mantle and can be easily transported up into the crust.
Crustal Excess 40Ar*
Is there only evidence for excess 40Ar* in the mantle, gleaned from rocks (basalts and ultramafic xenoliths) and minerals
(olivine, pyroxene, plagioclase, and diamonds) that were formed in, or ascended from, the mantle? Patterson, et al.,
envisage noble gases from the mantle (and the atmosphere) migrating and circulating through the crust, so there should be
evidence of excess 40Ar* in crustal rocks and minerals.219 In fact, noble gases in CO 2-rich natural gas wells support such
migration and circulationthat is, the isotopic signatures clearly indicate a mantle origin for the noble gases, including
amounts of excess 40Ar* in some CO2-rich natural gas wells exceeding those in the mantle-derived mid-ocean ridge
basalts.220, 221, 222, 223, 224 Staudacher also notes that the quantities of excess 40Ar* in the continental crust can be as
much as five times that found in such mantle-derived midocean ridge basalts, strongly suggesting that excess 40Ar* in
crustal rocks and their constituent minerals could well be the norm rather than the exception, thus making all K-Ar (and ArAr) dating questionable.225It has now been established that some diamonds can form in the crust during high-grade
metamorphism and via shock metamorphism during meteorite or asteroid impact.226, 227, 228 The pressures and
temperatures of high-grade metamorphism had been regarded as insufficient to produce diamonds, but the key ingredient
was found to be volatile N2-CO2-rich fluids. Noble gas data on these diamonds are not yet available, due to their size and
rarity, but such data have been definitive in establishing the crustal origin of carbonado diamonds.229 Nevertheless, they
still contain excess40Ar*.Dalrymple, referring to metamorphism and anatexis of rocks in the crust, commented, If the rock is
heated or melted at some later time, then some or all the 40Ar may escape and the K-Ar clock is partially or totally
reset.230 In other words,40Ar* escapes to migrate in the crust where it may then be incorporated in other minerals as
excess 40Ar*, just as 40Ar* degassing from the mantle does. Thus, for example, excess 40Ar* has been recorded in many
minerals (some of which contain no 40K) in crustal rocks, such as quartz, plagioclase, pyroxene, hornblende, biotite, olivine,
beryl, cordierite, tourmaline, albite, and spodumene in pegmatites, metamorphic rocks, and lavas.231, 232 And it is not just
K-Ar dating analyses that detect excess 40Ar*, as Lanphere and Dalrymple used the 40Ar/39Ar method to confirm the
presence of excess 40Ar* in feldspars and pyroxenes.233 Indeed, in a recent study, 128 Ar isotopic analyses were obtained
from ten profiles across biotite grains in amphibolite-granulite facies metamorphic rocks, and apparent 40Ar/39Ar ages within
individual grains ranged from 161 to 514 Ma.234 The investigators concluded that these observations cannot be solely due
to radiogenic build-up of 40Ar*, but must be the result of incorporation by diffusion of excess 40Ar* from an external source,
namely, 40Ar* from the mantle and other crustal rocks and minerals. Indeed, Harrison and McDougall were able to calculate
a well-defined law for 40Ar diffusion from hornblende in a gabbro due to heating.235 They also found that the excess 40Ar*
which had developed locally in the intergranular regions of the host gabbro reached partial pressures in some places of at
least 10-2 atm.This crustal migration of 40Ar* is known to cause grave problems in attempted regional geochronology studies.

In the Middle Proterozoic Musgrave Block of northern South Australia, Webb found a wide scatter of K-Ar mineral ages
ranging from 343 Ma to 4493 Ma due to inherited (or excess) 40Ar*, so that no meaningful interpretation could be drawn from
the rocks (granulite, gneiss, pseudotrachylite, migmatite, granite, and diabase).236 Of the diabase dikes which gave
anomalous ages, he concluded that
The basic magmas probably formed in or passed through zones containing a high partial pressure of 40Ar*, permitting
inclusion of some of the gas in the crystallizing minerals.
Likewise, when Baski and Wilson attempted to argon date Proterozoic granulite-facies rocks in the Fraser Range (Western
Australia) and Strangways Range (central Australia), they found that garnet, sapphirine, and quartz in those rocks contained
excess 40Ar* that rendered their argon dating useless because of ages higher than expected.237 They also concluded that
the excess 40Ar* was probably incorporated at the time of formation of the minerals, and their calculations suggested a
partial pressure of ~0.1 atm Ar in the Proterozoic lower crust of Australia, which extends over half the continent.
In a detailed 40Ar/39Ar dating study of high-grade metamorphic rocks in the Broken Hill region of New South Wales
(Australia), Harrison and McDougall found evidence of widely distributed excess 40Ar*.238 The minerals most affected were
plagioclase and hornblende, with step heating 40Ar/39Ar age spectra yielding results of up to 9.588 Ga. Such unacceptable
ages were produced by excess 40Ar* release, usually at temperatures of 350650C and/or 9301380C, suggesting the
excess 40Ar* is held in sites within the respective mineral lattices with different heating requirements for its release. There
are three principal trapping sites for Ar in solids: structural holes, edge dislocations, and lattice vacancies. (Argon is also
known to be held sometimes in some minerals in fluid inclusions.) Clearly, this study shows that at crustal temperatures,
which are less than 930C, some excess 40Ar* will always be retained in those trapping sites in minerals where it is obviously
held more tightly, thus rendering K-Ar and 40Ar/39Ar dating questionable. Harrison and McDougall were only able to produce
a viable interpretation of the data because they made assumptions about the expected age of the rocks and of a presumed
subsequent heating event (based on Pb-Pb and Rb-Sr dating), the latter being the time when they conjecture that
accumulated 40Ar* was released from minerals causing a significant regional Ar partial pressure of ~3x10 -4 atm to
develop.239
Mantle-Crust Domains and Excess 40Ar*
Harte and Hawkesworth have identified domains within the mantle and crust and described the interaction between them, all
of which is relevant to the migration and circulation of argon (and thus excess 40Ar*) from the lower mantle to the crust and
to lavas extruded on the earths surface.240 The six domains are physically distinct units which show wide differences in
average physical and chemical properties, as well as apparent age, structure, and tectonic behavior. They are the lower
mantle (below 670 km), upper mantle, continental mantle lithosphere, oceanic mantle lithosphere, continental crust, and
oceanic crust, and each is a distinct geochemical reservoir. Each domain may provide material for magmatic rocks, and
particular geochemical features of magmas may be associated with particular domains. Thus the convecting upper mantle
which comes to the surface at mid-ocean ridges may be identified as the source of most geochemical features of midocean
ridge basalts, including their excess 40Ar* content. Similarly, the convecting lower mantle is regarded as the primordial or
bulk earth geochemical reservoir, which may also contribute excess 40Ar* to mid-ocean ridge basalts, but is more important
for its contribution to ocean island basalts (for example, Hawaii) and other plume-related basalts (continental alkali basalts
and continental flood basalts). However, considerable complexity may be added to the deeper mantle geochemical structure
as a result of localized accumulation of subducted oceanic lithosphere.Porcelli and Wasserburg have proposed a steadystate upper mantle model for mass transfer of rare gases, including argon.241 The rare gases in the upper mantle are
derived from mixing of rare gases from the lower mantle, subducted rare gases, and radiogenic nuclides produced in situ.
Porcelli and Wasserburg claim that all of the 40Ar in the closed-system lower mantle has been produced by 40K decay in the
lower mantle, but this claim is based on the assumption of a 4.5 Ga earth. In any case, they contradict themselves, because
they also state, The lower mantle is assumed to have evolved isotopically approximately as a closed system with the in
situ decay of 129I, 244Pu, 238U, 232Th, and 40K adding to the complement of initial rare gases.242 In other words, they admit
that some of the 40Ar must be primordial and not derived from radioactive 40K. They then go on to claim that in the upper
mantle, 40K decay further increases the radiogenic 40Ar from the lower mantle by a factor of ~3, but again this presupposes a
4.5 Ga earth and doesnt allow for primordial 40Ar that could well also be in the upper mantle if it is admitted to be in the
lower mantle.In the case of the continental and oceanic lithospheric domains, the lack of convective stirring means that
different geological processes and events may implant in each domain a variety of geochemically distinct materials that will
remain isolated from one another. Therefore, these domains do not have a single set of geochemical characteristics; thus
identification of geochemically defined sources with particular physically defined crust-mantle domains is complex, and the
geochemical definition of particular reservoirs cannot be regarded as simply definition of major physical entities.
Nevertheless, excess 40Ar* will be added to these domains by the passage of basaltic magma plumes from the upper mantle
to the earths surface.Furthermore, the processes of oceanic lithosphere formation from the convecting upper mantle in
association with mid-ocean ridge activity mean that its isotopic characteristics everywhere will be largely similar to those of
the convecting upper mantle and mid-ocean ridge basalts, including the addition of excess 40Ar*. The corollary to this is that
the oceanic crust is formed as part of these same processes. However, the oceanic crust generally has a thin veneer of
sediments over it, and thick wedges of sediments adjacent to the domains of continental crust, whereas sections of oceanic
crust are hydrothermally altered. The compositions of these components of the oceanic crust may, therefore, include a
considerable contribution from continental detritus and ocean water, so that this oceanic crustal material may give rise to a
distinct geochemical reservoir, the fate of which during subduction back into the upper mantle becomes critically important if
it contributes to island arc volcanics, plume-related intra-place magmas, and mantle-derived xenoliths.
The complexity of continental crustal material is well known through direct observation, and the mantle lithosphere attached
to it may be expected to show a similar complexity. Nevertheless, it is evident that excess 40Ar* also resides in the
continental mantle lithosphere, as indicated by xenoliths.243 Likewise, there is evidence of excess 40Ar* in crustal magmatic
rocks (for example, gabbros, pegmatites, migrating through metamorphic terrains, and in natural gas in sedimentary
reservoirs.244, 245, 246, 247, 248, 249
Mt. Ngauruhoe in its Tectonic Framework
The presence, therefore, of excess 40Ar* in the recent andesite flows at Mt. Ngauruhoe is to be expected. The Taupo
Volcanic Zone is a volcanic arc and marginal basin of the Taupo-Hikurangi arc-trench system (see fig. 1 again), which is a
southward extension of the Tonga-Kermadec arc into the continental crustal environment of New Zealands North
Island.250 Geophysical investigations indicate that the Pacific Plate is being obliquely subducted beneath the Australian
Plate on which most of New Zealands North Island sits, and that the volcanoes of the Taupo Volcanic Zone, including
Ngauruhoe in the Tongariro Volcanic Center, are about 80 km directly above the subducting Pacific Plate, a zone of
earthquakes revealing where the movement is taking place.251 Friction along the plane of contact is believed to cause
melting to produce pockets of magma, which then feed via conduits to the volcanoes above. Thus the recent andesite flows
at Mt. Ngauruhoe are calcalkaline island arc volcanics.The tectonic and geochemical framework of the Ngauruhoe andesite

flows within the mantle-crust domains of Harte and Hawkesworth is that of subducting oceanic crust (derived from the
convecting upper mantle), carrying with it the wedge of continental sedimentary detritus which has accumulated at the
continental margin and in the adjacent trench to the east of the coastline.252 Attached beneath the subducting oceanic crust
is its associated oceanic mantle lithosphere, and together they are being thrust downwards into the upper mantle. Above the
subducting plate are the continental crust and continental mantle lithosphere of the overriding plate, the continental crust
being at the contact plane at shallow depths near the trench, and then the attached continental mantle lithosphere beneath
at a depth of about 35 km.253 Thus the geochemical reservoir from which the Ngauruhoe andesite magma has been drawn
is potentially a mixture of melted oceanic crust, continental sedimentary detritus and continental crust, and possibly
continental mantle lithosphere, or even upper mantle.
Genesis of the Mt. Ngauruhoe Andesite Magma and its Excess 40Ar*
One of the easier investigations of the petrogenesis of these volcanic rocks of the Taupo Volcanic Zone was that of Stipp
and of Ewart and Stipp.254, 255 They analyzed samples that had been systematically collected, including not only the lavas
and the pyroclastics, but also the Permian to Jurassic interbedded greywackes, siltstones, and shales (the potential crustal
source rocks), which are spatially related to, and underlie, the volcanics. Of primary interest were Sr, Rb, and K contents,
and 87Sr/86Sr and 87Rb/86Sr ratios. Three possibilities for the origin of the calc-alkaline andesite magma were under
investigation: fractional crystallization of a basalt magma under oxidizing conditions; some form of hybridization between
basaltic and acidic magmas, possibly followed by fractional crystallization; and derivation of a primary andesite magma from
the upper mantle. Ewart and Stipp regarded their Sr isotopic data as more consistent with the production of the andesites by
partial assimilation of sedimentary material by basaltic magma (derived from the upper mantle), the adjacent greywackes,
siltstones, and shales being the most likely sedimentary material, and the unassimilated gneissic xenoliths probably
representing the basement rocks to those sediments.256 However, they admitted that the data did not exclude the
possibility of a primary andesitic magma derived directly from the upper mantle, provided that some assimilation of crustal
material modified it prior to eruption.Subsequent investigations by Cole favored the alternate petrogenetic model of a
primary andesitic magma.257 He suggested that the subducting oceanic crust assimilated the greywacke-siltstone-shale
and overlying sediments east of the Taupo Volcanic Zone to produce amphibolite, which subsequently broke down to
produce phlogopite eclogite below 90 km. This in turn partially melted at 150200 km, and the resultant magma fractionated
in the upper mantle or lower crust to produce andesite. However, based on rare-earth element geochemistry, Cole,
Cashman, and Rankin modified that petrogenetic model, suggesting that while the andesite magma genesis was probably
associated in the upper mantle with the downgoing slab and some crustal contamination occurred, the andesite does not
appear to have had an eclogite parent.258 This would then suggest that the melting associated with the subducting slab to
generate the andesite magma occurred at a depth of less than 90 km.Graham and Hackett agreed with this conclusion,
demonstrating from geophysical evidence that the top of the subducting slab is at a depth of about 80 km below Ngauruhoe,
and that the crust there is probably less than 20 km thick.259 Thus the upper mantle wedge between would consist only of
plagioclase-peridotite and spinelperidotite. At 80 km depth the hydrated amphibolite assemblage of the upper portion of the
subducting slab of oceanic crust and oceanic mantle lithosphere would have started to dehydrate, thus liberating water and
possibly other volatile constituents into the overlying upper mantle wedge, significantly lowering its melting point. Graham
and Hackett then showed that the geochemical evidence requires the andesite magma for the Ngauruhoe lava flows to have
been generated from an original low-alumina basalt magma produced in the upper mantle wedge by anatexis of the
asthenosphere (uppermost mantle) and/or subcontinental mantle lithosphere probably catalyzed by hydrous, metasomatic
fluids from the subducting slab.Some specific geochemical enrichment then appears to have occurred as a result of this
mantle metasomatism and continental crustal contamination during ascent and storage of the magma. Graham and Hackett
used least squares geochemical modeling to show how the andesite magma could be generated from such a parent basalt
magma by a process of combined assimilation of crustal material (addition of 6% assimilant) and fractional crystallization
(30% removal of crystals).260 Furthermore, the presence of xenoliths in the Ngauruhoe andesite flows, particularly the
vitrified meta-greywacke and gneissic xenoliths, indicate conclusively that the assimilant was most likely a partial melt of
gneiss, originally the adjacent greywacke-siltstone-shale sediments.261, 262These processes responsible for the generation
of the andesite magma did not diminish the excess 40Ar* content of the resultant flows. Though the amount of excess 40Ar* is
not high when compared with that found in mid-ocean ridge basalts, it is nonetheless significant that the excess 40Ar* was
still present in the lavas upon eruption and cooling. The evidence indicates that the parent basaltic magma was generated in
the upper mantle where the excess 40Ar* in the geochemical reservoir is now known to be upwards of 150 times more than
the atmospheric content, relative to 36Ar. The subsequent crustal contamination and fractional crystallization to form the
andesite magma during ascent, and the degassing of the magma during eruption and lava flow and cooling, did not remove
all the excess 40Ar*, a small portion of which was left to be trapped in the congealed lava and its constituent minerals.This
model for the generation of the andesite magma in the post-Flood world is, of course, based on the plate tectonics model for
global tectonics through earth history. Even though the postulated plate movements today are extremely slow, and thus
extrapolated back over millions of years by uniformitarians, a catastrophic model for plate tectonics in the context of the
Flood is entirely compatible with both Scripture and the scientific data.263 Plate movements are regarded as occurring
catastrophically during the Flood and then rapidly slowing down to todays rates in the post-Flood era.
Conclusions
The fact that there is even some excess 40Ar* in these recent andesite flows, and that it appears to have ultimately come
from the upper mantle geochemical reservoirwhere it is regarded as leftover primordial argon not yet fully expelled by the
process of outgassing that is supposed to have occurred since the initial formation of the earthhas very significant
implications.First, this is clearly consistent with a young earth, where the very short time scale since the creation of the earth
has been insufficient for all the primordial argon to be released yet from the earths deep interior. Furthermore, it would also
seem that even the year-long global catastrophic Flood, when large-scale convection and turdecer occurred in the mantle,
was insufficient to expel all the deep earths primordial argon.264Second, this primordial argon is, in part, excess 40Ar not
generated by radioactive decay of 40K, which has then been circulated up into crustal rocks where it may continue migrating
and building up to partial pressure status regionally. Because the evidence clearly points to this being the case, when
samples of crustal rocks are analyzed for K-Ar dating the investigators can never really be sure that whatever 40Ar* is in the
samples is from in situ radioactive decay of 40K since the formation of the rocks, or whether some or all of it is from the
excess 40Ar* geochemical reservoirs in the lower and upper mantles. This could even be the case when the K-Ar analyses
yield dates compatible with other radioisotopic dating systems and/or with fossil dating based on evolutionary
assumptions. And there would be no way of knowing because the 40Ar* from radioactive decay of 40K cannot be
distinguished analytically from primordial40Ar that is not from radioactive decay, except, of course, by external assumptions
about the ages of the samples.Therefore, these considerations call into question all K-Ar dating, whether model ages or
isochron ages, and all40Ar/39Ar dating, as well as fossil dating that has been calibrated against K-Ar dates. Although
seemingly insignificant in themselves, the anomalous K-Ar model ages for these recent andesite flows at Mt. Ngauruhoe,

New Zealand, lead to deeper questions. Why is there excess 40Ar* in these rocks? From where did it come? Answers to
these questions in turn point to significant implications that totally undermine such radioactive dating and that are instead
compatible with a young earth.
Future Research
Further research is very definitely warranted. The most pressing need is to attempt to quantify how much primordial 40Ar
there is today in the upper mantle, as well as how much has circulated into crustal rocks, how much is in natural gas
reservoirs, and how much might have escaped into the atmosphere during 6,0007,000 years (including at accelerated
rates during the Flood). It might then be possible to quantify how much primordial 40Ar there was in the mantle at the time of
the earths creation. From these calculations and associated modeling exercises there might develop quantifiable evidence
for the earths youth.Additionally, further research is needed to quantify how much excess 40Ar* is in all the crustal rocks
and minerals that have been, and are, subject to K-Ar and 40Ar/39Ar dating. This would include what are regarded as mantle
xenoliths and xenocrysts (for example, diamonds). It is helpful to show on the one hand that such dating is questionable,
but on the other hand there are still many dates that are concordant, that is, they agree with other uniformitarian dating
systems and schemes. So ultimately we need to explain this agreement when other dates are discordant and anomalous.
There may, in fact, be some pattern or systematic way in which excess 40Ar has been trapped in rocks and occluded in
minerals at different levels (depths and relative ages) in the geological record. If so, then K-Ar and 40Ar/39Ar dating would
irrevocably be discredited.
The Fallacies of Radioactive Dating of Rocks
Basalt Lava Flows in Grand Canyon
by Dr. Andrew A. Snelling on June 28, 2006
Deep in the eastern Grand Canyon,
Arizona, is a stacked sequence of
ancient basalt lava flows known by
geologists as the Cardenas Basalt
(Figures 1 and 2).
These were once molten lavas that
were successively erupted onto the
earths surface through volcanic vents
and fissures and flowed over already
deposited layers of siltstones. They
quickly hardened into the dense,
black rock called basalt (Figure 3). Further thick sediment layers
were soon deposited on top of them. The position of the
Cardenas Basalt lava flows in the overall succession of rock
layers in Grand Canyon can be seen in the generalized geologic
block diagram of Grand Canyon (Figure 4). This article explores
the fallacious assumptions used by most secular geologists who
apply a uniformitarian belief system (slow-and-gradual geologic
processes) when dating the rocks of Grand Canyon.
Figure 1 The stacked sequence of Cardenas Basalt lava flows in
eastern Grand Canyon as seen from Desert View.
Determining the Ages of Rocks
Can the exact age of these basalt lavas be determined by their
position in the Grand Canyon sequence of rock layers or by
physically examining them? No! However, most geologists
classify the Cardenas Basalt as Precambrian; that is, it is older
than the so-called Cambrian rock layers such as the Tapeats Sandstone and Bright Angel Shale that contain fossils of
marine creatures such as trilobites (Figure 4). But the rock itself looks like many other basalts found all around the
world (Figure 5). With the application of radioactivity to the dating of rocks, these geologists believe they now have a virtually
foolproof method for determining exact ages.
Figure 2 Grand Canyon map showing the location of
outcrops of the Cardenas Basalt lavas.
Molten lavas, which successively erupted onto the
earths surface through volcanic vents and fissures,
flowed over already deposited layers of siltstones. They
quickly hardened to the dense, black rock called
basalt (see fig. 3). Further thick sediment layers were
soon deposited on top of them.Certain parent elements
such as potassium, rubidium, uranium and samarium are
radioactive and change by decay over time (very slowly
according to todays laboratory measurements) into
daughter elements argon, strontium, lead and
neodymium, respectively. This radioactive decay is like
the ticking of a clock, except instead of seconds these
radioactive-decay clocks are said to measure in millions
of years. By analyzing these parent and daughter
elements in a rock and applying todays measured rates
of radioactive decay, geologists confidently claim they can calculate how long it took the radioactive decay of the parents to
produce the measured quantities of daughters. This is then said to be the rocks exact (or absolute) age.
How Old Are the Cardenas Basalt Lavas?
These radioactive dating methods have been used to calculate an absolute age of 1,10366 million years for the Cardenas
Basalt lavas. (The number after the symbol refers to the error margins in the age determination so that 1,10366 million
years means that the age is between 1,037 and 1,169 million years.) So it would seem! However, a closer examination of
the results from all such studies reveals the fallacies of the radioactive dating methods.

Figure 3 Close view of the outcropping Cardenas Basalt


lava flows.
The claimed age of 1,10366 million years was obtained
using the rubidium-strontium isochron method with 10
samples and has been regarded as the best radioactive
dating result for any Grand Canyon rock unit.
Nevertheless, potassium-argon model ages for each of
15 individual Cardenas Basalt samples range from
57712 to 1,01337 million years, while the potassiumargon isochron age obtained using 14 samples is only
51630 million years. This is less than half the rubidiumstrontium isochron age of 1,11181 million years
obtained using 19 samples. It is also less than the
claimed Cambrian age of the Tapeats Sandstone that
sits on top of, and well above the Cardenas Basalt
lavas (Figure 4). Worse still, the samarium-neodymium
isochron age obtained using 8 samples is 1,588170 million yearsmore than three times the potassium-argon isochron
age of 51630 million years!
So what is the correct age of the Cardenas Basalt lavas?
(a) 51630 million years (the potassium-argon isochron age)
(b) 1,11181 million years (the rubidium-strontium isochron age)
(c) 1,588170 million years (the samarium-neodymium isochron age)
(d) None of the above
How can we know for sure which is the correct age when there is no independent cross-check?
Used by permission from ICR.
Figure 4 Geologic block diagram of Grand Canyon
rock layers showing the Cardenas Basalt lava flows in
the tilted Precambrian rock sequence and the recent
basalt lavas that flowed down the Canyon walls from
volcanic cones on the north rim.
Faster Radioactive Decay?
In any case, could it be that these radioactive clocks
have previously ticked at faster rates than they do
today? There is much convincing evidence to support
this possibility (see Further Reading at end of article).
Thus, radioactive decay of samarium was faster than
that of rubidium, which was faster than that of
potassium. Such accelerated radioactive decay rates
would
mean
that
these
basalt lavas may instead be only thousands of years old. Certainly the
radioactive dating methods simply cannot be trusted.Of course, those who
want the earth to be billions of years old, in order to accommodate their
belief in the evolution of the earth and all life on it, insist that these
radioactive rock dating methods still give ages of millions of years instead of
only the few thousand years required by the strict creation timescale.
However, what ages do these same methods yield when applied to rocks for
whose formation there are independent cross-checks?
Figure 5 View of a typical sample of the Cardenas Basalt under the geologic microscope showing its different colored
constituent minerals.
Recent Basalt Lava Flows and Their Age
In the western Grand Canyon are some other basalt lava flows (Figure 6). There are up to 160 volcanic cones on the
plateau to the north of the Canyon rim out of which these basalt lavas flowed. The eruptions were so recent, occurring after
the Grand Canyon formed, that some of these basalt lava flows cascaded like molten waterfalls over the Canyon rim, down
the Canyon walls and into the Canyon, where they formed dams that temporarily blocked the flow of the Colorado
River (Figures 4 and 7).
Its even possible the American Indians may have witnessed some of this awesome spectacle.1 Yet, when applying the
widely accepted assumptions of secular geologists, these basalt flows yield potassium-argon model ages of around
500,0001 million years. Worse still, their rubidium-strontium isochron age is 1,143220 million years! This is identical to the
rubidium-strontium isochron age of 1,11181 million years for the Cardenas Basalt lavas near the bottom of Grand Canyon
strata sequence (Figure 4).
Figure 6 View of a typical sample of a recent basalt lava flow in the western
Grand Canyon showing its different colored constituent minerals.
The Fallacies of Radioactive Dating
So how can the youngest basalt lava flows in Grand Canyon, whose
eruption was possibly witnessed only thousands of years ago, yield the
same radioactive rubidium-strontium age of 1.1 billion years as some of the
oldest basalt lava flows at the bottom of the Canyon? Answer: The molten
rock that produced the young basalt lava flows came from deep inside the
earth, from what geologists call the earths mantle; so these lavas have
inherited this rubidium-strontium composition from their mantle source. That
is, their rubidium-strontium composition has nothing to do with their age, but
everything to do with their source!However, the molten rock that produced
the ancient Cardenas Basalt lavas also came from the same region of the earths mantle beneath the Grand Canyon.

Therefore, it could be equally argued that the same rubidium-strontium composition of the Cardenas Basalt lavas was also
inherited from the same mantle source and thus has nothing to do with their age! Of course, some radioactive decay could
have occurred in the mantle source region of both these basalt lavas, but again, such radioactive decay there in the mantle
would not provide the basis for calculating any date for when these basalt lavas flowed out onto the earths surface.
So whichever way we look at it, these radioactive dating methods which are based on the biased assumptions of secular
geologists completely fail to date these basalt lava flows in Grand Canyon. How can we be sure radioactive decay has
always occurred at the rates we measure today? And how can we know how much radioactive decay occurred in their
mantle source before these lavas erupted? If the youngest lavas inherited all of their supposed radioactive decay age, then
so could the ancient lavas. After all, we know the true age of the young lavas because their eruption was possibly
witnessed within human history. Where there is an independent cross-check, these radioactive dating methods fail
completely. So because of these fallacies, why should they be trusted to give ages for any rock?
Figure 7 View of a frozen recent basalt lava that flowed down the walls of the
western Grand Canyon.
The True Ages of Grand Canyon Basalt Lavas
As for the true ages of these basalt lava flows in Grand Canyon, it is not difficult
to determine them when the creation framework and timescale for the earths
history are accepted as our absolute authority. The Cardenas Basalt lavas
erupted not long before fossil-bearing sedimentary layers were deposited on top
of them. So they would appear to have flowed over the pre-Flood ocean floor,
perhaps even within a few years of the fountains of the great deep breaking up
the ocean floor at the beginning of the Flood about 4,500 years ago. (The
scriptural description of the pre-Flood world does not preclude lavas flowing
quietly onto the pre-Flood ocean floor, just as lavas flow onto the ocean floor
unnoticed today by people living across the earths land surfaces.) Then because
the Grand Canyon was carved out at the very end of the Flood year (or soon
thereafter), the basalt lavas that cascaded down the Canyon walls did so after the
Flood, and after the American Indians had migrated to the area after having been
dispersed from Babel. So the lavas would be fewer than 4,000 years old.

Radioisotope Dating of Rocks in the Grand Canyon


by Dr. Andrew A. Snelling on June 1, 2005
Originally published in Creation 27, no 3 (June 2005): 44-49.
The radioisotope methods, long touted as irrefutably dating the earth as countless millions of years old, have repeatedly
failed to give reliable and meaningful absolute ages for Grand Canyon rocks.
Shop Now
Rafting through Grand Canyon, northern Arizona, is a most exhilarating and enjoyable experience. Deep below the rim, the
crystalline basement rocks tower above the turbulent Colorado River. Official publications say these rocks are more than a
billion years old, but when the methods used to date them are carefully examined, a totally different story is discovered.
Figure 1: Crystalline rockslight-coloured and pink granites, and darker metamorphic
rockswithin the Inner Gorge of Grand Canyon.
Clearly visible in the walls of the Inner Gorge are spectacular light-coloured rocks, such as
the pink granites,1 which stand out starkly against the darker, metamorphic rocks2 (figures
1 and 2). The latter are former sedimentary and volcanic layers that have been
transformed (metamorphosed) by heat and pressure during intense geologic upheavals in
the past.Among these volcanic layers are distinctive dark-coloured rocks called
amphibolites3 (figure 3). These were once flows of basalt lava, up to tens of metres thick
(figure 4). Some outcrops reveal round pillow structures, showing that the basalt lavas
erupted under water.
How Old are the Rocks?
Based on radiometric dating, long-age geologists say that the basalt lavas erupted 1,745
million years ago4 and were metamorphosed some 1,700 million years ago.5 Many
people, including many scientists, accept these dates as absolute truth. They believe that
when different radio-dating methods are used on the same rocks, they will all yield the
same age. But the quest to test this belief by sampling rocks from deep within Grand
Canyon has found it is not true.
Figure 3: Outcrop of the black, metamorphosed basalt flows called amphibolites.
Figure 4: Closer view of an outcrop
of amphibolite where a single 1.5 m
(5 ft) wide, original basalt flow (now
tilted vertically) can be identified.
Dating Amphibolite Samples
Figure 5: The
thin vertical
amphibolite layer (darker rock) just
upstream of Clear Creek.
During several raft trips through
Grand Canyon, many samples of
these Brahma amphibolites were

collected from various outcrops in the Inner Gorge.6 These included seven samples from a single amphibolite body (figure
5).7
All the samples were sent to two well-respected commercial laboratories for radioisotope testing.8 Both laboratories use
standard, best-practice procedures on state-of-the-art equipment and routinely provide accurate and repeatable
measurements of the required isotopes.
It is important to realize that the laboratories do not measure the age of the rocks but the isotopes in them at the present
time. Geologists calculate an age using the measured amount of a daughter isotope (e.g. argon) and its corresponding
parent isotope (e.g. potassium).
However, before this calculation can be made, it is necessary to assume how much of the daughter and parent were
present when the rock formed. It is also necessary to assume that no isotopes were gained or lost over time and that the
rate of radioactive decay has remained constant at the very slow rate measured today.
Figure 2: Schematic geologic diagram of the rock layers in Grand Canyon. The crystalline rocks of the
Inner Gorge are below the horizontal strata of the main canyon walls (after Austin, ref. 13).
The problem is that we dont know whether these assumptions are reasonable (because they are not
provable), and it is especially awkward for metamorphic rocks. Geologists overcome this problem by
interpreting the result. For example, the calculated age could be taken as the date of
metamorphism, or it could be the age of the original volcanic (or sedimentary) rock, or something in
between, or something else.
Different Methods Dont Agree
The calculated ages for all the individual samples from the same geologic formation using
the same dating method turned out to be vastly different (see box, Calculating the ages, below), even
for those closely spaced samples from the same outcrop of the same lava flow. The results are not
even close to each other, although the samples should all have given the same age.Furthermore, the
ages calculated for these Grand Canyon rocks using three different isochron methods also disagreed
greatly with each other. Even when the error margins are taken into account, the three different dating
methods give completely different ages that cannot be explained away.Indeed, none of the isochron
ages corresponds to the date for any theorized geologic eventneither the original lava eruptions
nor the subsequent metamorphism. Clearly, the calculated ages are useless for dating any event.
Calculating the ages
The so-called model1 potassium-argon (K-Ar) ages calculated for each of the 27 amphibolite samples from Grand
Canyon ranged from 405.1 10 Ma (million years) to 2,574.2 73 Ma. That is a six-fold difference, for samples that should
be of similar age.Note that the error estimates (the numbers) are small compared with the age. They are also small
compared with the variation in ages between samples. This means that the laboratory testing was precise. However, as the
results show, the error estimates say nothing about the accuracy of the ages of the rock samples.Furthermore, the seven
samples from the small amphibolite unit near Clear Creek, which should be even closer in age because they belong to the
same metamorphosed basalt lava flow, yielded K-Ar model ages ranging from 1,060.4 28 Ma to 2,574.2 73 Ma (figure
6). This includes two samples only 0.84 m (2 ft 9 in) apart that yielded K-Ar model ages of 1,205.3 31 and 2,574.2 73
Ma. Clearly, there is a problem with the assumptions on which the K-Ar ages are calculated.The isotopic results other than
potassium-argon (K-Ar), namely rubidium-strontium (Rb-Sr), samarium-neodymium (Sm-Nd) and lead-lead (Pb-Pb), were
used to calculate ages for the rock formation utilizing isochrons.2 Three ages altogether were obtained, one for each
isotopic system.The best isochron plots are where the straight line of best-fit falls within the analytical errors (the values)
for each data point. Routinely, if the data set is large, a few outlying data points are ignored if they dont plot on the line.
Geologists justify this, saying that some geochemical alteration in the past disturbed the radioisotopes in those samples.
The best-fit isochron plots for these amphibolites yielded a Rb-Sr age of 1240 84 Ma from 19 of the 27 samples, a SmNd age of 1,655 40 Ma from 21 samples, and a Pb-Pb age of 1,883 53 Ma from 20 samples (figure 7). 3Note that the
quoted age error margins (the values) are relatively small, due to the excellent statistical fit of these isochrons to the
data. In spite of this, the three different radioisotope methods give three very different agesthat is the isochron
discordance is pronounced. Figure 8 graphically illustrates how that, even when the calculated error margins are taken into
account, the different radioisotope dating methods yield vastly different ages that cannot be reconciled.
References
A model age is calculated by assuming a value for the original isotopic composition of the molten liquid from which the rock
solidified. In the case of K-Ar, it is assumed that when the rock formed, there was no Ar in it derived from radioactive decay
of K.BackAn isochron is a graphical plot of the isotopic compositions of the samples. It allows an isochron age to be
calculated from a straight line plotted through the graph of the results. The Isoplot computer program, developed by Dr Ken
Ludwig at the University of California Berkeley Geochronology Center, was used. See: Ludwig, K.R., Isoplot/Ex (Version
2.49): The Geochronological Toolkit for Excel, University of California Berkeley, Berkeley Geochronology Center, Special
Publication No. 1a, 2001. The method effectively requires multiple assumptions, namely that the initial isotopic ratio of each
sample was the same as the ratio of every other sample in the group.BackIt is important to note that geologists routinely use
only 610 samples for plotting isochrons and calculating isochron ages, so the isochrons obtained here from 1921 samples
are exceptional. Furthermore, all the results not included in the isochron age calculations still plotted very close to the lines
of best fit.Back
The Rule, Not the Exception
Some might want to dismiss these conflicting dating results as an isolated abnormality. They might claim that they are due
to the uncertain effects of metamorphism and later alteration, especially erosion and weathering. But these are not isolated
results. They are further confirmation of the repeated failure of all the radioisotope dating methods to successfully date
Grand Canyon rocks.Austin, S.A. and Snelling, A.A., Discordant potassium-argon model and isochron ages for Cardenas
Basalt (Middle Proterozoic) and associated diabase of eastern Grand Canyon, Arizona; in: Walsh, R.E. (Ed.), Proceedings
of the Fourth International Conference on Creationism, Creation Science Fellowship, Pittsburgh, Pennsylvania, pp. 3551,
1998.,9It is not just creationists who are discovering these dating failures. Other geologists are also reporting that different
methods on the same rock unit give conflicting radioisotope dates.10 But in their reports those geologists include tenuous
interpretations to try to explain away the abnormal amounts of daughter isotopes. It seems they are trying to avoid the
inescapable conclusion that the radioisotope methods simply do not yield reliable ages.11

Conclusion
The radioisotope methods, long touted as
irrefutably dating the earth as countless
millions of years old, have repeatedly
failed to give reliable and meaningful
absolute ages for Grand Canyon rocks.
Irreconcilable disagreement within, and
between, the methods is the norm, even
at outcrop scale. In fact, when carefully
examined, this radioisotopic evidence is
consistent with the view that these rocks
are young (see box, Accelerated
radioactive
decay
in
the
past,
below).These results are a devastating
blow to the concept of long ages,
foundational to uniformitarian geology
and evolutionary biology. It is entirely
feasible that the basalt lava flows, now
deep in Grand Canyon, erupted onto the
ocean floor during Creation Week and
were metamorphosed in the upheaval
that produced dry land, just six thousand
years ago.12
Figure 7: Isochron plots for the Brahma
amphibolites. (A) Rb-Sr (B)
Sm-Nd (C) Pb-Pb. The
crosses and ellipses are the
data
points
(sample
analyses) and their sizes are
proportional
to
the

analytical errors.
Figure 6: Sketched plan view
of the extent of the thin
amphibolite
layer
just
upstream of Clear Creek
showing the locations of the
dated samples and their
calculated K-Ar ages.
Figure 8: Present half-lives versus isochron ages for the different
radioisotopes dating the Brahma amphibolites.
Accelerated radioactive decay in the past
Research by the RATE project1 has uncovered much evidence that radioisotope decay rates were accelerated in a global
catastrophic event in the recent past.2 This evidence includes the patterns of discordances between dates from the
different radioisotope systems.3,4,5,6For example, if accelerated radioisotope decay occurred, then alpha-decaying
radioisotopes would yield older isochron ages than beta-decaying radioisotopes. This is exactly the pattern in the Brahma
amphibolites in Grand Canyon (figure 8).Because the different radioisotope pairs are supposed to be dating the same
geologic (rock formation) event, different dates mean that the parent radioisotopes decayed at different rates over the
same time period. In other words, the decay of the parent radioisotopes was accelerated by different amounts, the decay of
those yielding older ages (the alpha-decayers) having been accelerated more. This again matches theory.Obviously, if
radioisotope decay was accelerated, say during the Flood, then the radioisotope decay clocks could never be relied upon
when they date rocks as millions and billions of years old. Indeed, there are several independent lines of irrefutable
evidence7,8,9which indicate that the rates of decay of these long-age radioisotopes were grossly accelerated during some
event in the past, up to millions of times faster than their currently measured rates. Thus, it is entirely plausible that the rocks
are only a few thousand years old.
Radioisotope Dating of Rocks in the Grand Canyon
by Dr. Andrew A. Snelling on April 14, 2006
Was Dr. Andrew Snelling irresponsible for pointing out dating discrepancies at the Grand Canyon?
Shop Now
Andrew Snellings article in the June 2005 issue is irresponsible in that it uses technical jargon that an untrained lay-person
would believe to be truth when he (as a Ph.D in Geology) should understand the fact that different radioisotope systems
behave differently during different geologic processes. The worst system to use is K-Ar in a rock that has been
metamorphosed. Ar is a noble gas that resides weakly in the crystal latice of minerals. It has been shown many times that it
is easy to lose Argon during metamorphism or subsequent disturbances and occasionally there is excess Argon that gives
spuriously older ages for rocks. Rb-Sr is another method that is difficult to interpret when there is evidence of thermal
events. These elements have been shown to be mobile (acting as an open system) during tectonic events. These are just
two examples of the problems....there are many more!Regardless, all the ages reported are rather old. I dont think you can
compare different systems that have independent decay constants to prove that it is a flawed system, and you cant use this
to prove that constants have accelerated over time.Please try to explain the behavior of the individual isotope systems in the
future. Also, if you asked any geochronologist interested in dating the crystallization age of the Precambrian rocks in the
Grand Canyon, not a single one would even consider using the K-Ar system.
C.H.
USA

There is nothing irresponsible with my article in Creation, volume 27, number 3, 2005, on the radioisotope dating of rocks in
Grand Canyon (pages 4449). The article reported that there were problems with the different radioisotope systems in
obtaining the true age of these metamorphosed basalts in the Precambrian crystalline basement of the Grand Canyon. How
can it be irresponsible for me to be showing exactly what C.H. has admitted, namely, the fact that different radioisotope
systems behave differently during different geologic processes. I found it rather interesting that C.H. claimed that all the
ages reported are rather old, when in fact most of the ages reported in the article are younger than the true age claimed
for these rocks in the relevant geologic literature. All the problems with the different radioisotope systems C.H. has
highlighted have already been extensively reported in my chapter in the first volume of the Radioisotopes and the Age of the
Earth (RATE) project published in 2000 and referenced in the Creation article, and yes, all these problems are reported in
the standard textbooks in geochronology. The article was designed to corroborate these problems, and it did. And I agree
with C.H. that none of these radioisotope systems are reliable in dating these rocks, which was the conclusion of the article.
Greetings,
My wife and I had the pleasure of hearingKen Ham speak several times at the Indiana Homeschool Convention a couple of
weekends ago and later in the week hearing Mike Riddle in Fort Wayne. Thank you so much for all that you do! We were
staying with a couple in Indianapolis for the convention, and as we were talking my wifes cousins husband stated that he
saw no conflict between creation and evolution. Because of your organization I was able to talk to him intelligently and then
refered him to your website for more in-depth information. Just a couplke of years ago I would have just sat there and
listened to him without knowing what to say. What a difference!
Oh well, it looks like the evolutionists have the upper hand on you now, though. As you are aware, they just found another
missing link in the Arctic. :) As soon as I read the article on Fox News I checked out your website and, sure enough, you
already had a response. I dont know whether to laugh at or cry for these people. Thanks again for your wisdom!
Steve
USA
There was no irresponsibility in using technical jargon when the article was labeled semi-technical. Besides, if technical
jargon wasnt used others would accuse us of misleading people by not using the correct terminology! How can it be
irresponsible to provide copious footnotes and side boxes that are well referenced so that people reading the article would
know the statements made in it are backed by fuller explanations and documentation that they can consult if they want or
need to?
Any comparison of the different radioisotope systems to prove that the constants have accelerated over time was only
tangential to the article. Besides, proof is never claimed in the article. But the demonstrated inconsistencies between the
different radioisotope systems is evidence consistent with accelerated radioactive decay, just as there are many other lines
of evidence well referenced in the article that are consistent with the same possibility.
It is clear that C.H. has unfortunately not read the available scientific literature on dating of the crystalline Precambrian rocks
in the Grand Canyon, because he would find there geochronologists have attempted to date some of the Precambrian
crystalline rocks using the K-Ar system. Indeed, several of the references in the article document those attempts. One of
those attempts is highlighted in a forthcoming article in the inaugural edition of Answers magazine, which will again
demonstrate how the radioisotope systems are flawed and thus unable to yield reliable ages for Grand Canyon crystalline
rocks. C.H. has only confirmed what we already knew: that there are many problems in trying to use the different
radioisotope systems to date rocks!
(Dr.) Andrew Snelling

Radioactive Dating in Conflict!


Fossil Wood in Ancient Lava Flow Yields Radiocarbon
by Dr. Andrew A. Snelling on December 1, 1997
Originally published in Creation 20, no 1 (December 1997): 24-27.
Shop Now
When miners were sinking a ventilation shaft for the new Crinum Coal Mine in Central Queensland in 1993 (see map below)
they unearthed a rare find. After digging through the thin surface sands and clays, followed by basalt, 21 metres (almost 69
feet) down they found pieces of wood entombed in the bottom basalt flow.1 Below the basalt were layers of claystone,
siltstone, and sandstone with interbedded coal seams.2
Fossil wood in ancient basalt
The wood was in three statesash, charred, and intact.1 Those on-site at the time
speculated that there had been two distinct trees, partly standing, still organic in nature,
and thus not petrified. The imprint of a leaf was also discovered within the basalt, which
was also regarded as remarkable, remembering that the enclosing rock was once
molten lava erupted at 10001200C (about 18002200F).So how could these tree
trunks have survived being engulfed by molten lava? At approximately four metres (13
feet) thick, the basalt flow is relatively thin,1,3 and thus cooling would have been rapid
(perhaps days, but a few weeks at most4). This is verified by the observed internal
structure of the basalt flow.1,5 Since the tree trunks were engulfed at the bottom of the
flow, cooling may have been immediate, with any water present in the wood aiding
extremely rapid encapsulation and thus preservation.The local geological context
makes the basalt flow approximately 30 million years old,1,3 in keeping with other
basalt flows in the region all regarded as of Tertiary age (in the conventional
terminology). Since the tree trunks were entombed in the basalt lava, the wood is thus
supposedly at least 30 million years old. Also, what looked like the tree roots were found in the siltstone below the
basalt,3 suggesting the trees when alive were rooted into the siltstone and thus growing on a land surface that was then
covered by basalt lava. This siltstone belongs to the Permian German Creek coal measures, conventionally believed to be
around 255 million years old.6Collection of samplesSmall fragments of some of the wood samples were kindly sent to us,
and a subsequent mine visit took place in late August 1994.7 The pieces of wood recovered by the miners were examined
and photographed, as too was the leaf imprint, but access to the ventilation shaft was not possible, nor were samples of the
enclosing basalt available, having long been dumped with all the other rubble and waste rock. However, an exploratory hole
had been drilled close to where the shaft was eventually dug. In the relevant drill core, at the bottom of the lowermost basalt

flow, pieces of fossil wood still containing organic carbon were present encased in the basalt, right at the boundary of the
basalt flow with the siltstone below. This drill core was subsequently sent to us once permission was granted by the mining
company.7After visiting the mine site, nearby outcrops of the same basalt flows were investigated and sampled. This was to
make sure we at least had some samples of the basalt, just in case permission to have the drill core wasnt forthcoming.

Charred fossil wood.

Intact fossil wood.

Basalt with holes from former gas bubbles.

Fossil tree with roots in siltstone.

Laboratory work
Tiny portions of the same piece of fossil wood encased in the basalt in the drill core were sent for radiocarbon (14C) analyses
to two reputable laboratoriesGeochron Laboratories in Cambridge, Boston (USA), and the Antares Mass Spectrometry
laboratory at the Australian Nuclear Science and Technology Organisation (ANSTO), Lucas Heights near Sydney (Australia).
Neither laboratory was told exactly where the samples came from to ensure that there would be no resultant bias. Both
laboratories use the more sensitive accelerator mass spectrometry (AMS) technique for radiocarbon analyses, Geochron
being a commercial laboratory and Antares being a major research laboratory. Also, tiny fragments of the initial wood
samples provided to us, from the pieces of wood that had been found during sinking of the ventilation shaft, were sent off for
radiocarbon analysesone set of different fragments to each laboratory.Pieces of the basalt samples from the outcrop and
the drill core were also sent to analytical laboratories, for major, minor, and trace element analyses to establish the character
of these rocks, but mainly for radioactive dating analyses. Potassium-argon (K-Ar) dating was performed on the two
outcrop samples by the AMDEL laboratory in Adelaide (Australia), while one of the two outcrop samples and two drill core
samples, one being in contact with the fossil wood, were dated by Geochron Laboratories.
Results
The radiocarbon (14C) results are listed in Table 1.8 It is immediately evident that there was detectable radiocarbon in all
wood samples, so that the laboratories staff had neither hesitation nor difficulties in calculating 14C ages. When
subsequently questioned regarding the limits of the analytical method for the radiocarbon and any possibility of
contamination, staff at both laboratories (Ph.D. scientists) were readily insistent that the results, with one exception,9were
within the detection limits and therefore provided quotable finite ages!8 Furthermore, they pointed to the almost identical
13C results (last column in Table 1), consistent with the carbon being organic carbon from wood, and indicating no
possibility of contamination. So the results in Table 1 are staunchly defended by the laboratories as valid, indicating an age
of perhaps 44,00045,500 years for the wood encased in the basalt retrieved from the drill core.In stark contrast to the age
of the wood are the potassium-argon (K-AR) ages of the basalt (see Table 2).8 It is readily apparent that there are
significant variations in the results, as evident in the calculated ages of the outcrop 2 sample provided by each laboratory.
The problem of obtaining consistently acceptable K-AR ages is also highlighted by the observation that both outcrop and
both drill core samples probably represent the same basalt flow in each respective location (hence the calculated average
ages in the last column of Table 2)10 The staff of both laboratories (again Ph.D. scientists) defended their analytical
results,8,11 and did not hesitate to affirm that these basalt samples are, according to their radioactive K-AR dating, around
45 million years old.
Lab code

14

Geochron
Wood in Drill Core ANSTO

GX-20798-AMS
OZB472

>35,620
44,700950

-25.7
-25.78

Other Wood

Geochron

GX-20087-AMS

29,544759

-25.1

Other Wood

ANSTO

OZB473

37,8003,450

-26.16

Sample

Lab

C age (BP) years

13CPDB15

Conclusions
While the quality and accuracy of the analytical work undertaken by all the laboratories involved is unquestionably
respected, all the calculated ages are mere interpretations based on unproven assumptions about constancy of radioactive
decay rates, and on the geochemical behaviour of these elements (and their isotopes) in the unobservable past. To youngearth creationists the geological context of these fossil wood fragments in the basalt lava flow clearly indicates that these
represent post-Flood trees overwhelmed by a post-Flood volcanic eruption nearby, and thus both the fossil wood and the
basalt are less than 4,500 years old.12
Nevertheless, within the conventional (uniformitarian) framework of interpretation, a clear-cut conflict can be seen between
these two radioactive dating methods. Normally fossil wood found in such an ancient basalt would not be radiocarbon
dated, because the wood would be considered far too old for any radiocarbon to be left in it.13 Yet here these radioactive
dating methods are again demonstrated to be unreliable and clearly useless at determining the true age of the wood and
basalt.14 Therefore, any published results from these dating methods should not be seen as casting any doubts
whatsoever on the reliability of the creation chronology so carefully provided for us .
Basalt sample

Lab

Lab code

K-Ar
age
years)

(millionAverage
years)

K-Ar

age

(million

Outcrop 1

AMDEL

G878300G/95

44.9 1.1

Outcrop 2

AMDEL
Geochron

G878300G/95
R-11800

47.9
39.1 1.5

Drill Core

Geochron

R-11798

58.3 2.0

Drill Core Enclosing


Wood
Geochron

R-11799

36.7 1.2

+4.0
1.6

43.9

4.8

47.510.8

Table 2: Potassium-argon (K-AR) dates on basalt samples. Return to text.

The age of Australian Uranium


A case study of the Koongarra uranium deposit
by Dr. Andrew A. Snelling on June 1, 1981
Originally published in Creation 4, no 2 (June 1981): 44-57.
The objective of this study is to examine the published U/Pb dating results obtained from mineral samples from the
Koongarra uranium deposit, Northern Territory.
Summary
The results have been obtained on the basis of the assumption that the U/Pb system must have always been closed. The
evidence presented serves to demonstrate conclusively that the U/Pb system in this uranium deposit has been so open that
it is impossible to be sure of the precise history of U and Pb in any sample selected for dating, making any age
determinations meaningless. The Th/Pb age of the deposit, indicated by the published results, is zero years, a contrasting
and significant result.
The Koongarra uranium deposit
Located 225 km (140 miles) east of Darwin in the Northern Territory (Fig. 1), the Koongarra uranium deposit is one of four
large high grade deposits now being developed in the area, known as the Alligator Rivers Province. The deposit lies beneath
a gently sloping erosion surface just below the sandstone cliffs of
the Mount Brockman Mesa, an outlier from a regionally extensive
sandstone plateau (Fig. 2). This thick, flat-lying sandstone
unconformably overlies the ore-hosting schists, and in the recent
past served as a cover rock for the Koongarra orebodies.
The uranium mineralisation occurs in two distinct orebodies
separated by a barren gap (Fig. 3). Both orebodies consist of
primary zones of partially coalescing lenses containing
pitchblende veins, essentially stratabound within the steeply
dipping host Lower Proterozoic schists (Fig. 2). With the erosion
of the sandstone cover, the top of the No. 1 orebody has been
exposed to surficial weathering processes, resulting in dispersion
of uranium downslope in the weathered rock to form a fan of
secondary mineralisation (see Fig. 2).
Minor galena (PbS), (FeS2), and chalcopyrite (CuFeS2) are
associated with pitchblende in the primary zones, along with rare
native gold and other sulphides.
U/Pb dating of the Koongarra deposit

Figure 2.
Figure 3.
The uranium ore at Koongarra has been dated by the U/Pb method (1,2). (See Appendix 3). Both pitchblendes and galenas
were handpicked from drill core and isotopically analysed. Table 1 lists the analytical results and the dates calculated from
them, while Fig. 4 shows the data plotted on the conventional Concordia diagram (the concordia is the curve along which
the 206Pb age equals the 207Pb age). Note that in Table 1 the ages as calculated actually range from 447 million A.G.Yr.
(A.G.Yr = arbitrary geologic years) to 1282 million A.G.Yr., and that the 206Pb age < 207Pb age < 207Pb206Pb age, just as
would be predicted on the basis of neutron reaction (3). But the age of the pitchblende was concluded to be 870 million
A.G.Yr. due to one concordant point on the concordia diagram (Fig. 4) amidst a wild scatter of the other four data points.
In order to account for the wild scatter of the data, the Pb isotopic ratios of the pitchblendes and associated galenas were
plotted on the conventional 207/204 vs 206/204 diagram (Fig. 5). The Pb ratios all plot on or between the two radiating lines
which represent the limiting fields of the data on the diagram. These lines are supposed to correspond to ages of
approximately 1800 million A.G.Yr (upper line), defined by galenas only, and 860 million A.G.Yr (lower line), defined by
pitchblendes only. Thus it was concluded by Hills & Richards (1,2) that the galenas associated with the pitchblendes
suggested an older 1800 million A.G.Yr source for their Pb isotopes. Consequently what I term arbitrary line fitting through
the wild scatter of pitchblende data points on the concordia diagram (Fig. 4 insert) confirmed the desired conclusion that
these pitchblendes are contaminated with the Pb produced by an 1800 million A.G.Yr old generation of pitchblendes which
has long since been destroyed.
A critical comment and revision

The conclusion that the 870 million A.G.Yr old pitchblendes in the samples studied, are contaminated with Pb from a
previous generation of now destroyed pitchblendes, is really another way of saying that both U and Pb have migrated with
time in the deposit.The geochronologists have in reality suggested that the present pitchblendes formed 870 million A.G.Yr
ago from migrating U released by the destruction of previous pitchblendes which did not leave a trace, apart from the minor
Pb contamination of the present pitchblendes, and effectively resetting the U/Pb isotopic clock. This is of course a
statement that this U/Pb, system was open. A situation which is contrary to the first basic assumption of radioactive dating
that the system must be closed. In spite of data manipulation to explain away discrepancies, there is no way that any
geochronologists could be sure that all the Pb in the pitchblendes is not derived from contamination thus producing an
erroneous large age.There are 4 main lines of evidence that the system at Koongarra has been open (4,5,6,7) and the
implications of these are that the attempts to date Koongarra using U/Pb isotopes are invalid.

Figure 4, Figure 5, Figure 6.


(a) Ore textures
Figure 7.
Figure 8.
Mineralogical and textural studies of the
ore under both optical and electron
microscopes indicate that there have been
at least three remobilisations of the
uranium during the history of the ore. Lead
has likewise been mobile. This is shown diagrammatically in Fig. 6 as several
generations of pitchblende and galena. Plates 16 illustrate examples of the ore textures as seen under the microscopes,
the descriptions accompanying the plates indicate how the textures have been interpreted.
(b) Mineral chemistry
Pitchblende compositions are never uniform. Electron microprobe analyses of pitchblendes, that is, micro-analyses of
volumes of pitchblende between 5 and 10 cubic microns (Table 2), reveal that pitchblende compositions, particularly U, Pb
and Ca contents, change not only from grain to grain within any one sample regardless of which generation of pitchblende it
is, but even at the microscopic level within pitchblende grains themselves. Fig. 8 illustrates how Pb and Ca are both
substituting for U in the UO2 cubic lattice.
(c) Supergene alteration
Supergene alteration (oxidation) of pitchblendes has not only occurred where the zone of surficial weathering has
intersected the top of the No. 1 orebody, but at depth within the primary ore. Pitchblendes have been replaced by a wide
variety of colourful secondary uranium minerals (Table 3), their identities depending on the chemistries of the immediate
environment and the circulating groundwaters (Fig. 2 and 8). The nett result has been dispersion of uranium over distances
of up to 50 metres or more in the weathered zone with complete destruction of pitchblendes, and yet another remobilisation
of both U and Pb in the primary ore zones with in situ replacement of pitchblendes (Plates 79) and deposition of supergene
pitchblendes and uranyl silicates from U in solution (Plates 1012). The analyses in Table 4 show how the U and Pb
contents decrease as pitchblende is altered to uranyl silicates, which absorbed the migrating U and Pb.

Plate 1. Remobilisation and redeposition of pitchblende (white mineral). Photomicrograph shows pitchblende veins (top and
bottom) partially destroyed by dissolution of uranium which has been redeposited as scattered veinlets and shapeless
masses of a new generation of pitchblende (middle). (magnification 10X)
Plate 2. Pitchblende (light grey) has been dissolved and redeposited as thin veinlets and shapeless masses within a chlorite
(dark grey) matrix which is also replacing the main pitchblende grain. (magnification 120X)

Plate 3. Two generations of pitchblendegrains (lighter grey), and more oxidised supergene veins and patches (darker
grey). The small scattered white grains are galens. (magnification 200X)
Plate 4. Two generations of pitchblendegrains (white, top of photomicrograph) and later think supergene encrustations
(mid grey) around quartz grains (dark grey). The very bright mineral (bottom) is galena which has similarly been dissolved
and redeposited. (magnification 200X)

Plate 5. Remobilised pitchblende (light grey) deposited as scattered grains with a chlorite (dark grey) matrix. A remobilised
galena vein (white-grey) cuts across the pitchblende-chlorite association. (magnification 50X)
Plate 6. An enlarged view of pitchblende (dark grey) sub-grains veinlets which both cross-cut and separate the pitchblende
sub-grains. The PB in the galena is supposed to have migrated from the pitchblende where it was supposedly produced by
radioactive decay. (magnification 50X)

Plate 7. Kasolite (white) and uranophane (grey) replacing a former pitchblende vein. Note that the former vein shape, even
the sub-grains, have essentially been preserved. (magnification 210X; scale bar 50 microns)
Plate 8. Globular pitchblende mass (black shape just to the left of centre) being altered marginally to sklodowskite (grey
concentric sheath). (magnification 2X; scale bar 3mm)

Plate 9. Kasolite (light grey) and sklodowskite (dark grey) replacing a former pitchblende vein. (magnification 210X)
Plate 10. Supergene colloform banded pitchblende (grey) deposited in what was originally a void. The banding is produced
by a time sequence of pitchblende deposition. (magnification 840X)

Plate 11. A sklodowskite (white) vein composed of radiating aggregates of needle-shaped crystals. (magnification 220X;
scale bar 50 microns)
Plate 12. Uranophane (white) veinlets deposited between quartz (grey) grain boundaries. (magnification 220X; scale bar 50
microns)
(d) Uranium/daughter disequilibrium
There are two methods of measuring the grade of a uranium ore sample:
by assaying for U directly using standard chemical or related techniques; and
by measuring the radioactivity given off by the ore sample, the quantity of such radioactivity being related to the U content.
However, because the radioactivity measured is actually the gamma radiation given off by the daughter element bismuth214 (214Bi) far down the 238U decay chain, any addition or removal of daughter elements between 238U and214Bi will result in a
discrepancy between the above two measurements of the U content of the ore sample. To assess this possibility the two
measurements are compared:
Three possibilities arise:
(i) Ratio = 1. The ore sample is said to be in equilibrium since the two measurements agree, implying that the U and its
daughter elements are in equilibrium; neither have apparently migrated.
(ii) Ratio > 1. The ore sample is said to be in disequilibrium and since the U content is greater than the daughter element
content either U has been added to the sample or daughter elements removed.
(iii) Ratio < 1. Again the ore is said to be in disequilibrium but now the U content is less than the daughter element content
implying either U removal or daughter element addition to the sample.
Measurements on ore samples from Koongarra indicate that the ore is in overall disequilibrium (Fig. 10 and Table 6). High
resolution gamma-ray spectroscopy was then used to determine which daughter elements of 238U have been mobilised.
These investigations showed that even though the high grade pitchblende ore is near equilibrium, radium-226 ( 226Ra) and
radon-222 (222Rn) (a gas) have migrated from the ore into the surrounding host rocks, the ore being generally depleted
in 226Ra and 222Rn, and the immediate host rocks being relatively enriched in these two elements (Table 6). Fig. 10
schematically illustrates the movements of isotopes caused by present day circulating groundwaters. This demonstrated
migration of 226Ra from the pitchblende ore is significant in that it removes the lower two-thirds of the 238U decay chain from
the pitchblende, thus resulting in 206Pb relocation and spurious 206Pb ages (8).
Figure 9.
Figure 10.
(e) Conclusions
The above evidence conclusively demonstrates that the U/Pb system, including its
intermediate daughter products, especially Ra and Rn, has been so open with
repeated large scale migrations of the elements that it is impossible to be sure of the
precise status/history of any piece of pitchblende selected for dating. Even though
geochronologists take every conceivable precaution when selecting pitchblende
grains for dating, in the light of the above evidence, no one could be sure that the U
and Pb they are measuring is original and unaffected by the gross element
movements observed and measured. Those pitchblende grains dated have always
contained Pb, both within their crystal lattices and as microscopic inclusions of
galena, making it impossible to be sure that all the Pb was generated by radioactive
decay from U. In addition, the pitchblende grains dont have uniform compositions
so that dating of sub-sections of any grain would tend to yield widely divergent
U/Pb ratios and therefore varying ages within that single grain. A logical extension
of these data & conclusions is to suggest as others already have (9,10). that U/Pb
ratios may have nothing to do with the age of a mineral. So that in spite of the
popular dating results looking sensible, the evidence clearly indicates that these
dates are meaningless.
.
Table 5 Analyses of iron and manganese oxides.
CAS 165

CAS 114/1

CAS 114/2

CAS 95/1

CAS 95/2

CAS 95/3

UO2

2.81

1.63

1.05

0.36

2.83

1.91

PbO

12.42

4.41

0.30

5.03

8.16

3.34

CaO

0.20

0.09

1.d.

0.04

0.15

0.12

SiO2

2.49

3.11

6.28

2.87

2.54

3.20

Fe(FeO) 5.50

8.71

81.46

0.47

11.09

58.16

MnO2

80.35

1.96

88.52

73.53

27.70

77.48

MgO

0.12

0.37

Al2O3

0.15

1.23

P2 O5

0.33

1.d.

V2 O3

1.d.

Total

101.50

99.90

2.09

93.14

0.29

0.52

0.22

2.70

0.82

1.75

1.d.

1.d.

0.31

0.65

0.26

100.59

100.29

96.66

Appendix 1
A zero age for the Koongarra uranium deposit
In their U/Pb dating of the Koongarra uranium deposit (1, 2). the geochronologists overlooked one critical aspect
the232Th/208Pb relationship. In Table 1, taken directly from the published results (1), the %Th, 208Pb proportion, and208Pb/207Pb
and 208Pb/204Pb ratios are given, but the 208Pb ages are conspicuously absent. Their omission is significant.
The 204Pb content of the pitchblende is regarded as common or original Pb since it is not derived from any parent element
via radioactive decay. Because this so-called common Pb correction has to be applied to the raw data before calculation of
the U/Th/Pb ages. This, of course, is again another way of saying that not all the Pb is derived by radioactive decay, but that
it was with the U and Th in the beginning. The standard used to correct the data in Table 1was the Mt Isa Pb* standard
with an isotopic composition:
1.44% 204Pb 23.20% 206Pb

22.48% 207Pb

52.88% 208Pb
When this common Pb correction is applied to the data in Table 1 (11). most of the 208Pb has resulted from common Pb
contamination. In fact, in samples J804/1, J804/b and J807 all the 208Pb is due to contamination and none resulted
from 232Th decay, thus resulting in 208Pb ages of 0 yrs for these samples. The remaining two samples yield 208Pb ages (11) of
275 million A.G.Yr (J801) and 61 million A.G.Yr (J809), both considerably less than all the other Pb ages. Since
the 232Th/208Pb age is taken as the standard isotopic Pb clock (3), the consensus radiometric age of the Koongarra uranium
deposit is 0 yrs, since this is the only isotopic Pb date supported directly by a majority of samples. One suspects that
the 207Pb ages were left out of the published data because of the uncomfortable implications!
*It should be noted in passing also that the choice of the Mt Isa Pb standard is based on one of several possible theories of
element nucleogenesis and Pb evolution (1,2,3), making the whole age calculation procedure rather subjective, since it is
based on additional assumptions.
Appendix 2
All radiometric dating methods are based on three assumptions:
The physico-chemical system must have always been closed. Thus nothing inside the system could have been removed,
and nothing outside the system added to it.
The system must initially have contained none of its daughter elements.
The decay rate or half-life of the radioactive parent element must have always been the same, that is, constant.
Appendix 3
The U/Pb method is the most important of all the radiometric dating techniques, not only because it was first historically, but
because other methods are calibrated against it. U/Pb dating actually involves the measurement of several long parentdaughter decay chains, namely:
uranium-238 (238U) decays to produce lead-206 (206Pb) with a half-life of 4.5 billion A.G.Yr.
uraniun-235 (235U) decays to produce lead-207 (207Pb) with a half-life of 0.7 billion A.G.Yr.
thorium-232 (232Th) decays to produce lead-208 (208Pb) with a half-life of 14.1 billion A.G.Yr.
Helium Diffusion Rates Support Accelerated Nuclear Decay
by Dr. Russell Humphreys, Dr. Andrew A. Snelling, Dr. John Baumgardner, and Dr. Steve Austin on February 2, 2011
Abstract
Two decades ago, Robert Gentry and his colleagues at Oak Ridge National Laboratory reported surprisingly high amounts
of nuclear-decay-generated helium in tiny radioactive zircons recovered from Precambrian crystalline rock, the Jemez
Granodiorite on the west flank of the volcanic Valles Caldera near Los Alamos, New Mexico (Gentry, Glish, & McBay 1982).
Up to 58% of the helium (that radioactivity would have generated during the alleged 1.5 billion year age of the granodiorite)
was still in the zircons. Yet the zircons were so small that they should not have retained the helium for even a tiny fraction of
that time. The high helium retention levels suggested to us and many other creationists that the helium simply had not had
enough time to diffuse out of the zircons, and that recent accelerated nuclear decay had produced over a billion years worth
of helium within only the last few thousand years, during Creation and/or the Flood. Such acceleration would reduce the
radioisotopic time scale from megayears down to months. However, until a few years ago nobody had done the
experimental and theoretical studies necessary to confirm this conclusion quantitatively. There was only one (ambiguously
reported) measurement of helium diffusion through zircon (Magomedov 1970). There were no measurements of helium
diffusion through biotite, the black mica surrounding the zircons. In 2000 the RATE project (Humphreys 2000) began
experiments to measure the diffusion rates of helium in zircon and biotite specifically from the Jemez Granodiorite. The
data, reported here, are consistent with data for a mica related to biotite (Lippolt & Weigel 1988), with recently reported data
for zircon (Reiners, Farley, & Hickes 2002) and with a reasonable interpretation of the earlier zircon data (Magomedov
1970). We show that these data limit the age of these rocks to between 4,000 and 14,000 years. These results support our
hypothesis of accelerated nuclear decay and represent strong scientific evidence for the young world of Scripture.
Keywords: helium, diffusion, radioisotopes, age, dating, nuclear decay, zircon, biotite, accelerated decay
This paper was originally published in the Proceedings of the Fifth International Conference on Creationism, pp. 175195
(2003) and is reproduced here with the permission of the Creation Science Fellowship of Pittsburgh (www.csfpittsburgh.org).
Introduction
A significant fraction of the earths radioactive elements, particularly uranium and thorium, appear to be in the granitic rock of
the upper continental crust. Uranium and thorium tend to be localized in the granites inside special minerals such
as zircon (zirconium silicate, ZrSiO4). Zircon has high hardness, high density, and high melting point, often forming
microscopic, stubby, prismatic crystals with dipyramidal terminations (fig. 1) commonly grayish, yellowish, or reddish brown.
Atoms of uranium and thorium within cooling magma replace up to 4% of the normal zirconium atoms within the lattice

structure of zircon as it is crystallizing. The radioactive zircon crystals often become embedded in larger crystals, such as
mica, as magma cools and solidifies.As the uranium and thorium nuclei in a zircon decay, they produce helium. For
example, uranium-238 (238U) emits eight alpha particles as it decays through various intermediate elements to lead-206
(206Pb). Each alpha particle is a helium-4 nucleus (4He), consisting of two protons and two neutrons. Each explosively
expelled 4He nucleus eventually comes to a stop, either within the zircon or in the surrounding material. There it quickly
gathers two electrons and becomes a neutral helium atom.
Fig. 1. Zircons from the Jemez Granodiorite. Photo by R. V. Gentry.
Helium is a lightweight, fast-moving atom that does not form chemical
bonds with other atoms. It can diffuse through solids relatively fast,
meaning that helium atoms wiggle through the spaces between atoms in a
crystal lattice and spread themselves out as far from one another as
possible. For the same reason it can leak rapidly through tiny holes and
cracks, making it ideal for leak detection in laboratory vacuum systems.
The diffusion and leakage rates are so great that believers in the billions of
years had expected most of the helium produced during the alleged 4.5
billion years of the earths existence to have worked its way out of the
crust and into the earths atmosphere long ago.But the helium is not in the
earths atmosphere! When non-specialists hear that, they usually assume
that (A) helium has risen to the top of the atmosphere as it would in a
balloon, and (B) most of the helium has then leaked from the top of the
atmosphere into space. However, assumption (A) is wrong, because
unconfined helium spreads throughout the atmosphere from top to bottom,
like any other gas.On assumption (B), the simple kinetic theory of gases says the loss of neutral helium atoms into space
would be much too small to account for the missing helium. In 1957 Melvin Cook, a creationist chemist, pointed out this
problem in the prestigious journal Nature.1 In 1990 Larry Vardiman, a creationist atmospheric scientist, calculated that even
after accounting for such slow leakage into space, the earths atmosphere has only about 0.04% of the helium it should have
if the earth were billions of years old.2Until the 1970s, uniformitarians (see next section) had no good answer. However in
recent decades they have been trying to evolve one. Satellite data3,4 show that ions (electrically charged atoms) of helium
(and other gases) move back and forth along the earths magnetic lines of force above much of the atmosphere. Some
space plasma physicists theorize that storms of particles from the sun blow the helium ions loose from the lines of force
outward into space frequently enough to balance the [assumed] outgassing from the earths crust.5 The theory is very
complex, and no creationist expert in the field has yet reviewed it to see whether it is well founded.Rapid helium leaks into
space are essential to uniformitarians, but slow leaks are not essential to creationists. If the leakage turns out to be slow, it
would bolster our case here. But fast leakage would not damage our case. The next section offers evidence for a much
simpler explanation of the missing atmospheric helium: most of the radiogenic (nuclear decay generated) helium has not
entered the earths atmosphere. It is still in the earths crust and mantlemuch of it still in the zircons. In this paper we
argue that the helium has not had enough time (less than 14,000 years) to escape the zircons, much less the crust.
The Helium is Still in the Zircons
In the 1970s, geoscientists from Los Alamos National Laboratory began drilling core samples at Fenton Hill, a potential
geothermal energy site just west of the volcanic Valles Caldera in the Jemez Mountains near Los Alamos, New Mexico (fig.
2). There, in borehole GT-2, they sampled the granitic Precambrian basement rock, which we will refer to as the Jemez
Granodiorite. It has a radioisotopic age of about 1.5 billion years, as measured by
various methods using the uranium, thorium, and lead isotopes in the zircons
themselves.6 The depths of the samples varied from near the surface down to 4.3
kilometers, with temperatures from 20C to 313C. The Los Alamos team sent some of
these core samples to Oak Ridge National Laboratory for isotopic analysis.Most of the
zircons were in biotite,7 a black mica common in granitic rock. At Oak Ridge, Robert
Gentry, a creationist physicist, crushed the samples (without breaking the much harder
zircon grains), extracted a high-density residue (because zircons have a density of 4.7
grams/cm3), and isolated the zircons by microscopic examination, choosing crystals
about 5075 m long. The zircon masses were typically on the order of micrograms.
The Oak Ridge team then heated the zircons to 1000C in a mass spectrometer and
measured the amount of helium 4 liberated. In 1982 they published the data
in Geophysical Research Letters.8 Table 1 details their results.The first column
itemizes the samples analyzed. The second and third columns show the depth and
temperature of each sample in situ. The fourth column shows the volume (at standard
temperature and pressure) of helium liberated in the lab per microgram of zircon.
Fig. 2. Drilling rig at Fenton Hill, New Mexico. Photo by Los Alamos National
Laboratory.
The fifth column is the ratio of the observed quantity of helium Q (total number of
helium atoms in the crystal) to the calculated quantity Q0 that the zircons would have accumulated and retained if there had
been no diffusion. The Los Alamos team measured the amount of radiogenic lead in zircons 2.9 km deep in the same
borehole and same granodiorite,9 and the Oak Ridge team confirmed those figures with their ion microprobe.10Because the
various decay chains generate an average of 7.7 helium atoms per lead atom produced, Gentry and his colleagues were
able to calculate Q0 from the amount of lead in the zircons. In doing so, they compensated for the estimated loss of alpha
particles emitted from near the edges of the zircons out into the surrounding material.The Oak Ridge team estimated that
uncertainties in calculating Q0 might limit the accuracy of the ratio Q/Q0 to 30%. The sixth column of the table shows the
resulting estimated errors in the ratios.Samples 1 through 6 came from the granodiorite, but sample zero came from larger
zircons in a surface outcrop of an entirely different rock unit. For that rock unit U/Th/Pb information was not available,
making an estimate of Q0 not feasible. Lacking a ratio, we cannot use sample zero in the calculations.Samples 5 and 6 had
the same amount of helium. Gentry and his colleagues noted that helium emerged from those samples in shorter bursts
than the other samples, indicating a different distribution of helium within those zircons. In section 6, we will show that the
amount of helium from sample 5 is just about what would be expected from the trend in the cooler samples. But we allow for
the possibility of its error being considerably larger than the cooler samples.According to the thermal behavior outlined in the
next section, we would ordinarily expect that the hotter sample 6 would have much less helium than sample 5. The fact that
the helium content did not decrease suggests that some additional effect may have occurred which limited the outflow of

helium from the zircon. In section 6 we suggest a likely explanation.The above considerations suggest that we can use
samples 1 through 5 in a theoretical analysis with ordinary diffusion. We will treat sample 6 as a special case.
Samples 1 through 3 had helium retentions of 58, 27, and 17%. The fact that these percentages are high confirms that a
large amount of nuclear decay did indeed occur in the zircons. Other evidence strongly supports much nuclear decay having
occurred in the past.11 We emphasize this point because many creationists have assumed that old radioisotopic ages are
merely an artifact of analysis, not really indicating the occurrence of large amounts of nuclear decay. But according to the
measured amount of lead physically present in the zircons, approximately 1.5 billion years worthat todays ratesof
nuclear decay occurred. Supporting that, sample 1 still retains 58% of all the alpha particles (the helium) that would have
been emitted during this decay of uranium and thorium to lead.
Table 1. Helium retentions in zircons from the Jemez Granodiorite.
It is the uniformitarian assumption of
Depth (m) Temperature (C) He (109cm3/g) Q/Q0
Error
invariant decay rates, of course, that leads to
the usual conclusion that this much decay
0 0
20
82

required
1.5
billion
1 960
105
86
0.58
0.17
years. Uniformitarianism is the prevalent
belief of this age that all things continue as
2 2170
151
36
0.27
0.08
they were from the beginning denying the
possibility of any physical interventions by a
3 2900
197
28
0.17
0.05
designer
into the natural realm.
Uniformitarians interpret scientific data to
4 3502
239
0.76
0.012
0.004
support their idea of cosmic and
biological evolution during billions of years of
5 3930
277
~0.2
~0.001
imagined time. We maintain that their
6 4310
313
~0.2
~0.001
interpretations
are
a
distortion
of
observational data all around us. In this
paper we will include their assumption of billions of years of time and solely natural
processes in the uniformitarian model we construct for diffusion.Getting back to the helium
data, notice that the retention levels decrease as the temperatures increase. That is
consistent with ordinary diffusion: a high concentration of helium in the zircons diffusing
outward into a much lower concentration in the surrounding minerals, and diffusing faster in
hotter rock. As the next section shows, diffusion rates increase strongly with temperature.In
later sections, we will show that these large retentions are quite consistent with diffusion
taking place over thousands of years, not billions of years.
How Diffusion Works
Fig. 3. Helium atom moving through crystal.
If the reader is not very familiar with diffusion and wants to know more, we recommend a
very clear little book, Atomic Migration in Crystals, written for nonexperts.12 Fig. 3, adapted
from that book,13 illustrates how an atom diffuses through a solid crystal lattice of other
atoms. Fig. 3(a) shows a helium atom initially at position A, surrounded by a cell of lattice
atoms. The lattice atoms repel the helium atom, tending to confine it to the center of the
cell, where the repulsion balances out in all directions. Heat keeps the atoms of the lattice
vibrating at its various resonant frequencies.The vibrating atoms continually bump into the
helium atom, jostling it from all sides. The higher the temperature, the more vigorous the
jostling.Every now and then, the lattice atoms will bump the helium atom hard enough to push it into the activated
position B, midway between cells. The lattice atoms must give the helium enough kinetic energy to overcome the repulsive
potential energy barrier between the cells, which we have shown in Fig. 3(b). This required amount of kinetic energy, E, is
called the activation energy. If the lattice atoms have given any more energy than E to the helium atom, it will not stop at
position B. Instead, it will continue on to position C at the center of the adjacent cell. The helium atom has thus moved from
one cell to the next.If there is an initially high concentration of helium atoms in one part of the crystal,
these random motions will eventually spreadthat is, diffusethe helium more uniformly though the
crystal and out of it. Let us define C(x, y, z, t) as the concentration, the number of helium atoms per
unit volume, at position (x, y, z) at time t. Many textbooks show that when diffusion occurs, the time rate
of change of C is proportional to the sharpness of the edges of the distribution of helium, or more mathematically,
proportional to the Laplacian of C, 2C:
(1a)
where
(1b)
Equation (1a), called the diffusion equation, occurs frequently in many branches of physics, for example to describe heat
conduction in solids. Specialists in the diffusion of atoms through materials call it Ficks Second Law of Diffusion. The
factor D, the diffusion coefficient (or diffusivity), has dimensions of cm2 (or m2) per second. (Most of the diffusion literature
still uses centimeters and calories instead of meters and joules). Very often it turns out that at high temperatures, the
diffusion coefficient depends exponentially on the absolute temperature T (degrees kelvin
above absolute zero):
(2)
where R is the universal gas constant, 1.986 calories per mole-kelvin (8.314 joules per molekelvin). The constant D0 is independent of temperature. The intrinsic activation
energy E0 typically is between 10 and 100 kilocalories per mole (about 40 and 400 kilojoules per mole). Section 10
discusses how these quantities are related to the geoscience concept of closure temperature, and it shows why the
concept is irrelevant to our conclusions.
If the crystal has defects such as vacancies in the crystal lattice, impurities, dislocations, or grain boundaries, then the
diffusion coefficient equation will have a second term related to the defects:
(3)
The defect parameters (D1 and E1) are almost always smaller than the
intrinsic parameters (D0 and E0):
(4)

Fig. 4(a). Typical Arrhenius plot.


The typical Arrhenius plot in Fig. 4(a) shows how the diffusion
coefficient D of equation (3) depends on the inverse of the absolute
temperature, 1/T. Because the plot uses a logarithmic scale for D and a
linear scale for 1/T, each term of equation (3) manifests itself as a
straight line in the temperature region where it is dominant. (Plotting
with Tinstead of 1/T would make the lines curved instead of straight.)
The slopes are proportional to the activation energies E0 and E1. The
intercepts with the vertical axis, where 1/T is zero, are the
parameters D0and D1.The intrinsic line has a steep slope and a high
intercept, while the defect line has a shallow slope and a low intercept.
Starting on the right-hand side of the graph, at low temperatures, let us
increase the temperature, moving to the left. When the temperature is
high enough, we reach a region, the knee, where the two terms of
equation (3) are about equal. To the left of that region, at high
temperatures, the intrinsic properties of the crystal dominate the
diffusion. To the right of the knee, at lower temperatures, the defects
dominate. Because defects are very common in natural crystals, this two-slope character is typical.14

Fig. 4(b). Increasing defects slides the defect line upward.


Fig. 5. Interpretations of russian zircon data (blue, pink and orange
symbols and lines) compared with Nevada zircon data (green). The
ordinate is D (not D/a2).
For a given type of mineral, the location of the knee can vary greatly. It
depends on the value of D1, which depends on the amount of defects in
the particular crystal. The more defects there are, the higher D1 is. If we increase the number of defects, the defect line
moves upward (keeping its slope constant) on the graph, as Fig. 4(b) illustrates.
In the case of zircons containing radioisotopes, the main cause of defects is radiation damage, so highly radioactive
(metamict) zircons will have a large value of D1, causing the defect line to be higher on the graph than for a lowradioactivity zircon.
Early Zircon Data Were Ambiguous
In 1970 Sh. A. Magomedov, a researcher in Dagestan (then part of the Soviet Union) published diffusion data for radiogenic
lead and helium in highly metamict (radiation-damaged) zircons from the Ural Mountains.15 These were the only helium-inzircon diffusion data we could find during an extensive literature search we did in 1999.
Magomedov was mainly interested in lead diffusion, so he did not list his helium data explicitly in a table. Instead he showed
them in a small graph, along with data for lead diffusion and electrical conductivity, . His label for the ordinate was
ambiguous: ln(D, ). In scientific literature ln with no further note usually means the natural logarithm (base e). The
common logarithm (base 10) is usually shown as log. If we assume Magomedov was reporting ln eD, the resulting diffusion
coefficients would be very high, as the triangles and dotted line near the top of Fig. 5 show. The RATE book shows that
interpretation.16 Another interpretation is that Magomedov was reporting ln e(D/ a2), where a is the effective radius of his
zircons, about 75 m. As Fig. 5 shows (circles and thin solid line near middle), that still gives rather high diffusion rates in the
temperature range of interest to us.Based on those supposed high rates, we assumed in our first theoretical model17 that
the zircons were a negligible impediment to helium outflow, compared to the minerals around them. But in 2001 we received
a preprint of a paper18 listing new helium diffusion data in zircons from several sites in Nevada. Fig. 5 shows some of that
data (Fish Canyon Tuff sample FCT-1) as a line of solid dots. These data were many orders of magnitude lower than our
interpretation of Magomedovs graph. The Russian data would agree with the Nevada data if we re-interpret Magomedovs
label as meaning log10D, the common logarithm of D. Fig. 5 shows that interpretation near the bottom (squares and thick
solid line). The small difference between the high-slope intrinsic parts of the Russian and Nevada data is easily attributable
to site-to-site differences in composition. The nearly horizontal part of the Russian data is probably a defect line due to
much radiation damage (see end of previous section).The new data and our new interpretation of the old data imply that
zircon is not a negligible impediment to helium diffusion. In this paper we have changed our theoretical model to account for
that fact. As we will show in later sections, our new interpretation of the Russian data is still five orders of magnitude too high
for uniformitarian models. But it is quite compatible with creationist models and time scales.
Data for Minerals from the Jemez Granodiorite

Measurements of noble gas diffusion in a given type of naturally occurring mineral often show significant differences from
site to site, caused by variations in composition. For that reason it is important to get helium diffusion data on zircon and
biotite from the same rock unit (the Jemez Granodiorite) which was the source of Gentrys samples. Accordingly, in 2000 the
RATE project commissioned such experimental studies.Los Alamos National Laboratory was kind enough to give us core
samples of granodiorite from the same borehole, GT-2, and from
a similar depth, about 750 meters. The geology laboratory at the
Institute for Creation Research extracted the biotite using heavy
liquids and magnetic separation. Using similar methods,
Activation Laboratories, Ltd., in Ontario, Canada, extracted the
zircons and chose three of them for isotopic analysis. Appendix A
gives those results, which agreed fairly well with the lead-lead
dates published by Los Alamos National Laboratory for the same
site.19 We reserved the rest of the zircons, about 0.35
milligrams, for diffusion measurements.Through a small mining
company, Zodiac Minerals and Manufacturing, we contracted
with Kenneth A. Farley of the California Institute of Technology
(Division of Geological and Planetary Sciences) to measure the
diffusion coefficients of the zircon and biotite from the Jemez site.
He is a recognized expert on helium diffusion measurements in
minerals, having many publications related to that field. As we
wished, Zodiac did not tell Farley they were under contract to us,
the goals of the project, or the sites of the samples. We have
encouraged him to publish his measurements and offered to
send him the geologic site information if he does so. Appendices
B and C list his data in detail.Figs. 6(a) and 6(b) are Arrhenius
plots of the most relevant data for zircon and biotite, respectively.
The zircon data are from the Jemez Granodiorite in New Mexico,
the Fish Canyon Tuff in Nevada, and the Ural Mountains in
Russia (the re-interpreted Magomedov data). The first two
studies are for essentially the same size crystals (average length
~60 m, a 30 m, section 6). The Russian study was for
crystals ~150 m long. The biotite data are from the Jemez
Granodiorite. Those, and similar data we obtained (see Appendix
B) for biotite from the Beartooth Gneiss in Wyoming, are the only
data for that mineral we know of. For comparison to the biotite
data, we have also included published data for muscovite, a
different mica.20Notice that all the sets of zircon data agree fairly
well with each other at high temperatures. At 390C (abscissa =
1.5), the Russian data have a knee, breaking off to the right into
a more horizontal slope for lower temperatures. That implies a
high number of defects (see section 4), consistent with the high
radiation damage Magomedov reported. The Nevada and New
Mexico data go down to 300C (abscissa = 1.745) with no strong knee, implying that the data are on the intrinsic part of the
curve. A least-squares fit of equation (2) to the New Mexico (Jemez Granodiorite) zircon data gives the following diffusion
parameters:
Fig. 6(a). Observed diffusion coefficients in zircons. The ordinate is D (not D/a2).
Fig. 6(b). Observed diffusion coefficients in two types of mica. The ordinate is D (not D/a2).

(5a)
However, there appears to be a slight decrease of slope in the data below 450C. Later on we will need a fit at temperatures
below that. The best-fit parameters from 440C down to 300C are:

(5b)
Because the New Mexico zircons are radioactive, they must have some defects and should have a knee at some lower
temperature than 300C. We have recently requested that Farley get additional data from 100C to 300C. But as of
February 2003, we do not have reliable data for that range.The muscovite and biotite data are consistent with each other. In
the low temperature range of interest, the New Mexico biotite has a somewhat higher diffusion coefficient than the zircons.
That means the biotite, while not being negligible, did not impede the helium outflow as much as the zircon did.
A New Creation Model
We need a theoretical framework in which we can interpret the diffusion data of the previous section. As we mentioned at
the end of section 4, in our first creation model we wrongly assumed that the
zircons were a negligible impediment to the helium diffusion. In this section
we construct a new creation model.As before, the creation model starts with
a brief burst of accelerated nuclear decay generating a high
concentration C0of helium uniformly throughout the zircon (like the
distribution of U and Th atoms), but not in the surrounding biotite. After that
the helium diffuses out of the zircon into the biotite for a time t. As in our
previous model, we chose t = 6,000 years. The time is short enough that the
additional amount of helium generated by normal nuclear decay would be
small compared to the initial amount. We assume the temperatures to have
been constant at todays values. We will show in section 7 that this
assumption is generous to uniformitarians.
Figure 7. Spherical approximation of the zircon-in-biotite system.

Because the biotite diffusion coefficients are not too different from the zircon coefficients, we should have a model
accounting for two materials. Diffusion in zircon is isotropic, with helium flowing essentially at the same rate in all three
directions. Diffusion in biotite is not isotropic, because most of the helium flows two-dimensionally along the cleavage planes
of the mica. But accounting for anisotropy in the biotite would be quite difficult, so we leave that refinement to the next
generation of analysts. To keep the mathematics tractable, we will assume spherical symmetry, with a sphere of zircon of
effective radius a inside a spherical shell of material having an outer radius b, as Fig. 7 shows. Then the concentration C will
depend only on time and the distance r from the centerLet us consider the values we should assign to a and b.
Magomedovs zircons were between 100 and 200 m long,21 for an average length of about 150 m. He assigned the
crystals an effective radius of half the average length, or 75 m. Gentry selected zircons between about 50 m and 75 m,
for an average that we will round off to 60 m. Half of that gives us an effective radius for our analysis of the Jemez zircons:
(6)
Biotite in the Jemez Granodiorite is in the form of flakes averaging about 0.2 mm in thickness and about 2 mm in diameter.
Because the cleavage planes are in the long direction, and diffusion is mainly along the planes, the diameter is the relevant
dimension for diffusion. That gives us an outer radius of:
(7)
Because b is more than 32 times larger than a, the disk-like (not spherical) volume of biotite the helium enters is more than
1,000 (~32 squared) times the volume of the zircon. This consideration affects the boundary conditions we choose for r = b,
and how we might interpret sample 6 (see section 2), as follows.Suppose that helium could not escape the biotite at all.
Then as diffusion proceeds, C would decrease in the zircon and increase in the biotite, until the concentration was the same
throughout the two materials. After that C would remain essentially constant, at about 0.001 C0. The fraction Q/Q0 remaining
in the zircon would be about 0.001, which is just what Gentry observed in sample 6.So a possible explanation for sample 6
is that diffusion into the surrounding materials (feldspar, quartz), and leakage (along grain boundaries) was slow enough
(during the relatively short time t) to make the outflow of helium from the biotite negligible. For that sample, the temperature
and diffusion coefficient were high enough for helium to spread uniformly through both zircon and biotite during that time.Our
measurements (see Appendix B) showed that the helium concentration in the Jemez biotite at a depth of 750 m was small,
only about 0.32 109 cm3 (at STP) per microgram. Taking into account the difference in density of biotite and zircon (3.2
g/cm3 and 4.7 g/cm3), that corresponds to almost exactly the same amount of helium per unit volume as sample 6 contained.
That suggests the zircon and biotite were near equilibrium in sample 6, thus supporting our hypothesis.At lower
temperatures, for helium retentions greater than 0.001, C in the biotite would be lower than C in the zircon. In that case the
boundary at r = b would not significantly affect the outflow of helium from the zircon. We will assume this was approximately
true for sample 5 also, but not for sample 6. To simplify our analysis for samples 1 through 5, we will assume the usual
boundary condition, that the concentration C(r) falls to zero at radius r = b:
(8)
For the initial conditions, we assume that the concentration is a constant, C0, inside the zircon, and zero outside it:
At
(9a)
and
(9b)
After time zero, there also must be continuity of both C and helium flow at r = a. We need a solution to the diffusion equation
(1), in its radial form, for the above boundary conditions. In 1945, R. P. Bell published such a solution for the corresponding
problem in heat flow.22 His solution, which is mathematically complex, allows for different diffusion coefficients in the two
regions. We will simplify the solution considerably by making the diffusion coefficients the same in both regions. Because the
diffusion coefficient of biotite is somewhat higher than that of zircon at the temperatures of interest, our solution will have
slightly slower (no more than 30% slower) helium outflows and correspondingly longer times than the real situation. But
because uniformitarians need to increase the time anyhow, they should not object to this approximation.
With the above simplification, Bells equation reduces to one given by Carslaw and Jaeger.23 After making the simple
changes required to go from heat flow to atomic diffusion,24 and accounting for notation differences (note meanings
ofa and b), we get the following solution:

(10)
where D is the diffusion coefficient of zircon. Next we need to determine the fraction Q/Q0 of helium retained in the zircon
after diffusion takes place for time t. First, note that Q(t) and Q0 are the volume integrals of C(r, t) and C0 in the zircon:
(11a)
(11b)
Volume integrating equation (10) as required by equation (11a) and dividing by equation (11b) gives the fraction of helium
retained in the zircon after time t elapses:

(12)
where we define the function Sn as follows:

(13)
To solve equation (12), let us rewrite it in terms of a new variable, x, and a new function F(x) as follows:

(14a)
where

(14b)
and

(14c)
Now we can use software like Mathematica25 to find the roots of equation (14a), that is, to find the values of x for whichF(x)
will give us particular values of the retention fraction Q/Q0. When the latter and b/a are large, the series in equation (14b)
does not converge rapidly. For our value of b/a, 33.3, it was necessary to go out to N = 300 to get good accuracy. Table 2
lists the resulting values of x, and the values of D necessary to get those values from equation (14c) using a time of 6,000
years, t = 1.892 1011 seconds. The estimated errors in D result from the reported errors in Q/Q0.
In summary, the fifth column shows the zircon diffusion coefficients that would be necessary for the Jemez zircons to retain
the observed fractions of helium (third column) for 6,000 years at the temperatures listed in the second column.
Table 2. New creation model.
This new model turns out to be amazingly
T(C) Q/Q0
x
D(cm2/sec)
Error (%)
close to our previous creation modelwithin
0.5% for sample 1 and 0.05% for the others
1 105
0.580.17
5.9973 104 3.2103 1018 +122 67
despite the different assumptions and
2 151
0.270.08
2.4612 103 1.3175 1017 +49
30
equations. This strongly suggests there is an
underlying (but not obvious) physical
3
17
3 197
0.170.05
4.0982 10
2.1937 10
+39
24
equivalence between the two models, and that
the small differences are merely due to the
2
16
4 239
0.0120.004 3.3250 10
1.7798 10
+33
18
numerical error of the calculations. Thus our
previously published predictions26 of diffusion
5 277
~0.001
1.8190 101 9.7368 1016

coefficients are valid, but they should be reinterpreted to apply to zircon, not biotite.
We will compare the data not only to this new model, but also to a uniformitarian model, which we describe in the next
section.
Uniformitarian Model
In the RATE book,27 we outlined a simple model appropriate for the uniformitarian view, with its billions of years, of the
history of the rock unit:
. . . steady low-rate radioactive decay, He production, and He diffusion for 1.5 billion years at todays temperatures in the
formation.
Our assumption of constant temperatures is generous to uniformitarians. Two geoscientists from Los Alamos National
Laboratory constructed a theoretical model of the thermal history of the particular borehole (GT-2) we are concerned
with.28 They started by assuming a background vertical geothermal gradient of 25C/km. That means initial conditions with
absolute temperatures 16% to 31% lower than today for samples 1 through 6, putting them in the low-slope defect range of
diffusion. Their model then has an episode of Pliocene-Pleistocene volcanism starting to increase the temperature several
megayears ago. It would peak about 0.6 Myr ago at temperatures roughly 50 to 120C above todays values, depending on
depth. After the peak, temperatures would decline steadily until 0.1 Myr ago, and then level off at todays values.Later
studies29,30 add a more recent pulse of heat and have past temperatures being higher, 110C to 190C more than todays
levels just 24,000 years ago, and higher before that.31 This would put the samples well into the high-slope intrinsic range
of diffusion.The effect of such heat pulses would be great. For several million years, the diffusion coefficients would have
been about two to three orders of magnitude higher than todays values. During the previous 1.5 billion years, supposedly at
lower temperatures than today, the diffusion rates would have been on the defect line (fig. 4a) and therefore not much
below todays levels. Thus the long time at lower temperatures would not compensate for high losses during the few million
years at higher temperatures. This makes our assumption of constant temperatures at todays values quite favorable to the
uniformitarian scenario.As we will see, the long uniformitarian time scale requires zircon diffusion coefficients to be about a
million times slower than the measured biotite coefficients. That means the biotite would not be a significant hindrance to the
helium flow in the uniformitarian model, and the results would not be much different than those for a bare zircon. For
continuous production of helium, the concentration C in the zircon would reach its steady-state level relatively quickly (see
section 10) and remain at that level for most of the alleged 1.5 billion years. Again we assume a spherical zircon of radius a.
Carslaw and Jaeger32 give the corresponding solution for heat flow. Converting to the notation for atomic diffusion shows us
how the steady-state concentration C in the zircon depends on the radius r from the center:

(15)
Table 3. Uniformitarian model.
Q/Q0

D(cm2/sec)

Error (%)

1 105

0.580.17

2.1871 1023

30

2 151

0.270.08

4.6981 1023

30

T(C)

3 197

0.170.05

7.4618 1023

30

4 239

0.0120.004

1.0571 1021

30

5 277

~0.001

1.2685 10

20

Here Q0 is the total amount of helium that would be produced in time t. That is, Q0/t is the helium production rate. As
before, D is the diffusion coefficient of zircon, and a is the effective radius. Using equation (11a) to integrate equation (15)
and dividing by Q0 gives us the fraction of heliumQ/Q0 in the zircon in the steady-state condition:

(16)
Table 3 gives us the zircon diffusion coefficients required to give the observed retentions for a = 30 m and t = 1.5 billion
years = 4.73 1016seconds.
The same reasoning on sample 6 applies for this model as for the creation model, except that it is less likely the helium
could remain totally sealed in the biotite for over a billion years. For the other samples, this model is exactly the same as our
previously published evolution model.33
Comparing Data and Models
Fig. 8. The zircon data line up very well with the creation model.
The ordinate is D(not D/a2).
Fig. 8 shows the zircon data from the Jemez Granodiorite, along
with the two models. The zircon data are fully consistent with the
creation model. These new data are also quite consistent with all
published zircon data, as Fig. 6(a) shows. As of this writing
(February, 2003) we do not have reliable data on the Jemez
zircons below 300C. But notice that the data have the same
slope as the Creation model points for samples 3, 4, and 5, and
the data nearly touch point 5. That allows us to use equation
(14c) to roughly estimate values for the time t for those three
points:

(17)
Using a/b = 0.03, the values of D/a2 extrapolated down from the
best-fit experimental parameters of Equation 5b, and the values
of x and errors from Table 2 gives us the following times for diffusion to have occurred:
The errors above do not include the statistical errors in extrapolating the fit to the zircon diffusion coefficient data down to
the lower temperatures required. Actual data for temperatures below 300C would eliminate the extrapolation error.
Table 4. Time for diffusion.
In the meantime we can say the data of Table 4,
x
D/b2(sec1)
Time t(years) Error (years)
considering the estimates of error, indicate an
age between 4,000 and 14,000 years. This is far
3 4.0982 103 1.2672 1015 10389
+4050 2490
short of the 1.5 billion year uniformitarian age!It
4 3.3250 102 1.6738 1014 6392
+2110 1150
looks as if the retention data require points 1 and
2 of the creation model to be on a defect line,
5 1.8190 101 1.2311 1013
4747

similar to the Russian data for radiation-damaged


zircons. The similarity gives us good reason to
hope that the low-temperature zircon data, when they come in, will come close to those model points as well.
Table 5. Billion-year lower bounds versus observed retentions.
Sample T(C)

ObservedD/a2 (sec1) Extrapolated from data

Helium retentionsQ/Q0
After 1.5 billion years Observed

12

1.0007 106

0.170

197

1.4080 10

239

1.8597 1011

7.5764 108

0.012

277

10

0.001

1.3679 10

1.0368 10

The data offer no hope for the uniformitarian model. It is unlikely that the zircon data will continue down on the intrinsic line
for five more orders of magnitude. It is certain (because all natural zircons have defects) that at some lower temperature
there will be a knee, where the data will break off horizontally to the right into a shallow-slope defect line. But even if that
were not to be the case, the intrinsic line would pass well above the uniformitarian model.We can also use these observed
data to estimate what helium retentions Gentry should have found if the zircons were really 1.5 billion years old. If no helium
could leak out of the biotite during that time, then all of the samples would have had retentions of about 0.001, much less
than samples 1 through 4 [see section 6 between equations (7) and (8)]. However, we know that helium can diffuse through
the surrounding materials, quartz and feldspar. By assuming those are negligible hindrances, we can use the extrapolated
data in equation (16) to get lower bounds on the retentions. Table 5 shows the results:
In summary, the observed diffusion rates are so high that if the zircons had existed for 1.5 billion years at the observed
temperatures, samples 1 through 5 would have retained much less helium than we observe. That strongly implies they
have not existed nearly so long a time.
Closing Some Loopholes
One response to these data from uniformitarians might be this: assert that temperatures in the Jemez Granodiorite before
the Pliocene-Pleistocene volcanism were low enough to make the diffusion coefficients small enough to retain the helium.
We discussed that possibility in section 7, but here we point out how low such temperatures are likely to be.Until we have
reliable low-temperature data for the Jemez Granodiorite zircons, we must reason indirectly from the other data we have.
The only published low-temperature zircon data, the Russian data by Magomedov,34 show a defect line (fig. 6a). The line is

rather high, probably because those zircons had many defects due to the high radiation damage Magomedov reported. But
the slope of the defect line is similar to the slope of points 1, 2, and 3 in both the creation and uniformitarian models of the
retention data (fig. 8). Since the high-temperature Jemez zircon data agree well with the creation model, there is good
reason to suppose the low-temperature data will also conform to that model. In that case, the parameters of the zircon
defect line would be:
(18)
Because E1 is small, the slope of the defect line is small. These numbers would mean that to get the diffusion coefficients
low enough to meet uniformitarian needs, say on the order of 1023 cm2/sec, the pre-Pliocene temperature in the granodiorite
would have to have been about 190C, near that of liquid nitrogen. No uniformitarian we know would advocate an earth
that was cryogenic for billions of years! Of course these values are only preliminary estimates, and perhaps the actual
defect line of the Jemez zircons would require less severe cooling. But it demonstrates how zircons would need
unrealistically low temperatures to retain large amounts of helium for uniformitarian eons of time.A second uniformitarian line
of defense might be to claim that the helium 4 concentration in the biotite or surrounding rock is presently about the same as
it is in the zircons. (Such a scenario would be very unusual, because the major source of 4He is U or Th series radioactivity
in zircons or a few other minerals like titanite or apatite, but not biotite.) The scenario would mean that essentially no
diffusion into or out of the zircons is taking place. However, ourmeasurements (Appendix B) show that except for possibly
samples 5 and 6, the concentration of helium in the biotite [section 6, between equations (7) and (8)] is much lower than in
the zircons. Diffusion always flows from greater to lesser concentrations. Thus helium must be diffusing out of the zircons
and into the surrounding biotite. Moreover, the Los Alamos geothermal project made no reports of large amounts of helium
(commercially valuable) emerging from the boreholes, thus indicating that there is not much free helium in the formation as
a whole.A third uniformitarian defense could be that the Oak Ridge team somehow made a huge mistake, that the actual
amounts of helium were really many orders of magnitude smaller than they reported. But as Appendix C reports, our
experimenter Kenneth Farley, not knowing how much he should find and going up to only 500C, got a partial (not
exhaustive) yield of 540 nanomoles of helium per gram of zircon, or in Gentrys units, 11 10 9 cm3/ g. That is on the same
order of magnitude as Gentrys results in Table 2, which reports the total (exhaustive) amount liberated after heating to
1000C until no more helium would emerge. Thus our experiments support Gentrys data.
Closure Temperature Doesnt Help Uniformitarians
Some uniformitarians try to use the geoscience concept of closure temperature to claim that zircons below that
temperature are permanently closed systems, losing no significant helium by diffusion. They fail to understand that even well
below that temperature, zircons can re-open and lose large amounts of helium. Here we explain closure temperature and
re-opening, and show that in the uniformitarian scenario, the Jemez Granodiorite zircons would reopen early in their history.Consider a hot zircon cooling down in newly formed granite. If the cooling
rate is constant, then the seminal article by Martin Dodson35 on closure temperature shows that the
diffusion coefficient D (of helium moving out of the zircon) decreases exponentially with a time
constant given by:
where T is the absolute temperature, dT/dt is the cooling rate, R is the gas constant, and E0 is the
activation energy in the intrinsic region (section 3).
In the uniformitarian scenario, nuclear decay produces helium at a nearly constant rate. At the beginning, when the zircon is
very hot, helium diffuses out of the crystal as fast as nuclear decay produces it. But as the zircon cools, it will eventually
reach a temperature below which the loss rate becomes less than the production rate. That point is essentially what
Dodson meant by the closure temperature. He showed that for a constant cooling rate the closure temperature Tc is
where A is a dimensionless constant (55 for a sphere), D0 is the intrinsic intercept in Fig. 4(a),
a is the effective radius of the crystal, and is the diffusion time constant given by equation (19).
Since depends on the cooling rate, hence affecting Tc somewhat, geoscientists imply some
conventional cooling rate when they specify a closure temperature. In Appendix C Kenneth
Farley assumes a cooling rate of 10C per million years and finds that the closure temperature
of the Jemez Granodiorite zircons is 128C.That temperature is below the borehole
temperatures of samples 2 through 5 (Table 1). Most of our samples wereabove the closure temperature, so they would
always have been open systems, losing helium. However, even if they had reached closure temperature, the analysis below
shows they would not have remained closed for long.After the zircon cools below the closure temperature, helium begins to
accumulate in it, as Fig. 9 shows. Later, as the temperature levels off to that of the surrounding rock, the diffusion
coefficient D becomes constant. (The case of changing long-term temperatures is harder to analyze, but there will still be a
time of re-opening.) As the amount of helium in the zircon increases, Ficks laws of diffusion (section 3) say the loss rate
also increases. Eventually, even well below the closure temperature, the loss rate approaches the production rate, an
event we call the re-opening of the zircon. Then the amount of helium in the zircon will level off at a steady-state value,
which we called Q in equation (16). After that, the zircon will again lose helium as fast as nuclear decay produces it.Let us
estimate the closure interval, the length of time tci the zircon remains closed before re-opening. As we remarked just below
Equation 15, the helium production rate is Q0/t, where t is the uniformitarian age of the zircon, 1.5 billion years. Assuming a
linear rise as a first approximation, the production rate multiplied by tci is roughly equal to the steady-state value of Q, which
is the right-hand side of our equation (16) multiplied by Q0:

(21)
Solving for tci gives us the approximate closure interval:

(22)
If the closure interval were long compared to the age of the zircon, then the zircon would indeed be a closed system. But
would that be the case in the uniformitarian view of the Jemez zircons? Using the effective radius of the zircons, 30 m, and
the measured values of D (fig. 8) in equation (22) gives us tci values between a few dozen years and a few thousand
years, depending on the temperature of the sample in the borehole. Those times are very small compared to the
uniformitarian age of 1.5 billion years.So even if the zircons had cooled rapidly and reached closure temperature early in

their history, our measured diffusion rates say they would have re-opened shortly after that. During most of the alleged eons
the zircons would have been an open system. They would be losing as much helium as the nuclear decay produced. Thus
closure temperature does not help uniformitarians in this case, because the closure interval is brief.
Conclusion
The experiments the RATE project commissioned in 2000 have clearly confirmed the numerical predictions of our creation
model (updated slightly in section 6), which we published beforehand.36 Other experimental data published since 2000
agree with our data. The data also clearly reject the uniformitarian model. The data and our analysis show that over a billion
years worth of nuclear decay have occurred very recently, between 4,000 and 14,000 years ago. This strongly supports our
hypothesis of recent episodes of highly accelerated nuclear decay.These diffusion data are not precise enough to reveal
details about the acceleration episodes. Were there one, two, or three? Were they during early Creation Week, after the Fall,
or during the Flood? Were there only 500 to 600 million years worth of acceleration during the year of the Flood, with the
rest of the acceleration occurring before that? We cannot say from this analysis. However, the fact that these zircons are
from a Precambrian rock unit sheds some light on various creationist models about when strata below the Cambrian formed.
We can say that the diffusion clock requires a large amount of nuclear decay to have taken place within thousands of
years ago, after the zircons became solid. At whatever time in creation history Precambrian rocks came into existence,
these data suggest that 1.5 billion years worth of nuclear decay took place after the rocks solidified not long ago.Our most
important result is this: Helium diffusion casts doubt on uniformitarian long-age interpretations of nuclear data and
strongly supports the young world of Scripture.
Acknowledgments
Many people and institutions have contributed to collecting and interpreting these data. In particular, we would like to thank
Robert Gentry, Bill Hoesch, Yakov Kapusta, Roger Lenard, Majdah al-Quhtani, and Phil Legate. We would also like to thank
supporters of the RATE project for their generous contributions and prayers.
Appendix A: Isotopic Analysis of Jemez Zircons
Here we summarize a report by Dr. Yakov Kapusta (Activation Laboratories, Ltd., in Ontario, Canada) on an isotopic
analysis he made on three zircons from Los Alamos National Laboratories core sample GT-2480 from borehole GT-2 in the
Jemez Granodiorite at a depth of 750 m.
Dr. Kapusta separated zircons from the core sample using heavy liquids and magnetic separation. He picked three crystals
from the concentrate for analysis. Table A1 shows his results and notes.Mass fractionation correction of
0.15%/amu0.04%/amu (atomic mass unit) was applied to single-collector Daly analyses and 0.12%/amu 0.04% for
dynamic Faraday-Daly analyses. Total procedural blank less than 0.6 pg for Pb and less than 0.1 pg for U. Blank isotopic
composition: 206Pb/204Pb = 19.10 0.1, 207Pb/204Pb = 15.71 0.1,208Pb/204Pb = 38.65 0.1. Age calculations are based on the
decay constants of Steiger and Jger.37 Common-Pb corrections were calculated by using the model of Stacey and
Kramers38 and the interpreted age of the sample. The upper intercept of the concordia plot of the 206Pb/238U and 207Pb/238U
data was 1439.3 Ma 1.8 Ma. (The published Los Alamos radioisotope date for zircons from a different depth, 2900 m, was
1500 20 Ma39)
Appendix B: Diffusion Rates in Biotite
Below are two reports by Kenneth Farley (with our comments in brackets) on his measurements of helium diffusion in biotite
from two locations. As far as we know, these are the only helium-in-biotite diffusion data that have been reported. The first
sample, BT-1B, was from the Beartooth Gneiss near Yellowstone National Park. The second sample, GT-2, was from the
Jemez Granodiorite, borehole GT-2, from a depth of 750 m. The geology laboratory at the Institute for Creation Research
extracted the biotite for from both rock samples by crushing, magnetic separation, and density separation with heavy liquids.
Farley sieved both samples to get flakes between 75 and 100 microns in diameter. Taking half of the average diameter to
get an effective radius of 44 microns, we plotted the resulting diffusion coefficients for the GT-2 sample in Fig. 6(b). We
plotted the muscovite data in Fig. 6(b) using the effective radius recommended in the report,40 130 microns.
Table A1. Uranium-lead analysis of three zircons
Concentrations
#

Mass (g) (a)

U
(ppm)

Ratios

Pb (ppm)

Pb(c) (pg) (b)

206

Pb / 204Pb (c)

208

Pb / 206Pb (d)

206

Pb / 238U (e)

Error (2%)

z1 0.8

612

106.1

13.6

241.2

0.633

0.102828

.50

z2 1.0

218

59.6

1.4

2365.1

0.253

0.236433

.23

z3 1.7

324

62.7

1.7

3503.6

0.218

0.172059

.11

Ratios
#

207

Pb
(e)

Ages
/235U

Error (2%)

207

Pb /206Pb (e) Error (2%)

206

Pb /238U

207

Pb /235U

z1 1.2744

.56

0.08989

.23

631.0

834.4

1423.2

0.912

z2 2.9535

.26

0.09060

.12

1368.1

1395.7

1438.2

0.887

z3 2.1456

.13

0.09044

.07

1023.4

1163.6

1434.9

0.828

Notes:
Sample weights are estimated by using a video monitor and are known to within 40%.
Total common-Pb in analyses
Measured ratio corrected for spike and fractionation only.
Radiogenic Pb.

207

Pb /206Pb

Correlation coefficent

Corrected for fractionation, spike, blank, and initial common Pb.


Results of He Diffusion on Zodiac biotite, BT-1B
(Beartooth Gneiss) October 18, 2000
Kenneth A. Farley
Experiment:
Approximately 10 mg of biotite BT-1B, sieved to be between 75 and 100 m, was subjected to step heating. Steps ranged in
temperature from 50C to 500C in 50C increments, with an estimated uncertainty on T of < 3C. Durations ranged from 6
to 60 minutes, with longer durations at lower temperatures; uncertainty on time is < 1% for all steps. After the ten steps the
partially degassed biotite was fused to establish the total amount of He in the sample. He was measured by isotope dilution
quadrupole mass spectrometry, with an estimated precision of 2%. He diffusion coefficients were computed using the
equations of Fechtig and Kalbitzer41 assuming spherical geometry.
DataTable B1:
(In a later addendum to this report, Farley told us that the total amount of helium liberated was about 0.13 10 9 cm3(at
STP) per microgram of biotite.]
Interpretation:
He diffusion from this biotite defines a remarkably linear Arrhenius profile, fully consistent with thermally activated volume
diffusion from this mineral. The first two data points lie slightly below the array; this is a common feature of He release
during step heating of minerals and has been attributed to edge effects on the He concentration profile.42,43Ignoring those
two data points, the activation energy and diffusivity at infinite T based on these data are 25.7 kcal/mol and 752 respectively.
At a cooling rate of 10C/Myr, these parameters correspond to a closure temperature of 39C.
(After this Farley added a Recommendations section wherein he discussed the possibility of vacuum breakdown of the
biotite at high temperatures, the relevant effective radius for biotite [probably half the sieved flake diameter], and the source
of helium in the biotite [probably uranium and thorium in zircons that had been in the flakes before separation]. We decided
none of these questions were important enough to investigate in detail for now, since this sample was not from a site we
were interested in at the time. It merely happened to be on hand at the ICR geology laboratory, making it ideal for an initial
run to look for possible difficulties in experimental technique.)
Results of Helium Diffusion experiment on Zodiac biotite, GT2
(Jemez Granodiorite) March 24, 2001
Kenneth A. Farley
Experiment:
Approximately 10 mg of biotite GT2, sieved to be between 75 and 100 m, was subjected to step heating. Steps ranged in
temperature from 50C to 500C in 50C increments, with an estimated uncertainty on T of < 3C. Durations ranged from 7
to 132 minutes, with longer durations at lower temperatures; uncertainty on time is < 1% for all steps. After 11 steps of
increasing T, the sample was brought back to lower temperature, and then heated in 6 more T-increasing steps. After the 17
steps the partially degassed biotite was fused to establish the total amount of He in the sample. He was measured by
isotope dilution quadrupole mass spectrometry, with an estimated precision of 2% (steps 12 and 13 are much more
uncertain owing to low gas yield). He diffusion coefficients were computed using the equations of Fechtig and
Kalbitzer44 assuming spherical geometry.
Table B2. Diffusion of helium from biotite sample GT-2.
DataTable B2:
Step TempC Minutes Cumulative Fraction lne(D/a2)
(In a later addendum to this report, Farley told us that
the total amount of helium liberated was about 0.32
1
50
61
1.61E05
32.72
109 cm3 [at STP] per microgram of biotite.)
2
50
60
2.79E05
32.01
Interpretation:
He diffusion in this sample follows a rather strange
3
100
60
2.39E04
27.32
pattern, with a noticeable curve at intermediate
temperatures. I have no obvious explanation for this
4
150
61
1.91E03
23.18
phenomenon. Because biotite BT-1B did not show this
curve, I doubt it is vacuum breakdown. I ran more steps,
5
200
61
4.70E03
21.54
with a drop in temperature after the 500C step, to see if
the phenomenon is reversible. It appears to be, that is,
6
250
31
6.81E03
20.59
the curve appears again after the highest T step, but the
7
300
31
9.69E03
19.92
two steps (12, 13) that define this curve had very low
gas yield and high uncertainties. It is possible that we
8
350
16
1.35E02
18.63
are dealing with more than one He source (multiple
grain sizes or multiple minerals?). (We think it is likely
9
400
15
2.44E02
17.03
there were some very small helium-bearing zircons still
embedded in the biotite flakes, which would be one
10
450
9
4.90E02
15.05
source. The other source would be the helium diffused
11
500
7
1.07E01
13.13
out of larger zircons no longer attached to the flakes.)
This sample had about twice as much helium as BT-1B.
12
225
132
1.07E01
22.12
Note that despite the strange curvature in GT2, the two
biotite samples have generally similar He diffusivity
13
275
61
1.07E01
21.07
overall.(The similarity Farley remarks upon made us
decide that the biotite data were approximately correct.
14
325
61
1.07E01
19.70
Because these data below 300C were also about an
15
375
60
1.10E01
18.07
order of magnitude higher than our creation model, we
supposed that zircon might be a more significant
16
425
55
1.24E01
16.15
hindrance to helium loss than biotite, so we turned our
attention to zircon. It turned out that our supposition was
17
475
61
1.99E01
14.22
correct, which makes it less important to have exact
biotite data.)
Fusion
8.00E01
Appendix C: Diffusion Rates in ZirconBelow is a report
by Kenneth Farley (again with our comments in
Total
1.00000
brackets) on his measurements of helium diffusion in
zircons extracted by Yakov Kapusta from Los Alamos National Laboratories core sample GT-2480 from borehole GT-2 in the
Jemez Granodiorite at a depth of 750 m. Appendix A gives Kapustas radioisotopic analysis of three of the zircons. The rest,

unsorted by size and labeled as sample YK-511, were forwarded to Farley for diffusion analysis. In Fig. 8, we have assumed
an effective radius of 30 microns (or length 60 microns) and plotted the points (numbers 1544) which Farley concludes
below are the most reliable. These points only go down to 300C. In later publications we hope to report similar
measurements down to 100C.
Report on Sample YK-511
(Jemez Granodiorite) May 14, 2002, Kenneth A. Farley
We step heated 0.35 mg of zircons from the large vial supplied by Zodiac. We verified that the separate was of high purity
and was indeed zircon. The step heat consisted of 45 steps so as to better define the He release behavior. The first 15 steps
were monotonically increasing in temperature, after that the temperature was cycled up and down several times.
Results:
(See Table C1 on next page.) The first 14 steps lie on a linear array corresponding to an activation energy of ~ 46 kcal/mol
and a closure temperature of ~183C assuming a cooling rate of 10 C/Myr. However steps 15 to 44 (shown in Figs. 6(a)
and 8), which were cycled from low to high temperature and back, lie on a shallower slope, corresponding to Ea= 34.5
kcal/mol and Tc = 128C. This change in slope from the initial run-up to the main body of the experiment is occasionally
observed and attributed to either:
A rounded He concentration profile in the zircons, such that the initial He release is anomalously retarded. In other words,
the He concentration profile is shallower than the computational model used to estimate diffusivities assumes. This effect
goes away as the experiment proceeds and the effects of the initial concentration profile become less significant. This
rounding could be due to slow cooling or possibly to recent reheating.The change in slope might be due to changes in the
zircons during the heating experiment. For example, it is possible that annealing of radiation damage has occurred. This
sample has a very high He yield (540 nmol/g) so radiation damage is likely. However the zircons were only marginally within
the window where radiation damage is thought to anneal in zircons, so this hypothesis is deemed less likely.
Consideration of geologic history and/or further experiments are necessary to firmly distinguish between these
possibilities.
Table C1. Diffusion data for sample YK-511.
Step TempC

Helium 4 (nmol/g)

Time (sec)

Fraction

Cumulative Fraction

D/a2 (sec1)

300

5.337083

3660

0.001259

0.001259

3.78E11

300

1.316732

3660

0.000311

0.001570

2.10E11

300

0.935963

3660

0.000221

0.001791

1.77E11

325

3.719775

3660

0.000878

0.002669

9.34E11

350

7.910044

3660

0.001867

0.004536

3.21E10

375

18.12294

3660

0.004278

0.008815

1.36E09

400

36

3660

0.008498

0.017313

5.29E09

425

73.10049

3660

0.017256

0.034569

2.13E08

450

106.0761

3660

0.025040

0.059609

5.85E08

10

460

78.89137

1860

0.018623

0.078232

1.27E07

11

470

96.99925

1860

0.022897

0.101130

2.08E07

12

480

117.2479

1800

0.027677

0.128807

3.40E07

13

490

146.8782

1860

0.034671

0.163479

5.38E07

14

500

171.5538

1800

0.040496

0.203976

8.46E07

15

453

149.5962

7200

0.035313

0.239290

2.31E07

16

445

66.45767

7260

0.015687

0.254978

1.16E07

17

400

9.589814

6840

0.002263

0.257241

1.86E08

18

420

10.64711

3600

0.002513

0.259755

3.98E08

19

440

23.19366

3660

0.005475

0.265230

8.69E08

20

460

52.3035

3660

0.012346

0.277577

2.05E07

21

480

102.7062

3660

0.024244

0.301821

4.38E07

22

325

0.357828

3660

8.45E05

0.301906

1.61E09

23

350

0.718240

3660

0.000170

0.302075

3.23E09

24

375

1.690889

3660

0.000399

0.302475

7.62E09

25

400

4.246082

3660

0.001002

0.303477

1.92E08

26

425

3660

0.001888

0.305365

3.64E08

27

450

21

3660

0.004957

0.310323

9.70E08

28

460

22.0839

1860

0.005213

0.315536

2.05E07

29

470

33

1800

0.007789

0.323326

3.26E07

30

480

45

1860

0.010622

0.333948

4.47E07

31

490

62.39899

1800

0.014729

0.348678

6.75E07

32

500

82.65262

1800

0.019510

0.368189

9.59E07

33

475

120.222

7260

0.028379

0.396569

3.80E07

34

445

45

7260

0.010622

0.407191

1.53E07

35

400

5.879406

7260

0.001387

0.408579

2.05E08

36

300

0.075983

3660

1.79E05

0.408597

5.26E10

37

320

0.685076

21660

0.000162

0.408759

8.02E10

38

340

1.122111

18060

0.000265

0.409024

1.58E09

39

360

1.986425

14460

0.000469

0.409493

3.49E09

40

380

3.413768

10860

0.000806

0.410299

8.01E09

41

400

5.752365

7260

0.001357

0.411657

2.03E08

42

420

6.126626

3660

0.001446

0.413103

4.30E08

43

440

13.67016

3600

0.003226

0.416330

9.85E08

44

460

30.37821

3660

0.007171

0.423501

2.19E07

Conclusion
The most reasonable conclusion from the data is that the main body of the experiment, steps 1544, yields the best
estimate of the closure temperature, about 130C. This is somewhat cooler than we have observed before in zircons though
the database is not large. Radiation damage may be important in the He release kinetics from this He-rich sample.

METEORITES
Radioisotope Dating of Meteorites: I. The Allende CV3 Carbonaceous Chrondrite
by Dr. Andrew A. Snelling on April 16, 2014
Abstract
Meteorites have been used to date the earth with a 4.55 0.07 Ga Pb-Pb isochron called the geochron. They appear to
consistently yield 4.564.57 Ga radioisotope ages, adding to the uniformitarians confidence in the radioisotope dating
methods. The Allende CV3 carbonaceous chondrite meteorite, which fell to earth in Mexico February 8, 1969, is one of the
most-studied meteorites. Many radioisotope dating studies in the last 40 years have used the K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, UTh-Pb, Lu-Hf, Mn-Cr, W-Hf, Al-Mg, and I-Xe methods to yield an abundance of isochron and model ages for this meteorite
from chondrules, Ca-Al inclusions, whole-rock samples, and matrix, mineral, and other fractions, all of which components
are described. The age data were all tabulated and plotted on frequency versus age histogram diagrams, and strongly
cluster at 4.564.57 Ga, dominated by Pb-Pb isochron and model ages. The earliest (1976) and the latest (2010)
determined Pb-Pb isochron ages at 4.553 0.004 Ga and 4.56718 0.0002 Ga respectively are essentially the same,
testimony to that techniques supremacy as the uniformitarians ultimate, most reliable dating tool. Apart from scatter of the
U-Pb, Th-Pb, Rb-Sr, and Ar-Ar ages, no systematic pattern was found in the Allende isochron and model ages similar to the
systematic pattern of isochron ages found in Precambrian rock units during the RATE project. If such evidence in earth
rocks is applied to meteorites, then Allende seems to have no apparent evidence in it of past accelerated radioisotope
decay. However, if accelerated radioisotope decay did occur on the earth, then it could be argued every atom in the universe
would be similarly affected at the same time. Furthermore, meteorites are regarded as primordial material left over from the
formation of the solar system.Todays measured radioisotope composition of Allende may reflect a geochemical signature of
that primordial material, which included atoms of all elemental isotopes. So if some of the daughter isotopes were already in
the Allende meteorite when it was formed, and if the parent isotopes in the meteorite were also subject to subsequent
accelerated radioisotope decay at the same time it occurred in earth rocks, then the 4.56718 0.0002 Ga Pb-Pb isochron
age for the Allende cannot be its true real-time age, which according to the creation paradigm is only about 6000 real-time

years. The results of further studies of more radioisiotope ages data for many more meteorites will confirm or adjust these
tentative interim suggestions based on these Allende radioisotope ages data.Keywords: radioisotope dating, meteorites,
Allende, carbonaceous chondrite, matrix, chondrules, Ca-Al inclusions, organic carbon, K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, U-Th-Pb,
Lu-Hf, Mn-Cr, W-Hf, Al-Mg, I-Xe, isochron ages, model ages, discordant radioisotope ages, accelerated radioisotope decay,
primordial material, geochemical signature.
Introduction
In 1956 Claire Patterson at the California Institute of Technology in Pasadena published his now famous paper entitled Age
of meteorites and the earth (Patterson 1956). In it he reported lead (Pb) isotope analyses of three stony and two iron
meteorites, and plotted the results on a Pb-Pb isochron diagram (fig. 1). The age obtained from the excellent isochron
passing through these data points was 4.55 0.07 Ga. Patterson also found that the lead isotope analysis of a modern
ocean sediment sample plotted on the same isochron. He thus argued that, because the lead in the modern ocean sediment
sample was presumably an accumulation of leads mixed from many earth sources over the time since the earth formed, the
modern ocean sediments lead isotope composition represented an averaged and integrated lead isotopic composition for
the earth itself. And therefore, because the lead isotope composition of the modern ocean sediment plotted on this meteorite
Pb-Pb isochron, the earth must be the same age as the meteorites, namely, 4.55 0.07 billion years. Of course, implicit in
this conclusion were the assumptions that the meteorites all formed at the same time and they represented primordial earth
materials, because it was also assumed the meteorites are remnant materials from when the solar system formed.
Fig. 1. The Pb-Pb isochron obtained by
Patterson (1956) using three chondrite
meteorites (red squares) and two iron
meteorites (green square). The Pb isotopic
composition of modern ocean sediment (blue
circle) falls on the isochron, suggesting that
meteorites and the earth are related and of the
same age, which is why this isochron has been
called the geochron. The dashed lines are
growth by radioisotope decay curves (after
Dalrymple 2004).
Pattersons Pb-Pb isochron thus became
known as the geochron, that is, the isochron
that defined the age of the earth. And apart
from some minor adjustments due to
refinements in determinations of the U-Pb
decay constants, this claimed age for the earth
and this basis for it have stood unchallenged
for more than five decades. Indeed, the continued radioisotope dating of meteorites, many meteorites analyzed many times
by many different radioisotope methods, has only strengthened the claim that all the meteorites are about 4.564.57 billion
years old (Dalrymple 1991, 2004), which is thus argued to be the upper maximum for the earths age, assuming all the
meteorites represent primordial earth materials, leftover material from when the solar system formed. This has provided
uniformitarians with apparently impregnable evidence for a 4.55 billion-year-old earth, and great confidence in the supposed
reliability of their increasingly sophisticated radioisotope dating methodsPerhaps the most well-known meteorite is Allende, a
carbonaceous chondrite (CV3) meteorite (Krot et al. 2005) which has often been described as the best and most-studied
meteorite in history (Morris 2007; Norton 2002). Therefore, it is important to fully investigate just what ages, both isochron
and model ages, the different radioisotope dating methods have yielded for this meteorite and its various components, and
to explore what they might mean and their potential significance within the creation framework for the history of the earth
and the solar system.
The Allende CV3 Carbonaceous Chondrite Meteorite
On Saturday, February 8, 1969, between 1:05 and 1:10 a.m. (U.S. Central Standard Time) a brilliant fireball was observed
over much of northern Mexico and adjacent areas of Texas and New Mexico (Clarke et al. 1970). The most spectacular
phenomena were centered around the city of Hidalgo del Parral in the south-central part of the state of Chihuahua, Mexico
(fig. 2). The fireball approached from the south-southwest, moved across the sky, and as it neared its terminal point the
brilliant
light
was
accompanied by tremendous
detonations and a strong air
blast. In those five minutes
the fireball broke up in the
atmosphere, showering more
than two tonnes (2.2 tons) of
a
rare
carbonaceous
chondrite
meteorite,
consisting of thousands of
pieces with fusion crusts (a
coating formed by their
passage
through
the
atmosphere), over an area of
about 300 km2 (115.8 mi2)
(Clarke et al. 1970; Norton
2002).
Fig. 2. Location in northern
Mexico of the Allende
meteorite strewnfield, with
the direction of the fireball
approach indicated (after
Clarke et al. 1970). Click
image for larger view.

The original stone is believed to have been approximately the size of an automobile (at least 10 m 3 [13 yd3]) traveling
towards the earth at more than 15 km/sec (9.3 mi/sec). The roughly elliptical strewnfield extending in a southwest-northeast
direction is estimated to have measured approximately 8 km (4.9 mi) by 50 km (31.6 mi) (fig. 3), making it the largest stony
meteorite strewnfield that has been studied in any detail (Clarke et al. 1970). It is estimated that approximately 23 tonnes
(2.23.3 tons) of specimens were collected from that strewnfield over the next 25 years, making it the greatest total weight
ever recovered in a stony meteorite strewnfield up to that time (Norton 2002). The first specimen was recovered in the
northern section of the field in the small town of Pueblito de Allende located in the obscure Valley of the Rio de Allende (fig.
3), so this meteorite became known as Allende. The largest fragment recovered was discovered at the northern end of the
strewnfield just to the east of Sierra de Almoloya (fig. 3) and weighed approximately 110 kg, (242 lb) but had been broken
into many fragments on impact (Clarke et al. 1970).
Fig. 3. Plan of the Allende meteorite strewnfield in the state of Chihuahua, Mexico, showing the locations where some of the
larger meteorite specimens were recovered and their weights (green triangles) (after Clarke et al. 1970). Click image for
larger view.
Individual Allende specimens are fusion crusted (fig. 4). A view of a polished slab surface reveals that this meteorite consists
of a jumble of refractory Ca-Al-rich inclusions and round objects called chondrules set in a dark, carbonaceous matrix (fig.
5). Chondrules are unique to stony meteorites which are thus called chondrites, so because of its carbonaceous matrix
Allende is classified as a carbonaceous or C chondrite. Further classification involves typing according to where the first
meteorite or prototype in the category was found and whose characteristics are used to define the group. Thus Allende is a
CV chondrite, the V designating the town
of Vigarano in Emilia, Italy, in which two
stones weighing a total of 16 kg (35.2 lb)
fell in 1910 whose characteristics are used
to define this group (Norton 2002). Then
the last parameter to group chondrites is a
petrologic (petrographic)-chemical typing
scheme that provides a guide to the
degree of thermal and aqueous alteration
(Van Schmus and Wood 1967). Chondrite
types 3-6 show progressive stages of
thermal metamorphism (Norton, 2002),
with this sequence from type 3 (commonly
called unequilibrated) to type 6 (commonly
called equilibrated) representing an
increasing degree of chemical equilibrium
and textural recrystallization (Krot et al.
2005). Type 3 chondrites are widely
considered the least modified by
secondary processes, but even those
cannot be considered truly pristine, as
most show evidence of moderate amounts
of hydrothermal alteration (Norton 2002).
Nevertheless, of the Type 3 chondrites, the
reduced CV3.0 chondrites, which are the
least metamorphosed (Sears et al. 1991),
come closest to being pristine. Therefore,
Allende as a CV3 carbonaceous chondrite
is regarded as close to pristine, and thus
the intensive study of it has been
warranted.

Source: Norton (2002)


Fig. 4. A fragment of the Allende CV3 carbonaceous chondrite with a partial
fusion crust and lead-gray interior. The white specks are Ca-Al inclusions (CAIs).
This specimen measures about 12.7 cm (5 in) horizontally.
Source: Wikipedia http://www.en.wikipedia.org
Fig. 5. A polished surface of a slab of the Allende CV3 carbonaceous chondrite.
The large circular inclusions, many of them rimmed, are chondrules, and the
convoluted white inclusions are Ca-Al inclusions (CAIs), all set in a lead-gray
matrix.
The Components of the Allende CV3 Carbonaceous Chondrite Meteorite
It is useful to have some knowledge of the constituent components and minerals
within the Allende meteorite, because in dating studies radioisotope analyses
have been made on whole samples and separated components. The latter

procedure is done because it is argued that some components may have formed earlier in the solar nebula. Regardless, the
integrity of radioisotope dating results depend on the suitability and integrity of the samples, minerals, or components being
analyzed, so a knowledge of the constituents in meteorites like Allende is vital when assessing the quality of radioisotope
dates. Krot et al. (2005) describe the CV chondrites as characterized by:
Millimeter-sized chondrules with mostly porphyritic textures, most of which are Mg-rich and ~50% of which are surrounded
by coarse-grained igneous rims;
High matrix/chondrule ratios (0.5-1.2);
Unique presence of abundant salite-hedenbergite (pyroxene) andradite (garnet) nodules in the matrix;
High abundance of millimeter-to-centimeter-sized refractory Ca-Al-rich inclusions (CAIs) and amoeboid olivine aggregates
(AOAs); and
Common occurrence of igneous melilite-spinel-pyroxene anorthite CAIs.
To this list of components could also be added Fe-Ni metal alloys and sulfides.
Matrix
Matrix material is best defined as the optically opaque mixture of grains 10 nm to 5 m in size, and distinguishable by their
shapes and textures, that rims chondrules, inclusions and other components, and fills in the interstices between them (Scott
and Krot 2005). CV chondrites like Allende have much higher densities (~3.5 g/cm 3) than other carbonaceous chondrites,
essentially the same densities as ordinary chondrites (Norton 2002). This is, in part, because the matrix is much more
compactthere is less pore spaceand the water content is less than 2 wt%. The interior is also different from other
carbonaceous chondrites, the matrix being medium gray rather than black (fig. 5). CV chondrites contain on the average
about 42 vol. % matrix, but Allende has a higher average value of about 60 vol. % matrix.
Examination of matrix material in thin section shows it to be black and opaque (fig. 6). High magnification of ultra-thin
sections reveals the dominant mineral to be olivine, seen as a uniform distribution of tiny, euhedral crystals dispersed in
opaque material (Norton 2002). Like the matrix in ordinary chondrites, the olivine is Fe-rich with an average fayalite content
of 50% (Fa50). In Allende the matrix olivine grains
contain abundant inclusions of Ca-Fe-rich pyroxene,
nepheline, and pentlandite [(Fe,Ni)9S8], with magnetite
(Fe3O4) also present, and even at times some poorly
graphitized carbon (Scott and Krot 2005). The opaque
components of the matrix itself consist of pentlandite
[(Fe,Ni)9S8], troilite (FeS) and minor grains of Ni-rich
metal alloy, awaruite (Ni3Fe). Contributing to the
opacity of the matrix is evenly distributed carbonbased material that thinly coats much of the olivine.
CV3 chondrites are carbon-poor compared to the
other water-bearing carbonaceous chondrites,
averaging less than 1 wt% carbon. Allende contains
only 0.29 wt% carbon.
Source: Norton (2002)
Fig. 6. A typical view of the Allende CV3
carbonaceous chondrite under crossed-polarized light
in thin section under a petrological microscope showing the array of components. These include the rounded chondrules of
many different structures and mineralogies, the amoeboid olivine aggregates (AOAs), dark inclusions, individual mineral
grains and convoluted Ca-Al inclusions (CAIs), all set in a black matrix of tiny opaque mineral grains. The field is 22 mm
(0.866 in) wide horizontally. The elliptical chondrule with the dark center just below center and right is 3 mm (0.118 in) in its
longest dimension.
Chondrules
Chondrules are igneous-textured, rounded particles composed mainly of olivine and low-Ca pyroxene crystals with tiny
grains of Fe-Ni metal and sulfides set in a feldspathic glass or microcrystalline matrix (Krot et al. 2009; Scott and Krot 2005).
The most obvious distinction between CV3 chondrites and all other carbonaceous chondrite groups is the texture of the
chondrule fields. Chondrules fill an average of 44 vol. %, whereas in Allende it is about 30 vol. % (Norton 2002). The
chondrules are much larger on the average than those in the other groups, being typically between 0.5 and 2.0 mm (0.019
and 0.0787 in) or more. Occasionally, especially in Allende, much larger chondrules are found, some exceeding 4 or 5 mm
(0.157 or 0.196 in) or diameter with a record being a whopping 25 mm (1 in)!
The most common, almost 94% of the chondrules in Allende (Scott and Krot 2005), are porphyritic olivine chondrules made
up of small tightly packed euhedral and subhedral olivine grains remarkably uniform in size (>90 modal % olivine) (fig. 7)
(Norton 2002; Scott and Krot 2005). Some contain elongated prisms of clinoenstatite showing polysynthetic twinning (fig. 8).
Commonly seen within these chondrules are opaque grains of Fe-Ni metal alloys, iron sulfide, and magnetite. Surrounding
the chondrules are halos of dark brown material composed of serpentinite, a product of aqueous alteration of olivine
followed by dehydration (fig. 9). Mixed with the serpentinite is opaque FeS and NiFe 2S. The olivine within many of the
chondrules is almost pure forsterite averaging (Fa 6), but larger crystals frequently show a zoning with increasing Fe content
towards their edges and may be caused by reaction with the Fe-rich matrix (Norton 2002).
Source: Norton (2002).
Fig. 7. A porphyritic olivine chondrule under crossedpolarized light in thin section under a petrological
microscope with anhedral olivine grains showing
bright interference colors. A large deeply embayed
blue olivine grain occupies half of the interior of this
exceptionally large chondrule, measuring 4 mm
(0.157 in) in diameter, in the Allende CV3
carbonaceous chondrite.
Source: Norton (2002)
Fig. 8. A porphyritic olivine-pyroxene chondrule
measuring 1.8 mm (0.070 in) across under crossedpolarized light in an Allende CV3 carbonaceous

chondrite thin section under a petrological microscope. It contains small colorful equant crystals of olivine interspersed with
large lath-shaped blades of white clinoenstatite (pyroxene).
Source: Norton (2002)
Fig. 9. A large granular olivine chondrule measuring 3.7 mm (0.14 in) across with tiny anhedral olivine crystals under
crossed-polarized light in an Allende CV3 carbonaceous chondrite thin section under a petrological microscope. The entire
chondrule is surrounded by brown serpentinite produced by hydrous alteration of olivine. The blue chondrule to the upper
right is a porphyritic pyroxene chondrule.
Barred olivine chondrules are rarer in CV3 chondrites compared to the porphyritic chondrules, but still comprise up to 6% of
the chondrules (Scott and Krot 2005). The bars may be contiguous with the rim material as one crystal (fig. 10) or several
sets of parallel prisms with different orientations and thus different birefringent colors may compose the chondrule (fig. 11).
All are set in a glassy mesostasis. Often surrounding the barred olivine chondrules is a corona of loosely packed tiny
olivine crystals that is in optical continuity with the chondrule.
Porphyritic pyroxene chondrules are even rarer than barred olivine chondrules. Their individual crystals of low-Ca pyroxene
are usually large, tightly grouped and lack rim material. These excentroradial pyroxene chondrules are very rare, amounting
to less than 0.2 vol. % (Scott and Krot 2005).
Source: Norton (2002)
Fig. 10. A barred olivine chondrule from the Allende CV3
carbonaceous chondrite under crossed-polarized light in thin
section under a petrological microscope. The olivine bars and
double rim are in optical continuity, and the dark object below,
interrupting the thick rim, is a missing section with the mounting
glass in extinction. Several circular blebs of Fe-Ni metal appearing
black are in the rim above. The field of view is 1.9 mm (0.074 in) in
the longest dimension.

Source: Norton (2002)


Fig. 11. A polysomatic barred olivine chondrule under
crossed-polarized
light
under
a
petrological
microscope showing several sets of parallel olivine
bars oriented at different angles to each other
resulting in different interference colors and extinction
points. A thick rim encloses the polysomatic bars except at the upper edge where the rim has been disrupted. The horizontal
field of view in this Allende CV3 carbonaceous chondrite thin section is 2.1 mm (0.082 in).
All of the chondrules so far described are textually similar to chondrules found in ordinary chondrites. There are chondrules
in CV3 chondrites that are extremely rare in other chondrite groups but rather common in CV3 chondrites such as Allende
(Norton 2002). In particular, in thin sections in transmitted light some remarkably round forms ranging in size from 0.1 mm to
2 mm (0.039 in to 0.078 in) stand out. They have an overall dark gray color with white needle-like crystals arranged around
the perimeter roughly aligned radially with the center. These are anorthite-forsterite-spinel chondrules. The white crystals are
Ca-rich plagioclase or anorthite prisms getting progressively smaller from edge to center (fig. 12). Spinel (MgA l2O4) is found
as tiny dark, opaque crystals throughout the chondrule and within the anorthite prisms, with forsterite (olivine) occurring
between the prisms.
Source: Norton (2002)
Fig. 12. An anorthite-forsterite-spinel chondrule from the
Allende CV3 carbonaceous chondrite in thin section
under a petrological microscope. The large white
crystals along the edge are anorthite (plagioclase).
Spinel is scattered through the chondrule as very small
opaque crystals. The darker blades are forsterite
(olivine). Note that the forsterite and plagioclase are
radially arranged with respect to the center. A sparse
olivine aggregate is in the lower left and upper right
corners. Viewed in crossed-polarized light, the
chondrules diameter is 1.6 mm (0.062 in).
Amoeboid Olivine Aggregates
Amoeboid olivine aggregates (AOAs) are irregularly
shaped objects with fine grain sizes (520 m) that
occupy up to a few percent of the carbonaceous
chondrites (Scott and Krot 2005). Scattered more or less uniformly in the matrix, these are aggregates of micrometer-sized
crystals composed primarily of olivine, with lesser amounts of pyroxene and feldspathoids (Norton 2002). In the leastmetamorphosed carbonaceous chondrites such as Allende, the olivine aggregates consist of forsteritic olivine (Fa <1-3) with
minor diopside and anorthite, and in some cases spinel. With increasing alteration and metamorphism, the olivine and spinel
(where present) become increasingly FeO-rich.

Source: Norton (2002)


Fig. 13. An amoeboid olivine aggregate from the
Allende CV3 carbonaceous chondrite in thin section
under a petrological microscope viewed in crossedpolarized light. Note the pseudo-pod-like arms
projecting into the matrix. The aggregate is 0.8 mm
(0.314 in) at its narrowest.
Olivine aggregates can be associated with olivinerich chondrules, forming a surrounding halo, or they
can appear amoeboid-like as singular, irregular
crystalline olivine masses with projecting arms
(Norton 2002) (fig. 13). Like the olivine chondrules,
the olivine in aggregates is nearly pure forsterite.
Often enclosed within the aggregates are spherical
nodules of high temperature oxides and silicates
(such as melilite). The aggregates do not show
characteristics of having been melted but accreted
around the nodules, consolidating into aggregates within the matrix.
Dark Inclusions
These are relatively common in CV3 chondrites and they remain perhaps the most enigmatic of the inclusions (Norton
2002). They are usually revealed in cut slabs as angular in shape and much larger than chondrules. To the eye, they appear
fine-grained and featureless (fig. 14a), but under the microscope some show very small, roughly equant chondrules
averaging about 0.1 mm (0.039 in) in diameter or less, superficially resembling a fragment of a CO3 chondrite (fig. 14b). The
matrix around the fragments sometimes shows signs of alteration, and the composition is very close to that of the CV3
chondrite matrix.
Source: Norton (2002)
Fig. 14. (a) A xenolithic dark inclusion about 32 mm (1.25 in) long in an Allende CV3 carbonaceous chondrite slab. (b)
Photomicrograph of the xenolithic inclusion in (a) in thin section under a petrological microscope showing tiny equant
chondrules in a
dark matrix. Around the edge of the inclusion is an altered zone that separates
the CV3 texture from the inclusion texture. This mimics a fragment of a CO3
(carbonaceous Orgueil type 3) chondrite.
Calcium-Aluminum Inclusions (CAIs)
A very different inclusion is prominent in CV3 chondrites. They are non-igneous
(interpreted as evaporative residues or gas-solid condensates by uniformitarians)
or igneous (interpreted as possibly melted condensates by uniformitarians)
objects rich in Ca, Al, and Ti giving their name Ca-Al inclusions, or CAIs (Krot et
al. 2009; MacPherson 2005). They stand out conspicuously on broken surfaces
of Allende and other CV3 chondrites, appearing as whitish or pinkish irregular
masses contrasting against the black matrix. They are larger than most
chondrules. Although they had been known for many years, they became the
subject of intense research after the fall of Allende. While all CV3 chondrites
contain these white inclusions, Allende seems to possess more than any other
with about 510 vol. % (Norton 2002; Scott and Krot 2005). There are two types
based on texture: fine-grained and coarse-grained. The fine-grained variety is
composed of crystals less than 1 m in size, too small to see clearly with a light
microscope. Their irregular shapes terminate in lobate forms (fig. 15a). Under
crossed-polarized light they have very low or no birefringence, remaining a grayblue-silvery tone. The coarse-grained types have much larger crystals and some
appear to have reaction rims (fig. 15b).
Source: Norton (2002)
Fig. 15. (a) A large fine-grained Ca-Al-inclusion (CAI) 5.5 mm (0.216
in) long from an Allende CV3 carbonaceous chondrite specimen in
thin section viewed in crossed-polarized light under a petrological
microscope. It contains melilite (a silica-poor feldspathoid), spinel,
and fassaite (a variety of the pyroxene augite). Melilite is the white
bordering material. (b) A coarse-grained CAI in another Allende CV3
carbonaceous chondrite specimen viewed in thin section under
crossed-polarized light under a petrological microscope. The white
areas are melilite that has been partially altered to andradite (the CaFe garnet) and grossular (the Ca-Al garnet), appearing dark gray to
black. The faint rims around the andradite are diopside (pyroxene).
The inclusion is 1.7 mm (0.066 in) long.
The mineralogy of these CAIs is complex, being composed in part of
highly refractory oxides such as melilite (Ca2Al2SiO7), spinel
(MgAl2O4), and perovskite (CaTiO3), and the silicates clinopyroxene
and anorthite (CaAl2Si2O8), giving CAIs a ceramic-like chemistry and
mineralogy (MacPherson 2005). This mineralogy is comparable to
that of the anorthite-forsterite-spinel chondrules. There is also a long
list of unusual minerals found in CAIs that are rich in refractory
elements, some of which are not otherwise seen in meteorites. The
rims of some of the coarse-grained CAIs consist of two layers of
olivine plates and laths (Fa5-50 in Allende) and an outer layer of
andradite (garnet) and hedenbergite (pyroxene), showing that they

were involved in reactions with Mg-rich olivine that condensed on them while still in a molten state (Norton 2002; Scott and
Krot 2005).
Radioisotope dating is claimed to show the CAIs to be the oldest of any solar system material (MacPherson 2005; Norton
2002; Scott and Krot 2005). The other significant feature of the CAIs is their isotopic composition. Generally, the elemental
isotopic ratios of the elements in both the earths crustal rocks and the meteorites are very nearly the same. However, a
closer look at the chondrite meteorites, especially the carbonaceous C chondrites, shows the presence of isotopic
anomalies, deviations in the isotopic ratios. For example, some of the apparently first-formed Alrich minerals (the primary
minerals found in CAIs) such as anorthite (CaAl2Si2O8), melilite (Ca2Al2SiO7), and spinel (MgAl2O4), contain excess 26Mg in
their crystal lattices, which suggests that probably the apparently extinct radioisotope 26Al originally occupied those places.
The importance of such isotopic anomalies in meteorites is emphasized in the literature (MacPherson 2005; Norton 2002;
Scot and Krot 2005). They are regarded as furnishing the best clues to conditions existing within the supposed solar nebula
at the time the meteorite components were presumed to be forming.
The most important of these anomalies is that among the oxygen isotopes 16O, 17O, and 18O. The relative abundances of the
three oxygen isotopes in a sample are compared by calculating the difference between the 17O/16O ratio of the sample
(referred to as 17O) and a standard source, and then plotting this difference against the difference between the 18O/16O ratio
(18O) and the standard. The reference standard for comparing oxygen isotope ratios is terrestrial ocean water, referred to
as standard mean ocean water or SMOW.
There is a natural sorting process which causes the proportions of the three oxygen isotopes to change as a result of the
differences in their masses. This occurs in many mass-dependent physical processes, such as chemical reactions,
crystallization, vaporization and condensation, and diffusion. This mass fractionation (mass separation of the three oxygen
isotopes) is proportional to their mass differences, so for any given process, the 18O/16O ratio changes twice as much as
the 17O/16O ratio. Thus when the 17O deviations are plotted against the 18O deviations for terrestrial samples the values
reflect this fractionation and plot along a line with a slope of ~0.5. This is the terrestrial mass-fractionation line from which all
oxygen isotope ratios for different meteorite groups are compared. Any process that separates isotopes according to their
masses will produce such a distribution line. The C chondrites show a large, widely-dispersed deviation off this terrestrial
line. Most striking is the position of the CV3 chondrites, which plot along a line with a slope of ~1. These data come primarily
from chondrules and high temperature (refractory) minerals found in Allende CAIs, which are enriched in 16O relative to the
other two heavier oxygen isotopes (MacPherson 2005; Norton 2002; Scott and Krot 2005). So another process apart from
mass fractionation must be responsible, potentially suggesting that the oxygen isotopic compositions of Allende could have
resulted from a mixture of two distinctly different materials.
Fe-Ni Metal Alloys and Sulfides
The form and occurrence of metallic grains in carbonaceous C chondrites are quite distinctive, tending to be rounded or
ovoid in shape, and globules with somewhat roughened surfaces (Wood 1967). Two kinds of metallic grains are found in
chondritesgrains composed of refractory elements (Ir, Os, Ru, Mo, W, and Re) associated with CAIs, and grains
composed predominantly of Fe, Co, and Ni associated with chondrules (Scott and Krot 2005). Most refractory nuggets have
been studied in Allende CAIs. The metallic Fe-Ni grains found in chondrules in most type 3 chondrites such as Allende
(CV3) typically contain concentrations of 0.11% Cr, Si, and P.
Most of the metallic grains in Allende (~0.5 wt%) are awaruite (Ni 3Fe), an Fe-Ni alloy similar to taenite (Norton 2002). Minor
grains are found among the opaque components of the matrix itself, along with pentlandite [(Fe,Ni) 9S8] and troilite (FeS). In
Allende even the matrix olivine grains contain abundant inclusions of pentlandite, with magnetite (Fe 3O4) also present (Scott
and Krot 2005). Opaque grains of Fe-Ni metal alloys, iron sulfide, and magnetite are also commonly seen even distributed
within the porphyritic olivine chondrules (Norton 2002; Wood 1967).
Organic Carbon
Some carbonaceous C chondrites are rich in carbon at 1.56 wt%, but others are not (Scott and Krot 2005). Allende in its
matrix contains carbon, but only 0.29 wt% (Norton 2002). This is referred to as organic carbon, a term that does not mean
biogenic carbon, that is, carbon compounds made by living organisms. Instead, they are carbon compounds that have their
counterparts in terrestrial organic carbon present in living organisms, but which can be formed by non-biogenic processes.
Many of the analyses to identify the carbon compounds in carbonaceous meteorites were performed on the Murchison CM
chondrite, which like Allende fell in 1969. Of the total molecular carbon in Murchison, about 70% is in the form of insoluble
molecular material not completely identified. The remaining 30% is soluble organic compounds. The list of these organic
compounds identified includes non-volatile (>C10) aliphatic hydrocarbons (saturated and unsaturated open long-chain
hydrocarbons), amino acids, aromatic hydrocarbons, carboxylic acids, dicarboxylic and hydroxyl-carboxylic acids, nitrogen
heterocycles, and aliphatic amines and amides (Cronin, Pizzarello, and Cruikshank 1988). Specific amino acids of identical
composition and structure were found to be optically racemic, with equal numbers of left-handed and right-handed optical
forms, which strongly suggested an extraterrestrial origin. Some 74 amino acids have been identified, of which eight are
involved in protein synthesis in terrestrial life-forms. An additional 11 are also biogenic, though less commonly seen in
biological systems. The majority of the 74 amino acids have no counterparts on earth and are thus truly extraterrestrial.
Though none of these many organic
compounds are now thought to be the result
of life processes in space, they showed that
complex organic molecules could apparently
form beyond earths environment.
Radioisotope Dating of the Allende CV3
Carbonaceous Chondrite Meteorite
The first radioisotope dating studies on
Allende were published in 1973a wholerock Pb-Pb model age of 4.528 Ga (Huey
and Kohman 1973); two whole-rock Pb-Pb
model ages of 4.51 Ga and 4.52 Ga (Tilton
1973); Rb-Sr model ages for individual
whole-rock samples, chondrules, and Ca-Al
inclusions (CAIs) ranging from 3.59 Ga to
4.63 Ga (Gray, Papanastassiou, and
Wasserburg 1973); and a Pb-Pb model age
of 4.496 Ga (Tatsumoto, Knight, and Allgre
1973). Subsequently, the first statisticallysignificant isochron date was obtained by

Tatsumoto, Unruh, and Desborough (1976) using some 20 isotopic analyses of the matrix, magnetic separates, aggregates,
and chondrules that yielded a Pb-Pb isochron age of 4.553 0.004 Ga (fig. 16).
Fig. 16. The Pb-Pb isochron for the Allende CV3 carbonaceous chondrite obtained by Tatsumoto, Unruh, and Desborough
(1976). This analysis used 28 mineral fractions and chondrules separated from the meteorite (green circles). The isochron
passes through the composition of troilite from the Canyon Diablo iron meteorite (yellow circles). Where data are too tightly
clustered to be shown individually, the number of data represented by a single symbol is indicated. The data for one
chondrule was not used to calculate the isochron age, but the result is not significantly affected by its exclusion (after
Dalrymple 2004).
Numerous radioisotope dating studies of Allende have been undertaken in the ensuing decades, and many major studies
have even been published in the last five or so yearsAmelin et al. (2009), Amelin et al. (2010), Bouvier, Vervoort, and
Patchett (2008), Burkhardt et al. (2008), Connelly et al. (2008), Connelly and Bizzarro (2009), Hans, Kleine, and Bourdon
(2013), Jacobsen et al. (2008), Krot et al. (2009), and Trinquier et al. (2008)indicating the scientific community still regards
this meteorite as significant to their attempts to understand
the origin and formation of the solar system and the earth
by only natural processes. The currently accepted best
estimate of the age of the Allende CV3 carbonaceous
chondrite meteorite is 4.56718 0.0002 Ga based on a
Pb-Pb isochron obtained from Pb isotopic analyses of
fractions of an Allende CAI (fig. 17) (Amelin et al. 2010).
Fig. 17. Isochron plot of the Pb isotopic data for the
Allende CAI SJ101 which yields the current best estimate
for the age of the Allende CV3 carbonaceous chondrite
meteorite at 4.56718 0.0002 Ga (after Amelin et al.,
2010). The regression includes residues after acid
washing of the fractions 1, 1+, 7, 8, and 9, and numerically
recombined residues and third washes of the fractions 3, 4
and 5. The age uncertainty in brackets includes
uncertainty of the 238U/235U ratio.
Creationists have commented little on the radioisotope
dating of meteorites. Woodmorappe (1999) only very
briefly discussed the claimed use of meteorites to date the
earths age via the Pb-Pb isochron method, but did not
elaborate on what ages have been obtained on meteorites
by all the various radioisotope methods and any
significance or otherwise that can or should be attributed
to those ages.
Similarly, Snelling (2005, 2009) only briefly touched on the
meteorite Pb-Pb isochron used to date the age of the earth, the so-called geochron, and other radioisotope methods used to
date meteorites, but he also mentioned the similar bulk geochemistry of meteorites to the earths mantle in the context of
crustal rocks potentially inheriting their isotopic ratios from mantle rocks. Snelling (2005) also suggested that future work
should focus on compiling the copious radioisotope data on meteorites to examine the claimed concordance of the isochron
ages for them, especially since the meteorite radioisotope data could be relevant to the question of just how much apparent
accelerated radioisotope decay may have occurred during the earths history.
Morris (2007) also highlighted the significance of meteorites in the determination of the age of the earth, and then focused
on Allende as an example of a well-studied meteorite analysed by many radioisotope dating methods. However, he only
discussed the radioisotope dating results from one, older paper (Tatsumoto, Unruh, and Desborough 1976). In that
discussion he focused on the U-Th-Pb model ages published in that paper to report that they are very discordant (which has
been well-documented in the literature for some time), and therefore to demonstrate that such dating methods are
unreliable. He also stated that no isochron was possible. In so saying he completely ignored the excellent Pb-Pb isochron
age of 4.553 0.004 Ga based on some twenty isotopic analyses of the matrix, magnetic separates, aggregates and
chondrules reported in the same paper (fig. 16), as well as the U-Pb concordia isochron age of 4.548 0.025 Ga based on
those same samples.
Therefore, in order to thoroughly investigate the radioisotope dating of the Allende CV3 carbonaceous chondrite meteorite
all the relevant literature was searched. When papers containing radioisotope dating results for the Allende meteorite were
found, the reference lists were scanned to find further relevant papers. In this way a comprehensive set of papers, articles,
and abstracts on the radioisotope dating of Allende was collected. While it cannot be claimed that all the papers, articles,
and abstracts which have ever been published containing radioisotope dating results for Allende have thus been obtained,
the cross-checking undertaken between these publications does indicate the data set obtained is very comprehensive.
All the radioisotope dating results from these papers, articles and abstracts were then compiled and tabulated. Because of
the huge number of radioisotope dates, and because of the various portions of the Allende CV3 chondrite which have been
analysed, the results were tabulated in four categoriesisochron ages for some or all components (table 1), model ages for
chondrules (table 2), model ages for CAIs (table 3), and model ages for wholerock, matrix and other samples (table 4).
The data in these tables were then plotted on frequency versus age histogram diagrams, with the same color coding being
used to show the ages obtained by the different radioisotope dating methodsthe isochron ages for some or all
components (fig. 18), the model ages for chondrules (fig. 19), the model ages for CAIs (fig. 20), and the model ages for
whole-rock, matrix and other samples (fig. 21).
Table 1. Isochron ages for some or all components of the Allende CV3 carbonaceous chondrite meteorite, with the details
and literature sources.
Sample

Method

Age

inclusions

KAr

4.82

chondrule(s)

RbSr

4.38

Error +/ Notes

Source

Type

Decayers
Sample 12, White
0.08

Jessberger et al. 1980 isochron age


Shimoda et al. 2005

isochron age

Podosek et al. 1991,


Using D'Orbigny plagioclaseHans,
Kleine,
and
Sr initial
Bourdon 2013
isochron age

inclusions

RbSr

4.56

inclusions

RbSr

4.63

0.23

Measured Data

Hans,
Kleine,
Bourdon 2013

and
isochron age

inclusions

RbSr

4.55

0.14

Corrected Data

Hans,
Kleine,
Bourdon 2013

and
isochron age

4.78

0.62

Three samples plotted with 5


other samples from 5 other
meteorites
Patchett et al. 2004

4.553

0.004

Tatsumoto, Unruh, and


Desborough 1976
isochron age

Pb206Pb 4.565

0.004

Chen and Tilton 1976

wholerock samples LuHf

isochron age

Decayers
inclusions

PbPb

wholerock samples

207

isochron age

inclusions

204Pb
207
Pb

4.559

0.004

Chen and Wasserburg


1981
isochron age

inclusions (CaAl)

PbPb

4.566

0.002

Manhes, Gopel,
Allgre 1988

inclusions

207

Pb206Pb 4.566

0.002

chondrules (9)

PbPb

4.5668

0.0016

inclusions (CaAl)

PbPb

4.567

0.006

inclusions (CaAl)

PbPb

4.5666

0.0017

CaAl

3D linear regression,
samples with E

and
isochron age

Allegre, Manhes, and


Gopel1995
isochron age
Amelin et al. 2002

isochron age

six
Amelin et al. 2002

isochron age

Amelin et al. 2002

isochron age

inclusions (CaAl)

207

Pb206Pb 4.5685

0.0005

Three UCLA residues plotted


with seven E6O inclusions
(Amelin et al. 2002)
Bouvier et al. 2007

inclusions (CaAl)

207

Pb206Pb 4.568

0.0094

Three UCLA residues

0.0001

Acidleached residues from


a CaAl inclusion plus aBouvier, Vervoort, and
leachate
Patchett 2008
isochron age

inclusions (CaAl)

PbPb

4.56759

chondrules

207

Pb206Pb 4.56545

0.00045

inclusions (CaAl)

207

Pb206Pb 4.56772

0.00093

inclusion (CaAl)

207

Pb206Pb 4.56859

inclusion (CaAl)

207

Pb206Pb 4.56781

inclusion (CaAl)

207

inclusions (CaAl)

isochron age

Bouvier, Vervoort, and


Patchett 2008
isochron age

Connelly et al. 2008

isochron age

Six points (px + melilite)

Connelly et al. 2008

isochron age

0.00061

F2 (px + mel + plag)

Connelly et al. 2008

isochron age

0.00079

TS32 (px + mel + plag)

Connelly et al. 2008

isochron age

Pb Pb 4.5707

0.00110

TS33 (px + mel + plag)

Connelly et al. 2008

isochron age

207

Pb206Pb 4.56744

0.00034

CAIs AJEF & A43 residues &


px second wash
Jacobsen et al. 2008

isochron age

inclusions (CaAl)

207

Pb206Pb 4.5676

0.00036

AJEF residues & px second


wash
Jacobsen et al. 2008

isochron age

inclusions (CaAl)

207

Pb206Pb 4.5675

0.0014

A43 residues & px second


wash
Jacobsen et al. 2008

isochron age

chondrules

PbPb

4.56532

0.00081

Ten chondrules, plus AmelinConnelly and Bizzarro


& Krot (2007) chondrule data 2009
isochron age

chondrules

207

Pb206Pb 4.56537

0.00045

Combined Amelin et al. 2002


and Connelly et al. 2008 dataKrot et al. 2009

isochron age

inclusions

PbPb

0.0007

Third washes of sample


batch A004 in the data set Amelin et al. 2010

isochron age

0.00063

Complementary residues of
sample batch A004 in the
data set
Amelin et al. 2010

isochron age

0.00019

Residues of sample batches


A003 and A005 in the data
set
Amelin et al. 2010

isochron age

inclusions

inclusions

206

PbPb

PbPb

4.56772

4.56606

4.56726

inclusions

PbPb

4.56709

0.00062

All
data
residues)

inclusions

PbPb

4.56632

0.00076

All residues in the data set

inclusions

PbPb

4.56726

inclusions

PbPb

4.56679

inclusions

(washes

and
Amelin et al. 2010

isochron age

Amelin et al. 2010

isochron age

0.00019

Sample batches A003 and


A005 only, A005 #6 excluded Amelin et al. 2010

isochron age

0.00022

All residues and second


washes in the data set
Amelin et al. 2010

isochron age

All residues in batches A003


and A005, and recombined
residues and third washes in
batch A004 (#2 ecluded)
Amelin et al. 2010

isochron age

PbPb

4.56718

0.00021

UPb

4.914

0.080

Tatsumoto, Unruh, and


Desborough 1976
isochron age

UPb

4.553

0.07

Tatsumoto, Unruh, andisochron


age
Desborough 1976
(concordia)

inclusions

UPb

4.548

0.025

Tatsumoto, Unruh, andisochron


age
Desborough 1976
(concordia)

inclusions (CaAl)

UPb

4.56805

0.00065

3D linear
samples

inclusions (CaAl)

UPb

4.568

0.009

3D linear isochron,
samples with E

Amelin et al. 2002

isochron age

inclusions (CaAl)

UPb

4.568

0.015

3D linear planar, six samples


with E
Amelin et al. 2002

isochron age

inclusions

UPb

4.56774

0.00033

Third washes and residues

Amelin et al. 2010

isochron age

0.0001

ResiduesA003,
A005
measured;
A004
recombinedforced through
O
Amelin et al. 2010

isochron age

isochron age

inclusions
matrix

and

inclusions
matrix

and

inclusions

UPb

4.56704

regression,

six
Amelin et al. 2002

isochron age

six

inclusions

UPb

4.566

0.01

ResiduesA003,
A005
measured;
A004
recombinedunconstrained Amelin et al. 2010

inclusions (CaAl)

SmNd

4.80

0.07

Coarsegrained, melilite andPapanastassiou, Ngo,isochron age (4


pyroxene separates
and Wasserburg 1987 point)

inclusions (CaAl)

SmNd

4.80

0.18

Coarsegrained, melilite andPapanastassiou, Ngo,isochron age (2


pyroxene separates
and Wasserburg 1987 point)

inclusions (CaAl)

SmNd

4.53

0.09

Coarsegrained, melilite andPapanastassiou, Ngo,isochron age (2


pyroxene separates
and Wasserburg 1987 point)

0.1

One sample plotted with 33


other samples from 8 otherAmelin and Rotenberg
meteorites
2004
isochron age

0.0004

Six unleached fractions of a


CaAl inclusion, relative toBouvier, Vervoort, and
D'O E60 PbPb age
Patchett 2008
isochron age

0.0003

Six unleached fractions of a


CaAl inclusion, relative toBouvier, Vervoort, and
D'O PbPb age
Patchett 2008
isochron age

isochron age
isochron age

chondrule

SmNd

4.588

Calibrated by PbPb Ages

inclusions (CaAl)

inclusions (CaAl)

AlMg

AlMg

4.5671

4.5692

inclusions (CaAl)

HfW

4.5678

0.0022

Six separates of CaAl


inclusions relative to St.
Marguerite
Kleine et al. 2005

inclusions (CaAl)

HfW

4.4678

0.0022

Six separates from MS1

Kleine et al. 2005

inclusions (CaAl)

HfW

4.5683

0.0007

Numerous mineral separates


from
CaAl
inclusions,
relative to PbPb age of D'O,
Sah & NWA
Burkhardt et al. 2008

inclusions

I/Xe

4.5670

0.0002

Dark

Hohenburg et al. 2001 isochron age

inclusions

I/Xe

4.5678

0.0002

Dark

Hohenburg et al. 2001 isochron age

isochron age

bulk sample

MnCr

4.5681

0.0009

One bulk sample plotted with


five
other
carbonaciousShukolyukov
chondrites
Lugmair 2006

fractions (9)

MnCr

4.5625

0.0016

Nine fractions including CAI,


whole rock, and chondrule Trinquier et al. 2008

isochron age

chondrules

MnCr

4.56791

0.00076

Relative to the PbPb age of


D'Orbigny
Yin et al. 2009

isochron age

chondrules

MnCr

4.56742

0.00083

Relative to the PbPb age of


LEW 86010
Yin et al. 2009

isochron age

and
isochron age

Table 2. Model ages for chondrules in the Allende CV3 carbonaceous chondrite meteorite, with the details and literature
sources.
Sample

Method

Age

Error +/ Notes

Source

Type

40

4.56

0.05

Sample 5, Fine barred

Jessberger et al. 1980

plateau age

Decayers
chondrule(s)

Ar39Ar

ferromagnesian
chondruleA

KAr

4.63

0.01

Sample 3915

Herzog et al. 1980

total fusion
model age

ferromagnesian
chondruleA

KAr

4.36

0.01

Sample 3915

Herzog et al. 1980

total fusion
model age

ferromagnesian
chondruleA

KAr

4.44

0.01

Sample 3915

Herzog et al. 1980

total fusion
model age

ferromagnesian
chondruleA rim

KAr

4.76

0.02

Sample 3915

Herzog et al. 1980

total fusion
model age

ferromagnesian
chondruleB

KAr

3.97

0.01

Sample 3915

Herzog et al. 1980

total fusion
model age

ferromagnesian
chondruleB

KAr

4.08

0.01

Sample 3915

Herzog et al. 1980

total fusion
model age

ferromagnesian
chondruleC

KAr

4.43

0.02

Sample 3915

Herzog et al. 1980

total fusion
model age

ferromagnesian
chondruleC

KAr

4.40

0.02

Sample 3915

Herzog et al. 1980

total fusion
model age

chondrule

KAr

4.62

0.07

Sample 3, Monosomatic

Jessberger et al. 1980

model age

chondrule

KAr

4.63

0.07

Sample 4, Barred

Jessberger et al. 1980

model age

chondrule

KAr

4.26

0.16

Sample 5, Fine barred

Jessberger et al. 1980

model age

chondrule

KAr

4.53

0.06

Sample 6, Granular

Jessberger et al. 1980

model age

chondrule

KAr

4.51

0.07

Sample 7, Granular

Jessberger et al. 1980

model age

chondrule

KAr

4.56

0.03

Sample 8, Black

Jessberger et al. 1980

model age

olivine condrule

RbSr

4.33

0.04

B17

Gray, Papanastassiou, and


Wasserburg 1973
model age

olivine condrule

RbSr

4.59

0.03

B12

Gray, Papanastassiou, and


Wasserburg 1973
model age

olivine condrule

RbSr

4.28

0.02

B22

Gray, Papanastassiou, and


Wasserburg 1973
model age

pyroxene
condrule

RbSr

4.33

0.02

A5

Gray, Papanastassiou, and


Wasserburg 1973
model age

pyroxene
condrule

RbSr

4.53

0.02

D4

Gray, Papanastassiou, and


Wasserburg 1973
model age

pyroxene
condrule

RbSr

4.23

0.02

B19

Gray, Papanastassiou, and


Wasserburg 1973
model age

pyroxene
condrule

RbSr

4.49

0.02

E1 s2

Gray, Papanastassiou, and


Wasserburg 1973
model age

pyroxene
condrule

RbSr

4.5

0.02

E1 s3

Gray, Papanastassiou, and


Wasserburg 1973
model age

chondrule

RbSr

4.08

CaAlrich

Tatsumoto, Unruh,
Desborough 1976

and
model age

RbSr

4.17

CaAlrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.34

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.36

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.36

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

2.64

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.49

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.03

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.38

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.17

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.10

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.17

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.35

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.19

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.22

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule

RbSr

4.08

Mgrich

Tatsumoto, Unruh,
Desborough 1976

and

chondrule
chondrule

RbSr

4.37

H5

Shimoda et al. 2005

model age

chondrule

RbSr

9.24

H293 inner

Shimoda et al. 2005

model age

chondrule

RbSr

4.04

H293 outer

Shimoda et al. 2005

model age

chondrule

RbSr

3.54

H343

Shimoda et al. 2005

model age

chondrule

RbSr

1.78

H344

Shimoda et al. 2005

model age

chondrule

RbSr

4.40

H356 inner

Shimoda et al. 2005

model age

chondrule

RbSr

1.96

H356 outer

Shimoda et al. 2005

model age

chondrule

RbSr

4.29

H350

Shimoda et al. 2005

model age

chondrule

RbSr

4.31

H353

Shimoda et al. 2005

model age

chondrule

RbSr

4.13

H355

Shimoda et al. 2005

model age

chondrule

RbSr

4.45

H4

Shimoda et al. 2005

model age

chondrule

RbSr

4.22

H13

Shimoda et al. 2005

model age

chondrule

RbSr

4.28

H18

Shimoda et al. 2005

model age

chondrule

RbSr

4.12

H44

Shimoda et al. 2005

model age

chondrule

RbSr

3.99

H280

Shimoda et al. 2005

model age

chondrule

RbSr

4.08

H297

Shimoda et al. 2005

model age

chondrule

RbSr

4.00

H298

Shimoda et al. 2005

model age

chondrule

RbSr

4.34

H305

Shimoda et al. 2005

model age

chondrule

RbSr

4.22

H336

Shimoda et al. 2005

model age

chondrule

RbSr

4.71

H347

Shimoda et al. 2005

model age

model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age

chondrule

RbSr

4.80

H351

Shimoda et al. 2005

model age

chondrule

RbSr

4.70

H357

Shimoda et al. 2005

model age

chondrule

RbSr

4.80

NH3

Shimoda et al. 2005

model age

chondrule

RbSr

3.89

H349

Shimoda et al. 2005

model age

207

Pb206Pb

4.573

Chen
1976

and

chondrule

207

Pb206Pb

4.583

Chen
1976

and

chondrule

207

Pb206Pb

4.533

Chen
1976

and

chondrule

207

Pb206Pb

4.568

Chen
1976

and

chondrule

4.552

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.569

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.567

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.544

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.553

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.540

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.503

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.488

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.546

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.528

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.554

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.554

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.518

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.568

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.533

Tatsumoto, Unruh,
and
Desborough
1976
model age

Decayers

chondrule N1

chondrule N1

chondrule N2

chondrule N2

chondrule N3

chondrule N3

chondrule N17

chondrule N17

chondrule N18

chondrule N18

chondrule N19

chondrule N19

chondrule N20

chondrule N20

chondrule N34

207

207

207

207

207

207

207

207

207

207

207

207

207

207

207

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Pb206Pb

Tilton
model age
Tilton
model age
Tilton
model age
Tilton
model age

chondrule N40

207

Pb206Pb

4.521

Tatsumoto, Unruh,
and
Desborough
1976
model age
Tatsumoto, Unruh,
and
Desborough
1976
model age

chondrule N41

207

Pb206Pb

4.538

207

Pb206Pb

4.5705

0.0032

Residue 1

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.57

0.0036

Residue 2

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5702

0.004

Residue 3

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5635

0.0037

Residue 4

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.568

0.0019

Residue 5

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5554

0.0037

Residue 6

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5602

0.0021

Residue 7

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5503

0.0095

Residue 8

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5512

0.0071

Residue 9

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5616

0.0018

Residue 10

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.565

0.0065

Residue 11

Amelin
2007

and

chondrules (2)

Krot
model age

207

Pb206Pb

4.5653

0.0011

Residue 12

Amelin
2007

and

chondrule

Krot
model age

207

Pb206Pb

4.5645

0.0016

Residue 13

Amelin
2007

and

chondrules (2)

Krot
model age

207

Pb206Pb

4.5671

0.0017

Residue 14

Amelin
2007

and

chondrules (2)

Krot
model age

207

Pb206Pb

4.5627

0.0026

Residue 15

Amelin
2007

and

chondrules (6)

Krot
model age

207

Pb206Pb

4.5625

0.0018

Residue 16

Amelin
2007

and

chondrules (4)

Krot
model age

207

Pb206Pb

4.561

0.0024

Residue 17

Amelin
2007

and

chondrules (4)

Krot
model age

Pb206Pb

4.5662

0.0036

Residue 18

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.5736

0.0031

Residue 19

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.5681

0.0031

Residue 20a

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.5694

0.0076

Residue 20c

Amelin
2007

and

207

Krot
model age

207

Pb206Pb

4.5656

0.0033

Residue 21

Amelin
2007

and

chondrules (12)

Krot
model age

207

Pb206Pb

4.5661

0.0045

Residue 22

Amelin
2007

and

chondrules (8)

Krot
model age

207

Pb206Pb

4.5681

0.0034

Residue 23

Amelin
2007

and

chondrules (19)

Krot
model age

207

Pb206Pb

4.5641

0.0023

Residue 24

Amelin
2007

and

chondrules (21)

Krot
model age

chondrules
(multiple 1)
chondrules
(multiple 2)
chondrules
(multiple 3)
chondrules
(multiple 4)

207

Pb206Pb

4.5673

0.0024

Residue 25

Amelin
2007

and

chondrules (16)

Krot
model age

207

Pb206Pb

4.5663

0.0018

Residue 26

Amelin
2007

and

chondrules (9)

Krot
model age

207

Pb206Pb

4.5659

0.002

Residue 27

Amelin
2007

and

chondrules (14)

Krot
model age

207

Pb206Pb

4.5656

0.0017

Residue 28

Amelin
2007

and

chondrules (20)

Krot
model age

Pb206Pb

4.5631

0.0017

Residue 29

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.5613

0.002

Residue 30

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.565

0.0028

Residue 31

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.5646

0.003

Residue 32

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.5635

0.0022

Residue 33

Amelin
2007

and

207

Krot
model age

Pb206Pb

4.5659

0.0017

Residue 35

Amelin
2007

and

207

Krot
model age

207

Pb206Pb

4.5423

0.013

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (12)

Krot
model age

207

Pb206Pb

4.5677

0.0073

Acid leachate 21 HCl

Amelin
2007

and

chondrules (12)

Krot
model age

207

Pb206Pb

4.5318

0.0073

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (8)

Krot
model age

207

Pb206Pb

4.3609

0.0098

Acid leachate 21 HCl

Amelin
2007

and

chondrules (8)

Krot
model age

207

Pb206Pb

4.5577

0.0049

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (19)

Krot
model age

207

Pb206Pb

4.5677

0.011

Acid leachate 21 HCl

Amelin
2007

and

chondrules (19)

Krot
model age

207

Pb206Pb

4.559

0.009

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (21)

Krot
model age

207

Pb206Pb

4.5871

0.0356

Acid leachate 21 HCl

Amelin
2007

and

chondrules (21)

Krot
model age

207

Pb206Pb

4.5629

0.0082

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (16)

Krot
model age

207

Pb206Pb

4.5767

0.0201

Acid leachate 21 HCl

Amelin
2007

and

chondrules (16)

Krot
model age

207

Pb206Pb

4.5529

0.0065

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (9)

Krot
model age

207

Pb206Pb

4.5671

0.0065

Acid leachate 21 HCl

Amelin
2007

and

chondrules (9)

Krot
model age

207

Pb206Pb

4.5653

0.0054

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (14)

Krot
model age

207

Pb206Pb

4.5748

0.0132

Acid leachate 21 HCl

Amelin
2007

and

chondrules (14)

Krot
model age

207

Pb206Pb

4.5112

0.2005

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (20)

Krot
model age

207

Pb206Pb

4.5684

0.0074

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (20)

Krot
model age

chondrules

207

Pb206Pb

4.5309

0.0019

Connelly et al. 2008model age

chondrules

207

Pb206Pb

4.5275

0.0031

Connelly et al. 2008model age

chondrules

207

Pb206Pb

4.5575

0.0034

Connelly et al. 2008model age

chondrules
(multiple)
chondrules
(multiple)
chondrules
(multiple)
chondrules
(multiple)
chondrules
(multiple)
chondrules
(multiple)

chondrules

207

Pb206Pb

4.5637

0.0018

Connelly et al. 2008model age

chondrules

207

Pb206Pb

4.5642

0.0062

Connelly et al. 2008model age

chondrules

207

Pb Pb

4.5625

0.0016

Connelly et al. 2008model age

chondrules

207

Pb206Pb

4.5652

0.0012

Connelly et al. 2008model age

chondrules

207

Pb206Pb

4.5642

0.0015

Connelly et al. 2008model age

chondrules

207

4.5436

0.0036

Connelly et al. 2008model age

chondrules

206

206

Pb Pb

PbPb

4.5666

0.001

Average of Amelin and Krot


(2007) chondrule data7
sets of 2 washes and 1Connelly
residue
Bizzarro 2009

and
model age

0.0008

Average of eight residues


model ages in Amelin andConnelly
Krot (2007) chondrule data Bizzarro 2009

and
model age

chondrules

PbPb

4.5659

232Th208Pb

3.328

Chen
1976

and

chondrule

232Th208Pb

4.622

Chen
1976

and

chondrule

232Th208Pb

3.271

Chen
1976

and

chondrule

232Th208Pb

3.862

Chen
1976

and

chondrule

11.7

Tatsumoto, Unruh,
and
Desborough
1976
model age

10.40

Tatsumoto, Unruh,
and
Desborough
1976
model age

9.55

Tatsumoto, Unruh,
and
Desborough
1976
model age

7.52

Tatsumoto, Unruh,
and
Desborough
1976
model age

5.35

Tatsumoto, Unruh,
and
Desborough
1976
model age

6.99

Tatsumoto, Unruh,
and
Desborough
1976
model age

chondrule N1

chondrule N2

chondrule N3

chondrule N17

chondrule N18

chondrule N19

232Th208Pb

232Th208Pb

232Th208Pb

232Th208Pb

232Th208Pb

232Th208Pb

Tilton
model age
Tilton
model age
Tilton
model age
Tilton
model age

chondrule N20

232Th208Pb

6.40

Tatsumoto, Unruh,
and
Desborough
1976
model age

238U206Pb

5.104

Chen
1976

and

chondrule

238U206Pb

4.688

Chen
1976

and

chondrule

238U206Pb

3.312

Chen
1976

and

chondrule

238U206Pb

4.410

Chen
1976

and

chondrule

5.34

Tatsumoto, Unruh,
and
Desborough
1976
model age

chondrule N1

238U206Pb

Tilton
model age
Tilton
model age
Tilton
model age
Tilton
model age

chondrule N1

238U206Pb

5.44

Tatsumoto, Unruh,
and
Desborough
1976
model age

chondrule N2

238U206Pb

4.40

Tatsumoto, Unruh,model age

and
1976

chondrule N2

chondrule N3

chondrule N3

chondrule N17

chondrule N17

chondrule N18

chondrule N18

chondrule N19

chondrule N19

chondrule N20

chondrule N20

chondrule N34

chondrule N40

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

238U206Pb

Desborough

4.54

Tatsumoto, Unruh,
and
Desborough
1976
model age

5.04

Tatsumoto, Unruh,
and
Desborough
1976
model age

5.07

Tatsumoto, Unruh,
and
Desborough
1976
model age

6.50

Tatsumoto, Unruh,
and
Desborough
1976
model age

6.50

Tatsumoto, Unruh,
and
Desborough
1976
model age

4.26

Tatsumoto, Unruh,
and
Desborough
1976
model age

3.91

Tatsumoto, Unruh,
and
Desborough
1976
model age

5.02

Tatsumoto, Unruh,
and
Desborough
1976
model age

5.01

Tatsumoto, Unruh,
and
Desborough
1976
model age

5.73

Tatsumoto, Unruh,
and
Desborough
1976
model age

5.69

Tatsumoto, Unruh,
and
Desborough
1976
model age

6.05

Tatsumoto, Unruh,
and
Desborough
1976
model age

6.66

Tatsumoto, Unruh,
and
Desborough
1976
model age
Tatsumoto, Unruh,
and
Desborough
1976
model age

chondrule N41

238U206Pb

6.35

238U206Pb

4.536

Residue 1

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.423

Residue 2

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.536

Residue 3

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.483

Residue 4

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.614

Residue 5

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.659

Residue 6

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.569

Residue 7

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

5.526

Residue 8

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

5.147

Residue 9

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.791

Residue 10

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

5.18

Residue 11

Amelin
2007

and

chondrules (2)

Krot
model age

238U206Pb

4.733

Residue 12

Amelin
2007

and

chondrule

Krot
model age

238U206Pb

4.875

Residue 13

Amelin
2007

and

chondrules (2)

Krot
model age

238U206Pb

4.59

Residue 14

Amelin
2007

and

chondrules (2)

Krot
model age

238U206Pb

4.824

Residue 15

Amelin
2007

and

chondrules (6)

Krot
model age

238U206Pb

4.798

Residue 16

Amelin
2007

and

chondrules (4)

Krot
model age

238U206Pb

4.85

Residue 17

Amelin
2007

and

chondrules (4)

Krot
model age

238U206Pb

5.369

Residue 21

Amelin
2007

and

chondrules (12)

Krot
model age

238U206Pb

5.956

Residue 28

Amelin
2007

and

chondrules (20)

Krot
model age

238U206Pb

4.879

Residue 22

Amelin
2007

and

chondrules (8)

Krot
model age

238U206Pb

5.462

Residue 23

Amelin
2007

and

chondrules (19)

Krot
model age

238U206Pb

5.345

Residue 24

Amelin
2007

and

chondrules (21)

Krot
model age

238U206Pb

5.336

Residue 25

Amelin
2007

and

chondrules (16)

Krot
model age

238U206Pb

6.945

Residue 26

Amelin
2007

and

chondrules (9)

Krot
model age

238U206Pb

5.604

Residue 27

Amelin
2007

and

chondrules (14)

Krot
model age

238U206Pb

4.525

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (12)

Krot
model age

238U206Pb

4.281

Acid leachate 21 HCl

Amelin
2007

and

chondrules (12)

Krot
model age

238U206Pb

4.95

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (8)

Krot
model age

238U206Pb

10.066

Acid leachate 21 HCl

Amelin
2007

and

chondrules (8)

Krot
model age

238U206Pb

4.151

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (19)

Krot
model age

238U206Pb

4.842

Acid leachate 21 HCl

Amelin
2007

and

chondrules (19)

Krot
model age

238U206Pb

4.399

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (21)

Krot
model age

238U206Pb

3.926

Acid leachate 21 HCl

Amelin
2007

and

chondrules (21)

Krot
model age

238U206Pb

4.092

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (16)

Krot
model age

238U206Pb

4.756

Acid leachate 21 HCl

Amelin
2007

and

chondrules (16)

Krot
model age

238U206Pb

3.855

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (9)

Krot
model age

238U206Pb

4.175

Acid leachate 21 HCl

Amelin
2007

and

chondrules (9)

Krot
model age

238U206Pb

4.125

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (14)

Krot
model age

238U206Pb

4.208

Acid leachate 21 HCl

Amelin
2007

and

chondrules (14)

Krot
model age

238U206Pb

1.799

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (20)

Krot
model age

238U206Pb

3.986

Acid leachate 21 HNO3

Amelin
2007

and

chondrules (20)

Krot
model age

235U207Pb

4.730

Chen
1976

and

chondrule

235U207Pb

4.615

Chen
1976

and

chondrule

235U207Pb

4.114

Chen
1976

and

chondrule

235U207Pb

4.519

Chen
1976

and

chondrule

4.78

Tatsumoto, Unruh,
and
Desborough
1976
model age

chondrule N1

235U207Pb
235

chondrule N1

207
235

chondrule N2

207
235

chondrule N2

207
235

chondrule N3

207
235

chondrule N3

207
235

chondrule N17

207
235

chondrule N17

207
235

chondrule N18

207
235

chondrule N18

207
235

chondrule N19

207
235

chondrule N19

207
235

chondrule N20

207
235

chondrule N20

207
235

chondrule N34

207
235

chondrule N40

207
235

chondrule N41

207

Tilton
model age
Tilton
model age
Tilton
model age
Tilton
model age

U
Pb

4.82

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.51

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.54

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.70

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.70

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

5.05

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

5.04

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.46

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.33

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.69

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.69

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.86

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.89

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

4.96

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

5.10

Tatsumoto, Unruh, and Desborough 1976

model age

U
Pb

5.04

Tatsumoto, Unruh, and Desborough 1976

model age

Table 3. Model ages for Ca-Al inclusions (CAIs) in the Allende CV3 carbonaceous chondrite meteorite, with the details and
literature sources.
Sample

Method

Age

Error +/

Notes

Source

Type

Decayers
inclusion

40

Ar39Ar

4.397

0.015

#15 white center

Dominik et al. 1978

total apparent
age

inclusion

40

Ar39Ar

4.232

0.009

#16 rim fragment

Dominik et al. 1978

total apparent
age

inclusion

40

Ar39Ar

4.517

0.025

#14 gray core

Dominik et al. 1978

plateau age

inclusion

40

Ar39Ar

4.529

0.018

#15 white pocket

Dominik et al. 1978

plateau age

inclusion

40

Ar Ar

4.476

0.023

#16 white rim

Dominik et al. 1978

plateau age

inclusion

40

Ar39Ar

4.98

0.06

Sample 17

Jessberger and Dominik


1979a, b
plateau age

inclusion

40

Ar39Ar

4.89

0.03

Sample 18

Jessberger and Dominik


1979a, b
plateau age

40

Ar39Ar

5.43

0.04

Sample 17

Jessberger and Dominiktotal


1979a
age

fusion

inclusion

40

Ar39Ar

5.37

0.10

Sample 18

Jessberger and Dominiktotal


1979a
age

fusion

inclusion

40

Ar39Ar

4.68

0.12

Sample 19

Jessberger and Dominiktotal


1979a
age

fusion

inclusion

40

Ar39Ar

5.27

0.18

Sample 20

Jessberger and Dominiktotal


1979a
age

fusion

inclusion

40

Ar39Ar

4.62

0.10

Sample 21

Jessberger and Dominiktotal


1979a
age

fusion

inclusion
inclusion

40

Ar39Ar

4.47

0.01

Sample 10, White fine

Jessberger et al. 1980

plateau age

inclusion

40

Ar39Ar

4.55

0.03

Sample 11, White fluffy

Jessberger et al. 1980

plateau age

inclusion

40

Ar Ar

4.50

0.02

Sample 13, White coarse

Jessberger et al. 1980

plateau age

inclusion

40

Ar39Ar

4.98

0.06

Sample 17, White coarse

Jessberger et al. 1980

plateau age

inclusion

40

Ar Ar

5.43

0.04

Sample 17, White coarse

Jessberger et al. 1980

plateau age

inclusion

40

Ar39Ar

4.92

0.03

Sample 18, White coarse

Jessberger et al. 1980

plateau age

inclusion

40

Ar39Ar

5.37

0.10

Sample 18, White coarse

Jessberger et al. 1980

plateau age

inclusion

40

Ar39Ar

4.68

0.12

Sample 19, White coarse

Jessberger et al. 1980

plateau age

inclusion

40

Ar39Ar

5.27

0.18

Sample 20, White coarse

Jessberger et al. 1980

plateau age

inclusion

40

4.62

0.10

Sample 21, White coarse

Jessberger et al. 1980

plateau age

inclusion

KAr

4.20

0.13

#14 gray core

Dominik et al. 1978

model age

inclusion

KAr

4.40

0.11

#15 white pocket

Dominik et al. 1978

model age

inclusion

KAr

4.23

0.10

#16 white rim

Dominik et al. 1978

model age

inclusion

KAr

5.12

0.03

Sample 17

Jessberger and Dominik


1979a, b
model age

inclusion

KAr

5.08

0.08

Sample 18

Jessberger and Dominik


1979a, b
model age

inclusion

KAr

4.92

0.11

Sample 19

Jessberger and Dominik


1979a
model age

inclusion

KAr

5.54

0.09

Sample 20

Jessberger and Dominik


1979a
model age

inclusion

KAr

5.26

0.10

Sample 21

Jessberger and Dominik


1979a
model age

inclusion

KAr

4.57

0.41

Sample 23

Jessberger and Dominik


1979a
model age

inclusion CaAl

KAr

5.11

0.2

Sample 3839 finegrained Herzog et al. 1980

total
fusion
model age

inclusion CaAl

KAr

5.09

0.1

Sample 3839 finegrained Herzog et al. 1980

total
fusion
model age

39

39

39

39

Ar Ar

inclusion CaAl

KAr

4.68

0.01

Sample 3839 finegrained Herzog et al. 1980

total
fusion
model age

inclusion CaAl

KAr

4.80

0.02

Sample 3839 finegrained Herzog et al. 1980

total
fusion
model age

inclusion CaAl

KAr

4.34

0.10

Sample 3839 finegrained Herzog et al. 1980

total
fusion
model age

inclusion CaAl

KAr

4.25

0.12

Sample 3839 finegrained Herzog et al. 1980

total
fusion
model age

inclusion

KAr

4.28

0.01

Sample 3915

Herzog et al. 1980

total
fusion
model age

inclusion

KAr

4.52

0.07

Sample 9, White fine

Jessberger et al. 1980

model age

inclusion

KAr

4.33

0.08

Sample 10, White fine

Jessberger et al. 1980

model age

inclusion

KAr

4.44

0.07

Sample 11, White fluffy

Jessberger et al. 1980

model age

inclusion

KAr

4.82

0.06

Sample 12, White comp. of


coarse
Jessberger et al. 1980

model age

inclusion

KAr

4.49

0.03

Sample 13, White coarse

Jessberger et al. 1980

model age

CaAl
aggregateslow
Rb
RbSr

4.18

0.15

D2

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregateslow
Rb
RbSr

3.93

0.26

A8

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregateslow
Rb
RbSr

3.97

0.05

A4

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregateslow
Rb
RbSr

4.21

0.07

B30 a

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregateslow
Rb
RbSr

4.39

0.06

B30 b

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.58

0.02

D3

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.29

0.02

B29 a

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.55

0.02

B29 b

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.63

0.03

B29 c

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

3.69

0.02

B29 d

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

3.65

0.02

B29 e

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.35

0.02

B32

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.3

0.02

B31 s1

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.27

0.02

B31 s2

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.32

0.02

B31 s3

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.56

0.03

B31 m1

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.36

0.02

A10

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

3.59

0.02

B33 s1

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

3.59

0.02

B33 s2

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.49

0.03

A1

Gray, Papanastassiou,
and Wasserburg 1973 model age

CaAl
aggregates
high Rb

RbSr

4.16

0.02

A9

Gray, Papanastassiou,
and Wasserburg 1973 model age

Decayers
inclusion

207

Pb206Pb 4.555

White

Chen and Tilton 1976

model age

inclusion

207

Pb206Pb 4.556

Pink

Chen and Tilton 1976

model age

inclusion

207

Pb206Pb 4.564

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.561

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.544

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.557

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.498

Finegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.490

Finegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.565

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.561

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.408

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.422

Finegrained

Chen and Wasserburg


1981
model age

inclusion

207

Pb206Pb 4.471

Finegrained

Chen and Wasserburg


1981
model age

CaAl
G2/6

inclusion
4.5657

0.0026

Manhes, Gopel,
Allgre 1988

and

PbPb

CaAl inclusion
Br9/6
PbPb

0.0031

Manhes, Gopel,
Allgre 1988

and

4.5683

CaAl inclusion
48C/7
PbPb

0.0007

Manhes, Gopel,
Allgre 1988

and

4.566

CaAl
A1/7

0.0009

Manhes, Gopel,
Allgre 1988

and

4.5651

inclusion
PbPb

model age
model age
model age
model age

207

Pb206Pb 4.5657

0.0026

Allgre, Manhes,
Gopel 1995

and

inclusion
inclusion

207

Pb206Pb 4.5683

0.0031

Allgre,

andmodel age

Manhes,

model age

Gopel 1995
207

Pb206Pb 4.5660

0.0007

Allgre, Manhes,
Gopel 1995

and

inclusion

207

Pb206Pb 4.5651

0.0009

Allgre, Manhes,
Gopel 1995

and

inclusion

model age
model age

CAI F2, Type


B2, >3.15
PbPb

4.567

0.0007

Assuming primordial Pb

Amelin et al. 2002

model age

CAI TS32, Type


CTA, >3.15
PbPb

4.5669

0.0008

Assuming primordial Pb

Amelin et al. 2002

model age

CAI TS33, Type


B1, >3.15
PbPb

4.5676

0.0009

Assuming primordial Pb

Amelin et al. 2002

model age

CAI F2, Type


B2, 2.853.15
PbPb

4.5663

0.0008

Assuming primordial Pb

Amelin et al. 2002

model age

CAI TS32, Type


CTA, 2.853.15 PbPb

4.5678

0.0008

Assuming primordial Pb

Amelin et al. 2002

model age

CAI TS33, Type


B1, 2.853.15
PbPb

4.5667

0.0007

Assuming primordial Pb

Amelin et al. 2002

model age

CAI F2, Type


B2, >3.15
PbPb

4.5675

0.001

Assuming common Pb

Amelin et al. 2002

model age

CAI TS32, Type


CTA, >3.15
PbPb

4.5679

0.0017

Assuming common Pb

Amelin et al. 2002

model age

CAI TS33, Type


B1, >3.15
PbPb

4.5686

0.0018

Assuming common Pb

Amelin et al. 2002

model age

CAI F2, Type


B2, 2.853.15
PbPb

4.5671

0.0015

Assuming common Pb

Amelin et al. 2002

model age

CAI TS32, Type


CTA, 2.853.15 PbPb

4.5682

0.001

Assuming common Pb

Amelin et al. 2002

model age

CAI TS33, Type


B1, 2.853.15
PbPb

4.5678

0.0019

Assuming common Pb

Amelin et al. 2002

model age

model age

207

Pb206Pb 4.5679

0.0005

Mean of six ages on three


inclusions
with
Canyon
Diablo troilite
Amelin et al. 2002

inclusions

207

Pb206Pb 4.5336

0.0011

F2 plagioclase rich

Connelly et al. 2008

model age

inclusions

207

Pb206Pb 4.5664

0.0008

F2 melilite rich

Connelly et al. 2008

model age

inclusions

207

Pb206Pb 4.5664

0.0008

F2 pyroxene rich

Connelly et al. 2008

model age

inclusions

207

Pb206Pb 4.5663

0.0016

TS32 plagioclase rich

Connelly et al. 2008

model age

inclusions

207

Pb Pb 4.5678

0.0008

TS32 melilite rich

Connelly et al. 2008

model age

inclusions

207

Pb206Pb 4.5669

0.0008

TS32 pyroxene rich

Connelly et al. 2008

model age

inclusions

207

Pb Pb 4.5580

0.0020

TS33 plagioclase rich

Connelly et al. 2008

model age

inclusions

207

Pb206Pb 4.5668

0.0007

TS33 melilite rich

Connelly et al. 2008

model age

inclusions

207

Pb206Pb 4.5676

0.0009

TS33 pyroxene rich

Connelly et al. 2008

model age

207

Pb206Pb 4.56759

0.0001

Acidleached residues (3) +Bouvier, Vervoort, and


leachate
Patchett 2008
model age

207

Pb206Pb 4.5668

0.0004

AJEF Pyroxene

Jacobsen et al. 2008

model age

207

Pb206Pb 4566.6

0.0007

AJEF Metilite coarse

Jacobsen et al. 2008

model age

207

Pb206Pb 4.5677

0.0006

AJEF 1.8 A m
mediumcoarse

Jacobsen et al. 2008

model age

207

Pb206Pb 4.5696

0.0009

AJEF Pyroxene first wash

Jacobsen et al. 2008

model age

207

Pb206Pb 4.5659

0.001

AJEF
wash

Jacobsen et al. 2008

model age

inclusions (Ca
Al)

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

206

206

Pyroxene

nm

second

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)
inclusions (Ca
Al)

207

Pb206Pb 4.5667

0.0004

A43 Pyroxene from 0.41.1


A
Jacobsen et al. 2008

model age

207

Pb206Pb 4.5682

0.0039

A43 0.41.1 A mag med


coarse
Jacobsen et al. 2008

model age

207

Pb206Pb 4.5656

0.0027

A43 0.41.1 A mag med


fine
Jacobsen et al. 2008

model age

207

Pb206Pb 4.5659

0.0014

A43 0.41.1 A mag

model age

207

Pb206Pb 4.568

0.003

MNHNA
residue
Canyon Diablo troilite

withBouvier, Vervoort, and


Patchett 2008
model age

Pb206Pb 4.5715

0.0005

MNHNA
leachate 1

residue

207

withBouvier, Vervoort, and


Patchett 2008
model age

Pb206Pb 4.5482

0.0005

MNHNA
leachate 2

residue

207

withBouvier, Vervoort, and


Patchett 2008
model age

207

Pb206Pb 4.5222

0.003

MNHNB
residue
Canyon Diablo troilite

withBouvier, Vervoort, and


Patchett 2008
model age

Pb206Pb 4.5188

0.029

MNHNB
leachate 1

residue

207

withBouvier, Vervoort, and


Patchett 2008
model age

Pb206Pb 4.4914

0.0046

MNHNB
leachate 2

residue

207

withBouvier, Vervoort, and


Patchett 2008
model age

207

Pb206Pb 4.5657

0.0023

UCLAA
residue
Canyon Diablo troilite

withBouvier, Vervoort, and


Patchett 2008
model age

Pb206Pb 4.566

0.0023

UCLAA
leachate 1

residue

207

withBouvier, Vervoort, and


Patchett 2008
model age

Pb206Pb 4.5603

0.0033

UCLAA
leachate 2

residue

207

withBouvier, Vervoort, and


Patchett 2008
model age

207

Pb206Pb 4.5655

0.0023

UCLAB residue withBouvier, Vervoort, and


Canyon Diablo troilite Patchett 2008
model age

207

Pb206Pb 4.5659

0.0023

UCLAB residue withBouvier, Vervoort, and


leachate 1
Patchett 2008
model age

207

Pb206Pb 4.57

0.0036

UCLAB residue withBouvier, Vervoort, and


leachate 2
Patchett 2008
model age

207

Pb206Pb 4.5646

0.0029

UCLAC residue withBouvier, Vervoort, and


Canyon Diablo troilite Patchett 2008
model age

207

Pb206Pb 4.5653

0.003

UCLAC residue withBouvier, Vervoort, and


leachate 1
Patchett 2008
model age

207

Pb206Pb 4.5685

0.0042

UCLAC residue withBouvier, Vervoort, and


leachate 2
Patchett 2008
model age

Jacobsen et al. 2008

inclusions

PbPb

4.56857

0.00087

Mean of A004 third


washes
Amelin et al. 2010

model age

inclusions

PbPb

4.56702

0.00022

Mean of A003 and


A005 residues
Amelin et al. 2010

model age

inclusions

PbPb

4.56664

0.00033

Mean
of
residues

Amelin et al. 2010

model age

inclusion

232

White

Chen and Tilton 1976

model age

inclusion

232

Th Pb 4.063

Pink

Chen and Tilton 1976

model age

inclusion

238

U206Pb 4.023

White

Chen and Tilton 1976

model age

inclusion

238

U206Pb 4.644

Pink

Chen and Tilton 1976

model age

inclusion

238

U206Pb 4.73

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

238

U206Pb 5.31

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

238

U206Pb 4.83

Coarsegrained

Chen and Wasserburg


1981
model age

Th208Pb 3.773
208

A004

inclusion

238

U206Pb 13.0

Finegrained

Chen and Wasserburg


1981
model age

inclusion

238

U206Pb 5.0

Finegrained

Chen and Wasserburg


1981
model age

inclusion

238

U206Pb 4.56

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

238

U206Pb 4.59

Coarsegrained

Chen and Wasserburg


1981
model age

inclusion

238

U206Pb 16.7

Finegrained

Chen and Wasserburg


1981
model age

238

U206Pb 4.301

AJEF Pyroxene first


wash
Jacobsen et al. 2008

model age

238

U206Pb 4.34

AJEF
Pyroxene
second wash
Jacobsen et al. 2008

model age

238

U206Pb 4.761

AJEF 1.8 A m + nm
mediumcoarse
Jacobsen et al. 2008

model age

238

U206Pb 4.827

A43 Pyroxene from


0.41.1 A
Jacobsen et al. 2008

model age

238

U206Pb 4.741

A43 0.41.1 A mag


medcoarse
Jacobsen et al. 2008

model age

238

U206Pb 4.663

A43 0.41.1 A mag


medfine
Jacobsen et al. 2008

model age

238

U206Pb 4.858

A43 0.41.1 A mag

Jacobsen et al. 2008

model age

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca

inclusion
Al)

(Ca
207

inclusion

235U Pb4.385

White

Chen and Tilton 1976

model age

inclusion

235U207Pb4.583

Pink

Chen and Tilton 1976

model age

CAI F2, Type


B2, >3.15
UPb

4.564

0.029

Amelin et al. 2002

model age

CAI TS32, Type


CTA, >3.15
UPb

4.604

0.01

Amelin et al. 2002

model age

CAI TS33, Type


B1, >3.15
UPb

4.67

0.016

Amelin et al. 2002

model age

CAI F2, Type


B2, 2.853.15
UPb

4.762

0.038

Amelin et al. 2002

model age

CAI TS32, Type


CTA, 2.853.15 UPb

4.573

0.009

Amelin et al. 2002

model age

CAI TS33, Type


B1, 2.853.15
UPb

4.716

0.02

Amelin et al. 2002

model age

inclusions (Ca
Al)
UPb
inclusions (Ca
Al)
UPb

4.5666

4.5675

concordia
intercept)

age

(upper

0.0013

CAIs AJEF & A43


residues & px second
washes
Jacobsen et al. 2008

concordia
intercept)

age

(upper

0.0012

CAIs AJEF & A43


residues & px second
washes
Jacobsen et al. 2008

0.0001

Relative
to
Shallowater absolute
IXe age
Pravdivtseva et al. 2003 model age

0.0002

Relative
to
Shallowater absolute
IXe age
Pravdivtseva et al. 2003 model age

0.0002

Relative
to
Shallowater absolute
IXe age
Pravdivtseva et al. 2003 model age

0.0003

Relative
to
Shallowater absolute
IXe age
Pravdivtseva et al. 2003 model age

Calibrated by PbPb Ages

dark inclusions

dark inclusions

dark inclusions

dark inclusions

IXe

IXe

IXe

IXe

4.5645

4.5649

4.5641

4.5652

inclusions (Ca
Al)
IXe
inclusions (Ca
Al)
IXe
inclusions (Ca
Al)
IXe

4.5629

4.563

4.5623

0.0002

Relative
to
Shallowater absolute
IXe age
Pravdivtseva et al. 2003 model age

0.0002

Relative
to
Shallowater absolute
IXe age
Pravdivtseva et al. 2003 model age

0.0002

Relative
to
Shallowater absolute
IXe age
Pravdivtseva et al. 2003 model age

Table 4. Model ages for whole-rock, matrix and other samples of the Allende CV3 carbonaceous chondrite meteorite, with
the details and literature sources.
Sample

Method

Age

Error +/

Notes

Source

Type

Decayers
whole rock

40

whole rock

Ar39Ar

4.57

0.03

Sample 2

Jessberger et al. 1980

plateau age

KAr

4.43

0.10

Finegrained

Dominik and Jessberger 1979 model age

matrix

KAr

3.34

0.02

Sample 3915

Herzog et al. 1980

total fusion
model age

matrix

KAr

3.58

0.02

Sample 3915

Herzog et al. 1980

total fusion
model age

matrix

KAr

3.43

0.04

Sample 3915

Herzog et al. 1980

total fusion
model age

matrix

KAr

3.63

0.05

Sample 3915

Herzog et al. 1980

total fusion
model age

matrix

KAr

3.99

0.01

Sample 3915

Herzog et al. 1980

total fusion
model age

granular
material

KAr

6.29

0.01

Sample 3915

Herzog et al. 1980

total fusion
model age

granular
material

KAr

4.11

0.02

Sample 3915

Herzog et al. 1980

total fusion
model age

anorthite

KAr

14.2

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

anorthite

KAr

8.5

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

anorthite

KAr

6.2

1.1

Sample 3529Z

Herzog et al. 1980

total fusion
model age

anorthite

KAr

14.5

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

pyroxene

KAr

8.5

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

pyroxene

KAr

5.3

0.9

Sample 3529Z

Herzog et al. 1980

total fusion
model age

pyroxene

KAr

8.2

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

melilite

KAr

6.0

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

melilite

KAr

5.4

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

finegrained
material

KAr

8.8

1.3

Sample 3529Z

Herzog et al. 1980

total fusion
model age

finegrained
material

KAr

6.8

0.9

Sample 3529Z

Herzog et al. 1980

total fusion
model age

finegrained
material

KAr

7.6

0.9

Sample 3529Z

Herzog et al. 1980

total fusion
model age

vein

KAr

7.97

0.6

Sample 3655A

Herzog et al. 1980

total fusion
model age

vein

KAr

5.10

0.17

Sample 3655A

Herzog et al. 1980

total

fusion

model age
vein

KAr

4.30

0.03

Sample 3655A

Herzog et al. 1980

total fusion
model age

vein

KAr

4.86

0.04

Sample 3655A

Herzog et al. 1980

total fusion
model age

vein

KAr

7.12

0.9

Sample 3655A

Herzog et al. 1980

total fusion
model age

spinel

KAr

6.46

0.6

Sample 3655A

Herzog et al. 1980

total fusion
model age

spinel

KAr

7.33

1.2

Sample 3655A

Herzog et al. 1980

total fusion
model age

spinel

KAr

6.13

0.46

Sample 3655A

Herzog et al. 1980

total fusion
model age

spinel

KAr

9.37

Sample 3655A

Herzog et al. 1980

total fusion
model age

melilite

KAr

10.90

1.3

Sample 3655A

Herzog et al. 1980

total fusion
model age

melilite

KAr

16.24

1.3

Sample 3655A

Herzog et al. 1980

total fusion
model age

melilite

KAr

10.07

1.34

Sample 3655A

Herzog et al. 1980

total fusion
model age

melilite

KAr

14.80

1.34

Sample 3655A

Herzog et al. 1980

total fusion
model age

pyroxene

KAr

13.50

1.3

Sample 3655A

Herzog et al. 1980

total fusion
model age

pyroxene

KAr

9.3

1.3

Sample 3655A

Herzog et al. 1980

total fusion
model age

pyroxene

KAr

8.87

1.34

Sample 3655A

Herzog et al. 1980

total fusion
model age

rim

KAr

4.09

0.02

Sample 3655A

Herzog et al. 1980

total fusion
model age

rim

KAr

4.01

0.02

Sample 3655A

Herzog et al. 1980

total fusion
model age

rim

KAr

4.18

0.01

Sample 3655A

Herzog et al. 1980

total fusion
model age

inner part of
vein
KAr

16.83

1.34

Sample 3655A

Herzog et al. 1980

total fusion
model age

finegrained
material

KAr

9.18

1.32

Sample 3655A

Herzog et al. 1980

total fusion
model age

matrix

KAr

3.80

0.09

Sample 1

Jessberger et al. 1980

model age

whole rock

KAr

4.43

0.09

Sample 2

Jessberger et al. 1980

model age

4.52

0.04

A6 a

Gray, Papanastassiou,
Wasserburg 1973

and

total meteorite RbSr

4.55

0.04

A6 b

Gray, Papanastassiou,
Wasserburg 1973

and

total meteorite RbSr

4.56

0.03

A6 s1

Gray, Papanastassiou,
Wasserburg 1973

and

total meteorite RbSr

4.59

0.03

A6 s2

Gray, Papanastassiou,
Wasserburg 1973

and

total meteorite RbSr

4.56

0.04

A6 s3

Gray, Papanastassiou,
Wasserburg 1973

and

total meteorite RbSr


RbSr

4.48

Tatsumoto,
Unruh,
Desborough1976

and

whole rock

RbSr

4.49

Tatsumoto,
Unruh,
Desborough1976

and

whole rock

model age
model age
model age
model age
model age
model age
model age

RbSr

4.84

Tatsumoto,
Unruh,
Desborough1976

and

matrix

RbSr

4.60

Tatsumoto,
Unruh,
Desborough1976

and

matrix

RbSr

3.33

White

Tatsumoto,
Unruh,
Desborough1976

and

aggregate

RbSr

4.41

Pinkishwhite

Tatsumoto,
Unruh,
Desborough1976

and

aggregate
whole rock

RbSr

4.56

WR1

Shimoda et al. 2005

model age

whole rock

RbSr

4.58

WR2

Shimoda et al. 2005

model age

model age

model age
model age
model age
model age

Decayers
207

whole rock

206
207

whole rock

206
207

whole rock

206
207

whole rock

206
207

aggregate

206
207

whole rock

206
207

matrix

206
207

matrix

206
207

matrix

206
207

matrix

206
207

matrix

206
207

magnetic

206
207

magnetic

206
207

magnetic

206
207

aggregate

206
207

aggregate

206
207

aggregate

206
207

aggregate

206
207

matrix

206
207

matrix

206
207

A003 1W1

206

A003 1W1 +
extra HBr

207
206

Pb
Pb

4.528

0.04

Huey and Kohman 1973

Pb
Pb

4.496

0.01

Tatsumoto, Knight, and Allgre


1973
model age

Pb
Pb

4.51

Tilton 1973

model age

Pb
Pb

4.52

Tilton 1973

model age

Pb
Pb

4.562

Chen and Tilton 1976

model age

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.496

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.494

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.481

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.489

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.506

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.524

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.467

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.492

Pb
Pb

Tatsumoto,
Unruh,
Desborough1976

and

4.483

Pb
Pb

White

Tatsumoto,
Unruh,
Desborough1976

and

4.557

Pb
Pb

White

Tatsumoto,
Unruh,
Desborough1976

and

4.562

Pb
Pb

Pinkish

Tatsumoto,
Unruh,
Desborough1976

and

4.547

Pb
Pb

Pinkish

Tatsumoto,
Unruh,
Desborough1976

and

4.533

Pb
Pb

C3 same sample

Tatsumoto,
Unruh,
Desborough1976

and

4.49

Pb
Pb

C3 same sample

Tatsumoto,
Unruh,
Desborough1976

and

4.52

Pb
Pb

4.54569

0.0027

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.54968

0.00186

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

White

model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age

207

A004 2W1

206
207

A004 3W1

206
207

A004 4W1

206
207

A004 5W1

206
207

A005 6W1

206
207

A005 7W1

206
207

A005 8W1

206
207

A005 9W1

206
207

A003 1W2

206

A003 1W2 +
extra HBr

207
206
207

A004 2W2

206
207

A004 4W2

206
207

A004 5W2

206
207

A005 6W2

206
207

A005 7W2

206
207

A005 8W2

206
207

A005 9W2

206
207

A004 2W3

206
207

A004 3W3

206
207

A004 4W3

206
207

A004 5W3

206

A003
1R
Mediumfine
whole rock

207

A003
1R+
(aliquot 2 +
extra HBr)

207

A004
2R
Coarse
pyroxenerich

207

A004
3R
Coarse
melilite
and
feldsparrich
A004
4R
Coarse whole

206

206

206

207
206
207
206

Pb
Pb

4.55967

0.00861

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.55548

0.00507

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.54045

0.00276

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.53429

0.00209

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.55282

0.00555

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.55808

0.00235

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.53446

0.00244

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.55494

0.00164

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56563

0.00143

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56748

0.00159

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56847

0.00114

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56926

0.00187

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56756

0.00101

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.50331

0.00358

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56954

0.00137

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56872

0.00206

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.57021

0.00101

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.57036

0.00175

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56837

0.00033

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56905

0.00047

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56758

0.00103

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56679

0.00025

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56697

0.00024

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56697

0.00029

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56643

0.00017

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

4.5667

0.00018

Pb
Pb

Assuming primordial isotopicAmelin et al. 2010


composition of initial Pb

model age

rock
A004
5R
Mediumfine
whole rock

207

A005
6R
Mediumfine
dark

207

A005
7R
Mediumfine
light

207

A005
8R
whole rock

206

206

206

Pb
Pb

4.56671

0.00023

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56781

0.00044

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Pb
Pb

4.56712

0.0003

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010

model age

Fine

207
206

A005 9R Very fine


whole rock

207

A004
2R+W3
recombined

207

A004
3R+W3
recombined

207

A004
4R+W3
recombined

207

A004
5R+W3
recombined

207

A003
Frac.1
recombined

207

A003
Frac.1+
recombined

207

A004
Frac.2
recombined

207

A004
Frac.3
recombined

207

A004
Frac.4
recombined

207

A004
Frac.5
recombined

207

A005
Frac.6
recombined

207

A005
Frac.7
recombined

207

A005
Frac.8
recombined

207

A005
Frac.9
recombined

207

206

206

206

206

206

206

206

206

206

206

206

206

206

206

206
207

A004 3W1

206
207

A004 5W1

206
207

A005 7W1

206
207

A005 9W1

206
207

A003 1W2

206

A003 1W2 + extra


HBr

207
206
207

A004 2W2

206

Pb
Pb

4.56714

0.00028

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56725

0.00033

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56787

0.00068

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56701

0.00021

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.5673

0.00025

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56699

0.00049

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56314

0.00077

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56413

0.00064

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56717

0.0015

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56595

0.00066

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56453

0.00077

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56047

0.00088

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.5297

0.00287

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56594

0.00079

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56079

0.00094

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56455

0.00077

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.75219

0.00657

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

6.14348

0.07377

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.80634

0.00291

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.87254

0.00258

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.65712

0.00142

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.65956

0.00195

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.61145

0.00096

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

207
206

A004 4W2

207
206

A004 5W2

207
206

A005 6W2

207
206

A005 7W2

207
206

A005 8W2

207
206

A005 9W2

207
206

A004 2W3

207
206

A004 3W3

207
206

A004 4W3

207

A004 5W3

206

A003 1R Medium
fine whole rock

207

A003 1R+ (aliquot


2 + extra HBr)

207

A004 2R Coarse
pyroxenerich

207

A004 3R Coarse
melilite
and
feldsparrich

206

206

206

207
206

A004 4R Coarse
whole rock

207

A004 5R Medium
fine whole rock

207

A005 6R Medium
fine dark

207

A005 7R Medium
fine light

207

A005
8R
whole rock

207

Fine

206

206

206

206

206

A005 9R Very fine


whole rock

207

A004
2R+W3
recombined

207

A004
3R+W3
recombined

207

A004
4R+W3
recombined

207

A004
5R+W3
recombined

207

A003
Frac.1
recombined

207

A003
Frac.1+
recombined

207

A004
Frac.2
recombined

207

A004

207

Frac.3

206

206

206

206

206

206

206

206

Pb
Pb

4.69912

0.00217

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.59785

0.00082

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

5.41676

0.04032

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.61215

0.00121

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.59428

0.00184

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.59091

0.00073

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.57146

0.00155

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56929

0.0003

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.57121

0.00044

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56825

0.00101

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56755

0.00022

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.5677

0.00021

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56751

0.00022

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56669

0.0001

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56722

0.00014

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56708

0.00013

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56822

0.00019

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56733

0.00021

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56733

0.00013

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56754

0.0002

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56856

0.00058

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56747

0.00016

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56823

0.00022

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.56746

0.00041

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

3.81054

0.0003

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

3.81241

0.00034

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb
Pb

4.13562

0.00058

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Pb

4.5844

0.00075

Assuming primordial isotopicAmelin et al. 2010 model age

recombined

206

Pb

A004
Frac.4
recombined

207

Pb
Pb

4.0668

0.00048

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

A004
Frac.5
recombined

207

Pb
Pb

4.8893

0.01526

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

A005
Frac.6
recombined

207

Pb
Pb

4.44588

0.02266

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

A005
Frac.7
recombined

207

Pb
Pb

4.61371

0.0008

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

A005
Frac.8
recombined

207

Pb
Pb

3.65461

0.00033

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

A005
Frac.9
recombined

207

Pb
Pb

4.64884

0.00089

Assuming primordial isotopic


composition of initial Pb
Amelin et al. 2010 model age

Th
Pb

4.391

206

206

206

206

206

206
232

aggregate

208

composition of initial Pb

Chen
1976

White

whole rock

10.40

9.86

Tatsumoto, Unruh,
and Desborough
1976
model age

16.49

Tatsumoto, Unruh,
and Desborough
1976
model age

9.15

Tatsumoto, Unruh,
and Desborough
1976
model age

232

matrix

Th
208
Pb
232

matrix

Th
208
Pb
232

matrix

Th
208
Pb
232

magnetic

208
232

magnetic

208
232

aggregate

208
232

model age

Tatsumoto, Unruh,
and Desborough
1976
model age

232

Th
208
Pb

and Tilton

Th
Pb

Tatsumoto,
Unruh,
Desborough 1976

and

4.81

Th
Pb

Tatsumoto,
Unruh,
Desborough 1976

and

10.1

Th
Pb

White

Tatsumoto,
Unruh,
Desborough 1976

and

5.15

Th
Pb

Pinkish

Tatsumoto,
Unruh,
Desborough 1976

and

5.44

White

Chen and Tilton 1976

model age
model age
model age

aggregate

208

aggregate

238

U206Pb4.349

238

U206Pb8.82

Tatsumoto,
Unruh,
Desborough 1976

and

whole rock

238

U206Pb6.44

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

238

U206Pb6.33

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

238

U206Pb6.42

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

238

U206Pb7.80

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

238

U206Pb5.45

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

238

U206Pb4.41

Tatsumoto,
Unruh,
Desborough 1976

and

magnetic

238

U206Pb7.75

Tatsumoto,
Unruh,
Desborough 1976

and

magnetic

238

U206Pb7.82

Tatsumoto,
Unruh,
Desborough 1976

and

magnetic

238

U206Pb4.96

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate

White

model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age

238

U206Pb4.92

White

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate

238

U206Pb5.73

Pinkish

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate

238

U206Pb5.50

Pinkish

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate
A003 1W1

238

U206Pb4.2287

A003 1W1 + extra HBr

238

A004 2W1

model age
model age
model age

Amelin et al. 2010

model age

U Pb4.2333

Amelin et al. 2010

model age

238

U206Pb5.0618

Amelin et al. 2010

model age

A004 3W1

238

U206Pb3.5471

Amelin et al. 2010

model age

A004 4W1

238

U206Pb4.1125

Amelin et al. 2010

model age

A004 5W1

238

U206Pb4.6855

Amelin et al. 2010

model age

A005 6W1

238

U Pb4.1773

Amelin et al. 2010

model age

A005 7W1

238

U206Pb4.029

Amelin et al. 2010

model age

A005 8W1

238

U Pb4.4346

Amelin et al. 2010

model age

A005 9W1

238

U206Pb4.1142

Amelin et al. 2010

model age

A003 1W2

238

U206Pb3.634

Amelin et al. 2010

model age

A003 1W2 + extra HBr

238

U206Pb3.6411

Amelin et al. 2010

model age

A004 2W2

238

U206Pb3.7778

Amelin et al. 2010

model age

A004 4W2

238

U Pb3.8166

Amelin et al. 2010

model age

A004 5W2

238

U206Pb3.9313

Amelin et al. 2010

model age

A005 6W2

238

U Pb11.6247

Amelin et al. 2010

model age

A005 7W2

238

U206Pb4.1284

Amelin et al. 2010

model age

A005 8W2

238

U206Pb3.8734

Amelin et al. 2010

model age

A005 9W2

238

U206Pb4.0317

Amelin et al. 2010

model age

A004 2W3

238

U206Pb4.3613

Amelin et al. 2010

model age

A004 3W3

238

U Pb4.3663

Amelin et al. 2010

model age

A004 4W3

238

U206Pb4.316

Amelin et al. 2010

model age

A004 5W3

238

U Pb4.3897

Amelin et al. 2010

model age

238

U206Pb4.772

Amelin et al. 2010

model age

238

U206Pb4.7727

Amelin et al. 2010

model age

238

U206Pb4.9657

Amelin et al. 2010

model age

238

U206Pb5.0338

Amelin et al. 2010

model age

238

U206Pb4.963

Amelin et al. 2010

model age

238

U206Pb5.0656

Amelin et al. 2010

model age

A005 6R Mediumfine dark

238

U206Pb4.7688

Amelin et al. 2010

model age

A005 7R Mediumfine light

238

U206Pb4.7881

Amelin et al. 2010

model age

A005 8R Fine whole rock

238

U206Pb4.8041

Amelin et al. 2010

model age

238

U206Pb4.8472

Amelin et al. 2010

model age

A004 2R+W3 recombined

238

U206Pb4.8047

Amelin et al. 2010

model age

A004 3R+W3 recombined

238

Amelin et al. 2010

model age

A003 1R
whole rock

206

206

206

206

206

Coarse

A004 3R Coarse melilite


and feldsparrich
A004 4R Coarse whole
rock
A004 5R
whole rock

206

Mediumfine

A003 1R+ (aliquot 2 + extra


HBr)
A004
2R
pyroxenerich

206

Mediumfine

A005 9R Very fine whole


rock

206

U Pb4.834

A004 4R+W3 recombined

238

U206Pb4.7986

Amelin et al. 2010

model age

A004 5R+W3 recombined

238

U206Pb4.8475

Amelin et al. 2010

model age

A003 Frac.1 recombined

238

U Pb4.5705

Amelin et al. 2010

model age

A003 Frac.1+ recombined

238

U206Pb4.5727

Amelin et al. 2010

model age

A004 Frac.2 recombined

238

U206Pb4.6835

Amelin et al. 2010

model age

A004 Frac.3 recombined

238

U Pb4.716

Amelin et al. 2010

model age

A004 Frac.4 recombined

238

U206Pb4.5766

Amelin et al. 2010

model age

A004 Frac.5 recombined

238

U Pb4.6935

Amelin et al. 2010

model age

A005 Frac.6 recombined

238

U206Pb8.5341

Amelin et al. 2010

model age

A005 Frac.7 recombined

238

U206Pb4.5683

Amelin et al. 2010

model age

A005 Frac.8 recombined

238

U206Pb4.6101

Amelin et al. 2010

model age

A005 Frac.9 recombined

238

U206Pb4.5388

Amelin et al. 2010

model age

aggregate

235

U Pb4.496

Chen and Tilton 1976

model age

235

U207Pb5.57

Tatsumoto,
Unruh,
Desborough 1976

and

whole rock

235

U207Pb5.03

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

235U207Pb

4.99

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

235U207Pb

5.02

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

235U207Pb

5.36

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

235U207Pb

4.79

Tatsumoto,
Unruh,
Desborough 1976

and

matrix

235U207Pb

4.45

Tatsumoto,
Unruh,
Desborough 1976

and

magnetic

235U207Pb

5.34

Tatsumoto,
Unruh,
Desborough 1976

and

magnetic

235U207Pb

5.34

Tatsumoto,
Unruh,
Desborough 1976

and

magnetic

235U207Pb

4.68

White

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate

235U207Pb

4.67

White

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate

235U207Pb

4.88

Pinkish

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate

235U207Pb

4.81

Pinkish

Tatsumoto,
Unruh,
Desborough 1976

and

aggregate
A003 1W1

235U207Pb

4.446

Amelin et al. 2010

model age

A003 1W1 + extra HBr

235U207Pb

4.450

Amelin et al. 2010

model age

A004 2W1

235U207Pb

4.709

Amelin et al. 2010

model age

A004 3W1

235U207Pb

4.218

Amelin et al. 2010

model age

A004 4W1

235U207Pb

4.404

Amelin et al. 2010

model age

A004 5W1

235U207Pb

4.580

Amelin et al. 2010

model age

A005 6W1

235U207Pb

4.434

Amelin et al. 2010

model age

A005 7W1

235U207Pb

4.389

Amelin et al. 2010

model age

A005 8W1

235U207Pb

4.503

Amelin et al. 2010

model age

206

206

206

207

White

235

A005 9W1

207

U
Pb

4.415

Amelin
2010

model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age

et

al.
model age

235

A003 1W2

207
235

A003 1W2 + extra HBr

207
235

A004 2W2

207
235

A004 4W2

207
235

A004 5W2

207
235

A005 6W2

207
235

A005 7W2

207
235

A005 8W2

207
235

A005 9W2

207
235

A004 2W3

207
235

A004 3W3

207
235

A004 4W3

207
235

A004 5W3

207
235

A003 1R Mediumfine whole rock

207
235

A003 1R+ (aliquot 2 + extra HBr)

207
235

A004 2R Coarse pyroxenerich

207
235

A004 3R Coarse melilite and feldsparrich

207
235

A004 4R Coarse whole rock

207
235

A004 5R Mediumfine whole rock

207
235

A005 6R Mediumfine dark

207
235

A005 7R Mediumfine light

207
235

A005 8R Fine whole rock

207
235

A005 9R Very fine whole rock

207
235

A004 2R+W3 recombined

207
235

A004 3R+W3 recombined

207
235

A004 4R+W3 recombined

207
235

A004 5R+W3 recombined

207
235

A003 Frac.1 recombined

207

U
Pb

Amelin
2010

et

4.257

U
Pb

Amelin
2010

et

4.261

U
Pb

Amelin
2010

et

4.310

U
Pb

Amelin
2010

et

4.324

U
Pb

Amelin
2010

et

4.362

U
Pb

Amelin
2010

et

6.131

U
Pb

Amelin
2010

et

4.430

U
Pb

Amelin
2010

et

4.343

U
Pb

Amelin
2010

et

4.398

U
Pb

Amelin
2010

et

4.505

U
Pb

Amelin
2010

et

4.505

U
Pb

Amelin
2010

et

4.490

U
Pb

Amelin
2010

et

4.512

U
Pb

Amelin
2010

et

4.629

U
Pb

Amelin
2010

et

4.629

U
Pb

Amelin
2010

et

4.686

U
Pb

Amelin
2010

et

4.705

U
Pb

Amelin
2010

et

4.685

U
Pb

Amelin
2010

et

4.715

U
Pb

Amelin
2010

et

4.629

U
Pb

Amelin
2010

et

4.634

U
Pb

Amelin
2010

et

4.639

U
Pb

Amelin
2010

et

4.651

U
Pb

Amelin
2010

et

4.638

U
Pb

Amelin
2010

et

4.645

U
Pb

Amelin
2010

et

4.635

U
Pb

Amelin
2010

et

4.649

U
Pb

Amelin
2010

et

4.562

al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age

235

A003 Frac.1+ recombined

207
235

A004 Frac.2 recombined

207
235

A004 Frac.3 recombined

207
235

A004 Frac.4 recombined

207
235

A004 Frac.5 recombined

207
235

A005 Frac.6 recombined

207
235

A005 Frac.7 recombined

207
235

A005 Frac.8 recombined

207
235

A005 Frac.9 recombined

207

U
Pb

Amelin
2010

et

4.563

U
Pb

Amelin
2010

et

4.598

U
Pb

Amelin
2010

et

4.606

U
Pb

Amelin
2010

et

4.563

U
Pb

Amelin
2010

et

4.597

U
Pb

Amelin
2010

et

5.445

U
Pb

Amelin
2010

et

4.564

U
Pb

Amelin
2010

et

4.573

U
Pb

Amelin
2010

et

4.553

al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age

Discussion
In all instances the investigators took extreme care in the preparation of whole-rock samples and the separation of
components for subsequent radioisotope analyses. Fusion crusts were removed before crushing whole-rock samples, and
often components were then hand picked. If components were separated by density differences in heavy liquids and/or by
their magnetic properties, care was always taken to inspect under a microscope the purity of the final products before
proceeding with the radioisotope analyses.
In whole-rock analyses the complete samples were digested and therefore the components were homogenized in solution
before the target radioisotopes were extracted. Whether it was component minerals or constituents such as chondrules or
CAIs that were separated, they too were similarly digested and homogenised in solution. In the literature the suitability and
integrity for radioisotope analyses of components such as chondrules and CAIs has never been questioned. This is because
such components have been found from microscope study to be distinct entities within the overall fabric of meteorites,
independent from other components and therefore self-contained units suitable for radioisotope analyses on their own.
Furthermore, the fairly uniform mineralogical composition of chondrules and CAIs insures that bulk samples of such
separated components from individual meteorites will be representative. Therefore, the integrity and quality of the tabulated
radioisotope dates has been carefully maintained.
The two most obvious observations from a quick glance at Figs. 1821 are that:
In spite of some outlying scattered data, the bulk of the results strongly cluster so that the mode in all four frequency
diagrams lies in the range 4.564.57 Ga, which would appear to convincingly suggest the Allende CV3 chondrite has an
apparent age of between 4.56 and 4.57 Ga. However, this mode disappears entirely if the Pb-Pb ages (and the radioisotope
systems calibrated to them) are omitted. The some outlying scattered data refers to >90% of the non-Pb-Pb and non-PbPb-calibrated data. Consequently, the only dating method that shows strong clustering is Pb-Pb, and that clustering occurs
between 4.56 and 4.57 Ga, although the wide scatter in the results from the other dating methods does seem to roughly
group around an apparent age of 4.56 and 4.57 Ga.
As already noted the dominant radioisotope system used to obtain this apparent age is Pb-Pb, via both isochron (multisample) and model (single sample) methods. Indeed, if the Pb-Pb and ages calibrated with the Pb-Pb ages are ignored,
there is no strong clustering at all.

Fig. 18. Frequency versus radioisotope ages histogram diagram for the isochron ages for some or all components of the
Allende CV3 carbonaceous chondrite meteorite, with color coding being used to show the ages obtained by the different
radioisotope dating methods. Click image for larger view.

Fig. 19. Frequency versus radioisotope ages histogram diagram for the model ages for chondrules in the Allende CV3
carbonaceous chondrite meteorite, with color coding being used to show the ages obtained by the different radioisotope
dating methods. Click image for larger view.

Fig. 20. Frequency versus radioisotope ages histogram diagram for the model ages for Ca-Al inclusions (CAIs) in the
Allende CV3 carbonaceous chondrite meteorite, with color coding being used to show the ages obtained by the different
radioisotope dating methods. Click image for larger view.

Fig. 21. Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock, matrix and other
samples of the Allende CV3 carbonaceous chondrite meteorite, with color coding being used to show the ages obtained by
the different radioisotope dating methods. Click image for larger view.
Furthermore, a perusal of Tables 14 readily reveals that
In spite of improvements in the technology and greater sophistication in the techniques to measure the Pb-Pb isotope ratios
even more precisely over the decades during which the listed studies were undertaken, there has really been no substantial
change in the estimated age for the Allende CV3 chondrite. Indeed, Tatsumoto, Unruh, and Desborough in 1976 determined
the first Pb-Pb isochron age of 4.553 0.004 Ga (fig. 16), and yet 34 years later Amelin et al. in 2010 reported the most
recently determined Pb-Pb isochron age of 4.56718 0.0002 Ga (fig. 17), a mere difference between these two
determinations of only 0.01418 billion years, or 14.18 million years!
The Pb-Pb model ages on the Allende whole-rock samples, chondrules, Ca-Al inclusions (CAIs), and its matrix all essentially
cluster strongly around the same 4.564.57 Ga mark, giving the uniformitarians much certainty that they have successfully
and firmly dated the Allende CV3 chondrite as 4.567 Ga old.
The same pattern apparent in the Pb-Pb isochron ages is also apparent in the Pb-Pb model ages, namely, that in spite of
improvements in the technology and greater sophistication in the techniques to measure the Pb-Pb isotope ratios even more
precisely over the decades during which the listed studies were undertaken, there has really been no substantial change in
the estimated age for the Allende CV3 chondrite. For example:
Huey and Kohman in 1973 obtained a Pb-Pb model age for an Allende whole-rock sample of 4.528 0.04 Ga, yet 37 years
later Amelin et al. in 2010 determined a Pb-Pb model age of 4.56755 0.0002 Ga, a mere difference between these two
determinations of only 0.03955 billion years, or 39.55 million years!
Similarly, Chen and Tilton in 1976 determined Pb-Pb model ages for Allende chondrules of 4.568 Ga and 4.573 Ga, yet 33
years later Connelly and Bizzarro in 2009 obtained Pb-Pb model ages of 4.5659 0.0004 Ga and 4.5666 0.001 Ga, again
only meager differences between these determinations of 0.0021 billion years or 2.1 million years and 0.0064 billion years or
6.4 million years respectively!
And again, Chen and Tilton in 1976 determined Pb-Pb model ages for Allende Ca-Al inclusions of 4.555 Ga and 4.556 Ga,
yet 34 years later Amelin et al. in 2010 obtained Pb-Pb model ages of 4.56664 0.0003 Ga and 4.56702 0.0002 Ga, again
only small differences between these determinations of 0.01164 billion years or 11.64 million years and 0.01102 billion years
or 11.02 million years respectively!
But what of the other radioisotope dating methods? As to be expected, the U-Pb isochron method also yielded similar ages
to those obtained by the Pb-Pb isochron method, because both methods involve the radioisotope decay of U to Pb.
However, the Rb-Sr and Sm-Nd isochron methods yielded some results within the 4.564.57 Ga mode, but also results
within 0.3 billion years either side of that mode (fig. 18). Not only are the Rb-Sr isochron ages on either side of the mode of
Pb-Pb isochron ages, but so are the model Rb-Sr ages generally in Figs. 1921. In contrast, the Sm-Nd isochron ages are
greater than the Pb-Pb isochron ages (fig. 18).
The other successful radioisotope methods are not really independent and thus objective, because they are calibrated
against the Pb-Pb method (see table 1) and therefore are automatically guaranteed to give ages identical to those obtained
by the Pb-Pb isochron method. Specifically, the Al-Mg method is calibrated against the Pb-Pb isochron age for the
DOrbigny achondrite meteorite (Bouvier, Vervoort, and Patchett 2008), the Hf-W method is calibrated against the Pb-Pb
isochron age for the St. Marguerite chondrite meteorite (Kleine et al. 2005) and the DOrbigny achondrite meteorite
(Burkhardt et al. 2008), the Mn-Cr method is calibrated against the Pb-Pb isochron age for the St. Marguerite chondrite
meteorite (Trinquier et al. 2008), and the DOrbigny and LEW 86010 achondrite meteorites (Yin et al. 2009), and the I-Xe
method is calibrated against the I-Xe age of the Shallowater achondrite meteorite (Hohenberg et al. 2001), which is

calibrated against the Pb-Pb isochron age for the St. Marguerite chondrite meteorite (Brazzle et al. 1999). Thus, as to be
expected, all the dates obtained by these methods which are calibrated against these Pb-Pb isochron ages all plot in the
4.564.57 Ga mode with the clustered Pb-Pb dates (figs 18 and 20).
This strong clustering around the mode of 4.564.57 Ga is thus an example of high precision (versus accuracy), with only
the Pb-Pb (and calibrated methods) shows precision in ages. So in the eyes of uniformitarians the Pb-Pb isochron dating
method stands supreme as the ultimate, most reliable tool for determining the age of the Allende CV3 carbonaceous
chondrite meteorite.
The same is also true for the model ages obtained for Allende (figs. 1921 and tables 24). Whether it samples, the Pb-Pb
model ages have proven to be the most precise, the overwhelming numbers of analyses consistently clustering around the
4.564.57 Ga mark. Model ages obtained via the U-Pb, Rb-Sr, and Ar-Ar methods occasionally confirm the Pb-Pb model
ages, but they along with the Th-Pb method are also the major contributors to the many scattered model ages among the
data.
So having established the superior precision of the Pb-Pb radioisotope dating method, now we need to explore whether
there are any other patterns in all these isochron and model ages for the Allende CV3 carbonaceous chondrite meteorite
that might provide us with clues about the meaning and significance of these isochron and model ages within the creation
framework for the history of the earth and the solar system.
The major conclusion of the 19972005 RATE (radioisotopes and the age of the earth) project was that radioisotope decay
rates have not necessarily been constant throughout earth history, because there is evidence that there have been one or
more episodes of accelerated rates of radioisotope decay, particularly during the Flood only about 4350 years ago
(Vardiman, Snelling, and Chaffin 2005). While there were several lines of documented evidence that confirmed this
conclusion, the principal evidence was different isochron ages obtained from the same samples from the same rock units by
the different radioisotope dating methods (Snelling 2005; Vardiman, Snelling, and Chaffin 2005).
Furthermore, there was a consistent pattern to the isochron ages from the different methods that indicated that there was an
underlying systematic cause of these age differences, namely, an episode of accelerated radioisotope decay (Snelling 2005;
Vardiman, Snelling, and Chaffin 2005). For example, it was found that the -decaying radioisotopes U and Sm always gave
older ages than the -decaying K and Rb. And then between the -decayers, K with the shorter half-life (more rapid decay
today) and the lighter atomic weight, always yielded younger ages than the slower decaying and heavier Rb. While exactly
the same pattern was not confirmed among the -decaying U and Sm, both the half-lives and the atomic weights were still
believed to be the factors at work.
An example will best illustrate how this pattern of different isochron ages obtained from the same samples from the same
rock units by the different radioisotope dating methods provides evidence of the proposed accelerated radioisotope decay
event. Snelling, Austin, and Hoesch (2003) used the same samples from the Precambrian Bass Rapids diabase sill in the
Grand Canyon to obtain isochron dates for this rock unit of 841.5 164 Ma (K-Ar), 1060 24 Ma (Rb-Sr), 1250 130 Ma
(Pb-Pb), and 1379 140 Ma (Sm-Nd). Thus it was argued that if all the radioisotope clocks were set to zero when the
diabase sill crystallized and cooled, then the only way to reconcile these discordant radioisotope dates is if the different
parent radioisotopes then decayed over the same real time period from formation of the sill until today at different faster
rates than their rates today. It was also suggested that the parent radioisotopes decaying at different accelerated rates was
caused by their different atomic weights and abilities to decay (their half-lives) (Snelling, 2005; Vardiman, Snelling, and
Chaffin 2005).
The mechanism proposed for this past episode of accelerated radioisotope decay was small changes to the binding forces
in the nuclei of the parent radioisotopes (Vardiman, Snelling, and Chaffin 2005). These changes would thus have to have
affected every atom making up the earth, and by logical extension every atom of the universe at the same time physical
laws were created to governing the universe, to operate consistently through time and space. Therefore, we should expect
that this past episode of accelerated radioisotope decay had affected the asteroids from where many meteorites have come,
and that the meteorites may thus today yield the same pattern of different radioisotope ages from the different radioisotope
dating methods.
However, looking over Figs. 1821 again, there appears to be no identical systematic pattern of different radioisotope ages
from the different radioisotope dating methods, whether isochron or even model ages. This is more easily seen in Figs. 22
and 23, which are plots of the isochron ages versus the atomic weights and half-lives respectively, based on the data in
Table 1 in which the data has been grouped according to the mode of decay of the parent radioisotope. The sole K-Ar
isochron age is greater than any of the Rb-Sr isochron ages, and it is also greater than the sole Lu-Hf isochron age,
although the latter has a very broad error margin (see table 1). Yet this is the opposite of what should be expected, because
K has the shortest half-life today and has the lightest atomic weight among these three -decayers. At least the sole Lu-Hf
isochron age is predictably older than all the Rb-Sr isochron ages, expected because the -decaying Lu has a longer halflife today and has a heavier atomic weight than Rb. But the Sm-Nd isochron ages are both older and younger than, and
similar to, the Pb-Pb and U-Pb isochron ages (figs. 22 and 23). And among all the model ages (tables 24) there is so much
scatter no discernible systematic pattern is evident. In any case, model ages were not determined and compared during the
RATE project, primarily because model ages are dependent on, and thus subject to, the assumptions and deficiencies of the
models used to derive them. This is well documented in the literature, and is why model ages generally, apart from Pb-Pb
model ages, are regarded as unreliable (Faure and Mensing 2005). Of course, it is premature to draw too firm a conclusion
from these observations, because this is just one set of radioisotope ages for one meteorite. More sets of such data from
more meteorites are needed, and subsequent papers in preparation will supply these, enabling firmer conclusions to be
drawn.

Fig, 22. The isochron age yielded by five radioisotope systems in samples of some or all components from the Allende CV3
carbonaceous chondrite meteorite plotted against the atomic weights of the parent radioisotopes according to their mode of
decay.
Fig. 23. The isochron ages yielded by five
radioisotope systems in samples of some
or all components from the Allende CV3
carbonaceous chondrite meteorite plotted
against the present half-lives of the parent
radioisotopes according to their mode of
decay.
What is also clear is that even for earthbound rock units there are not yet enough
data sets of discordant radioisotope ages
derived by different radioisotope dating
methods. Those we so far have are
primarily for Precambrian rock units, and
even for those the pattern of discordance
and the amount of discordance are not
uniform or the same (Snelling 2005). For
example, Austin and Snelling (1998)
reported their radioisotope dating study of
the Precambrian Cardenas Basalt in the
Grand Canyon. The same samples yielded
isochron ages of 516 30 Ma (K-Ar), 1111
81 Ma (Rb-Sr) and 1588 170 Ma (SmNd). No Pb-Pb isochron age could be
derived from the U-Pb isotopic analytical
results. Now while these ages follow the
same discordance pattern as those for the
Precambrian Bass Rapids diabase sill
listed earlier, the amount of discordance is even greater, since the Rb-Sr isochron age is more than twice as large as the KAr isochron age, and the Sm-Nd isochron age is more than three times as large as the K-Ar isochron age. These two
Precambrian rock units are about the same 10601111 Ma Rb-Sr isochron age, are found in the same Grand Canyon
region, and within the creation framework of earth history are generally regarded as pre-Flood. One would have thought that
these two Precambrian rock units would have thus both equally suffered from the same accelerated radioisotope decay
episode during the Flood year about 4350 years ago, and therefore their radioisotope ages should be discordant by about
the same amount.Similarly, the discordance between the isochron ages of 1240 84 Ma (Rb-Sr), 1655 40 Ma (Sm-Nd),
and 1883 53 Ma (Pb-Pb) for the Precambrian Brahma amphibolites in the Grand Canyon is different again (Snelling 2005,
2008). Not only is the amount of discordance between the isochron ages different to the discordances between the isochron
ages for the other two Grand Canyon Precambrian rock units, but the pattern is also different. For the Bass Rapids diabase
sill the Sm-Nd isochron age is older than the Pb-Pb isochron age, whereas for the Brahma amphibolites the Sm-Nd isochron
age is younger than the Pb-Pb isochron age. So again, we need many more sets of discordant radioisotope ages data for
many more rock units spanning more of the geologic record before any further conclusions can be drawn.However, there is
still the issue of why the isochron and model ages for the Allende CV3 carbonaceous chondrite meteorite so definitely
cluster around an apparent radioisotope age of 4.564.57 Ga, with the current best Pb-Pb isochron and model ages of
4.56718 0.0002 Ga (fig. 17) and 4.56702 0.0002 Ga respectively (Amelin et al. 2010). Of course, it is premature to come
to any firm conclusions just based on this set of radioisotope dating data for this one meteorite, but it is nevertheless
worthwhile to begin considering possibilities, which can then be discussed in the light of more data sets in future papers.
Most meteorites are believed have been derived from asteroids via collisions between them breaking off fragments that then
hurtled towards the earth. So to be consistent, if the asteroids were also made on Day Four from this Day One primordial
material left over from the making of the planets and their satellites in the solar system, then this would imply the meteorites
could represent samples of this same primordial material. Similarly, this would have to also mean that at the beginning of
Day One the earth was also fashioned out of the same primordial material. Thus the 4.564.57 Ga Pb isotopic composition

of both this Allende meteorite and the bulk earth (as plotted on the geochron) may represent a geochemical signature from
this primordial material.This raises the obvious question about another aspect of the radioisotope dating technique. The
evidence of past accelerated radioisotope decay (non-constant radioisotope decay rates), that is, the inconsistent
radioisotope age data in earth rocks (Vardiman, Snelling, and Chaffin 2005), would appear to negate the assumption of
constant decay rates that enables the radioisotope clocks to be reading 4.564.57 Ga for the supposed elapsed real time
since the formation of the meteorites and the earth. However, another assumption necessary for these radioisotope clocks
to work is that all the daughter isotopes were only derived by radioisotope decay from the parent isotopes. But what if all the
isotopes were created at the beginning in the primordial material, including isotopes that subsequently formed by
radioisotope decay as daughter isotopes from parent isotopes? In other words, when the primordial material was created d
was it include in it 206Pb, 207Pb, and 208Pb atoms along with 238U, 235U, and 232Th atoms? It may be reasonable to
posit that He did, given that when created the primordial material had to have some initial isotopic ratios. Even the
conventional scientific community have assumed the initial material of the solar system had the primeval Pb isotopic ratios
as measured in the troilite (iron sulfide) in the Canyon Diablo iron meteorite (Faure and Mensing 2005). So if He did, then
the Pb isotopes we measure today are not all the product of radioisotope decay, and they therefore cannot be measuring the
elapsed real time.Following from this is one final consideration. How many atoms of the Pb isotopes were create in the
primordial material? And if the 4.564.57 Ga age for this meteorite is a geochemical signature of the primordial material
were created with some of the Pb isotopes we measure today already in it, then how many of the atoms of the measuredtoday Pb isotopes are due to past accelerated radioisotope decay? Was it most of them, or only some of them? So far we
dont know. What we do know is that here on the earth we dont find rocks that still have the 4.564.57 Ga Pb isotopic
signature in them, though most rocks that date back to the continental foundations laid down during Creation Week still
contain various large amounts of Pb isotopes in them and therefore yield an array of multi-Ga ages. Significantly, the earth
sample that plotted on Pattersons 1956 geochron (fig. 1) was a modern ocean sediment sample whose Pb isotopic
signature had been acquired by mixing and integration from many earth rocks over time. This may point to a third possible
process responsible for the Pb isotopic compositions in earth rocks, namely, inheritance and mixing in the earths mantle
and crust subsequent to the creation of the original earth on Day One of the Creation Week, as previously proposed by
Snelling (2000, 2005), primarily during the catastrophic geologic processes of the Day Three upheaval when the dry land
was formed and during the Flood (Snelling 2009).
The resultant conclusion from all these considerations and deliberations, based on the assumptions made, is that the
4.56718 0.0002 Ga age for the Allende CV3 carbonaceous chondrite meteorite obtained by Pb-Pb radioisotope isochron
dating of one of its Ca-Al inclusions (CAIs) (Amelin et al. 2010) is likely not its true real-time age. The assumptions on which
the radioisotope dating methods are based are simply unprovable, and in the light of the evidence for possible past
accelerated radioisotope decay in earth rocks and the possibility of an inherited primordial geochemical signature, these
assumptions are unreasonable. However, we are still left without a coherent explanation of what these radioisotope
compositions really mean within our young-age Creation-Flood framework for earth and universe history. We have some
possible clues already, and a clearer picture may yet emerge from continued investigations now in progress, for example, of
the radioisotope dating of many other meteorites.
Conclusions
There is no doubt that after decades of numerous careful radioisotope dating investigations of the Allende CV3
carbonaceous chondrite meteorite that its Pb-Pb isochron age of 4.56718 0.0002 Ga has been well established. This date
for Allende is supported by a very strong clustering of other Pb-Pb isochron and model ages in the 4.564.57 Ga range, as
well as being confirmed by Pb-Pb model ages and by both isochron and model age results by the U-Pb, and to a lesser
extent, the Rb-Sr and Sm-Nd methods. The Al-Mg, Hf-W, Mn-Cr, and I-Xe methods are all calibrated against the Pb-Pb
isochron method, so their results are not objectively independent. Thus the Pb-Pb isochron dating method stands supreme
as the ultimate, most precise tool for determining the age of the Allende CV3 carbonaceous chondrite meteorite.
There is no other discernible systematic pattern in the isochron and model ages for Allende, apart from scatter of the U-Pb,
Th-Pb, Rb-Sr, and Ar-Ar model ages particularly. The Allende ages do not follow the systematic pattern found in Grand
Canyon Precambrian rock units during the RATE project, where the -decay ages are younger than the -decay ages
according to the lengths of the half-lives and the atomic weights of the parent radioisotopes. Thus there appears to be no
evidence in the Allende CV3 carbonaceous chondrite meteorite similar to the evidence found in earth rocks of past
accelerated radioisotope decay.
Any explanation for the 4.56718 0.0002 Ga age for this Allende meteorite needs to consider the origin of meteorites. Most
meteorites appear to be fragments derived from asteroids via collisions, but even in the naturalistic paradigm the asteroids,
and thus the meteorites, are regarded as primordial material left over from the formation of the solar system. If some of the
daughter isotopes were inherited by the Allende meteorite when it was formed from that primordial material, and the parent
isotopes in the meteorite were also subject to subsequent accelerated radioisotope decay, then the 4.56718 0.0002 Ga
Pb-Pb isochron age for the Allende CV3 carbonaceous chondrite meteorite cannot be its true real-time age, which
according to the creation paradigm is only about 6000 real-time years.
However, these conclusions and the suggested explanation can at best be regarded as tentative and interim while their
confirmation or adjustment awaits the examination of more radioisotope dating data from many more meteorites. Such
studies are already in progress.
Radioisotope Dating of Meteorites II: The Ordinary and Enstatite Chondrites
by Dr. Andrew A. Snelling on August 20, 2014
Abstract
Meteorites date the earth with a 4.55 0.07 Ga Pb-Pb isochron called the geochron. They appear to consistently yield 4.554.57 Ga radioisotope ages, adding to the uniformitarians confidence in the radioisotope dating methods. About 82% of all
meteorite falls are chondrites, stony meteorites containing chondrules. Nearly 94% of chondrites are ordinary (O)
chondrites, which are subdivided into H, L, and LL chondrites based on their iron contents. Enstatite (E) chondrites
comprise only 1.4% of the chondrites. Many radioisotope dating studies in the last 45 years have used the K-Ar, Ar-Ar, RbSr, Sm-Nd, U-Th-Pb, Re-Os, U-Th/He, Mn-Cr, Hf-W, and I-Xe methods to yield an abundance of isochron and model ages
for these meteorites from whole-rock samples, and mineral and other fractions. Such age data for fifteen O and E chondrites
were tabulated and plotted on frequency versus age histogram diagrams. They generally cluster, strongly in some of these
chondrites, at 4.554.57 Ga, dominated by Pb-Pb and U-Pb isochron and model ages, testimony to that techniques
supremacy as the uniformitarians ultimate, most reliable dating tool. These ages are confirmed by Ar-Ar, Rb-Sr, Re-Os, and
Sm-Nd isochron ages, but there is also scatter of the U-Pb, Th-Pb, Rb-Sr, and Ar-Ar model ages, in some cases possibly
due to thermal disturbance. No pattern was found in these meteorites isochron ages similar to the systematic patterns of
isochron ages found in Precambrian rock units during the RATE project, so there is no evidence of past accelerated

radioisotope decay having occurred in these chondrites. This is not as expected, because if accelerated radioisotope decay
did occur on the earth, then it could be argued every atom in the universe would be similarly affected at the same time.
Otherwise, asteroids and the meteorites derived from them are regarded as primordial material left over from the formation
of the solar system, which .Dodays measured radioisotope compositions of these O and E chondrites may reflect a
geochemical signature of that primordial material, which included atoms of all elemental isotopes. So if some of the
daughter isotopes were already in these O and E chondrites when they were formed, then the 4.55-4.57 Ga ages for the
Richardton (H5), St. Marguerite (H4), Bardwell (L5), Bjurbole (L4), and St. Sverin (LL6) ordinary chondrite meteorites
obtained by Pb-Pb and U-Pb isochron and model age dating are likely not their true real-time ages, which according to the
creation paradigm is only about 6,000 real-time years. The results of further studies of more radioisotope ages data for
many more other meteorites should further elucidate these interim suggestions.
Keywords: meteorites, classification, ordinary (O) chondrites, H chondrites, L and LL chondrites, enstatite (E) chondrites,
radioisotope dating, Allegan, Forest Vale, Guarena, Richardton, St. Marguerite, Barwell, Bjurbole, Bruderheim, Olivenza, St.
Sverin, Abee, Hvittis, Indarch, St. Marks, St. Sauveur, K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, U-Th-Pb, U-Th/He, Re-Os, Mn-Cr, Hf-W,
I-Xe, isochron ages, model ages, discordant radioisotope ages, accelerated radioactive decay, asteroids, primordial
material, geochemical signature, inheritance and mixing
Introduction
Ever since 1956 when Claire Patterson at the California Institute of Technology in Pasadena reported a Pb-Pb isochron age
of 4.55 0.07 Ga for three stony and two iron meteorites, this has been declared the age of the earth (Patterson 1956).
Furthermore, many meteorites appear to have consistently dated around the same age (Dalrymple 1991, 2004), bolstering
the evolutionary communitys confidence that they have successfully dated the age of the earth and the solar system at
around 4.57 Ga. It has also strengthened their case for the supposed reliability of the increasingly sophisticated radioisotope
dating methods.Creationists have commented little on the radioisotope dating of meteorites, apart from acknowledging the
use of Pattersons geochron to establish the age of the earth, and that many meteorites give a similar old age. Morris (2007)
did focus on the Allende carbonaceous chondrite as an example of a well-studied meteorite analysed by many radioisotope
dating methods, but he only discussed the radioisotope dating results from one, older (1976) paper. Furthermore, he only
focused on the U-Th-Pb model ages published in that paper, apparently ignoring the excellent Pb-Pb isochron age of 4.553
0.004 Ga based on some twenty isotopic analyses of the matrix, magnetic separates, aggregates and chondrules reported
in that same paper, as well as the U-Pb concordia isochron age of 4.548 0.025 Ga based on those same samples.
In order to rectify this lack of engagement by the creationist community with the meteorite radioisotope dating data, Snelling
(2014) obtained as much radioisotope dating data as possible for the Allende CV3 carbonaceous chondrite meteorite (due
to its claimed status as the most studied meteorite), displayed the data, and attempted to analyse it. He found that both
isochron and model ages for the total rock, separated components, or combinations of these strongly clustered around a
Pb-Pb age of 4.564.57 Ga. However, while he then sought to discuss the possible significance of this clustering in terms of
various potential creationist models for the history of radioisotopes and their decay, drawing firm conclusions from the
radioisotope dating data for just this one meteorite was premature. This present contribution is therefore designed to
document the radioisotope dating data for more meteorites, the ordinary and enstatite chondrites, so as to continue the
discussion of the significance of these data.
The Classification of Meteorites
Meteorites have been classified into distinct groups and subgroups that show similar chemical, isotopic, mineral, and
physical relationships. Within the evolutionary community the ultimate goal of such a classification scheme is to group all
known specimens that apparently share a common origin on a single, identifiable parent body, or even a body yet to be
identified. This could be another planet, moon, asteroid, or other current solar system object, or one that is believed to have
existed in the past (for example, a shattered asteroid). However, several meteorite groups classified this way appear to have
come from a single, heterogeneous parent body, or even a single group may contain members that may have come from a
variety of similar but distinct parent bodies. So any meteorite classification system is not absolute, and is only as valid as the
criteria used to develop it.More than 24,000 meteorites are currently catalogued (Norton 2002), and this number is rapidly
growing due to the ongoing discovery of large concentrations of meteorites in the worlds cold and hot deserts (for example,
in Antarctica, and Australia and Africa, respectively). Traditionally meteorites have been divided into three overall categories
based on whether they are dominantly composed of rocky materials (stones or stony meteorites), metallic material (irons or
iron meteorites), or mixtures (stony-irons or stony-iron meteorites). These categories have been in use since at least the
early nineteenth century, but they are merely descriptive and do not have any genetic connotations. In reality, the term
stony-iron is a misnomer, as the meteorites in one group (the CB chondrites) have over 50% metal by volume and were
called stony-irons until their affinities with chondrites were recognized. Similarly, some iron meteorites also contain many
silicate inclusions but are rarely described as stony-irons.Nevertheless, these three categories are still part of the most
widely used meteorite classification system. Stony meteorites are traditionally divided into two other categorieschondrites
(meteorites that are characterized by containing chondrules and which apparently have undergone little change since their
parent bodies originally formed), and achondrites (meteorites that appear to have had a complex origin involving asteroidal
or planetary differentiation). Iron meteorites were traditionally divided into objects with similar internal structures
(octahedrites, hexahedrites, and ataxites), but these terms are now only used for descriptive purposes and have given way
to chemical group names. Stony-iron meteorites have always been divided into pallasites (which now comprise several
distinct groups) and mesosiderites (a textural term which is also synonymous with the name of a modern group).Based on
their bulk compositions and textures, meteorites have been more recently divided by Krot et al. (2005) into two major
categorieschondrites and non-chondritic meteorites. They also further subdivided the non-chondritic meteorites into the
primitive achondrites and igneously differentiated meteorites, the latter including the achondrites, stony-irons (pallasites and
mesosiderites), and the irons. Within all these categories the meteorites are grouped on the basis of their oxygen isotopes,
chemistry, mineralogy, and petrography.Weisberg, McCoy and Krot (2006) made only minor changes to this classification
scheme, which is illustrated in Fig. 1. Note that the three main categories have now been reduced just to chondrites,
primitive achondrites and achondrites, the main change being to simply rename the igneously differentiated meteorites the
achondrites. As in Krot et al.s (2005) classification scheme, the IAB and IIICD irons are included in the primitive achondrites
because of their silicate inclusions, while the rest of the groups of irons, the stony-irons, the martian and lunar meteorites
are included with the other achondrite groups in the achondrites.
The Chondrites
About 82% of all meteorite falls are chondrites (Norton 2002). As already noted, the chondrites derive their name from their
interior texture, which is unlike any found in terrestrial rocks. Dispersed more or less uniformly throughout these meteorites
are spherical, sub-spherical and sometimes ellipsoidal structures called chondrules. These range in size from about 0.1 to 4
mm (0,0039 to 0,15 in) diameter, with a few reaching centimeter size. Their abundance within a given chondrite can vary
enormously from only a few per cent of the total volume of the meteorite to as much as 70%, with fine-grained matrix

material dispersed between the chondrules. Most chondrules are rich in the silicate minerals olivine and pyroxene. The other
major components of chondrites are refractory inclusionsCa-Al-rich inclusions (CAIs) and amoeboid olivine aggregates
(AOAs)and Fe-Ni metal alloys and sulfides (Brearley and Jones 1998; Scott and Krot 2005; Snelling 2014).
The chondrites have been subdivided into three classescarbonaceous (C), ordinary (O), and enstatite (E) chondritesand
fifteen groups, including the rare R and K chondrites (fig. 1). The carbonaceous (C) chondrites, representing almost 4% of
all chondrites, are so named because their matrix is carbon-rich, containing various amounts of carbon in the form of
carbonates and complex organic compounds including amino acids (Cronin, Pizzarello, and Cruikshank 1988). Further
classification involves typing according to where the first meteorite or prototype in the category was found and whose
characteristics are used to define the groupfor example, CI where I denotes Ivuna, a town in Tanzania, CM where M
stands for Mighei in Ukraine, CV where V designates Vigarano in Italy, CO where the O stands for the town of Ornans in
France, CR where R denotes Renazzo in Italy, and CK where K designates Karoonda, a town in South Australia (Krot et al.
2009; Norton 2002).

Fig. 1. The classification system for meteorites (after Weisberg, McCoy, and Krot 2006). Click image for larger view.
Ordinary (O) chondrites are by far the most common type of meteorite to fall to earth. About 77% of all meteorites and nearly
94% of chondrites are ordinary chondrites. They have been
divided into three groupsH, L and LL chondritesthe
letters designating their different bulk iron contents and
different amounts of metal (Krot et al. 2005; Norton 2002):
H chondrites have High total iron contents and high metallic
Fe (1520% Fe-Ni alloys by mass) and smaller chondrules
than L and LL chondrites. About 42% of ordinary chondrite
falls belong to this group.
L chondrites have Low total iron contents (including 711%
Fe-Ni alloys by mass). About 46% of ordinary chondrite falls
belong to this group, which makes them the most common
type of meteorite to fall to earth.
LL chondrites have Low total iron and Low metal contents
(35% Fe-Ni alloys by mass, of which 2% is metallic Fe).
About 1012% of ordinary chondrite falls belong to this
group.
Fig. 2. Two histograms showing the Mg/Si and Ca/Si
compositions of chondrites (after Norton 2002; Von
Michaelis, Ahrens, and Willis 1969; Van Schmus and Hayes
1974). These atomic ratios differ significantly so that three
divisions or classes of chondrites are evidentthe enstatite
(E) chondrites, ordinary (O) chondrites, and carbonaceous
(C) chondrites. The data even allows each class to be
resolved into groupsenstatite chondrites into EH and EL;
ordinary chondrites into H, L, and LL; and carbonaceous
chondrites into CI, CM, CV, and CO.
Fig. 3. Plot of the weight percent oxidized iron (in minerals)
versus the weight percent iron metal plus FeS (unoxidized
iron) in chondrites observed to fall and recovered shortly

thereafter (after Mason 1962). A clear division of the three classes of chondrites is obvious, along with the three groups in
the ordinary chondritesH, L, and LL.
Fig. 4. Plot of the fayalite (Fa) content of olivine versus the ferrosilite (Fs) content of orthopyroxene in equilibrated ordinary
chondrites clearly reveals the existence of the three oxidation groupsH, L, and LL (after Keil and Fredriksson 1964; Norton
2002).The E chondrites comprise only 1.4% of the chondrites, and are obviously named after their primary silicate mineral,
enstatite. Enstatite is the Mg-rich end member of the orthopyroxene solid-solution series and makes up 6080 vol. % of
these meteorites (Krot et al. 2009; Norton 2002). E chondrites contain more metal phases than any other stony meteorite
class, with total iron contents varying between 22 and 33 wt %. Virtually all of their iron is in metal phases (1328 vol. %) or
as sulfides (517 vol. %). So like the ordinary (O) chondrites, the E chondrites are divided into two groups, EH and EL,
according to whether they have relatively High or Low total iron and metal contents. EH chondrites average about 30 vol. %
total iron of which about 5 vol. % is sulfides, whereas EL chondrites have about 25 vol. % total iron with 3.5 vol. %
sulfides.Of all the meteorites, the chondrites show the greatest similarities in composition, so there are only subtle chemical
differences between them. The lithophile elements (those with a strong affinity for oxygen that tend to concentrate in silicate
phases) Mg and Ca show the most distinct divisions among the chondrites. Fig. 2 provides histogram plots of Mg/Si and
Ca/Si abundances in the chondrite groups (Von Michaelis, Ahrens, and Willis 1969; Van Schmus and Hayes 1974), and
shows an obvious distinction between the chondrite groups. The E chondrites exhibit the lowest element/Si ratios, while the
C chondrites cluster among the highest ratios, and the ordinary chondrites fall in a tight cluster between the two.An even
more striking distinction among the chondrites is evident when oxidized Fe is plotted against Fe in the metal phase and FeS
(Mason 1962). Fig. 3 shows a clear distinction between the three classes of chondrites. The E chondrites form a tight cluster
exhibiting little oxidation, while the C chondrites display the greatest oxidation of their Fe. Again, the O chondrites fall in
between, with separate clusters for each of their constituent H, L and LL groups reflecting their respective Fe metal contents,
the H chondrites having the highest Fe metal content.The O chondrites can thus also be classified according to their range
of FeO/(FeO + MgO) molecular percentages in their two most common ferromagnesian minerals, olivine and pyroxene. For
meteorites in general the fayalite (Fe2SiO4) composition of olivine most commonly lies between 15 and 30% (Fa15-30), with
the olivine in a typical O chondrite in the H group having an Fa 18 composition. Like olivine, the orthopyroxene composition in
meteorites is measured as the mole percent of the Fe-bearing end member, ferrosilite (FeSiO 3). A typical pyroxene
composition for an L group O chondrite would be Fs 22.The enstatite and three groups of ordinary chondrites are
distinguished by their total iron content, both oxidized iron (combined in minerals) and metal (unoxidized iron), with the
normal variations found in the metal phase, total iron, fayalite (in olivine) and ferrosilite (in pyroxene) contents listed in Table
1. The H, L, and LL designations are as defined above, and are applied to both the O and E chondrites. From these data in
Table 1 it is evident that the more oxidized iron in minerals such as fayalite and ferrosilite, the less unoxidized iron there is
as metal in the bulk composition of these chondrite meteorites. Furthermore, as the oxidized iron increases in minerals so
their oxygen content also increases. So if the mole percent fayalite (Fa) in the olivine and the mole percent ferrosilite (Fs) in
the pyroxene are plotted against each other the three ordinary chondrite groups are clearly distinguished, because the H
chondrites are the least oxidized and the LL chondrites are the most oxidized of the ordinary chondrites (fig. 4).The
classification of chondrites based on chemical and mineralogical criteria is considered a primary classification because the
bulk chemistry of meteorites is a primary characteristic. However, meteorites within a particular chemical group, such as the
three groups within the ordinary (O) chondrite class, have remarkably similar bulk compositions, but under a hand lens and
microscope there are striking petrographic differences. Thus a classification system needs to take into account these
petrographic differences so that meteorites can be at least roughly classified by visual inspection. This requires secondary
properties be considered, that is, properties that formed from processes which modified the original primary petrographic
characteristics. Consequently, an effective classification of chondrites takes into account both their petrographic properties
and their chemical differences, using the petrographic differences to subdivide and further refine the chemical groups.
Table 1. The classification of the enstatite (E) chondrites (H, L) and ordinary (O) chondrites (H, L, LL) according to their total
iron content (after Norton 2002).The symbols H, L, and LL designate the chemical abundance of iron found in each, both as
metal (unoxidized) and iron combined in minerals (oxidized)H (High total iron), L (Low total iron), and LL (Low total iron
and Low iron). The fayalite content of olivine and the ferrosilite content of pyroxene are both distinguishing indicators of each
group.
Fig. 5 is a comprehensive classification chart
Metal
Total
IronFayalite
Ferrosilite
giving ten criteria proposed by Van Schmus and
Class
Group
(wt %) (wt %)
(Fa mole %) (Fs mole %)
Wood (1967) that with some modification is still
being used to determine the petrographic type
Enstatite EH & EL 1723 2233
1
0
of each chondrite group. Of the ten, most
H
1519 2530
1620
1420
involve precise chemical and mineral analyses.
However, fortunately, among the criteria
L
110
2023
2125
2030
(numbers 3, 4, 7, and 8) there are well-defined
properties that are readily observable through
Ordinary LL
13
1922
2632
3240
microscope study of thin sections so that the
petrographic type can be visually estimated with some confidence without chemical analyses. Criteria numbers 7 and 8 are
discussed here because they establish the features needed to understand the classification of the petrographic types to
which the remaining criteria refer.

Fig. 5. Chart showing the criteria for distinguishing petrographic types in chondrites (after Brearley and Jones 1998; Norton
2002; Sears and Dodd 1988; Van Schmus and Wood 1967). The ten criteria used in this scheme as they were originally
devised are displayed with the details that define each type for each criterion. The broken lines are intended to reflect the
lack of sharpness of the boundaries between two petrographic types.
Of the ten criteria, the chondrule texture and density (criterion number 7) is the most easily observed. Petrographic types
range from 1 to 6. In Type 1 chondrites chondrules are absent. Type 2 chondrites contain distinct chondrules but they are
sparsely distributed within a matrix that constitutes nearly 50% of the meteorite by volume. Types 36 show progressive
stages of thermal metamorphism. The chondrule boundaries became progressively indistinct as solid state recrystallization
occurred. This caused alteration of the original chondrule boundaries due to intergrowth of chondrules and the matrix. This
recrystallization does not represent heating to the point of fusion, but only sufficient heating to allow migration and
recombination of the mineral elements into new minerals. This solid state recrystallization occurred between 400 and 950C.
Ordinary chondrites show petrographic types from 3 to 6 (fig. 5). Often Types 5 and 6 O chondrites show brecciated
textures, composed of light clasts set against a dark matrix. It is not unusual to see more than one petrographic type in
these breccias. Typically the clasts show Type 5 or 6, while the matrix shows Type 3 or 4. In that case the entire
petrographic range is designated Type 36. Matrix texture (criterion number 8) is easily observed in thin sections.
Matrix textures in Type 1 and 2 chondrites are opaque (black) and very fine-grained with scattered recognizable crystal
fragments. Type 2 chondrites show small chondrules, enclosing only about 12% of the meteorites by volume. Type 3
chondrites are still unequilibrated and their matrix is still dark but chondrules are increased in number and take up 30% or
more of the volume. From Type 4 to 6, increasing thermal metamorphism in ordinary chondrites produced recrystallization of
the matrix in which the crystals grew from cryptocrystalline to near naked-eye visibility. This turned the matrix transparent,
giving the interior of the chondrite a white appearance.
In examining the homogeneity of olivine and pyroxene compositions (criterion number 1) (fig. 5), from the textures of the
ordinary chondrites it is assumed they all began in a relatively unmetamorphosed state designated Type 3. The parent
chondritic body from which the meteorite came is said to have been chemically unequilibrated; that is, its mineral
composition was heterogeneous, showing wide variations in chemical composition within each mineral. In particular, the two
most common minerals in chondrites, olivine and pyroxene, show wide variations in their Mg/Fe compositions (table 1 and
fig. 4). The minerals in unequilibrated Type 3 O chondrites were therefore not in equilibrium with their surroundings, the iron
composition in olivine and orthopyroxene varying from grain to grain by more than 5%. This variation was progressively
reduced through Type 4 until it reached nearly a singular composition at Type 5 where both have become more ferrous. All
the olivine and orthopyroxene then have similar iron compositions. Types 5 and 6 chondrites are both homogeneous and
equilibrated.
The other criteria are listed in Fig. 5 and the characteristics for each criterion are provided for each petrographic type. Most
are self-evident and require thin section examinations, whereas others require mineral or bulk chemical analyses. The
defining of these petrographic types adds to the classification of chondrite meteorites. The known petrographic types for the
chondrite groups are summarized in Fig. 6. Thus chemical types H, L, and LL ordinary (O) chondrites can have a
petrographic type between 3 and 6, labelled as H3H6, L3L6, and LL3LL6, respectively. Taken together, the
carbonaceous (C) chondrites vary from C16 and the enstatite (E) chondrites EH and EL 16.

Fig. 6. Chart summarizing the grouping of all chondrites into chemical and petrographic types (after Norton 2002). The
chemical types are claimed to represent different asteroid parent bodies, while the petrographic types refer to various states
of thermal metamorphism or aqueous alteration occurring on or within the parent bodies. The ordinary chondrites show
thermal metamorphism, while the carbonaceous chondrites can be divided into those that show aqueous alteration and
those that show thermal metamorphism. The blank boxes indicate the combinations that either do not exist or have yet to be
found.However, the exceptions are the Type 3 ordinary and carbonaceous chondrites, which have been sub-typed from 3.0
to 3.9 using a different set of criteria. This was found necessary because Type 3 ordinary chondrites appear to have gone
through an unusually large range of thermal metamorphism, more so than other types. Among the new criteria are
thermoluminesence sensitivity (tendency to emit light or infrared energy upon heating), percent matrix recrystallization,
variation of cobalt in the low nickel kamacite, variations of the fayalite in olivine, and the FeO/(FeO + MgO) ratio in the
matrix.While these details are all background information, their presentation is necessary for an understanding of the
identifications and designations of the meteorites investigated in this study. It is important to establish what the different
designations mean so that one can have confidence that within the groupings of the meteorites chosen for comparing their
radioisotope dates the meteorites are essentially the same chemically and mineralogically. This hopefully eliminates any
differences in radioisotope ages being due to chemical and/or mineralogical differences.
The Radioisotope Dating of the Ordinary and Enstatite Chondrites
Fig. 7. Hand specimen of the L4 chondrite Bjurbole (after Norton
2002). Its extreme friability makes it subject to crumbling, so that
the chondrules (the high relief, ovoid shapes) frequently fall out of
the surrounding matrix leaving cavities. The specimen is 5.3 cm
(2 in) in the largest dimension.
To thoroughly investigate the radioisotope dating of the ordinary
(O) and enstatite (E) chondrite meteorites all the relevant
literature was searched. The objective was to find chondrites that
have been dated by more than one radioisotope method, and a
convenient place to start was Dalrymple (1991, 2004), who
compiled lists of such data. Ordinary (O) chondrite meteorites
that were found to have been dated multiple times by more than
one radioisotope method included five H chondritesAllegan
(H5), Forest Vale (H4), Guarena (H6), Richardton (H5), and St. Marguerite (H4); three L chondritesBardwell (L5), Bjurbole
(L4) (fig. 7), and Bruderheim (L6) (fig. 8); and two LL chondritesOlivenza (LL5) and St. Sverin (LL6). Five E chondrite
meteorites were found to have been dated multiple times by more than one radioisotope methodAbee (EH4), Hvittis
(EL6), Indarch (EH4), St. Marks (EH5), and St. Sauveur (EH5). So this study focused on all fifteen of these meteorites.
When papers containing radioisotope dating results for these chondrites were found, the reference lists were also scanned
to find further relevant papers. In this way a comprehensive set of papers, articles and abstracts on radioisotope dating of
these chondrite meteorites was collected. While it cannot be
claimed that all the papers, articles and abstracts which have
ever been published containing radioisotope dating results for
these chondrites have thus been obtained, the cross-checking
undertaken between these publications does indicate the data
set obtained is very comprehensive.
All the radioisotope dating results from these papers, articles and
abstracts were then compiled and tabulated. For ease of viewing
and comparing the radioisotope dating data, the isochron and
model ages for some or all components of each meteorite were
tabulated separatelythe H chondrites in Tables 2 (isochron
ages) and 3 (model ages), the L chondrites (tables 4 and 5
respectively), the LL chondrites (tables 6 and 7 respectively), and
the E chondrites (tables 8 and 9 respectively).

Fig. 8. Photomicrograph of a cut surface of the L6 chondrite Bruderheim, which fell in Alberta, Canada, in 1960 (after Norton
2002). The limonite (yellow-brown hydrated iron oxides) staining of the matrix around the included metallic iron-nickel grains
demonstrates the effect of chemical weathering after meteorites fall to earth due to the reactions with water and atmospheric
oxygen. The horizontal field of view is 35 mm (1.3 in).
The data in these tables were then plotted on frequency versus age histogram diagrams, with the same color coding being
used to show the ages obtained by the different radioisotope dating methodsthe isochron and model ages for some or all
components of the H chondrites (figs. 9 and 10 respectively), of the L chondrites (figs. 11 and 12 respectively), of the LL
chondrites (figs. 13 and 14 respectively), and of the E chondrites (figs. 15 and 16 respectively).

Fig. 9. Frequency versus radioisotope ages histogram diagram for the isochron ages for some or all components of the H
chondrite meteorites (a) Allegan (H5), (b) Forest Vale (H4), (c) Guarena (H6), (d) Richardton (H5), and (e) St. Marguerite
(H4), with color coding being used to show the ages obtained by the different radioisotope dating methods. Click images for
larger view.

Fig. 10. Frequency versus radioisotope ages histogram diagram for the model ages for some or all components of the H
chondrite meteorites (a) Allegan (H5), (b) Forest Vale (H4), (c) Guarena (H6), (d) Richardton (H5), and (e) St. Marguerite
(H4), with color coding being used to show the ages obtained by the different radioisotope dating methods. Click images for
larger view.

Fig. 11. Frequency versus radioisotope ages histogram diagram for the isochron ages for some or all components of the L
chondrite meteorites (a) Bardwell (L5), (b) Bjurbole (L4), and (c) Bruderheim (L6), with color coding being used to show the
ages obtained by the different radioisotope dating methods. Click images for larger view.

Fig. 12. Frequency versus radioisotope ages histogram diagram for the model ages for some or all components of the L
chondrite meteorites (a) Bardwell (L5), (b) Bjurbole (L4), and (c) Bruderheim (L6), with color coding being used to show the
ages obtained by the different radioisotope dating methods. Click images for larger view.

Fig. 13. Frequency versus radioisotope ages histogram diagram for the isochron ages for some or all components of the LL
chondrite meteorites (a) Olivenza (LL5) and (b) St. Sverin (LL6), with color coding being used to show the ages obtained
by the different radioisotope dating methods. Click images for larger view.

Fig. 14. Frequency versus radioisotope ages histogram diagram for the model ages for some or all components of the LL
chondrite meteorites (a) Olivenza (LL5) and (b) St. Sverin (LL6), with color coding being used to show the ages obtained
by the different radioisotope dating methods. Click images for larger view.

Fig. 15. Frequency versus radioisotope ages histogram diagram for the isochron ages for some or all components of the E
chondrite meteorites (a) Abee (EH4), (b) Hvittis (EL6), (c) Indarch (EH4), (d) St. Marks (EH5), and (e) St. Sauveur (EH5),
with color coding being used to show the ages obtained by the different radioisotope dating methods. Click images for larger
view.

Fig. 16. Frequency versus radioisotope ages histogram diagram for the model ages for some or all components of the E
chondrite meteorites (a) Abee (EH4), (b) Hvittis (EL6), (c) Indarch (EH4), (d) St. Marks (EH5), and (e) St. Sauveur (EH5),
with color coding being used to show the ages obtained by the different radioisotope dating methods. Click images for larger
view.
Table 2. Isochron ages for some or all components of the H chondrite meteorites Allegan (H5), Forest Vale (H4), Guarena
(H6), Richardton (H5) and St. Marguerite (H4), with the details and literature sources.
Sample

Method

Reading

Err +/-

Note

Source

Type

206

4.557

0.008

Unruh, Hutchison, andisochron


Tatsumoto 1982
age

4.573

0.003

Brazzle et al. 1999

Mn-Cr

4.5613

0.0008

Polnau and Lugmairisochron


2000
age

Mn-Cr

4.5609

0.0008

Polnau and Lugmairisochron


2001
age

4.63

0.16

Sanz and Wasserburgisochron


1969
age

4.56

0.08

Wasserburg,
Papanastassiou,
Sanz 1969

Rb-Sr

4.46

0.08

Dalrymple 2004

Rb-Sr

4.48

0.08

Minster, Birck,
Allgre 1982

Allegan (H5)
whole rock samples plotted with
five Barwell samples (1 each)
from
three
other
chondrite
meteorites

Pb-207Pb

feldspar, temperature extractions


(8001800C)
I-Xe

isochron
age

Forest Vale (H4)

Guarena (H6)
two pyroxene samples plotted
with eighteen fractions of the
Olivenza chondrite
Rb-Sr

Rb-Sr

thirteen fractions plottedwhole


meteorite (3), phosphate (1) and
density splits (9)

one sample plotted with ten


analyses
from
five
other
meteorites
Sm-Nd

andisochron
age
isochron
age
andisochron
age

Jacobsen
andisochron
Wasserburg 1980
age

4.6

Richardton (H5)
seven chondrules

Rb-Sr

isochron
age

4.39

0.03

Evensen et al 1979

six silicate chondrules combined


with phosphates (4)
Rb-Sr

4.611

0.11

Rotenberg and Amelinisochron


2002
age

only phosphates (4)

Rb-Sr

2.284

0.88

Rotenberg and Amelinisochron


2002
age

only silicates (6)

Rb-Sr

4.62

0.14

Rotenberg and Amelinisochron


2002
age

chondrules and fragments

Pb-Pb

4.5627

0.0017

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

phosphate fractions

Pb-Pb

4.5507

0.0026

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

nine meteorite fragments

207

Pb-206Pb

4.545

0.01

Abranches, Arden, andisochron


Gale 1980
age

six mineral fractions

207

Pb-206Pb

4.5622

0.0012

Amelin 2001

204

Pb/206PbPb/206Pb

4.5512

0.0032

phosphates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5505

0.0008

phosphates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.551

0.001

phosphates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5583

0.0077

phosphates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5546

0.0036

phosphates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5507

0.0026

phosphates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5629

0.0016

silicates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5621

0.0042

silicates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206/PbPb/206Pb

4.5626

0.0014

silicates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5627

0.0017

silicates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.5631

0.0007

silicates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/206PbPb/206Pb

4.562

0.0008

silicates

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/238UPb/238U

4.5523

0.0031

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

Pb/238UPb/238U

4.5624

0.0022

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

all fractions

207
204

#2 excluded

207
204

radiogenic (206Pb-204Pb > 200)

207
204

all fractions plus troilite

207
204

all fractions plus primordial Pb

207

all fractions plus Gpel, Manhs,


and Allgre 1994 analyses

204
207
204

all fractions

207
204

all fractions plus troilite

207
204

all fractions plus primordial Pb

207
204

chondrules

207
204

chondrules plus troilite

207
204

chondrules plus primordial Pb

207
206

phosphates (5)

207
206

isochron
age

silicates (8)

207

nine meteorite fragments

U-Pb

4.549

0.007

Abranches, Arden, andisochron


Gale 1980
age

five phosphate fractions

U-Pb

4.551

0.0035

Amelin 2000

isochron
age

six mineral fractions

U-Pb

4.5623

0.0013

3D
linear
regression
Amelin 2001

isochron
age

phosphates (5)

U-Pb

4.5513

0.0029

3D
linearAmelin, Ghosh,
regression
Rotenberg 2005

andisochron
age

silicates (8)

U-Pb

4.563

0.001

3D
linearAmelin, Ghosh,
regression
Rotenberg 2005

andisochron
age

phosphates (5)

235

U-207Pb

4.538

0.004

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

silicates (8)

235

U-207Pb

4.562

0.026

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

phosphates (5)

232

Th-208Pb

4.486

0.14

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

silicates (8)

232

Th-208Pb

4.594

0.15

Amelin, Ghosh,
Rotenberg 2005

andisochron
age

four phosphate samples plotted


with nine other samples of four
other chondrite meteorites
Sm-Nd

4.182

0.51

Rotenberg and Amelinisochron


2001
age

twelve
chondrules
and
phosphates plotted with twentytwo other samples with seven
other meteorites
Sm-Nd

4.588

0.1

Amelin and Rotenbergisochron


2004
age

twelve
chondrules
and
phosphates plotted with seventyseven other samples of other
chondrite meteorites
Sm-Nd

4.547

0.11

Amelin and Rotenbergisochron


2004
age

five phosphates
chondrule fractions

Sm-Nd

4.624

0.12

Amelin and Rotenbergisochron


2004
age

Hf-W

4.5628

0.001

Kleine et al 2008

Mn-Cr

4.5563

0.0016

Polnau and Lugmairisochron


2001
age

4.588

0.002

Brazzle et al. 1999

isochron
age

and

seven

eleven fractions
whole
rock,
silicates,
chondrite fractions

and

feldspar, temperature extractions


(800-1800C)
I-Xe

isochron
age

St. Marguerite (H4)


whole rock residueCDT

206

Pb-207Pb

4.5624

0.0011

Bouvier et al 2007

isochron
age

chondrule residueCDT

206

Pb-207Pb

4.5655

0.0012

Bouvier et al 2007

isochron
age

206

Pb-207Pb

4.5633

0.0011

Bouvier et al 2007

isochron
age

206

Pb-207Pb

4.5643

0.0008

Bouvier et al 2007

isochron
age

206

Pb-207Pb

4.5617

0.0012

Bouvier et al 2007

isochron
age

207

Pb-204Pb

4.5627

0.0006

Bouvier et al 2007

isochron
age

Radioisotope Dating of Meteorites


II: The Ordinary and Enstatite
Chondrites 267 mean of a whole
rock and chondrule ages of
Gpel, Manhs, and Allgre
(1994) and Bouvier et al (2007)
Pb-Pb

4.5644

0.0034

Kleine et al 2008

isochron
age

one magnetic and three nonmagnetic fractions


Hf-W

4.5653

0.0006

Kleine et al 2002

isochron
age

four non-magnetic fractions and


the mean of three analyses of a
metal fraction
Hf-W

4.5669

0.0005

Kleine et al 2008

isochron
age

internal isochron recalculated

4.5665

0.0005

Kleine et al 2008

isochron
age

whole rock, silicate and chromite


fractions
Mn-Cr

4.5649

0.0007

Polnau and Lugmairisochron


2001
age

seven fractions of minerals and


chondrules
Mn-Cr

4.5629

0.001

Trinquier et al 2008

isochron
age

phosphate,
temperature
extractions (8001800C)
I-Xe

4.565

0.006

Brazzle et al 1999

isochron
age

feldspar, temperature extractions


(8001800C)
I-Xe

4.567

0.002

Brazzle et al 1999

isochron
age

pyroxene-olivine
leachate

residue

pyroxene-olivine residueCDT
chondrulepyroxeneolivine
residue
whole
rock,
chondrules,
phosphates, olivene leachates
ten point plotted

Hf-W

Table 3. Model ages for some or all components of the H chondrite meteorites Allegan (H5), Forest Vale (H4), Guarena
(H6), Richardton (H5), and St. Marguerite (H4), with the details and literature sources.
Sample

Method

Reading

Err +/-

4.511

Note

Source

Type

0.011

Trieloff et al 2003

plateau age

Allegan (H5)
whole rock (with 100%, 24 out of 26
extractions)
Ar-Ar
fragments of the meteorite

206

Pb-207Pb 4.5477

0.0019

Gpel, Manhs, and


Allgre 1994
model age

fragments of the meteorite

206

Pb-207Pb 4.5359

0.0019

Gpel, Manhs, and


Allgre 1994
model age

fragments of the meteorite

206

Pb-207Pb 4.5385

0.0015

Gpel, Manhs, and


Allgre 1994
model age

phosphate separates

206

Pb-207Pb 4.5502

0.0007

Gpel, Manhs, and


Allgre 1994
model age

phosphate separates

206

Pb-207Pb 4.5563

0.0008

Gpel, Manhs, and


Allgre 1994
model age

Forest Vale (H4) whole rock (width


40%, 19 out of 42 extractions)
Ar-Ar

4.522

0.008

Trieloff et al 2003

whole rock

Ar-Ar

4.52

0.03

Turner, Enright, and


Hennessey 1978
plateau age

fragment of meteorite

206

Pb-207Pb 4.6142

0.0042

Gpel, Manhs, and


Allgre 1994
model age

phosphate separate

206

Pb-207Pb

0.0007

Gpel, Manhs, and


Allgre 1994
model age

plateau age

Guarena (H6)
Ar-Ar

4.44

0.03

Turner, Enright, and


Hennessey 1978
plateau age

Ar-Ar

4.445

0.008

Trieloff et al 2003

plateau age

feldspar separate (9/14 extractions,


90% Ar)
Ar-Ar

4.472

0.013

Trieloff et al 2003

plateau age

pyroxene separate (9/14 extractions,


80% Ar)
Ar-Ar

4.46

0.013

Trieloff et al 2003

plateau age

mean of the whole rock, feldspar and


pyroxene ages
Ar-Ar

4.454

0.006

Trieloff et al 2003

plateau age

4.44

0.06

Dalrymple 2004

plateau age

whole rock (10/24 extractions, 80% Ar)

Ar-Ar
two phosphate separates

206

Pb-207Pb 4.5044

0.0005

Gpel, Manhs, and


Allgre 1994
model age

two phosphate separates

206

Pb-207Pb 4.5044

0.0005

Gpel, Manhs, and


Allgre 1994
model age

two phosphate separates

206

Pb-207Pb 4.5056

0.0005

Gpel, Manhs, and


Allgre 1994
model age

fragment of meteorite

206

Pb-207Pb 4.5172

0.0018

Gpel, Manhs, and


Allgre 1994
model age

phosphates (5)

235

U-207Pb

4.538

0.004

Amelin, Ghosh, andisochron


Rotenberg 2005
age

silicates (8)

235

U-207Pb

4.562

0.026

Amelin, Ghosh, andisochron


Rotenberg 2005
age

phosphates (5)

232

Th-208Pb 4.486

0.14

Amelin, Ghosh, andisochron


Rotenberg 2005
age

silicates (8)

232

Th-208Pb 4.594

0.15

Amelin, Ghosh, andisochron


Rotenberg 2005
age

Richardton (H5)
whole rock (width 100%, 32 out of 32
extractions)
Ar-Ar

4.595

0.011

Trieloff et al 2003

whole rock

Ar-Ar

4.5

0.03

Turner, Enright, and


Hennessey 1978
plateau age

chondrules (6) plus matrix

Rb-Sr

4.5

Evensen et al 1979 model age

chondrules (6) plus matrix

Rb-Sr

4.44

Evensen et al 1979 model age

chondrules (6) plus matrix

Rb-Sr

4.48

Evensen et al 1979 model age

chondrules (6) plus matrix

Rb-Sr

4.59

Evensen et al 1979 model age

chondrules (6) plus matrix

Rb-Sr

4.5

Evensen et al 1979 model age

chondrules (6) plus matrix

Rb-Sr

4.46

Evensen et al 1979 model age

chondrules (6) plus matrix

Rb-Sr

4.8

Evensen et al 1979 model age

plateau age

mean of five phosphate fractions

Pb-Pb

whole rock

207

whole rock

207

Pb- Pb 4.604

whole rock

207

Pb-206Pb 4.478

0.008

Huey and Kohman


1973
model age

whole rock

207

Pb-206Pb 4.519

0.015

Tatsumoto, Knight,
and Allgre 1973
model age

phosphates

207

Pb-206Pb 4.5514

0.0006

Gpel, Manhs, and


Allgre 1994
model age

phosphates

207

Pb-206Pb 4.5534

0.0006

Gpel, Manhs, and


Allgre 1994
model age

plagioclase

207

Pb-206Pb 4.5644

0.0064

Amelin 2001

model age

olivine

207

Pb- Pb 4.5646

0.0037

Amelin 2001

model age

pyroxene

207

Pb-206Pb 4.5621

0.0009

Amelin 2001

model age

pyroxene

207

Pb- Pb 4.5627

0.001

Amelin 2001

model age

pyroxene

207

Pb-206Pb 4.5649

0.0023

Amelin 2001

model age

pyroxene

207

Pb-206Pb 4.5623

0.0007

Amelin 2001

model age

phosphate fractions

207

Pb-206Pb 4.554

0.001

Rotenberg
Amelin 2001

and
model age

phosphate fractions

207

Pb-206Pb 4.555

0.001

Rotenberg
Amelin 2001

and
model age

phosphate fractions

207

Pb-206Pb 4.56

0.002

Rotenberg
Amelin 2001

and
model age

phosphate fractions

207

Pb-206Pb 4.552

0.001

Rotenberg
Amelin 2001

and
model age

silicate chondrules

207

Pb-206Pb 4.572

0.001

Rotenberg
Amelin 2002

and
model age

silicate chondrules

207

Pb-206Pb 4.559

0.003

Rotenberg
Amelin 2002

and
model age

silicate chondrules

207

Pb-206Pb 4.562

0.001

Rotenberg
Amelin 2002

and
model age

silicate chondrules

207

Pb-206Pb 4.561

0.001

Rotenberg
Amelin 2002

and
model age

silicate chondrules

207

Pb-206Pb 4.563

0.002

Rotenberg
Amelin 2002

and
model age

silicate chondrules

207

Pb-206Pb 4.56

0.001

Rotenberg
Amelin 2002

and
model age

207

Pb-206Pb 4.554

0.0013

Phosphate 1

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5554

0.0007

Phosphate 2

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5531

0.0008

Phosphate 3

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5597

0.0016

Phosphate 4

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5519

0.0007

Phosphate 5

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.4593

0.005

Troilite

Amelin, Ghosh, and


Rotenberg 2005
model age

meteorite fractions and fragments using


primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973

4.5539

0.0028

Pb-206Pb 4.633
206

206

206

Amelin 2000

model age

Tilton 1973

model age

Tilton 1973

model age

meteorite fractions and fragments using


primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Tatsumoto, Knight, and
Allgre 1973
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite
meteorite fractions and fragments using
primodial Pb of Richardson troilite

207

Pb-206Pb 4.5666

0.0065

Low-density
fraction 1

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5638

0.0045

Low-density
fraction 2

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5608

0.0261

Olivine

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5621

0.0008

Chondrule
fragment 1

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5627

0.0008

Chondrule
fragment 2

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5623

0.0007

Chondrule
Fragment 3

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5626

0.0007

Chondrule 3

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5613

0.0008

Chondrule 4

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5608

0.001

Chondrule 8

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5608

0.001

Chondrule 8

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5574

0.0014

Phosphate 1

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5581

0.0008

Phosphate 2

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5566

0.0008

Phosphate 3

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5705

0.0018

Phosphate 4

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.553

0.0008

Phosphate 5

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5674

0.0052

Low-density
fraction 1

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5674

0.0037

Low-density
fraction 2

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5864

0.0249

Olivine

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5629

0.0008

Chondrule
fragment 1

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5635

0.0008

Chondrule
fragment 2

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5629

0.0007

Chondrule
Fragment 3

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5641

0.0007

Chondrule 3

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5623

0.0008

Chondrule 4

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5632

0.0011

Chondrule 8

Amelin, Ghosh, and


Rotenberg 2005
model age

fractions, weighted averages, corrected


using primodial Pb (all fractions)
fractions, weighted averages, corrected
using primodial Pb (#2 excluded)
fractions, weighted averages, corrected
using primodial Pb (radiogenic [206Pb204Pb > 200])

207

Pb-206Pb 4.553

0.0028

phosphates

Amelin, Ghosh, and


Rotenberg 2005
model age

207

Pb-206Pb 4.5533

0.0039

phosphates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.5535

0.0026

phosphates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.5531

0.0023

silicates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.5621

0.0005

silicates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.562

0.0007

silicates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.5568

0.0053

phosphates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.5562

0.0081

phosphates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.556

0.0039

phosphates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.556

0.004

silicates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.5632

0.0007

silicates

Amelin, Ghosh, and


Rotenberg 2005
model age

207Pb-206Pb

4.5631

0.0028

silicates

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 1

206Pb-238U

4.628

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 2

206Pb-238U

4.596

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 3

206Pb-238U

4.647

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 4

206Pb-238U

4.845

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 5

206Pb-238U

4.548

Amelin, Ghosh, and


Rotenberg 2005
model age

Troilite

206Pb-238U

6.473

Amelin, Ghosh, and


Rotenberg 2005
model age

Low-density fraction 1

206Pb-238U

5.165

Amelin, Ghosh, and


Rotenberg 2005
model age

Low-density fraction 2

206Pb-238U

4.724

Amelin, Ghosh, and


Rotenberg 2005
model age

Olivine

206Pb-238U

5.519

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule fragment 1

206Pb-238U

4.576

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule fragment 2

206Pb-238U

4.593

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule Fragment 3

206Pb-238U

4.554

Amelin, Ghosh, and


Rotenberg 2005
model age

fractions, weighted averages, corrected


using primodial Pb (all fractions plus
Gpel, Manhs, and Allgre 1994
analyses)
fractions, weighted averages, corrected
using primodial Pb (all fractions)
fractions, weighted averages, corrected
using primodial Pb (chondrules)
fractions, weighted averages, corrected
using Richardton troilite Pb (all
fractions)
fractions, weighted averages, corrected
using Richardton troilite Pb (#2
excluded)
fractions, weighted averages, corrected
using Richardton troilite Pb (radiogenic
[206Pb-204Pb >200])
fractions, weighted averages, corrected
using Richardton troilite Pb (all fractions
plus Gpel, Manhs, and Allgre 1994
analyses)
fractions, weighted averages, corrected
using Richardton troilite Pb (all
fractions)
fractions, weighted averages, corrected
using
Richardton
troilite
Pb
(chondrules)

Chondrule 3

206Pb-238U

4.643

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule 8

206Pb-238U

4.625

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 1

207Pb-235U

4.577

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 2

207Pb-235U

4.568

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 3

207Pb-235U

4.582

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 4

207Pb-235U

4.646

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 5

207Pb-235U

4.551

Amelin, Ghosh, and


Rotenberg 2005
model age

Troilite

207Pb-235U

5.013

Amelin, Ghosh, and


Rotenberg 2005
model age

Low-density fraction 1

207Pb-235U

4.743

Amelin, Ghosh, and


Rotenberg 2005
model age

Low-density fraction 2

207Pb-235U

4.612

Amelin, Ghosh, and


Rotenberg 2005
model age

Olivine

207Pb-235U

4.837

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule fragment 1

207Pb-235U

4.566

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule fragment 2

207Pb-235U

4.572

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule Fragment 3

207Pb-235U

4.56

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule 3

207Pb-235U

4.587

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule 8

207Pb-235U

4.58

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 1

208Pb-232Th

4.607

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 2

208Pb-232Th

4.619

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 3

208Pb-232Th

4.683

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 4

208Pb-232Th

4.98

Amelin, Ghosh, and


Rotenberg 2005
model age

Phosphate 5

208Pb-232Th

4.604

Amelin, Ghosh, and


Rotenberg 2005
model age

Low-density fraction 1

208Pb-232Th

4.6

Amelin, Ghosh, and


Rotenberg 2005
model age

Low-density fraction 2

208Pb-232Th

4.626

Amelin, Ghosh, and


Rotenberg 2005
model age

Olivine

208Pb-232Th

5.127

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule fragment 1

208Pb-232Th

4.582

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule fragment 2

208Pb-232Th

4.57

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule fragment 3

208Pb-232Th

4.565

Amelin, Ghosh, and


Rotenberg 2005
model age

Chondrule 3

208Pb-232Th

4.32

Amelin, Ghosh, and


Rotenberg 2005
model age

208Pb-232Th

4.249

whole rock (30/38 extractions, 93% Ar)

Ar-Ar

4.532

two phosphate fractions

206

two phosphate fractions

Chondrule 8

Amelin, Ghosh, and


Rotenberg 2005
model age

St. Marguerite (H4)


0.016

Trieloff et al 2003

plateau age

Pb-207Pb 4.563

0.0006

Gpel, Manhs, and


Allgre 1994
model age

206

Pb-207Pb 4.5627

0.0007

Gpel, Manhs, and


Allgre 1994
model age

two phosphate fractions

206

Pb-207Pb 4.5627

0.0006

Gpel, Manhs, and


Allgre 1994
model age

fragment of meteorite

206

Pb-207Pb 4.5667

0.0016

Gpel, Manhs, and


Allgre 1994
model age

Table 4. Isochron ages for some or all components of the L chondrite meteorites Bardwell (L5), Bjurble (L4), and
Bruderheim (L6), with the details and literature sources.
Sample

Method

Reading

Err +/-

Note

Source

Type

isochron age

Barwell (L5)
two samples plotted with 16 other points
from six other meteorites
two whole rocks, chondrules and troilie
fractions, plus -150 mesh fraction

207

Pb-206Pb 4.521

0.01

Unruh 1982

207

Pb-206Pb 4.557

0.008

Unruh, Hutchison,
Tatsumoto 1982

whole rock and troilite plotted with other L


chondrites
U-Pb

Unruh
1980

4.55

and

and
isochron age

Tatsumoto
isochron age

Bjurble (L4)
two silicate fractions with seven other
samples of two other L chondrites
Rb-Sr

0.054

Rotenberg
2002

and

4.54

Amelin
isochron age

silicate (2) and phosphate (3) fractions


plotted with silicate (7) and phosphate (4)
fractions of two other L chrondrites
Rb-Sr

0.04

Rotenberg
2002

and

4.499

Amelin
isochron age

two pyroxene fractions

Pb-Pb

4.5542

0.0046

Amelin 2001

pyroxenes

Pb-Pb

4.5543

0.0033

Rotenberg
2001
Unruh
1980

isochron age
and

and

Amelin
isochron age

triolite sample plotted with seven other L


chondrites
U-Pb

4.55

Tatsumoto

three phosphate fractions plotted with ten


other samples of four other chondrite
meteorites
Sm-Nd

4.182

0.51

Rotenberg
2001

four chondrules and three phosphates


plotted with 27 other samples of seven
other chondrite meteorites
Sm-Nd

0.1

Amelin
2004

and

4.588

four chondrules and three phosphates


plotted with 82 other samples of seven
other chondrite meteorites
Sm-Nd

0.11

Amelin
2004

and

4.547

whole rock

I-Xe

4.566

0.22

Hohenberg and Kennedy


1981; Brazzle et al. 1999 isochron age

whole rock and fractions (eight total)

Rb-Sr

4.54

15 whole rock samples

207

15 whole rock samples

U-Pb

isochron age
and

Amelin
isochron age

Rotenberg
isochron age
Rotenberg
isochron age

Bruderheim (L6)

Pb-206Pb 4.482

triolite and whole rock plotted with seven


other L-chrondrites
U-Pb

4.536

Shima and Honda 1967

isochron age

Gale.
Arden,
Abranches 1980

and

0.017

isochron age

Gale.
Arden,
Abranches 1980

and

0.006

Unruh
1980

4.55

and

isochron age

Tatsumoto
isochron age

Table 5. Model ages for some or all components of the L chondrite meteorites Bardwell (L5), Bjurble (L4), and Bruderheim
(L6), with the details and literature sources.
Sample

Method

Reading

Err +/-

Note

Source

Type

Barwell (L5)
Ar-Ar

4.45

0.06

Turner, Enright, and


Hennessey 1978
plateau age

Ar-Ar

4.45

0.03

Turner, Enright, and


Cardogan, 1979
plateau age
Gale, Arden, and
Hutchinson 1972
model age

whole rock sample

207

Pb-206Pb 4.68

troilite sample

207

Pb-206Pb 4.548

whole rock sample

207

Pb-206Pb 4.561

207

Pb-206Pb 4.549

0.004

Unruh 1982

model age

207

Pb-206Pb 4.545

0.003

Unruh 1982

model age

207

Pb-206Pb 4.56

0.005

Unruh 1982

model age

207

Pb-206Pb 4.56

0.005

meteorite fragments

207

Pb-206Pb 4.5577

0.0018

Gpel, Manhs, and


Allgre 1994
model age

meteorite fragments

207

Pb-206Pb 4.5522

0.001

Gpel, Manhs, and


Allgre 1994
model age

phosphate separates

207

Pb-206Pb 4.5382

0.0007

Gpel, Manhs, and


Allgre 1994
model age

phosphate separates

207

Pb-206Pb 4.5384

0.0008

Gpel, Manhs, and


Allgre 1994
model age

207

Pb-208Pb 4.583

0.025

Unruh 1982

model age

207

Pb-208Pb 4.578

0.025

Unruh 1982

model age

silicate fractions of two whole rock


samples
silicate fractions of two whole rock
samples
contaminant-corrected concordant
model ages for two silicate fractions
of whole-rock samples

silicate fractions of two whole rock


samples
contaminant-corrected concordant
model ages for two silicate fractions
of whole rock samples

Unruh and Tatsumoto


1980
model age

0.004

Unruh and Tatsumoto


1980
model age

model age

two samples plotted with 16 other


points from six other meteorites
U-Pb

4.547

0.015

Unruh 1982

concordia age

triolite-corrected
data
for
two
samples plotted with 16 other data
points from six other meteorites
U-Pb

4.551

0.007

Unruh 1982

concordia age

U-Pb

4.564

0.005

Unruh 1982

concordia age

whole rock

K-Ar

4.32

K-Ar

4.394

0.016

Podosek and Hunekestep


630C extraction 1973b
age

model

whole rock

K-Ar

4.369

0.005

Podesek and Hunekestep


745C extraction 1973b
age

model

whole rock

K-Ar

4.413

0.009

Podosek & Hunekestep


850C extraction 1973b
age

model

whole rock

K-Ar

4.438

0.015

Podesek and Hunekestep


935C extraction 1973b
age

model

whole rock

K-Ar

4.334

0.029

1040C
extraction

Podosek and Hunekestep


1973b
age

model

whole rock

K-Ar

4.518

0.005

1125C
extraction

Podesek and Hunekestep


1973b
age

model

whole rock

K-Ar

4.496

0.004

1195C
extraction

Podosek and Hunekestep


1973b
age

model

whole rock

Bjurble (L4)
Geiss
and
Hess
1958; Wood 1967
model age

K-Ar

4.486

0.002

1370C
extraction

Podesek and Hunekestep


1973b
age

model

whole rock

K-Ar

4.513

0.005

1515C
extraction

Podesek and Hunekestep


1973b
age

model

whole rock
whole rock

K-Ar

4.51

high
temperature

Podesek and Huneke


1973b
plateau age

whole rock

Ar-Ar

4.51

triolite

207

Pb-206Pb 4.556

whole rock

207

Pb-206Pb 4.583

whole rock

207

Pb-206Pb 4.59

0.006

Unruh 1982

207

Pb-206Pb 4.5543

0.0033

Rotenberg
Amelin 2001

and
model age

silicate fractions

207

Pb-206Pb 4.556

0.004

Rotenberg
Amelin 2001

and
model age

silicate fractions

207

Pb-206Pb 4.552

0.003

Rotenberg
Amelin 2001

and
model age

phosphate fractions

207

Pb-206Pb 4.59

0.011

Rotenberg
Amelin 2001

and
model age

phosphate fractions

207

Pb-206Pb 4.661

0.011

Rotenberg & Amelin


2001
model age

phosphate fractions

207

Pb-206Pb 4.519

0.002

Rotenberg & Amelin


2001
model age

pyroxene

207

Pb-206Pb 4.5545

0.0012

Amelin 2001

model age

pyroxene

207

Pb-206Pb 4.5546

0.0018

Amelin 2001

model age

whole rock

U-Th/He

0.08

Turner 1969

plateau age

0.01

Unruh and Tatsumoto


1980
model age
Unruh and Tatsumoto
1980
model age
model age

Eberhardt and Hess


1960; Wood 1967
model age

4.2

Bruderheim (L6)
whole rock

207

Pb-206Pb 4.53

whole rock

207

Pb-206Pb 4.518

whole rock sample

207

Pb- Pb 4.605

triolite

207

Pb-206Pb 4.537

whole rock

207

Pb-206Pb 4.55

whole rock

207

Pb-206Pb 4.535

whole rock sample

206

Pb-238U

whole rock sample

206

whole rock sample

Gale, Arden, and


Hutchinson 1972
model age
Huey
1973

0.003

206

and

Tilton 1973

Kohman
model age
model age

Unruh and Tatsumoto


1980
model age

0.004

Unruh and Tatsumoto


1980
model age
Unruh 1982

model age

4.126

Tilton 1973

model age

Pb-238U

4.542

Tilton 1973

model age

206

Pb-238U

4.959

Tilton 1973

model age

whole rock sample

207

Pb-235U

4.447

Tilton 1973

model age

whole rock sample

207

Pb- U

4.592

Tilton 1973

model age

whole rock sample

207

Pb-235U

4.703

Tilton 1973

model age

235

0.004

Table 6. Isochron ages for some or all components of the LL chondrite meteorites Olivenza (LL5) and St. Sverin (LL6), with
the details and literature sources.
Sample

Method

Reading

Err +/-

4.63

0.16

Note

Source

Type

Olivenza (LL5)
18 fractionswhole rock, chondrules,
pyroxene, olivene, density splits and
leachates
Rb-Sr

Sanz and Wasserburg 1969 isochron age

Rb-Sr

4.53

0.16

Dalrymple 1991

isochron age

4.486

0.02

Minster and Allgre 1981

isochron age

Rb-Sr

4.56

0.15

These four on isochron


Gopalan and Wetherill 1969 isochron age

whole rock + heavy liquid separates +


whitlockite
Rb-Sr

4.61

0.15

Manhs, Minster, and Allgre


1978
isochron age

two whole rock samples with ten other


meteorites
Rb-Sr

4.486

0.02

Minster and Allgre 1981

isochron age

recalculated

Rb-Sr

4.51

0.15

Minster and Allgre 1981

isochron age

one sample plotted with five iron


meteorites
Re-Os

4.58

0.21

Luck, Birck, and Allgre 1980isochron age

one sample plotted with 21 other


meteorites
Re-Os

4.55

five whole rock, two metal, one FeS


analyses
Re-Os

4.68

0.15

Chen, Papanastassiou, and


Wasserburg 1998
isochron age

whole rock, px + ol, ap + met, 3


leachates
Pb-Pb

4.5759

0.009

Bouvier et al 2007

isochron age

phosphates

Pb-Pb

4.5549

0.0002

Bouvier et al 2007

isochron age

whole rock + plagioclase + whitlockite

206

Pb-207Pb 4.543

0.019

Manhs, Minster, and Allgre


1978
isochron age

208

Pb-206Pb 4.55

207

Pb-204Pb 4.558

one whole rock plotted with ten other


meteorites
Rb-Sr
St. Sverin (LL6)
four fractions
meteorites

ploted

with

other

whole rock + plagioclase + Canyon


Diablo triolite
phosphates (Manhs, Minster, and
Allgre1978; Chen and Wasserburg
1981; Gpel, Manhs, and Allgre
1994)

Luck and Allgre 1983

isochron age

Manhs, Minster, and Allgre


1978
isochron age

0.006

Tera and Carlson 1999

isochron age

whole rock, whitlockite, light and dark


fractions
Sm-Nd

4.55

0.33

Jacobson and Wasserburg


1984
isochron age

whole rock, silicates, and chromitespinel fractions


Mn-Cr

4.5546

0.0014

Glavin and Lugmair 2003

isochron age

feldspar, temperature extractions (800


1800C)
I-Xe

4.558

0.004

Brazzle et al 1999

isochron age

Table 7. Model ages for some or all components of the LL chondrite meteorites Olivenza (LL5) and St. Sverin (LL6), with
the details and literature sources.
Sample

Method

Reading

Err +/-

Note

Source

Type

Ar-Ar

4.49

0.06

Turner, Enright, and


Hennessey 1978
plateau age

one sample

K-Ar

4.38

0.06

Funkhouser, Kirsten,
and Schaeffer 1967 model age

whole rock sample

K-Ar

4.6

0.05

used Ar-Ar measurements

Podosek 1971

14 samples from drill


core
K-Ar

4.4

0.45

used Ar-Ar measurements

Schultz
1976

used
as
irradiated,
heating

Ar-Ar

4.56

0.05

Alexander, Davis, and


Lewis 1972
plateau age

Ar-Ar

4.5

0.03

Podosek and Huneke


1973a
model age

weighted
average
calculated from three
standards irradiated
Ar-Ar

4.504

0.02

Alexander and Davis


1974
model age

light and dark fractions

4.383

0.03

Hohenberg et al 1981 plateau age

Olivenza (LL5)

St. Sverin (LL6)

monitor,
stepwise-

Ar-Ar

and

plateau age
Signer
plateau age

Ar-Ar

4.42

0.03

plateau age

Ar-Ar

4.333

0.03

total age

Ar-Ar

4.359

0.03

total age

light
fraction
(Hohenberg et al 1981) Ar-Ar

4.4313

Ar-Ar

4.4053

dark
fraction
(Hohenberg et al 1981) Ar-Ar

4.4688

Ar-Ar

4.4424

single
separate

using
revised
constants

decayMin, Reiners,
Shuster 2013

one whitlockite analysis


only
one whitlockite analysis
only

plateau age
plateau age

using
revised
constants

decayMin, Reiners,
Shuster 2013

and
plateau age
plateau age

phosphate

one whitlockite analysis


only

and

206

Pb-207Pb 4.5536

0.0007

Gpel, Manhs, and


Allgre 1994
model age

206

Pb-207Pb 4.5571

0.0015

model age

206

Pb-207Pb 4.55

0.01

Manhs, Minster, and


Allgre 1978
model age

208

Pb-232Th 4.57

0.05

Manhs, Minster, and


Allgre 1978
model age

238

U-206Pb

4.52

0.04

Manhs, Minster, andconcordia


Allgre 1978
age

235

U-207Pb

4.54

0.02

concordia
age

five
phosphates
(merrillite) grains
U-Th/He

4.412

0.075

weighted mean of five oldest


grains out of fourteenMin, Reiners,
analyzed
Shuster 2013

four
phosphates
(merrillite) grains
U-Th/He

4.152

0.07

weighted mean of four oldestMin, Reiners,


grains out of five analyzed Shuster 2013

interior
sample

whole

U-Th/He

4.1

0.15

Eugster 1988

revised,
ADOR

anchored

I-Xe

4.556

0.04

Glavin and Lugmair


2003
model age

and
model age
and
model age

rock
to

model age

Table 8. Isochron ages for some or all components of the E chondrite meteorites Abee (EH4), Hvittis (EL6), Indarch (EH4),
St. Marks (EH5), and St. Sauveur (EH5), with the details and literature sources.
Sample

Method

Reading

Err +/-

Note

Source

Type

Rb-Sr

4.52

two whole rock samples plotted with 12 whole


rock samples from seven other meteorites
Rb-Sr

4.54

0.13

Goplan and Wetherill


1970
isochron age

nine whole rock and density fractions

4.51

0.1

Minster, Rickard, and


Allgre 1979
isochron age

one whole rock samples plotted with seven


other whole rock samples from three other Echondrite meteorites
Rb-Sr

4.516

0.029

Minster, Rickard, and


Allgre 1979
isochron age

one whole rock samples plotted with seven


other whole rock samples from three other Echondrite meteorites
Rb-Sr

4.508

0.037

Minster, Birck,
Allgre 1982

207

Pb-206Pb 4.505

0.008

Huey
1973

Pb-206Pb 4.578

0.007

Bogard, Unruh,
Tatsumoto 1983

and

207

Abee (EH4)
whole rock and three fractions

one whole rock sample plotted with 16 other


meteorites
10 fractions from three clasts plotted with
Manhs and Allgre, 1978 analyses of four
meteorites

Rb-Sr

Shima
1967

and

Honda
isochron age

and

and
isochron age

Kohman
isochron age

isochron age

Ar-Ar

4.547

0.006

Bogard, Dixon,
Garrison 2010

and

Hvittis (EL6) 795% heating steps

Ar-Ar

4.569

0.008

Bogard, Dixon,
Garrison 2010

and

795% heating steps

4.54

0.13

Gopalan and Wetherillisochron age

one whole rock sample plotted with 13 whole Rb-Sr

isochron age
isochron age

rock samples from other meterorites

1970

Indarch (EH4) three whole rock samples


plotted with five other whole rock samples
from three other meteorites
Rb-Sr

4.516

0.029

Minster, Rickard, and


Allgre 1979
isochron age

three whole rock E-chondrite samples plotted


with five other E-chondrite samples
Rb-Sr

4.508

0.037

Minster, Birck,
Allgre 1982

updated decay constant applied to Gopalan


and Wetherill 1970
Rb-Sr

4.46

0.08

Dalrymple 1991

Rb-Sr

4.52

0.15

Bogard, Dixon,
Garrison 2010

and

updated with newer decay constant

Rb-Sr

4.449

0.043

Bogard, Dixon,
Garrison 2010

and

updated with newer decay constant

Rb-Sr

4.5

0.13

Bogard, Dixon,
Garrison 2010

and

updated with newer decay constant

St. Marks (EH5) two whole rock samples


plotted with 12 other whole rock samples of
seven other meteorites
Rb-Sr

4.54

0.13

Goplan and Wetherill


1970
isochron age

three whole rock samples plotted with five


other whole rock samples from three other Echondrite meteorites
Rb-Sr

4.516

0.029

Minster,
Rickard,and
Allgre 1979
isochron age

nine fractions of whole rock

Rb-Sr

4.335

0.05

Minster, Rickard, and


Allgre 1979
isochron age

three whole rock samples plotted with five


other whole rock samples from three other Echondrite meteorites
Rb-Sr

0.037

Minster, Birck,
Allgre 1982

and

4.508

Rb-Sr

4.391

0.05

Bogard, Dixon,
Garrison 2010

and

updated with newer decay constant

one whole rock sample plotted with seven


other whole rock samples from three other Echondrite meteorites
Rb-Sr

4.516

0.029

Minster, Rickard, and


Allgre 1979
isochron age

nine fractions of whole rock

Rb-Sr

4.457

0.047

Minster, Rickard, and


Allgre 1979
isochron age

one whole rock sample plotted with seven


other whole rock samples from three other Echondrite meteorites
Rb-Sr

4.508

0.037

Minster et al 1982

updated with newer decay constant

4.514

0.047

Bogard, Dixon,
Garrison 2010

Pb-206Pb 4.577

0.004

Manhs
1978

and
isochron age
isochron age
isochron age
isochron age
isochron age

isochron age
isochron age

St. Sauveur (EH5)

Rb-Sr

one whole rock sample plotted with three


other E-chondrite meteorites

207

and

isochron age
and
isochron age

Allgre
isochron age

Table 9. Model ages for some or all components of the E chondrite meteorites Abee (EH4), Hvittis (EL6), Indarch (EH4), St.
Marks (EH5), and St. Sauveur (EH5), with the details and literature sources.
Sample

Method

Reading

Err +/-

Note

Source

Type

Ar-Ar

4.5

0.03

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 1, 1, 04

Ar-Ar

4.52

0.03

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 2, 2, 05

Ar-Ar

4.49

0.03

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06

Ar-Ar

4.1

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 1, 1, 04 (350C)

Ar-Ar

4.43

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 1, 1, 04 (450C)

Ar-Ar

4.5

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 1, 1, 04 (525C)

Abee (EH4)
plateau age
plateau age
plateau age
step-heating age
step-heating age
step-heating age

Bogard,
Unruh,
Tatsumoto 1983

and

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

4.52

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Ar-Ar

4.5

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 1, 1, 04 (875C)

Ar-Ar

4.53

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 1, 1, 04 (975C)

Ar-Ar

4.48

0.03

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 1, 1, 04 (1090C)

Ar-Ar

4.39

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 1, 1, 04 (1250C)

Ar-Ar

4.13

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 1, 1, 04 (1400C)

Ar-Ar

3.63

0.06

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 1, 1, 04 (1500C)

Ar-Ar

3.56

0.06

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (300C)

Ar-Ar

7.2

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (400C)

Ar-Ar

4.8

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (500C)

Ar-Ar

4.33

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (600C)

Ar-Ar

4.5

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (675C)

Ar-Ar

4.52

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (750C)

Ar-Ar

4.52

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (800C)

Ar-Ar

4.54

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (900C)

Ar-Ar

4.5

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (1000C)

Ar-Ar

4.43

0.03

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (1150C)

Ar-Ar

4.27

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (1300C)

Ar-Ar

4.03

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 2, 2, 05 (1550C)

Ar-Ar

3.99

0.09

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 3, 3, 06 (300C)

Ar-Ar

8.9

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 3, 3, 06 (400C)

Ar-Ar

5.4

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 3, 3, 06 (500C)

Ar-Ar

3.82

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 3, 3, 06 (600C)

Ar-Ar

4.15

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 3, 3, 06 (675C)

Ar-Ar

4.31

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 3, 3, 06 (725C)

Ar-Ar

4.38

0.02

Bogard,
Unruh,
Tatsumoto 1983

andstep-heating age

Clast 1, 1, 04 (600C)

Ar-Ar

4.38

0.02

Clast 1, 1, 04 (650C)

Ar-Ar

4.41

Clast 1, 1, 04 (725C)

Ar-Ar

Clast 1, 1, 04 (800C)

step-heating age

Ar-Ar

4.47

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06 (775C)

Ar-Ar

4.49

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06 (850C)

Ar-Ar

4.39

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06 (950C)

Ar-Ar

4.29

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06 (1050C)

Ar-Ar

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06 (1175C)

Ar-Ar

3.76

0.02

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06 (1325C)

Ar-Ar

3.58

0.09

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 3, 3, 06 (1500C)

207

Pb-208Pb

4.56

0.15

Bogard,
Unruh,
Tatsumoto 1983

and

Clast 1, 1, 01; fraction 1, 1, I

Pb-208Pb

4.72

0.05

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.56

0.1

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.45

0.07

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.7

0.1

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.54

0.13

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.41

0.03

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.56

0.1

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.7

0.06

Bogard,
Unruh,
Tatsumoto 1983

and

207

Pb-208Pb

4.55

0.05

Bogard,
Unruh,
Tatsumoto 1983

and

207

Clast 1, 1, 01; fraction 1, 1,


E1
Clast 1, 1, 01; fraction 1, 1,
E2
Clast 1, 1, 01; fraction 1, 1,
E2 (H2O-L)
Clast 2, 2, 02; fraction 2, 2, I
Clast 2, 2, 02; fraction 2, 2,
E
Clast 2, 2, 02; fraction 2, 2,
E (H2O-L)
Clast 3, 3, 07; fraction 3, 3, I
Clast 3, 3, 07; fraction 3, 3,
E
Clast 3, 3, 07; fraction 3, 3,
E (H2O-L)

step-heating age
step-heating age
step-heating age
step-heating age
step-heating age
step-heating age
step-heating age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age

Hvittis (EL6)
clast (whole rock)

Ar-Ar

4.47

Kinsey et al 1995

plateau age

Ar-Ar

4.544

0.018

Bogard, Dixon, and Garrison


2010
plateau age

7-95% heating steps

Ar-Ar

4.494

0.046

Bogard, Dixon, and Garrison


2010
plateau age

250C

Ar-Ar

4.2276

0.0131

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

300C

Ar-Ar

3.7347

0.0078

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

350C

Ar-Ar

3.6969

0.0054

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

400C

Ar-Ar

3.9411

0.0057

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

475C

Ar-Ar

4.2912

0.0049

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

550C

Ar-Ar

4.5024

0.0049

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

600C

Ar-Ar

4.5285

0.0048

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

650C

Ar-Ar

4.5295

0.0047

Bogard, Dixon, and Garrisonmodel (step heating)

2010

age

700C

Ar-Ar

4.5457

0.0044

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

725C

Ar-Ar

4.5535

0.0045

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

750C

Ar-Ar

4.5606

0.0044

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

775C

Ar-Ar

4.56

0.0046

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

825C

Ar-Ar

4.5557

0.0045

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

875C

Ar-Ar

4.5393

0.0046

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

925C

Ar-Ar

4.5112

0.0046

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

975C

Ar-Ar

4.4783

0.0052

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1025C

Ar-Ar

4.4222

0.005

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1100C

Ar-Ar

4.4498

0.0045

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1200C

Ar-Ar

4.4444

0.0043

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1300C

Ar-Ar

4.1384

0.0047

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1400C

Ar-Ar

4.4291

0.0283

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

29-83% five extractions

Ar-Ar

4.249

0.013

Bogard, Dixon, and Garrison


2010
model (plateau) age

8399% extractions

Ar-Ar

4.351

0.008

Bogard, Dixon, and Garrison


2010
model (plateau) age

525C

Ar-Ar

3.8659

0.0059

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

625C

Ar-Ar

4.0582

0.005

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

700C

Ar-Ar

4.1812

0.0044

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

775C

Ar-Ar

4.2169

0.0036

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

825C

Ar-Ar

4.255

0.004

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

875C

Ar-Ar

4.2538

0.0048

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

925C

Ar-Ar

4.2492

0.0038

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

975C

Ar-Ar

4.2275

0.0038

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1075C

Ar-Ar

4.2525

0.0035

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1125C

Ar-Ar

4.2977

0.0037

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1225C

Ar-Ar

4.2958

0.0048

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1325C

Ar-Ar

4.3511

0.0073

Bogard, Dixon, and Garrisonmodel (step heating)

Indarch (EH4)

2010

age

1450C

Ar-Ar

4.1725

0.0149

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

whole rock

K-Ar

4.2

0.1

Schaeffer and Stoenner 1965 model age

enstatite

K-Ar

4.45

0.16

Schaeffer and Stoenner 1965 model age

relative to Shallowater

I-Xe

4.56

Busfield, Turner, and Gilmour


2008
model age

St. Marks (EH5) 7191%


extraction
Ar-Ar

4.433

0.004

Bogard, Dixon, and Garrison


2010
model (maximum) age

5971% extraction

Ar-Ar

4.411

0.005

Bogard, Dixon, and Garrison


2010
model (maximum) age

950C

Ar-Ar

3.7429

0.0103

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1050C

Ar-Ar

4.0609

0.0051

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1125C

Ar-Ar

4.2363

0.004

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1200C

Ar-Ar

4.236

0.0058

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1275C

Ar-Ar

4.3382

0.0039

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1325C

Ar-Ar

4.4107

0.004

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1375C

Ar-Ar

4.4329

0.0031

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1450C

Ar-Ar

4.3504

0.0066

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

1600C

Ar-Ar

4.1247

0.0666

Bogard, Dixon, and Garrisonmodel (step heating)


2010
age

relative to Shallowater

I-Xe

4.56

Busfield, Turner, and Gilmour


2008
model age

I-Xe

4.56

Busfield, Turner, and Gilmour


2008
model age

St. Sauveur (EH5)


relative to Shallowater

Discussion
In contrast to the Allende CV3 carbonaceous chondrite meteorite (Snelling 2014), there have been fewer radioisotope
methods used on these meteorites and therefore fewer radioisotope ages obtained. However, even a cursory examination of
Figs. 916 reveals that there is still a clustering of radioisotope ages, both isochron and model ages, around 4.554.57 Ga.
And where there are larger numbers of radioisotope ages available the clustering around 4.554.57 Ga is very pronounced,
similar to the pattern found for Allende CV3 carbonaceous chondrite by Snelling (2014). Again this clustering is dominated
by Pb-Pb isochron and model ages, and Pb-Pb calibrated Mn-Cr, Hf-W, and I-Xe ages, but it is also supported by some UPb, Th-Pb, Rb-Sr, Sm-Nd, Ar-Ar, and Re-Os ages. There is also much scattering of K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, U-Pb, ThPb, Re-Os, and U-Th/He ages, though the pattern varies from meteorite to meteorite and depends on which methods were
applied to them.
The H Chondrites
The Richardton (H5) meteorite has been the most radioisotope dated of the H chondrites, and the clustering of both its PbPb isochron and model ages at 4.554.57 Ga is very strong (figs 9 and 10). St. Marguerite (H4) has only Pb-Pb isochron
ages that are around 4.564.57 Ga. As to be expected, by definition the Mn-Cr and Hf-W isochron ages coincide with the St.
Marguerites Pb-Pb isochron ages because they have been calibrated against the St. Marguerite meteorites Pb-Pb isochron
and model ages (figs. 9 and 10) (Gpel, Manhs, and Allgre 1994; Polnau and Lugmair 2001; Kleine et al. 2002, 2008;
Trinquier et al. 2008), and the I-Xe isochron ages similarly coincide with the Pb-Pb isochron ages (fig. 9), because they are
calibrated against the I-Xe isochron age of the Shallowater achondrite, which in turn is calibrated against the Pb-Pb ages of
several other meteorites (Brazzle et al. 1999; Gilmour et al. 2006, 2009). Thus the Richardton (H5) meteorites Mn-Cr, Hf-W,
and I-Xe isochron ages also coincide with its Pb-Pb isochron ages at 4.564.57 Ga for the same reasons. But Richardtons
Pb-Pb isochron and model 4.564.57 Ga ages are both supported by U-Pb isochron and model ages and a Th-Pb model
age (figs. 9 and 10). Nevertheless, there is also some scatter of U-Pb, Th-Pb, Sm-Nd, and Rb-Sr isochron ages for
Richardton (H5) either side of this strong 4.564.57 Ga clustering, plus two outlying Rb-Sr isochron ages and one outlying
Sm-Nd isochron age (table 2 and fig. 9). In contrast, there is considerable wide scatter of U-Pb, Th-Pb, Ar-Ar, and Rb-Sr
model ages for Richardton (H5) (table 3 and fig. 10). And there is no real pattern to this scatter. For both isochron and model
ages there are U-Pb, Th-Pb, Rb-Sr, Sm-Nd, and Ar-Ar ages respectively either side of the strong 4.564.57 Ga clustering,
although the U-Pb isochron ages are all lower (younger) than the clustering, while the U-Pb model ages are nearly all above
(older) than the clustering. Furthermore, when the U-Pb and Th-Pb model ages for the same sample fractions in the same
study are compared [for example, the Amelin, Ghosh, and Rotenberg (2005) data in table 3], the U-Pb model ages are for

some sample fractions older than the Th-Pb model ages, and for other sample fractions younger than the Th-Pb model
ages.The other three H chondrites in tables 2 and 3, and figs. 9 and 10, have only a few radioisotope age data for them.
Allegan (H5) has one Pb-Pb 4.56 Ga isochron age, one I-Xe isochron age (which via calibration agrees with the Pb-Pb
isochron age) and two Pb-Pb 4.56 Ga model ages. Forest Vale (H4) has only two Mn-Cr isochron ages that are calibrated
by St. Marguerites Pb-Pb isochron age (Polnau and Lugmair 2000, 2001), so by definition there is agreement. In contrast,
the Guarena (H6) meteorite only has one Sm-Nd isochron age which is older than 4.564.57 Ga, whereas the four Rb-Sr
isochron ages are scattered, with one at 4.56 Ga, one above 4.56 Ga, and two below 4.50 Ga. Among the model ages for
these three H chondrites, the Ar-Ar model ages are all younger than their corresponding Pb-Pb-Pb model ages (fig. 10).
However, whereas one Pb-Pb model age for Forest Vale (H4) is older than the other 4.56 Ga Pb-Pb model age, all four PbPb model ages for Guarena (H6) are younger than the 4.564.57 Ga target, and all six Ar-Ar model ages are much
younger. This could well be related to the classification of these meteorites on a scale of increasing thermal metamorphism
from H4 through to H6 (fig. 6) based on the observable and measurable criteria listed in fig. 5 (Norton 2002; Van Schmus
and Wood 1967). It is thus logical that the U-Pb and K-Ar systems in the more thermally metamorphosed (to higher
temperatures) Guarena (H6) meteorite have been disturbed, some of the daughter Ar gas particularly having been lost and
thus resulting in younger Ar-Ar measured ages.
The L Chondrites
All three meteorites investigated have only a few isochron and model ages via only a few radioisotope dating methods
(tables 4 and 5, and figs. 11 and 12). However, the clustering of the radioisotope ages is again around the 4.554.57 Ga
mark. This target date was achieved by both Pb-Pb and U-Pb isochron ages for Bardwell (L5) and Bjurble (L4), and by
both Pb-Pb and U-Pb model ages for Bardwell (L5), but only by Pb-Pb model ages for Bjurble (L4). One Rb-Sr, one I-Xe,
and two Sm-Nd isochron ages also cluster around the 4.554.57 Ga mark for Bjurble (L4), but one Rb-Sr isochron age is
slightly younger and one Sm-Nd isochron age is very much younger. The I-Xe isochron age, though initially designated as
the I-Xe standard (Hohenberg and Kennedy 1981), by definition agrees with Bjurbles Pb-Pb isochron age, because its I-Xe
isochron age has also been calibrated against the I-Xe isochron age of the Shallowater achondrite, which in turn is
calibrated against the Pb-Pb ages of several other meteorites (Brazzle et al. 1999; Gilmour et al. 2006, 2009). The other
patterns are that for both Bardwell (L5) and Bjurble (L4) the K-Ar and Ar-Ar model ages are consistently younger than the
target age, while one Pb-Pb model age for each meteorite is well above (much older) than the 4.554.57 Ga cluster.
The sole U-Th/He model age for Bjurble (L4), similar to its K-Ar and Ar-Ar model ages, is much younger than the 4.554.57
Ga mark (fig. 12), probably because these methods depend on daughter isotopes He and Ar that are noble (inert) gases of
light atomic weights and small sizes which therefore are prone to diffusing away from their parent radioisotopes and
escaping completely from the host mineral lattices. Indeed, it has been suggested that severe heating after meteorite
formation may cause total loss of both gases, which results in the He and Ar ages being younger and concordant, with both
clocks reset to zero at the time of the heating event after the meteorites formed (Lewis 1997). Furthermore, where U-Th/He
ages are younger than K-Ar and Ar-Ar ages, as here for the Bjurble (L4) chondrite (fig. 12), it has been suggested that
even moderate heating of the meteorites would have caused He to diffuse out of the mineral grains in a time too short for
major loss of the more slowly diffusing Ar.As a more thermally metamorphosed meteorite, Bruderheim (L6) displays a
slightly different pattern, consistent with disturbance of the radioisotope systems. Among the few isochron ages it is the two
U-Pb and single Rb-Sr isochron ages that cluster around the 4.554.57 Ga mark, while the single Pb-Pb isochron age is
less than 4.50 Ga (fig. 11). However, among the model ages it is the Pb-Pb model ages that cluster close to and just below
the 4.554.57 Ga mark, while the U-Pb model ages are widely scattered both well above (older) and well below (younger),
and around the 4.554.57 Ga mark.
The LL Chondrites
Only two LL chondrites have been radioisotope dated multiple times. For the Olivenza (LL5) meteorite only Rb-Sr isochron
and Ar-Ar model ages have been obtained (figs. 13 and 14), and these are somewhat scattered with respect to the 4.55
4.57 Ga mark, with one Rb-Sr isochron age above and the other ages below. In contrast, the St. Sverin (LL6) meteorites
age has been well constrained by one Rb-Sr, one Re-Os, three Pb-Pb, and one Sm-Nd 4.564.57 Ga isochron ages (fig.
13), and by three Pb-Pb, one Th-Pb, and one Ar-Ar 4.564.57 Ga model ages (fig. 14). These 4.564.57 Ga ages are also
supported by Mn-Cr isochron and I-Xe isochron and model ages, as is to be expected because of these methods being
calibrated against Pb-Pb isochron and model ages for other meteorites (Brazzle et al. 1999; Gilmour et al. 2006, 2009;
Kleine et al. 2008; Polnau and Lugmair 2001; Trinquier et al. 2008). Again there is also scatter, with Re-Os, Rb-Sr, and PbPb isochron ages above the 4.564.57 Ga mark, and Rb-Sr and Pb-Pb isochron ages below. In contrast, apart from one KAr model age above the 4.564.57 Ga mark (table 7), all the other K-Ar and Ar-Ar (bar one) model ages, and the U-Pb and
U-Th/He model ages, are below the 4.564.57 Ga mark. In fact, many of the Ar-Ar, K-Ar, and U-Th/He model ages are well
below the 4.564.57 Ga mark (table 7), with a clustering of such model ages centred around the 4.004.20 Ga mark. Two UTh/He ages are lower outliers. The probable explanation for this pattern would seem to be that these methods depend on
daughter isotopes that are noble (inert) gases of light atomic weights and small sizes which therefore are prone to diffusing
away from their parent radioisotopes and escaping completely from the host mineral lattices, especially due to a heating
event subsequent to meteorite formation (Lewis 1997). Conventionally this 4.004.20 Ga age clustering would be identified
as the age of such a re-heating event that reset the K-Ar, Ar-Ar, and U-Th/He model ages due to the loss of Ar and He gases
(Bogard 2011; Min, Reiners, and Shuster 2013; Trieloff et al. 2003).
The E Chondrites
Five E chondrites have been dated by more than one radioisotope methodfour by the Rb-Sr and Pb-Pb isochron methods
and one by the Ar-Ar and Rb-Sr isochron methods (fig. 15); two by the Ar-Ar and Pb-Pb model age methods, two by the KAr, Ar-Ar, and I-Xe model age methods, and one just by the I-Xe model age method (fig. 16). None of the methods produced
results on these meteorites that clustered at the 4.564.57 Ga mark, except for the I-Xe model ages for Indarch (EH4), St.
Marks (EH5), and St. Sauveur (EH5). These are by definition in agreement with this 4.564.57 age, because I-Xe ages are
always calibrated against the I-Xe age of the Shallowater achondrite (Busfield, Turner, and Gilmour 2008), which in turn is
calibrated against the Pb-Pb ages of several other meteorites that date at the 4.564.57 Ga mark (Brazzle et al. 1999;
Gilmour et al. 2006, 2009). However, some individual age determinations did produce results in the 4.564.57 Ga range
one Ar-Ar isochron age for Hvittis (EL6), one Rb-Sr isochron age for Indarch (EH4), and one Pb-Pb isochron age for St.
Sauveur (EH5) (fig.15); four Pb-Pb model ages for Abee (EH4), and four Ar-Ar model ages for Hvittis (EL6) (fig.16).
Otherwise, in general the Rb-Sr isochron ages for these meteorites are younger than the 4.564.57 Ga target age (except
for the one Rb-Sr isochron age for Indarch that is in that range), and younger than or equal to all of both the Pb-Pb and ArAr isochron ages, the one exception being an Abee (EH4) Pb-Pb isochron age (fig. 15). This pattern of isochron ages for
these two -decaying radioisotope systems is not consistent with that reported by the RATE project, which found that K-Ar
and Ar-Ar isochron ages were always younger than Rb-Sr isochron ages (Snelling 2005; Vardiman, Snelling and Chaffin
2005). Among the model ages the Ar-Ar (and the two K-Ar) model ages are widely scattered and invariably are younger than

both the 4.554.57 Ga target age and the Pb-Pb model ages, except for the four Hvittis (EL6) Ar-Ar ages in the 4.554.57
Ga range, and the four Abee (EH4) Ar-Ar model ages greater than 4.8 Ga (fig. 16). Both the K-Ar and Pb-Pb systems seem
to have been affected in the Abee (EH4) meteorite, as the Pb-Pb model ages are scattered between 4.42 and 4.72 Ga (table
9), which is much less that the scatter in the Ar-Ar model ages between 3.56 and 8.9 Ga. The usual explanation for
disturbance of the K-Ar (and Ar-Ar) system is a heating event sometime after formation of the meteorites parent body
(Bogard 2011; Min, Reiners, and Shuster 2013; Trieloff et al. 2003), so that might explain this pattern in these E chondrites.
The U-Pb system if usually perturbed by earth surface weathering, so perhaps the Abee (EH4) chondrite was affected by
water on the earths surface before it was recovered for study.
Comparisons to the RATE Study
One of the issues Snelling (2014) discussed in relation to the radioisotope ages that have been obtained for the Allende
CV3 carbonaceous chondrite is whether the meteorite yielded a pattern of isochron ages similar to that found for earth rocks
by the 19972005 RATE (radioisotopes and the age of the earth) project (Vardiman, Snelling, and Chaffin 2005). The major
conclusion of the RATE project was that radioisotope decay rates have not necessarily been constant throughout earth
history, because there is evidence that there have been one or more episodes of accelerated rates of radioisotope decay,
particularly during the Flood only about 4350 years ago (Vardiman, Snelling, and Chaffin 2005). While there were several
lines of documented evidence that confirmed this conclusion, the principal evidence was different isochron ages obtained
from the same samples from the same rock units by the different radioisotope dating methods (Snelling 2005; Vardiman,
Snelling, and Chaffin 2005). Furthermore, there was a consistent pattern to the isochron ages from the different methods
that indicated that there was an underlying systematic cause of these age differences, namely, an episode (or episodes) of
accelerated radioisotope decay (Snelling 2005; Vardiman, Snelling, and Chaffin 2005). For example, it was found that the decaying radioisotopes U and Sm always gave older ages than the -decaying K and Rb. And then between the -decayers,
K with the shorter half-life (more rapid decay today) and the lighter atomic weight, always yielded younger ages than the
slower decaying and heavier Rb. While exactly the same pattern was not confirmed among the -decaying U and Sm
radioisotopes, both their half-lives and atomic weights were still believed to be the factors at work.
The mechanism proposed for this past episode (or episodes) of accelerated radioisotope decay was small changes to the
binding forces in the nuclei of the parent radioisotopes (Vardiman, Snelling, and Chaffin 2005). These changes would thus
have to have affected every atom making up the earth, and by logical extension every atom of the universe at the same
time. The designer Himself is not bound by those physical laws which He can change at any time anywhere or everywhere.
Therefore, we should expect that this past episode(s) of accelerated radioisotope decay had affected the asteroids from
where many meteorites have come, and that the meteorites may thus today yield the same pattern of different radioisotope
ages from the different radioisotope dating methods.Snelling (2014) found no pattern of isochron ages similar to the patterns
found in the RATE study was yielded by the Allende CV3 carbonaceous chondrite, and the same is true of the isochron ages
for the fifteen H, L, LL, and E chondrites reported in this study, as already discussed above. In fact, the -decay isochron
ages (Rb-Sr, Re-Os) are sometimes older than the -decay (U, Sm) isochron ages. Furthermore, the E chondrites yielded
Rb-Sr isochron ages generally younger than or equal to their Ar-Ar isochron ages (table 8 and fig. 15), when the pattern in
the RATE studys rocks was the opposite because Rb has a longer half-life and a heavier atomic weight. In contrast, the ReOs isochron ages yielded by the St. Sverin (LL6) chondrite were greater than or equal to its Rb-Sr isochron ages (table 6
and fig. 13). While the RATE study didnt deal with Re-Os isochron ages, this pattern is arguably somewhat predictable from
the conclusions of the RATE study, because while the -decaying Re has a slightly shorter half-life (at 42.7 billion years)
than the -decaying Rb (at 48.8 billion years), it has a heavier atomic weight (187 for Re compared to 87 for Rb) (Faure and
Mensing 2005). So if the atomic weight is the dominant factor in the amount of accelerated radioisotope decay which
occurred, then the Re-Os isochron ages should be older than the Rb-Sr isochron ages, though the effect of the half-lives
may result in the occasional equality of their isochron ages.Extending this argument further, if the atomic weight is the
dominant factor in the amount of accelerated radioisotope decay which occurred, then among the -decaying parent
radioisotopes the Pb-Pb isochron ages should be older than the Sm-Nd isochron ages, because the parent U radioisotopes
have atomic weights of 238 and 235, whereas the parent Sm radioisotopes atomic weight is only 147. But again the heavier
atomic weight U radioisotopes have shorter half-lives (at 4.47 billion years and 0.704 billion years respectively) than Sm (at
106 billion years) (Faure and Mensing 2005). So the Pb-Pb isochron ages are mostly older (or equal to) the Sm-Nd isochron
age of the St Sverin (LL6) chondrite (table 6 and fig. 13), but this is not the case with the Bjurble (L4) chondrite (table 4
and fig. 11) and the Richardton (H5) chondrite (table 2 and fig. 9), where some Sm-Nd isochron ages are older than the PbPb isochron ages. So again the effect of the half-lives may result in the occasional equality or reversal in the pattern of their
isochron ages.It is therefore fairly obvious that there are no clear and consistent patterns in these meteorite isochron ages
comparable to the patterns of isochron ages obtained in the RATE study. However, there is a major difference between the
RATE study and these studies on meteorites, in that the RATE study investigated earth rocks that yielded isochron ages of
less than 3 Ga, whereas these meteorites come from elsewhere in the solar system and generally yield 45 Ga ages. So the
origin of these meteorites could well have a major bearing on the radioisotope ages they yield, as initially discussed by
Snelling (2014).
The Origin of Meteorites from Asteroids
There is unanimity among astronomers and planetary geologists that most meteorites come from asteroids, which are
primarily orbiting the Sun in the asteroid belt between Mars and Jupiter (Libourel and Corrigan 2014). There is also
unanimity among conventional scientists that the asteroids represent leftover precursors to the terrestrial planets (MercuryMars) (Michel 2014). They postulate that about 4.56 billion years ago the early solar system consisted of a rotating disk of
gas and dust, called the protoplanetary disk, revolving around the sun. Planets then supposedly formed from that disk, and
different populations of small bodies, in particular the main belt asteroids between the orbits of Mars and Jupiter, survived as
remnants of that era.According to current conventional models, the asteroid belt that remained at the end of the planetforming processes was probably very different from the current main belt, perhaps containing an earth mass or more of
material in planetary embryos with masses similar to the Moon or Mars, as well as tens, hundreds, or thousands of times
more bodies like the asteroid 4 Vesta and the dwarf planet 1 Ceres than are present in the main belt today (Michel 2014).
Throughout its history the asteroid belt appears to have been shaped by collisional processes, such as cratering, disruption,
and the generation of new asteroids as collisional fragments. The net result is that the total mass of the main asteroid belt
today is only about 4% of the Moons mass, or less than 1/1000th of the Earths mass (Libourel and Corrigan 2014). So it
would appear that the asteroid belt has been significantly depleted of asteroids since its early history.Orbital resonances,
when two bodies have orbital periods that are a simple integer ratio of each other, may lead to destabilization of the orbits of
small bodies (Libourel and Corrigan 2014). Within the main asteroid belt, objects that have orbital periods in resonance with
the orbital period of Jupiter are gradually ejected into different, random orbits, leading to the removal of asteroids from
regions within the main asteroid belt that are now empty. Another important resonance is that between asteroids and Saturn,
which has formed the inner boundary of the main asteroid belt, and which is responsible for delivering asteroids into planet-

crossing orbits. Once asteroids become Mars-crossers they are usually ejected from the main asteroid belt due to close
encounters with Mars gravitational field. If Mars-crossing asteroids fail to interact with Mars, then their orbital semi-major
axes are gradually reduced and they become Near Earth Asteroids (NEAs).Asteroids that are nudged by the gravitational
attraction of nearby planets or have significant inclination and eccentricity may collide with other bodies traveling along
different orbits (Libourel and Corrigan 2014). Even if the current impact probability appears low, collisions between asteroids
are not rare, and do not appear to have been rare in the past. Depending on the relative impact velocity between the bodies
and on their sizes, collisions result in 1) fragmentation of a parent asteroid into several large pieces, and/or 2) the formation
of fine, micron-sized asteroidal dust. A collision between large asteroids brings into play both fragmentation and gravitation
(Michel 2014). The asteroids are partially to totally shattered, and subsequent gravitational attraction between fragments
leads to reaccumulation, which finally forms an entire family of large and small objects (new asteroids). Accordingly, most of
the smaller asteroids are thought to be piles of rubble held together loosely by gravity (Michel and Richardson 2013;
Tsuchiyama 2014). The largest asteroids, those larger than 125 km (200 mi) in diameter, however, are probably primordial
objects that have never been disrupted (Asphaug 2009; Michel 2014). Asteroids are therefore currently thought to have
been quite mobile within the main belt. Due to asteroid collisions and other effects, main belt asteroids migrate, passing
through the orbital resonances to end up crossing the orbits of Mars, Earth, Venus, and even Mercury (Libourel and
Corrigan 2014). NEAs do not have stable orbits, so they have relatively short lifetimes. Once the orbits of asteroids whose
diameters exceed 100150 m (330490 ft) are within 7.5 million km (4.7 million mi) of the earths orbit, there is a greater
possibility of them colliding with the earth and impacting its surface. By definition, meteors are asteroids that enter the
earths atmosphere. Due to their high entry velocity (several kilometers per second) they are heated to high temperatures as
they are slowed by the atmosphere. This produces visible paths, and the meteors are then known as fireballs or shooting
stars. If the meteors survive their plunge through the atmosphere and land on the earths surface, they are classified as
meteorites.
From these orbital and dynamical arguments, it is believed that most meteorites have indeed come from the asteroid belt
and therefore are samples of asteroidal materials (Cloutis, Binzel, and Gaffey 2014; Libourel and Corrigan 2014). However,
linking meteorites to their parent asteroids is a complicated issue. There is a photographic technique that has been used to
get estimates of the orbital parameters (approximately) of meteors before they make contact with earths atmosphere. Two
cameras have to have synchronized shutters and photograph the object before it makes contact with the atmosphere. This
has been done for a few cases of meteors and the orbital elements suggest the objects did come from the asteroid region.
So there are some cases where this has been done and then fragments of the photographed objects were found, which has
established good indications of meteorites coming from asteroids. Furthermore, there have been a multitude of methods
used to investigate asteroids, such as earth-based radar imaging, optical and radar polarimetry, thermal-infrared
observations, reflectance spectroscopy, and thermal emission spectroscopy, but only the availability of meteorites of known
provenance has enabled additional confirmation of asteroid-meteorite links. Two recent examples confirming the asteroidmeteorite link are relevant to the H, L, LL, and E chondrites in this study.On October 6, 2008, the small, ~4 m (13 ft) wide
asteroid 2008 TC3 was discovered and predicted to hit the earth within ~19 hours (Goodrich, Bischoff, and OBrien 2014).
Early morning, October 7, 2008, eyewitnesses saw the fireball that resulted when the asteroid hit the earths atmosphere
above the Nubian Desert of northern Sudan, Africa. A few seconds later, at ~37 km (23 mi) above the earth, the asteroid was
shattered in the atmosphere by dynamic ram pressures in a series of explosions into fragments. Approximately 700 cmsized (275 in-sized) fragments were subsequently recovered and constitute what became known as the Almahata Sitta
meteorite. Study of their physical, chemical, and mineralogical properties has revealed that the fragments are remarkably
heterogeneous. The most abundant samples are ureilitic lithologies (see fig. 1 URE among the Primitive Achondrites) with
various olivine/pyroxene ratios, mineral compositions and grain sizes (Bischoff et al. 2010; Goodrich, Bischoff, and OBrien
2014). Among the chondritic samples, enstatite (E) chondrites are the most abundant, including both E chondrite subgroups
(EL and EH), though the EL subgroup dominates, with representatives of the various petrologic types (EL3, EL4, EL5, and
EL6), which are indistinguishable from previously known E chondrites. So far, several L and H group ordinary (O) chondrites
have also been analysed. It has been concluded that the 2008 TC3 asteroid was not solid rock, but consisted mostly of finegrained, highly porous matrix material, weakly cementing a small fraction of isolated, centimeter-sized fragments of denser
rocks that became the fallen meteorite.In June 2010 the Japanese spacecraft Hayabusa successfully returned to earth with
fine particles collected in September 2005 from the surface of Near Earth Asteroid 25143 Itokawa (Nakamura et al. 2011;
Tsuchiyama 2014). Measuring 30180 m (0.00110.007 in) in diameter, initial analyses of the mineralogy, micropetrology,
and elemental and isotopic compositions of the returned regolith particles from asteroid Itokawa indicate that these dust
particles are identical to thermally metamorphosed LL chondrites, particularly the LL5 and LL6 ordinary chondrites, such as
Olivenza and St. Sverin (respectively) in this study.
Conclusions
After decades of numerous careful radioisotope dating investigations of ordinary (O) chondrite meteorites (H, L, and LL
groups) and enstatite (E) chondrite meteorites their Pb-Pb isochron age of 4.554.57 Ga has been well established. This
date for these chondrite meteorites is supported for some of them by a strong clustering of their Pb-Pb isochron and model
ages in the 4.554.57 Ga range, as well as being confirmed by both isochron and model age results via the U-Pb method,
and to a lesser extent, by the Ar-Ar, Rb-Sr, Re-Os, and Sm-Nd methods. The Hf-W, Mn-Cr, and I-Xe methods are all
calibrated against the Pb-Pb isochron method, so their results are not objectively independent. Thus the Pb-Pb isochron
dating method stands supreme as the ultimate, most precise tool for determining the ages of the chondrite meteorites.There
are only two other discernible patterns in the isochron and model ages for these O and E chondrites, apart from scatter of
the U-Pb, Th-Pb, Rb-Sr, and Ar-Ar model ages particularly. These chondrite ages do not follow the systematic pattern found
in Grand Canyon Precambrian rock units during the RATE project. The -decay ages are not always older than the -decay
ages for particular meteorites, and among the -decayers the ages are not always older according to the increasing
heaviness of the atomic weights of the parent radioisotopes, but may have also been modified according to the lengths of
their half-lives. Thus there appears to be no consistent evidence in these O and E chondrite meteorites similar to the
evidence found in earth rocks of past accelerated radioisotope decay.Any explanation for the 4.554.57 Ga age for these O
and E chondrite meteorites needs to consider the origin of meteorites. Most meteorites appear to be fragments derived from
asteroids via collisions, but even in the naturalistic paradigm the asteroids, and thus the meteorites, are regarded as
primordial material left over from the formation of the solar system.If some of the daughter isotopes were thus inherited
by these O and E chondrite meteorites when they were formed from that primordial material, and the parent isotopes in the
meteorite were also subject to some subsequent accelerated radioisotope decay, then the 4.554.57 Ga Pb-Pb isochron
age for these O and E chondrite meteorites cannot be their true real-time age, which according to the creation paradigm is
only about 6000 real-time years.

However, these conclusions and the suggested explanation can at best be regarded as tentative and interim while their
confirmation or adjustment awaits the examination of more radioisotope dating data from many more meteorites.
Furthermore, further extensive studies of the radioisotope dating of many more earth rocks from all levels within the whole
geologic record are required to attempt to systematize the proportions of isotopes in each radioisotope dating system
measured today that are due to inheritance from the primordial material, past accelerated radioisotope decay, and mixing,
additions and subtractions in the earths mantle and crust through earth history, particularly during the Day Three Upheaval
and then subsequently during the Flood. Such studies are already in progress.
Acknowledgments
The invaluable help of my research assistant Lee Anderson, Jr., in compiling these radioisotope dating data into the tables
and then plotting the data in the color coded age versus frequency histogram diagrams is acknowledged.
Radioisotope Dating of Meteorites: III. The Eucrites (Basaltic Achondrites)
by Dr. Andrew A. Snelling on December 31, 2014
Abstract
Meteorites date the earth with a 4.55 0.07 Ga Pb-Pb isochron called the geochron. They appear to consistently yield
4.554.57 Ga radioisotope ages, adding to the uniformitarians confidence in the radioisotope dating methods. Achondrites,
meteorites not containing chondrules, account for about 8% of meteorites overall. About 3% of the witnessed falls of all
meteorite types are the achondrites known as eucrites, which makes them the fourth-most-common meteorite to fall.
Eucrites are similar to basalts and are believed to be space debris from the crust of main belt asteroid 4-Vesta. Many
radioisotope dating studies in the last 45 years have used the K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, U-Th-Pb, Lu-Hf, Mn-Cr, Hf-W, AlMg, I-Xe, and Pu-Xe methods to yield an abundance of isochron and model ages for these basaltic achondrites from wholerock samples, and mineral and other fractions. Such age data for 12 eucrites were tabulated and plotted on frequency
versus age histogram diagrams. They strongly cluster in many of these eucrites at 4.554.57 Ga, dominated by Pb-Pb and
U-Pb isochron and model ages, testimony to that techniques supremacy as the uniformitarians ultimate dating tool, which
they consider very reliable. These ages are confirmed by Rb-Sr, Lu-Hf, and Sm-Nd isochron ages, but agreement could be
due to calibration with the Pb-Pb system. There is also scatter of the U-Pb, Pb-Pb, Th-Pb, Rb-Sr, K-Ar, and Ar-Ar model
ages, in most cases likely due to thermal disturbances resulting from metamorphism or impact cratering of the parent
asteroid. No pattern was found in these meteorites isochron ages similar to the systematic patterns of isochron ages found
in Precambrian rock units during the RATE project, so there is no evidence of past accelerated radioisotope decay having
occurred in these eucrites, and therefore on their parent asteroid. This is not as expected, yet it is the same for all
meteorites so far studied. Thus it is argued that accelerated radioisotope decay must have only occurred on the earth, and
only the 500600 million years worth we have physical evidence for during the Flood. Otherwise, due to their 4.554.57 Ga
ages these eucrites and their parent asteroid are regarded as originally representing primordial material Todays
measured radioisotope compositions of these eucrites could reflect a geochemical signature of that primordial material,
which included atoms of all elemental isotopes. So if most of the measured daughter isotopes were already in these basaltic
achondrites when they were formed on their parent asteroid, then their 4.554.57 Ga ages obtained by Pb-Pb and U-Pb
isochron and model age dating are likely not their true real-time ages, which according to the creation paradigm is only
about 6000 real-time years. Further investigation of radioisotope ages data for meteorites in remaining groups of
achondrites, for lunar rocks, and for rocks from every level in the earths geologic record, should enable the interim ideas
presented here to be confirmed or modified.
Keywords: meteorites, classification, achondrites, eucrites, asteroids, 4-Vesta, radioisotope dating, Bereba, Cachari,
Caldera, Camel Donga, Ibitira, Juvinas, Moama, Moore County, Pasamonte, Serra de Mag, Stannern, Yamato 75011, K-Ar,
Ar-Ar, Rb-Sr, Sm-Nd, U-Th-Pb, Lu-Hf, Mn-Cr, Hf-W, Al-Mg, I-Xe, Pu-Xe, isochron ages, model ages, discordant radioisotope
ages, accelerated radioactive decay, thermal disturbance, resetting, primordial material, geochemical signature, mixing,
inheritance
Introduction
In 1956 Claire Patterson at the California Institute of Technology in Pasadena reported a Pb-Pb isochron age of 4.55 0.07
Ga for three stony and two iron meteorites, which since then has been declared the age of the earth (Patterson 1956).
Adding weight to that claim is the fact that many meteorites appear to consistently date to around the same age (Dalrymple
1991, 2004), thus bolstering the evolutionary communitys confidence that they have successfully dated the age of the earth
and the solar system at around 4.56 Ga. These apparent successes have also strengthened their case for the supposed
reliability of the increasingly sophisticated radioisotope dating methods.Creationists have commented little on the
radioisotope dating of meteorites, apart from acknowledging the use of Pattersons geochron to establish the age of the
earth, and that many meteorites give a similar old age. Morris (2007) did focus on the Allende carbonaceous chondrite as an
example of a well-studied meteorite analyzed by many radioisotope dating methods, but he only discussed the radioisotope
dating results from one, older paper (Tatsumoto, Unruh, and Desborough 1976).In order to rectify this lack of engagement
by the creationist community with the meteorite radioisotope dating data, Snelling (2014a) obtained as much radioisotope
dating data as possible for the Allende CV3 carbonaceous chondrite meteorite (due to its claimed status as the most studied
meteorite), displayed the data, and attempted to analyze them. He found that both isochron and model ages for the total
rock, separated components, or combinations of these strongly clustered around a Pb-Pb age of 4.564.57 Ga, the earliest
(Tatsumoto, Unruh, and Desborough 1976) and the latest (Amelin et al. 2010) determined Pb-Pb isochron ages at 4.553
0.004 Ga and 4.56718 0.0002 Ga respectively being essentially the same. Apart from scatter of the U-Pb, Th-Pb, Rb-Sr,
and Ar-Ar ages, no systematic pattern was found in the Allende isochron and model ages similar to the systematic pattern of
isochron ages found in Precambrian rock units during the RATE project that was interpreted as produced by an episode of
past accelerated radioisotope decay (Snelling 2005c; Vardiman, Snelling, and Chaffin 2005).Snelling (2014b) subsequently
gathered together all the radioisotope ages obtained for 10 ordinary (H, L, and LL) and five enstatite (E) chondrites and
similarly displayed the data. They generally clustered, strongly in the Richardton (H5), St. Marguerite (H4), Bardwell (L5),
Bjurbole (L4), and St. Sverin (LL6) ordinary chondrite meteorites, at 4.554.57 Ga, dominated by Pb-Pb and U-Pb isochron
and model ages, but confirmed by Ar-Ar, Rb-Sr, Re-Os, and Sm-Nd isochron ages. There was also scatter of the U-Pb, ThPb, Rb-Sr, and Ar-Ar model ages, in some cases possibly due to thermal disturbance. Again, no pattern was found in these
meteorites isochron ages indicative of past accelerated radioisotope decay.Snelling (2014a, b) then sought to discuss the
possible significance of this clustering in terms of various potential creationist models for the history of radioisotopes and
their decay. He favored the idea that asteroids and the meteorites derived from them are primordial material left over from
the formation of the solar system, which is compatible Thus he argued that todays measured radioisotope compositions of
all these chondrites may reflect a geochemical signature of that primordial material, which included atoms of all elemental
isotopes. So if some of the daughter isotopes were already in these chondrites when they were formed, then the 4.554.57
Ga ages for them obtained by Pb-Pb and U-Pb isochron and model age dating are likely not their true real-time ages,

which according to the creation paradigm is only about 6000 real-time years.However, Snelling (2014a, b) admitted that
drawing firm conclusions from the radioisotope dating data for just these 16 chondrite meteorites was premature, and
recommended further studies of more meteorites. This present contribution is therefore designed to further document the
radioisotope dating data for more meteorites, the basaltic achondrites or eucrites, so as to continue the discussion of the
potential significance of these data.
The Classification of Achondrite Meteorites
The most recent classification scheme for the meteorites is that of Weisberg, McCoy, and Krot (2006), which is reproduced
in Fig. 1. Based on their bulk compositions and textures, Krot et al. (2005) divided meteorites into two major categories,
chondrites (meteorites containing chondrules) and achondrites (meteorites not containing chondrules or non-chondritic
meteorites). They further subdivided the achondrites into primitive achondrites and igneously differentiated achondrites.
However, Weisberg, McCoy, and Krot (2006) simply subdivided all meteorites into three categorieschondrites, primitive
achondrites and achondrites (fig. 1).The non-chondritic meteorites contain virtually none of the components found in
chondrites. It is conventionally claimed that they were derived from chondritic materials by planetary melting, and that
fractionation caused their bulk compositions to deviate to various degrees from chondritic materials (Krot et al. 2005). The
degrees of melting that these rocks experienced are highly variable, and thus, these meteorites have been divided into the
two major categoriesprimitive and differentiated. However, there is no clear cut boundary between these categories.The
differentiated non-chondritic meteorites, or achondrites (fig. 1), are conventionally regarded as having been derived from
parent bodies that experienced large-scale partial melting, isotopic homogenization (ureilites are the only exception), and
subsequent differentiation. Based on abundance of FeNi-metal, these meteorites are commonly divided into three types
achondrites, stony-irons, and irons. Each of these types contains several meteorite groups and ungrouped members (fig. 1).
Several groups of achondrites and iron meteorites are likely to be genetically related and were possibly derived from single
asteroids or planetary bodies.The achondrites account for about 8% of meteorites overall, and the majority of them (about
two-thirds) are HED meteorites (howardites, eucrites, and diogenites), believed to have originated from the crust of asteroid
4-Vesta (Norton 2002) (fig. 1). Other types include martian, lunar, and several types thought to originate from as-yet
unidentified asteroids. These groups have been determined on the basis of, for example, their bulk Fe/Mn and 17O/18O ratios,
which are thought to be characteristic fingerprints for each parent body (Mittlefehldt et al. 1998).The achondrites represent
the products of classical igneous processes acting on the silicate-oxide system of asteroidal bodiespartial to complete
melting, differentiation, and magmatic crystallization (Mittlefehldt 2005). Iron meteorites represent the complimentary metalsulfide system products of this process. Thus the achondrites consist of materials similar to terrestrial basalts and plutonic
rocks, so they exhibit igneous textures, or igneous textures modified by impact and/or thermal metamorphism, and
distinctive mineralogies indicative of igneous processes.The HED meteorites are sometimes grouped with the angrites and
aubrites (fig. 1) and termed the asteroidal achondrites, because of all having been differentiated on parent asteroidal bodies.
The howardite-eucrite-diogenite (HED) meteorites have been traditionally classified into the one clan, because there is
strong evidence they originated on the same parent body, the asteroid 4-Vesta (Binzel and Xu 1993; Consolmagno and
Drake 1977; Drake 2001; Mandler and Elkins-Tanton 2013; McCord, Adams, and Johnson 1970; McSween et al. 2011;
McSween et al. 2013, 2014; Righter and Drake 1997). This was one of the first links made between meteorites and an
asteroid (Cloutis, Binzel, and Gaffey 2014; McCord, Adams, and Johnson 1970). Initially their spectroscopic similarity, which
suggested the HED achondrites are impact ejecta off 4-Vesta, was deemed dynamically dubious owing to the apparent lack
of a plausible pathway from Vesta to the earth. The discovery of the Vesta family of asteroids or Vestoids (Binzel and Xu
1993) extending from Vesta to resonance delivery zones solidified the link. This link has stood the test of time and has been
confirmed by the in situ results provided by the Dawn mission to this asteroid (McSween et al. 2014). Thus the HED clan
allows the confident association of specific types of igneous processes with an asteroid body of known size.

Fig. 1. The classification system for meteorites (after Weisberg, McCoy, and Krot 2006). (Click image for larger view.)
The HED clan is the most extensive suite of differentiated crustal rocks from an asteroid (Mittlefehldt 2005). Evidence that
these achondrites belong in the same clan includes their identical oxygen isotopic compositions (Clayton and Mayeda
1996), similarities in Fe/Mn ratios in pyroxenes, the occurrence of polymict breccias consisting of materials of eucritic and
diogenitic parentage (for example, the howardites), and the existence of rocks intermediate between diogenites and
cumulate eucrites (Krot et al. 2005). The suite of meteorites comprising the HED clan is composed of mafic and ultramafic
igneous rocks, most of which are breccias. The parent lithologies were mostly metamorphosed, which has obscured original
igneous zoning in most cases (Mittlefehldt 2005). The suite contains four main igneous lithologiesbasalt and cumulate
gabbro (eucrites), and orthopyroxenite and harzburgite (diogenites). When both eucrite and diogenite clasts are present in a
meteorite that is a polymict breccia, then it is a howardite. These lithologies are consistent with a postulated layered crust
model for the HED parent body, 4-Vesta (Mandler and Elkins-Tanton 2013; McSween et al. 2013, 2014; Righter and Drake
1997; Takeda 1997).
The Eucrites
The eucrites are the most common of the achondrites. About 3% of the witnessed falls of all meteorite types are eucrites,
which makes them the fourth most common meteorite to fall (Norton 2002). Of the HED meteorites, eucrites are by far the
most common, about 52%. Until the meteorite finds in Antarctica became available with their large cache of eucrites,
eucrites were defined as monomict breccias. However, the large number of eucrites recovered that show a wide variation of
lithic fragments, unlike the fragments in howardites, has prompted the acceptance of eucrites as either monomict or
polymict.The most obvious external characteristic of a freshly fallen eucrite is its very black and lustrous fusion crust

compared to the dull black crust of a chondrite, due to the intense heating of the outer surface during passage through the
earths atmosphere (Norton 2002). Eucrites are Ca-rich and this combined with the usually present small amount of Fe gives
these meteorites a wet look (fig. 2). Fusion crusts form in the final second or two of the ablation process as meteorites
pass rapidly through the earths atmosphere during the fireball stage. The fusion crusts then rapidly cool, so contraction
cracks often form, leaving the outer surface of the meteorites looking much like the crazing on pottery (figs. 2 and 3).The
similarities of eucrites chemically and petrographically to terrestrial basalts is frequently noted, but a broken face of a eucrite
exposes a light gray interior, which is unlike the dark gray to black interiors of terrestrial basalts. Eucrite textures are also
fine-grained, and often glomeroporphyritic due to clumps of phenocrysts set in the groundmass. This is typical of terrestrial
volcanic rocks that have cooled more slowly, producing glomerocrysts of interlocking plagioclase and pyroxene crystals. If
basaltic lava contains dissolved gases when it suddenly erupts onto the earths surface, the sudden reduction in pressure
releases the gases which quickly form bubbles that make their way to the top of the flow. The eucrite, Ibitira, one of the few
unbrecciated eucrites known, shows a remarkable vesicular texture (fig. 4), similar to that seen in a terrestrial basalt lava
flow as that just described. Microscopically, the resemblance of most eucrites to terrestrial basalts is also most striking (fig.
5a and b).
Fig. 5. Thin section photomicrographs in transmitted light under crossed polars of four typical eucrites (basaltic achondrites)
(after Krot et al. 2005; McSween et al. 2011).

(a) The unequilibrated noncumulate eucrite Pasamonte, showing the typical basaltic texture of plagioclase (light) and
pyroxene (colored). (Click image for larger view.)
(b) The metamorphosed (equilibrated) noncumulate eucrite Ibitira, showing a recrystallized texture with plagioclase (white)
and pyroxene (colored), with the round, dark areas in the center, bottom, and left being vesicles. (Click image for larger
view.)

(c) The
cumulate eucrite Serra de Mag, consisting of large crystals of plagioclase
(lighter material with straight twin lamellae) and mostly orthopyroxene with
complex augite exsolution lamellae (darker material with irregular,
sometimes worm-like exsolution lamellae). (Click image for larger view.)

(d) The cumulate eucrite Moore County, consisting of large crystals of plagioclase (lighter material with straight twin
lamellae) and colorful abundant orthopyroxene (with occasional exsolution lamellae) (scale bar is 2.5mm [0.09in]).

Mineralogically, the eucrites are quite simple. They consist almost entirely of plagioclase (3050%) and clinopyroxene (40
60%), the clinopyroxene usually dominating by 1020%. The plagioclase in eucrites is calcic, being primarily anorthite with
some bytownite, that is, within the range An 75-95, and igneous zoning is commonly preserved. The clinopyroxene is low-Ca
pigeonite, with a composition that varies widely from specimen to specimen, and even within a given specimen. A typical
pyroxene composition (wollastonite-enstatite-ferrosilite) in mole percent might be Wo1-25 En42-48 Fs43-52. Minor minerals include
chromite (FeCr2O4), Fe-Ni metal, ilmenite (FeTiO3) and troilite (FeS) as opaque minerals, orthopyroxene, and polymorphs of
silicaquartz, tridymite, and cristobalite.Eucrites are subdivided into three major subclassesthe noncumulate eucrites
(basaltic eucrites), the cumulate eucrites (cumulate gabbros), and the polymict eucrites (polymict breccias of basaltic and
cumulate eucrites) (Krot et al. 2005; Mittlefehldt 2005).Noncumulate (basaltic) eucrites are mostly fragmental breccias of fine
to medium grained, subophitic to ophitic basalts that are postulated to have formed originally as quickly cooled surface lava
flows. They are known as unequilibrated, unmetamorphosed or least-metamorphosed, noncumulate eucrites (such as
Pasamontesee fig. 5a), and are composed of pigeonite and plagioclase, with minor silica, ilmenite, and chromite, and
accessory phosphates, troilite, Fe-Ni metal, fayalitic olivine, zircon, and baddeleyite. As a result of their apparent fast cooling
their pyroxenes (pigeonite of Mg# ~7020) are zoned, and exsolution lamellae are only visible by TEM. However, most
noncumulate eucrites appear to have been subsequently metamorphosed, and are thus known as metamorphosed or
equilibrated noncumulate eucrites (such as Juvinas, Stannern, and Ibitirasee fig. 5b). They are highly abundant and so
are also collectively referred to as the ordinary eucrites. They are unbrecciated or monomict-brecciated, metamorphosed
basalts and contain homogeneous low-Ca pigeonite (Mg# ~4230) with fine exsolution lamellae of high-Ca pyroxene. The
pyroxenes were originally ferroan pigeonite (~Wo7-15 En29-43 Fs48-58) which exsolved augite during metamorphism. In most
eucrites, pyroxene Fe/Mg is uniform as a result of metamorphism, but original igneous zoning is preserved in very few.
Plagioclase is calcic, with most in the range An75-93, and igneous zoning is commonly preserved.
Cumulate eucrites are coarse-grained gabbros, many unbrecciated (such as Serra de Magsee fig. 5c, and Moore County
see fig. 5d), composed of pigeonite, plagioclase, and minor chromite with silica, ilmenite, Fe-Ni metal, troilite, and
phosphate as trace accessory phases. The original igneous pyroxene was pigeonite (~Wo7-16 En38-61 Fs32-46) which exsolved
augite and, in some, inverted to orthopyroxene. They contain orthopyroxene inverted from low-Ca clinopyroxene (Mg# ~67
58) and orthopyroxene inverted from pigeonite (Mg# ~5745). Plagioclase is generally more calcic than that typical for
basaltic eucrites, with most in the range An91-95.
Polymict eucrites are polymict breccias consisting of fragmental and melt-matrix breccias mostly of eucritic material, but they
also contain <10 vol.% of diogenitic component in the form of orthopyroxenite.
The Radioisotope Dating of the Eucrites
To thoroughly investigate the radioisotope dating of the eucrite achondrites all the relevant literature was searched. The
objective was to find eucrite achondrites that have been dated by more than one radioisotope method, and a convenient
place to start was Dalrymple (1991, 2004), who compiled lists of such data. The 12 eucrite achondrite meteorites that were
found to have been dated multiple times by more than one radioisotope methodBereba, Cachari, Caldera, Camel Donga
(fig. 2), Ibitira (figs. 4 and 5b), Juvinas, Moama, Moore County (fig. 5d), Pasamonte (figs. 3 and 5a), Serra de Mag (fig. 5c),
Stannern, and Yamato 75011thus became the focus of this study. When papers containing radioisotope dating results for
these eucrites were found, the reference lists were also scanned to find further relevant papers. In this way a
comprehensive set of papers, articles, and abstracts on radioisotope dating of these basaltic achondrite meteorites was
collected. While it cannot be claimed that all the papers, articles, and abstracts which have ever been published containing
radioisotope dating results for these eucrites have thus been obtained, the cross-checking undertaken between these
publications does indicate the data set obtained is very comprehensive.All the radioisotope dating results of these 12
eucrites were then compiled and tabulated. For ease of viewing and comparing the radioisotope dating data, the isochron
and model ages for some or all components of each of these 12 eucrites were tabulated separatelythe isochron ages in
Table 1 and the model ages in Table 2. The data in these tables were then plotted on frequency versus age histogram
diagrams, with the same color coding being used to show the ages obtained by the different radioisotope dating methods
the isochron ages for whole-rock samples and some or all components of each of these 12 eucrites (fig. 6), and the model
ages for whole-rock samples and components of each of these 12 eucrites (fig. 7).
Table 1. Isochron ages for whole-rock samples and some or all components of 12 eucrite achondrites, with the details and
literature sources.
Sample

Method

Date

Err +/-

Note

Source

Type

Rb-Sr

4.17

0.26

Birck
1978

4.08

0.26

Basaltic
Volcanism
Study Project 1981 isochron age

Bereba

eight
fractions
(whole rock and
mineral)

Rb-Sr

and

Allgre
isochron age

ten
other
meteorites
plus
three plagioclaseHans, Kleine,
samples
Bourdon 2013

one
whole-rock
sample plotted with
12 other samples
Rb-Sr

4.55

0.19

one sample plotted


with
14
other
meteorites
Lu-Hf

4.604

0.039

Blichort-Toft
2002

whole rock and


plagioclase samples Pb-Pb

0.004

Carlson, Tera,
Boctor 1988

and

4.522

plagioclase
and
whole
rock
+
leachate
Pb-Pb

Tera, Carlson,
Boctor 1997

and

4.52

three zircon grains


intercept concordia U-Pb

4.538

0.026

three zircon grains U-Pb

4.538

0.026

Lee et al. 2009

et

and
isochron age
al.
isochron age
isochron age

isochron age

isochron age

intercept concordia
three point analyses
of one apatite grain U-Pb

4.196

0.013

Zhou et al. 2011

isochron age

five point analyses


of
three
zircon
grains
U-Pb

4.552

0.021

Zhou et al. 2013

isochron age

whole
rock
+
mineral separates
Sm-Nd

4.79

one sample plotted


with
17
other
meteorites
Sm-Nd

4.464

0.075

Blichort-Toft
2002

one sample plotted


with seven other
meteorites
Hf-W

4.5632

0.0014

Kleine et al. 2004

4.55

0.0041

Hf-W

Carlson, Tera,
Boctor 1988

and
isochron age

et

al.
isochron age

isochron age

Kleine et al. 2005

isochron age

Cacheri
one sample plotted with 17
other meteorites
Lu-Hf

4.604

0.039

Blichort-Toft
2002

Pb-Pb

4.453

0.015

Tera, Carlson,
Boctor 1997

and

mean of three isochrons

206

Pb-204Pb4.453

0.015

Tera, Carlson,
Boctor 1997

and

six fractions

207

Pb-206Pb4.451

0.015

Tera, Carlson,
Boctor 1997

and

eight fractions

207

Pb-204Pb4.455

0.016

Tera, Carlson,
Boctor 1997

and

eight fractions

207

Pb-204Pb4.558

0.025

Zhou et al. 2013

isochron age

sixteen analyses
zircon grains

of

et

al.
isochron age
isochron age
isochron age
isochron age
isochron age

six

three analyses of one zircon


grain
U-Pb

4.546

0.01

Zhou et al. 2011

isochron age

three analyses of one zircon


grain
U-Pb

4.548

0.024

Zhou et al. 2013

isochron age

six fractions

3.99

0.21

Tera, Carlson,
Boctor 1997

0.075

Blichort-Toft
2002

et

4.464

one whole-rock sample


plotted with 14 other
meteorites
Lu-Hf

0.039

Blichert-Toft
2002

et

4.604

pyroxene and plagioclase


fractions
Pb-Pb

4.5161

0.0028

Galer
1996

four points on one zircon


grain
U-Pb

4.563

0.18

Zhou et al. 2013

two
whole
rocks,
plagioclase, and pyroxene
fractions
Sm-Nd

4.544

0.019

Wadhwa and Lugmair


1996
isochron age

four fractions

Mn-Cr

4.545

Ar-Ar

3.706

0.097

Kennedy et al. 2013

isochron age

Ar-Ar

3.685

0.085

Kennedy et al. 2013

isochron age

Ar-Ar

3.703

0.059

Kennedy et al. 2013

isochron age

Sm-Nd

one sample plotted with 17


other meteorites
Sm-Nd

and
isochron age
al.
isochron age

Caldera

and

al.
isochron age

Lugmair
isochron age
isochron age

Lugmair
and
Shukolyukov 1998
isochron age

Camel Donga
matrix
samples
pyroxenes, inverse

with

matrix
samples
pyroxenes, inverse

with

mean of two, inverse

U-Pb

4.512

0.011

Zhou et al. 2013

isochron age

Hf-W

4.546

0.005

Kleine et al. 2005

isochron age

Hf-W

4.545

0.0035

Kleine et al. 2005

isochron age

whole rock plus four mineral


separates
Rb-Sr

0.25

Birck
1978

and

4.52

isochron age

Ibitira samples plotted with


Juvinas
samples,
plus
whole-rock samples of eight
other meteorites
Rb-Sr

0.13

Birck
1978

and

4.57

one sample plotted with 14


other meteorites
Lu-Hf

4.604

0.039

Blichert-Toft
2002

thirty-five point analyses of


14 zircon grains
Ibitira

three whole-rock samples

207Pb206Pb

Allgre

Allgre
isochron age
et

al.
isochron age

Chen
and
Wasserburg 1985
isochron age

4.556

nine samples (pyroxenes) 207Pbexternal normalization (EN) 206Pb

4.55703

0.00082

Amelin, Wadhwa, and


Lugmair 2006
isochron age

nine samples (pyroxenes) 207PbEN and double spike (DS)


206Pb

4.55744

0.00055

Amelin, Wadhwa, and


Lugmair 2006
isochron age

pyroxene
residues

Pb-Pb

4.5561

0.0023

Iizuka et al. 2013

isochron age

residues of eight pyroxenerich and two whole-rock


fractions and washes of two
pyroxene-rich
and
one
plagioclase-rich
fractions(13-point isochron) Pb-Pb

4.5565

0.0013

Iizuka et al. 2014

isochron age

residues of one whole-rock


and
six
pyroxene-rich
fractions (7-point isochron) Pb-Pb

4.55675

0.00057

Iizuka et al. 2014

isochron age

residues of eight pyroxenerich and two whole-rock


fractions and washes of two
pyroxene-rich
and
one
plagioclase-rich
fractions
(13-point chord)
U-Pb

4.5569

0.0014

Iizuka et al. 2014

isochron age
(concordia)

residues of one whole-rock


and
six
pyroxene-rich
fractions (7-point chord)
U-Pb

4.556

0.0052

Iizuka et al. 2014

isochron age
(concordia)

leachates

and

whole rock plus five mineral


separates
Sm-Nd

4.46

0.02

Prinzhofer,
Papanastassiou, and
Wasserburg 1992
isochron age

two whole-rock samples


plus multiple plagioclase,
pyroxene and phosphate
separates
Sm-Nd

4.57

0.09

Nyquist et al. 1999

isochron age

two whole-rock samples


plus
phosphate
and
pyroxene separates
Sm-Nd

4.6

0.05

Nyquist et al. 1999

isochron age

two whole-rock samples


plus
pyroxene
and
plagioclase separates
Sm-Nd

4.41

0.07

Nyquist et al. 1999

isochron age

one sample plotted with 17


other meteorites
Sm-Nd

4.464

0.075

Blichert-Toft
2002

three fractions

Hf-W

4.549

0.012

Kleine et al. 2005

three fractions

Mn-Cr

4.557

0.003

Lugmair
and
Shukolyukov 1998
isochron age

4.5574

0.0025

Iizuka et al. 2014

after
Lugmair
and
Shukolyukov (1998) using
Pb-Pb age of DOrbigny
Mn-Cr

et

al.
isochron age
isochron age

isochron age

after Yin, Amelin, and


Jacobsen (2009) using the
Pb-Pb age of DOrbigny
Mn-Cr

4.5559

whole rock, pyroxene (2)


and plagioclase (2) fractions Al-Mg

Iizuka et al. 2014

isochron age

4.5614

Wadhwa et al. 2004

isochron age

whole rock, pyroxene (2)


and plagioclase (2) fractions Al-Mg

4.5607

Amelin, Wadhwa, and


Lugmair 2006
isochron age

relative to Shallowater

4.555

0.001

Claydon,
Crowther,
and Gilmour 2013
isochron age

one whole-rock sample


plotted with six other
meteorites
Rb-Sr

4.39

0.26

Papanastassiou and
Wasserburg 1969
isochron age

four mineral separates plus


whole rock
Rb-Sr

4.6

0.07

Allgre et al. 1975

isochron age

earlier
Ibitira

4.58

0.14

Birck
1978

isochron age

4.5

0.07

Quitte, Birck,
Allgre 2000

0.19

ten
other
meteorites
and
two
plagioclaseHans, Kleine,
samples
Bourdon 2013

I-Xe

0.0032

Juvinas

data

plotted

with
Rb-Sr

revised Allgre et al. (1975)


age
Rb-Sr
whole-rock and plagioclase
samples plotted with eleven
other samples
Rb-Sr

4.55

and

Allgre

adjusted to agree
with
otherPatchett
methods
Tatsumoto 1980

and
isochron age

and
isochron age

one sample plotted with


nine other meteorites
Lu-Hf

4.55

one sample plotted with 14


other meteorites
Lu-Hf

4.604

0.039

Blichert-Toft
2002

pyroxene and plagioclase


fractions
Pb-Pb

4.3209

0.017

Galer
1996

207Pb206Pb

4.556

0.012

eight samples

Tatsumoto and Unruh


1975
isochron age

207Pb206Pb

4.54

0.0007

nine samples

Manhes, Allgre, and


Provost 1984
isochron age

206Pb238U

4.531

0.003

upper intercept

Manhes, Allgre, and


Provost 1984
isochron age

206Pb238U

4.539

0.004

upper intercept

Manhes, Allgre, and


Provost 1984
isochron age

206Pb238U

4.545

0.0012

seven samples

Manhes, Allgre, and


Provost 1984
isochron age

207Pb235U

4.543

0.003

eight samples

Manhes, Allgre, and


Provost 1984
isochron age

232Th208Pb

4.47

0.03

nine samples

Manhes, Allgre, and


Provost 1984
isochron age

4.53

0.033

Zhou et al. 2013

twenty analyses of seven


zircon grains
U-Pb

and

et

and
isochron age
al.
isochron age

Lugmair
isochron age

isochron age

whole rock (2), plagioclase,


and pyroxene fractions
Sm-Nd

4.56

0.08

Lugmair
1974;
Lugmair,
Scheinin,
and Marti 1975
isochron age

one sample plotted with 17


other meteorites
Sm-Nd

4.464

0.075

Blichert-Toft
2002

five samples plotted with


seven other meteorites
Hf-W

4.5632

0.0014

Kleine et al. 2004

isochron age

Hf-W

4.5457

0.0036

Kleine et al. 2005

isochron age

Mn-Cr

4.5625

0.001

Lugmair
and
Shukloyukov 1998
isochron age

Mn-Cr

4.5642

0.0012

Schiller, Baker, andisochron age

five fractions

five fractions

et

al.
isochron age

Bizzarro 2010
whole rock, pyroxene (3),
and plagioclase (3) fractions Al-Mg

4.561

Wadhwa et al. 2004

isochron age

Moama
whole-rock
sample
plotted with 12 other
samples
Rb-Sr

4.55

one sample plotted with


nine other meteorites
Lu-Hf

4.55

one sample plotted with


14 other meteorites
Lu-Hf

4.604

0.039

Blichert-Toft
2002

whole
rock,
WR
leachate,
and
plagioclase
and
pyroxene separates
whole
rock,
WR
leachate,
and
plagioclase
and
pyroxene separates
whole
rock,
WR
leachate,
and
plagioclase
and
pyroxene separates

mean of the
isochron ages

Pb-Pb

0.19

ten other meteorites


and three plagioclaseHans, Kleine,
samples
Bourdon 2013
Patchett
Tatsumoto 1980
et

and
isochron age
and
isochron age
al.
isochron age

Pb-206Pb

4.439

0.097

Tera, Carlson,
Boctor 1997

and

207

Pb-204Pb

4.416

0.092

Tera, Carlson,
Boctor 1997

and

207

Pb-204Pb

4.423

0.094

Tera, Carlson,
Boctor 1997

and

206

4.426

0.094

Tera, Carlson,
Boctor 1997

and

mean Pb-Pb

4.52

0.05

Hamet et al. 1978

4.46

0.03

Jacobsen
and
Wasserburg 1984
isochron age

Sm-Nd

whole rock, plus two


pyroxene
and
plagioclase separates
Sm-Nd
whole rock, pyroxene
and
plagioclase
fractions
Sm-Nd

isochron age

isochron age

isochron age
isochron age
isochron age

plus Hamet et al. 1978


and
Jacobsen
andBoyet, Carlson, and
Wasserburg 1984 data Horan 2010
isochron age

4.594

0.079

one sample plotted with


eight other meteorites
Rb-Sr

4.557

0.253

Cumming 1969

one whole-rock sample


plotted with six other
meteorites
Rb-Sr

4.39

0.26

Papanastassiou and
Wasserburg 1969
isochron age

Moore County

plagioclase
sample
plotted with 12 other
samples
Rb-Sr

4.55

one sample plotted with


nine other meteorites
and adjusted to agree
with other methods
Lu-Hf

4.55

one sample plotted with


14 other meteorites
Lu-Hf

4.604

0.039

Blichert-Toft
2002

4.477

0.022

Tera, Carlson,
Boctor 1997

207

whole
rock
(3),
pyroxene
(2)
and
plagioclase (2) fractions
and leachates

Pb-204Pb

0.19

ten other meteorites


and two plagioclaseHans, Kleine,
samples
Bourdon 2013

Patchett
Tatsumoto 1980
et

isochron age

and
isochron age

and
isochron age
al.
isochron age
and
isochron age

207

Pb-206Pb

4.494

0.017

isochron age

206

204

Pb- Pb

4.481

0.02

isochron age

mean Pb-Pb

4.484

0.019

isochron age

0.04

Unruh,
Nakamura,
and Tatsumoto 1977 isochron age

0.025

Tera, Carlson,
Boctor 1997

and

4.456
4.457

0.025

Tera,

andisochron age

internal Sm-Nd 4.6

whole rock, pyroxene


plagioclase (2) fractions Sm-Nd
Sm-Nd

Carlson,

isochron age

Boctor 1997
one sample plotted with
17 other meteorites
Sm-Nd

4.464

0.075

Blichert-Toft
2002

et

al.
isochron age

plus Tera, Carlson, and


Boctor
1997
and
Blichert-Toft et al. 2002Boyet, Carlson, andisochron
data
Horan 2010
two

whole rock, pyroxene


and
plagioclase
fractions
Sm-Nd

4.542

fractions

Mn-Cr

4.549

one sample plotted with


five other meteorites
Rb-Sr

4.411

0.088

Shields, Pinson, and


Hurley 1965
isochron age

one sample plotted with


eight other meteorites
Rb-Sr

4.557

0.253

Cumming 1969

one whole-rock sample


plotted with six other
meteorites
Rb-Sr

4.39

0.26

Papanastassiou and
Wasserburg 1969
isochron age

two whole-rock samples


plotted with nine other
meteorites
Rb-Sr

4.33

0.49

Birck
1978

0.085

age

Lugmair
and
Shukolyukov 1998
isochron age

Pasamonte

and

isochron age

Allgre
isochron age

whole-rock
sample
plotted with 12 other
samples
Rb-Sr

4.55

one sample plotted with


nine other meteorites
and adjusted to agree
with other methods
Lu-Hf

4.55

one sample plotted with


14 other meteorites
Lu-Hf

4.604

0.039

Blichert-Toft
2002

revised
Unruh,
Nakamura,
and
M.
Tatsumoto 1977 age
Pb-Pb

4.573

0.011

Quitt, Birck,
Allgre 2000

whole rock and


mineral separates

4.53

0.03

Unruh,
Nakamura,
and Tatsumoto 1977 isochron age

six
sample
points
(whole-rock, plag, and
density fractions)
Sm-Nd

4.58

0.12

Unruh,
Nakamura,
and Tatsumoto 1977 isochron age

two samples
with
seven
meteorites

4.5632

0.0014

Kleine et al. 2004

0.19

Patchett
Tatsumoto 1980

13
207

Pb-206Pb

ten other meteorites


and three plagioclaseHans, Kleine,
samples
Bourdon 2013

et

and
isochron age

and
isochron age
al.
isochron age
and
isochron age

plotted
other
Hf-W

isochron age

Serra De Mag
whole-rock
sample
plotted with 12 other
samples
Rb-Sr

4.55

one sample plotted with


nine other meteorites
and adjusted to agree
with other methods
Lu-Hf

4.55

one sample plotted with


14 other meteorites
Lu-Hf

4.604

0.039

Blichert-Toft
2002

0.045

Tera, Carlson,
Boctor 1997

and

4.406

whole rock, two whole


rock leachates, and
pyroxene
and
plagioclase fractions

207

Pb-204Pb

0.19

ten other meteorites


and three plagioclaseHans, Kleine,
samples
Bourdon 2013

Patchett
Tatsumoto 1980
et

and
isochron age

and
isochron age
al.
isochron age
isochron age

Pb-206Pb

4.39

0.016

Tera, Carlson,
Boctor 1997

and

207

Pb-204Pb

4.4

0.037

Tera, Carlson,
Boctor 1997

and

206

4.399

0.035

Tera, Carlson,
Boctor 1997

andisochron age

mean Pb-Pb

isochron age
isochron age

4.41

0.02

Lugmair,
Scheinin,
and Carlson 1977
isochron age

one sample plotted with


17 other meteorites
Sm-Nd

4.464

0.075

Blichert-Toft
2002

one sample plotted with


seven other meteorites Hf-W

4.5632

0.0014

Kleine et al. 2004

three fractions

4.553

0.003

Lugmair
and
Shukolyukov 1998
isochron age

one whole-rock sample


plotted with six other
meteorites
Rb-Sr

4.39

0.26

Papanastassiou and
Wasserburg 1969
isochron age

eight fractions (wholerock, three plagioclase,


pyroxene, plus three
others)
Rb-Sr

3.3

0.5

Birck
1978

Sm-Nd

Mn-Cr

et

al.
isochron age
isochron age

Stannern

and

Allgre
isochron age

whole-rock
sample
plotted with 12 other
samples
Rb-Sr

ten other meteorites


and three plagioclaseHans, Kleine,
samples
Bourdon 2013

4.55

0.19

two samples
with
nine
meteorites

4.55

0.019

Patchett
Tatsumoto 1980

4.604

0.039

Blichert-Toft
2002

4.131

0.012

Tera, Carlson,
Boctor 1997

plotted
other
Lu-Hf

one sample plotted with


14 other meteorites
Lu-Hf
207

Pb-204Pb

isochron age
and
isochron age

et

al.
isochron age
and
isochron age

207

Pb- Pb

4.124

0.023

isochron age

206

Pb-204Pb

4.13

0.012

isochron age

mean Pb-Pb

4.128

0.016

isochron age

Sm-Nd

4.48

0.07

Lugmair and Scheinin


1975
isochron age

one sample plotted with


17 other meteorites
Sm-Nd

4.464

0.075

Blichert-Toft
2002

whole rock, magnetic,


and two non-magnetic
fractions
Hf-W

4.564

0.002

Kleine et al. 2005

whole rock and density


fractions 73 matrix
Rb-Sr

0.11

(decay
0.0139)

constantBansal, Shih,
Wiesmann 1985

and

4.52

whole rock and density


fractions 73 matrix
Rb-Sr

0.11

(decay
0.0142)

constantBansal, Shih,
Wiesmann 1985

and

4.42

whole rock and density


fractions 84B clast
Rb-Sr

0.11

(decay
0.0139)

constantBansal, Shih,
Wiesmann 1985

and

4.56

whole rock and density


fractions 84B clast
Rb-Sr

0.1

(decay
0.0142)

constantBansal, Shih,
Wiesmann 1985

and

4.46

73 matrix, five data


points,
old
decay
constant
Rb-Sr

4.56

0.06

Nyquist et al. 1986

isochron age

73 matrix, five data


points,
new
decay
constant
Rb-Sr

4.46

0.06

Nyquist et al. 1986

isochron age

73 matrix, nine data


points,
old
decay
constant
Rb-Sr

4.6

0.05

Nyquist et al. 1986

isochron age

73 matrix, nine data


points,
new
decay
constant
Rb-Sr

4.5

0.05

Nyquist et al. 1986

isochron age

three-point
isochron
based on whole-rock,
pyroxene,
and
plagioclase fractions

206

and

et

al.
isochron age

isochron age

Yamato 75011
isochron age
isochron age
isochron age
isochron age

five data points

U-Pb

4.552

0.017

Misawa et al. 2005

isochron age

73 matrix, 84B clast, 10


data points
Sm-Nd

4.55

0.14

Nyquist et al. 1986

isochron age

84B clast, seven data


points
Sm-Nd

4.54

0.21

Nyquist et al. 1986

isochron age

Discussion
In contrast to the Allende CV3 carbonaceous chondrite meteorite (Snelling 2014a), there have been fewer radioisotope ages
obtained for these eucrites (basaltic achondrites), even though all the radioisotope dating methods have been used on some
of them, and a few of the methods on others. Yet the outcome is similar to that found for the ordinary and enstatite
chondrites (Snelling 2014b).
Isochron Ages
A 4.554.57 Ga isochron age for Bereba, Ibitira, Juvinas, Pasamonte, Serra de Mag, and Yamato 75011 is clearly defined
by a strong clustering of their isochron radioisotope data, via the Pb-Pb, U-Pb, Sm-Nd, Rb-Sr, and Lu-Hf methods, though
not all these methods cluster for each of these eucrites (fig. 6). As expected, the Mn-Cr, Hf-W, Al-Mg, and I-Xe methods also
yield isochron ages that coincide with the Pb-Pb and/or U-Pb isochron ages, simply because the Mn-Cr method is calibrated
against the Pb-Pb isochron age of the Lewis Cliff 86010 angrite achondrite (Lugmair and Galer 1992; Lugmair and
Shukolyukov 1998), the Hf-W method is calibrated against the Hf-W and Mn-Cr isochron ages of the St. Marguerite H4
chondrite anchored to the weighted average of its U-Pb whole-rock ages (Gpel, Manhs, and Allgre 1994; Kleine et al.
2002, 2004, 2005; Polnau and Lugmair 2001), the Al-Mg method is calibrated against the Pb-Pb isochron age of the CAIs in
the CR chondrite Acfer 059 (Amelin et al. 2002; Amelin, Wadhwa and Lugmair 2006; Wadhwa et al. 2004), and the I-Xe
method is calibrated against the I-Xe isochron age of the Shallowater aubrite achondrite (Claydon, Crowther, and Gilmour
2013), which is calibrated against the I-Xe isochron and Pb-Pb model ages of phosphate grains from the Acapulco primitive
achondrite (Brazzle et al. 1999; Gpel, Manhs, and Allgre 1992, 1994; Nichols et al. 1994).
There is also considerable scattering of the isochron radioisotope age data for these and the other eucrites studied (fig. 6).
Many Rb-Sr isochron ages are younger than the 4.554.57 Ga clustering (for Bereba, Juvinas, Moore County, Pasamonte,
Stannern, and Yamato 75011), though some are older (for Juvinas and Yamato 75011). Similarly many Sm-Nd isochron
ages are younger than the 4.554.57 Ga clustering (for Bereba, Cacheri, Ibitira, Juvinas, Moama, Moore County, Serra de
Mag, and Stannern), while a few are older (for Bereba, Ibitira, and Moore County). Somewhat surprisingly, all the Pb-Pb
isochron ages for Bereba, Cacheri, Moama, Moore County, Serra de Mag, and Stannern, and one for Juvinas, are younger
than the 4.554.57 Ga clustering, much younger in the case of Stannern. In contrast, the U-Pb isochron (concordia) ages
are always in or close to that clustering.
No consistent pattern is evident of Rb-Sr and Sm-Nd isochron ages always being younger than the Lu-Hf and U-Th-Pb
isochron ages respectively in the order of the parents atomic weights or their decay rates (half-lives), according to their
and decay mode respectively. Such a pattern would be potentially indicative of a past episode of accelerated radioisotope
decay, as suggested by Snelling (2005c) and Vardiman, Snelling, and Chaffin (2005) from their radioisotope investigations
of earth rocks and minerals.
However, it also has to be taken into account that there is still disagreement over the values of the decay constants and halflives of, for example, 87Rb and 176Lu (Snelling 2014c, d), because of there being discrepancies between determinations
based on comparisons of the Rb-Sr and Lu-Hf ages of meteorites (primarily eucrites), lunar rocks, and earth minerals and
rocks with their U-Pb, K-Ar, and Ar-Ar ages. Indeed, the comparisons of ages involving meteorites (primarily eucrites) and
lunar rocks yield a slightly higher decay constant and a slightly faster decay rate (half-life) for both 87Rb and 76Lu than for age
comparisons involving earth rocks and minerals.
Thus if the different decay constants are used in calculating the Rb-Sr and Lu-Hf ages of eucrites, then the affect would be
only small (0.64.0%) and would still result in Rb-Sr ages that are apparently too young and Lu-Hf ages that are apparently
too old. This would not change the conclusion that there are no systematic isochron age differences based on the different
atomic weights of the parent radioisotopes that would be due to a past accelerated decay event. Since in most instances the
old decay constants have been used for calculating the isochron Rb-Sr ages but the meteorite decay constant has been
used for calculating the isochron Lu-Hf ages, recalculating the isochron Rb-Sr ages using the meteorite decay constant
would only bring them closer to agreement with the isochron Lu-Hf ages. This only serves to reinforce the lack of any
consistent pattern in the isochron ages obtained by the different radioisotope systems, and thus there is no evidence of a
past accelerated decay event.
Model ages
In contrast to the isochron ages for these eucrites (basaltic achondrites), there are many more model ages for them (fig. 7).
A 4.554.57 Ga model age for Bereba, Cacheri, Camel Donga, Ibitira, and Juvinas is very clearly defined by a strong
clustering of Pb-Pb and U-Pb model ages, and for Yamato 75011 supported by Sm-Nd and Rb-Sr model ages. As again
expected, the Pu-Xe, Al-Mg, and Mn-Cr model ages always plot in, or closely adjacent to, the strong clustering of Pb-Pb and
U-Pb model ages. This is because the Pu-Xe model ages are calibrated against the Pb-Pb isochron age of the Angra dos
Reis (ADOR) angrite achondrite (Lugmair and Galer 1992; Lugmair and Marti 1977; Miura et al. 1998; Shukolyukov and
Begemann 1996a, b), the Al-Mg model ages are calibrated against the Pb-Pb isochron age of the CAIs in the Allende CV3
chondrite (Jacobsen et al. 2008; Schiller, Baker, and Bizzarro 2010), and the Mn-Cr model age for Caldera is calibrated
against its Sm-Nd and Pb-Pb isochron ages (Galer and Lugmair 1996; Wadhwa and Lugmair 1996).
Scattering of the model ages is prolific, with K-Ar and Ar-Ar model ages generally being younger than the 4.554.57 Ga
clustering for all these eucrite meteorites, except for a few that are older for Ibitira, Moore County, Pasamonte, Serra de
Mag, and Stannern (fig. 7). Similarly the U-Pb model ages are either younger or older than the clustering, or in most
instances both, whereas for Camel Donga they are usually younger, but for Juvinas and Yamato 75011 they are invariably
older. Some of the Pb-Pb model ages for Bereba, Cacheri, Camel Donga, Ibitira, Juvinas, and Stannern are younger than
the 4.554.57 Ga clustering, sometimes much younger, and a few for Ibitira, Juvinas, and Pasamonte are much older. For
Yamato 75011 the Rb-Sr model ages are scattered either side of the strong clustering. Where available for Ibitira and
Pasamonte, the Th-Pb model ages are nearly all older.
Table 2. Model ages for whole-rock samples and some or all components of 12 eucrite achondrites, with the details and
literature sources.
Sample
Bereba

Method

Date

Error +/- Note

Source

Type

whole rock samples

K-Ar

2.8

Heymann, Mazor, and


Anders 1968
model age

pyroxene

K-Ar

Hampel et al. 1980

model age

K-Ar

3.3

Hampel et al. 1980

model age

plagioclase

K-Ar

3.31

Shukolyukov
Begemann 1996b

whole rock

Pb-Pb

4.44

Manhs et al. 1975

model age

Pb-Pb

4.44

Manhs et al. 1975

model age

4.415

Manhs et al. 1975

model age

Pb-Pb

4.536

Carlson Tera, and Boctor


1988
model age

Pb-Pb

4.536

Tera,
Carlson,
Boctor 1997

and
model age

mean of all Pb isotope


combinations on all samples
Pb-Pb

4.521

0.0004

Tera,
Carlson,
Boctor 1997

and
model age

whole rock sample

Pb-Pb

whole rock samples


whole
model

rock,

single

stage

and
model age

207

Pb-206Pb 4.534

0.016

Bukovanska and Ireland


1993
model age

207

Pb-206Pb 4.552

0.02

Zhou et al. 2013

model age

Spot 1-1

207

Pb-206Pb 4.515

0.025

Zhou et al. 2013

model age

Spot 1-2

207

Pb-206Pb 4.569

0.02

Zhou et al. 2013

model age

Spot 2

207

Pb-206Pb 4.574

0.026

Zhou et al. 2013

model age

Spot 3-1

207

Pb-206Pb 4.549

0.019

Zhou et al. 2013

model age

Spot 3-2

207

Pb- Pb 4.545

0.024

Zhou et al. 2013

model age

Spot 1-1

207

Pb-235U

4.425

0.034

Zhou et al. 2013

model age

Spot 1-2

207

Pb- U

4.572

0.03

Zhou et al. 2013

model age

Spot 2

207

Pb-235U

4.563

0.032

Zhou et al. 2013

model age

Spot 3-1

207

Pb-235U

4.637

0.037

Zhou et al. 2013

model age

Spot 3-2

207

Pb-235U

4.49

0.031

Zhou et al. 2013

model age

Spot 1-1

206

Pb-238U

4.229

0.089

Zhou et al. 2013

model age

Spot 1-2

206

Pb- U

4.579

0.085

Zhou et al. 2013

model age

Spot 2

206

Pb-238U

4.539

0.084

Zhou et al. 2013

model age

Spot 3-1

206

Pb- U

4.841

0.114

Zhou et al. 2013

model age

206

Pb-238U

4.369

0.083

Zhou et al. 2013

model age

4.498

0.016

Shukolyukov
Begemann 1996a

4.512

0.018

Miura et al. 1998

model age

zircons
weighted
analyses
grains

average of five
of three zircon

Spot 3-2

206

235

238

238

Pu-Xe

whole rock sample relative to


ADOR
Pu-Xe

and
model age

Cachari
whole rock sample

Ar-Ar

3.04

0.7

Bogard et al. 1985

model age

whole rock sample

Ar-Ar

3.47

0.4

Bogard et al. 1985

model age

whole rock sample

Pb-Pb

4.13

three analyses of one zircon


grain
sixteen analyses of six zircon
grains
Grain 1-1

Tera,
Carlson,
Boctor 1987

and
model age

207

Pb-206Pb 4.549

0.013

weighted average

Zhou et al. 2011

model age

207

Pb-206Pb 4.551

0.014

weighted mean

Zhou et al. 2013

model age

207

Pb-206Pb 4.539

0.017

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 1-2

207

Pb-206Pb 4.533

0.017

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 1-3

207

Pb-206Pb 4.556

0.017

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 2-1

207

Pb-206Pb 4.535

0.019

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 3-1

207

Pb-206Pb 4.579

0.016

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 3-2

207

Pb-206Pb 4.587

0.016

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 3-3

207

Pb-206Pb 4.596

0.016

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 4-1

207

Pb-206Pb 4.565

0.019

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 4-2

207

Pb-206Pb 4.546

0.021

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 5-1

207

Pb-206Pb 4.51

0.019

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 5-2

207

Pb-206Pb 4.516

0.019

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 5-3

207

Pb-206Pb 4.524

0.018

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 5-4

207

Pb-206Pb 4.563

0.018

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 5-5

207

Pb-206Pb 4.54

0.019

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 6-1

207

Pb-206Pb 4.558

0.019

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 6-2

207

Pb-206Pb 4.515

0.019

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 6-3

207

Pb-206Pb 4.564

0.02

ages using Canyon


Diablo troilite Pb
Zhou et al. 2013

model age

Grain 1-1

207

Pb-206Pb 4.539

0.018

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 1-2

207

Pb-206Pb 4.533

0.018

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 1-3

207

Pb-206Pb 4.557

0.016

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 2-1

207

Pb-206Pb 4.536

0.019

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 3-1

207

Pb-206Pb 4.58

0.016

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 3-2

207

Pb-206Pb 4.588

0.016

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 3-3

207

Pb-206Pb 4.596

0.016

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 4-1

207

Pb-206Pb 4.566

0.019

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 4-2

207

Pb-206Pb 4.547

0.021

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 5-1

207

Pb-206Pb 4.51

0.019

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 5-2

207

Pb-206Pb 4.516

0.019

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 5-3

207

Pb-206Pb 4.524

0.018

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 5-4

207

Pb-206Pb 4.563

0.018

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 5-5

207

Pb-206Pb 4.54

0.019

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 6-1

207

Pb-206Pb 4.559

0.019

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 6-2

207

Pb-206Pb 4.515

0.017

ages using terrestrial


common Pb
Zhou et al. 2013

model age

Grain 6-3

207

Pb-206Pb 4.565

0.02

ages using terrestrial


common Pb
Zhou et al. 2013

model age

207

Pb-206Pb 4.55

0.039

Zhou et al. 2013

model age

weighted averages of three


analyses of one zircon grain
Spot 7-1

207

Pb- Pb 4.529

0.013

Zhou et al. 2013

model age

Spot 7-2

207

Pb-206Pb 4.549

0.032

Zhou et al. 2013

model age

Spot 7-3

207

Pb- Pb 4.596

0.093

Zhou et al. 2013

model age

207

Pb-235U

4.568

0.033

Zhou et al. 2013

model age

weighted averages of three


analyses of one zircon grain

206

206

Spot 7-1

207

Pb- U

4.561

0.013

Zhou et al. 2013

model age

Spot 7-2

207

Pb-235U

4.562

0.029

Zhou et al. 2013

model age

Spot 7-3

207

Pb- U

4.565

0.081

Zhou et al. 2013

model age

206

Pb-238U

4.608

0.096

Zhou et al. 2013

model age

Spot 7-1

206

Pb-238U

4.555

0.01

Zhou et al. 2013

model age

Spot 7-2

206

Pb-238U

4.586

0.027

Zhou et al. 2013

model age

Spot 7-3

206

4.659

0.081

Zhou et al. 2013

model age

weighted averages of three


analyses of one zircon grain

235

235

238

Pb- U

whole rock sample relative to


ADOR
Pu-Xe

Shukolyukov
Begemann 1996a

and

4.498

glass sample
ADOR

4.517

Shukolyukov
Begemann 1996b

and

Pu-Xe

K-Ar

4.19

Shukolyukov
Begemann 1996b

and

whole rock sample


pyroxene fraction

Pb-Pb

4.525

0.0019

Galer and Lugmair 1996 model age

weighted mean of four points


on one zircon grain
Pb-Pb

4.542

0.08

Zhou et al. 2013

model age

relative

to

model age
model age

Caldera
model age

Spot 1-1

207

Pb-206Pb 4.591

0.034

Zhou et al. 2013

model age

Spot 1-2

207

Pb-206Pb 4.502

0.045

Zhou et al. 2013

model age

Spot 1-3

207

Pb-206Pb 4.49

0.03

Zhou et al. 2013

model age

Spot 1-4

207

Pb- Pb 4.573

0.029

Zhou et al. 2013

model age

Spot 1-1

207

Pb-235U

4.575

0.069

Zhou et al. 2013

model age

Spot 1-2

207

Pb- U

4.491

0.074

Zhou et al. 2013

model age

Spot 1-3

206

Pb-238U

5.131

0.156

Zhou et al. 2013

model age

Spot 1-4

206

Pb-238U

5.048

0.133

Zhou et al. 2013

model age

Spot 1-1

206

Pb- U

4.54

0.204

Zhou et al. 2013

model age

Spot 1-2

206

Pb-238U

4.467

0.207

Zhou et al. 2013

model age

Spot 1-3

207

Pb- U

4.68

0.05

Zhou et al. 2013

model age

Spot 1-4

207

Pb-235U

4.714

0.044

Zhou et al. 2013

model age

4.537

0.012

Wadhwa and Lugmairmodel age


1996

206

235

238

235

anchored to other meteorite Mn-Cr


dates

Pu-Xe

Shukolyukov
Begemann 1996a

4.513

and
model age

Camel Donga
matrix
samples
pyroxenes, weighted

with

matrix
samples
pyroxenes, weighted

with

Ar-Ar

3.704

0.079

Kennedy et al. 2013

plateau age

Ar-Ar

3.67

0.08

Kennedy et al. 2013

plateau age

weight mean of two

Ar-Ar

3.693

0.051

Kennedy et al. 2013

plateau age

pyroxene fraction

Pb-Pb

4.5109

0.001

Iizuka et al. 2013

model age

Zhou et al. 2013

model age

thirty-five point analyses of 14


zircon grains

207

Pb-206Pb 4.531

0.01

Spot 1-1

207

Pb-206Pb 4.525

0.014

Zhou et al. 2013

model age

Spot 1-2

207

Pb- Pb 4.579

0.012

Zhou et al. 2013

model age

Spot 2-1

207

Pb-206Pb 4.551

0.023

Zhou et al. 2013

model age

Spot 2-2

207

Pb- Pb 4.547

0.017

Zhou et al. 2013

model age

Spot 3

207

Pb-206Pb 4.552

0.051

Zhou et al. 2013

model age

Spot 4

207

Pb-206Pb 4.512

0.02

Zhou et al. 2013

model age

Spot 5-1

207

Pb-206Pb 4.572

0.033

Zhou et al. 2013

model age

Spot 5-2

207

Pb-206Pb 4.516

0.032

Zhou et al. 2013

model age

Spot 5-3

207

Pb- Pb 4.531

0.02

Zhou et al. 2013

model age

Spot 6

207

Pb-206Pb 4.479

0.034

Zhou et al. 2013

model age

Spot 7-1

207

Pb- Pb 4.546

0.014

Zhou et al. 2013

model age

Spot 7-2

207

Pb-206Pb 4.517

0.02

Zhou et al.. 2013

model age

Spot 7-3

207Pb-206Pb

4.557

0.013

Zhou et al. 2013

model age

Spot 7-4

207Pb-206Pb

4.53

0.021

Zhou et al. 2013

model age

Spot 8

207Pb-206Pb

4.484

0.022

Zhou et al. 2013

model age

Spot 9-1

207Pb-206Pb

4.546

0.019

Zhou et al. 2013

model age

Spot 9-2

207Pb-206Pb

4.542

0.023

Zhou et al. 2013

model age

Spot 9-3

207Pb-206Pb

4.536

0.018

Zhou et al. 2013

model age

Spot 9-4

207Pb-206Pb

4.471

0.03

Zhou et al. 2013

model age

Spot 9-5

207Pb-206Pb

4.507

0.032

Zhou et al. 2013

model age

Spot 9-6

207Pb-206Pb

4.514

0.025

Zhou et al. 2013

model age

Spot 9-7

207Pb-206Pb

4.535

0.037

Zhou et al. 2013

model age

Spot 10-1

207Pb-206Pb

4.519

0.022

Zhou et al. 2013

model age

Spot 10-2

207Pb-206Pb

4.481

0.022

Zhou et al. 2013

model age

Spot 10-3

207Pb-206Pb

4.525

0.019

Zhou et al. 2013

model age

Spot 11

207Pb-206Pb

4.534

0.046

Zhou et al. 2013

model age

Spot 12-1

207Pb-206Pb

4.566

0.019

Zhou et al. 2013

model age

Spot 12-2

207Pb-206Pb

4.537

0.017

Zhou et al. 2013

model age

Spot 12-3

207Pb-206Pb

4.536

0.018

Zhou et al. 2013

model age

Spot 12-4

207Pb-206Pb

4.57

0.015

Zhou et al. 2013

model age

Spot 12-5

207Pb-206Pb

4.551

0.026

Zhou et al. 2013

model age

Spot 13

207Pb-206Pb

4.511

0.022

Zhou et al. 2013

model age

Spot 14-1

207Pb-206Pb

4.478

0.013

Zhou et al. 2013

model age

Spot 14-2

207Pb-206Pb

4.479

0.016

Zhou et al. 2013

model age

206

206

206

206

weighted average

207Pb-206Pb

4.52

0.014

207Pb-235U

4.492

0.023

Spot 1-1

207Pb-235U

4.461

Spot 1-2

207Pb-235U

Spot 2-1

207Pb-235U

Spot 2-2

Spot 14-3

Zhou et al. 2013

model age

Zhou et al. 2013

model age

0.034

Zhou et al. 2013

model age

4.419

0.033

Zhou et al. 2013

model age

4.466

0.042

Zhou et al. 2013

model age

207Pb-235U

4.513

0.035

Zhou et al. 2013

model age

Spot 3

207Pb-235U

4.561

0.064

Zhou et al. 2013

model age

Spot 4

207Pb-235U

4.405

0.036

Zhou et al. 2013

model age

Spot 5-1

207Pb-235U

4.531

0.043

Zhou et al. 2013

model age

Spot 5-2

207Pb-235U

4.48

0.042

Zhou et al. 2013

model age

Spot 5-3

207Pb-235U

4.493

0.036

Zhou et al. 2013

model age

Spot 6

207Pb-235U

4.691

0.042

Zhou et al. 2013

model age

Spot 7-1

207Pb-235U

4.469

0.034

Zhou et al. 2013

model age

Spot 7-2

207Pb-235U

4.486

0.035

Zhou et al. 2013

model age

Spot 7-3

207Pb-235U

4.433

0.034

Zhou et al. 2013

model age

Spot 7-4

207Pb-235U

4.513

0.035

Zhou et al. 2013

model age

Spot 8

207Pb-235U

4.597

0.037

Zhou et al. 2013

model age

Spot 9-1

207Pb-235U

4.53

0.035

Zhou et al. 2013

model age

Spot 9-2

207Pb-235U

4.657

0.037

Zhou et al. 2013

model age

Spot 9-3

207Pb-235U

4.451

0.035

Zhou et al. 2013

model age

Spot 9-4

207Pb-235U

4.586

0.038

Zhou et al. 2013

model age

Spot 9-5

207Pb-235U

4.374

0.041

Zhou et al. 2013

model age

Spot 9-6

207Pb-235U

4.545

0.039

Zhou et al. 2013

model age

Spot 9-7

207Pb-235U

4.498

0.045

Zhou et al. 2013

model age

Spot 10-1

207Pb-235U

4.539

0.036

Zhou et al. 2013

model age

Spot 10-2

207Pb-235U

4.451

0.035

Zhou et al. 2013

model age

Spot 10-3

207Pb-235U

4.497

0.035

Zhou et al. 2013

model age

Spot 11

207Pb-235U

4.52

0.05

Zhou et al. 2013

model age

Spot 12-1

207Pb-235U

4.473

0.035

Zhou et al. 2013

model age

Spot 12-2

207Pb-235U

4.476

0.035

Zhou et al. 2013

model age

Spot 12-3

207Pb-235U

4.431

0.035

Zhou et al. 2013

model age

Spot 12-4

207Pb-235U

4.496

0.035

Zhou et al. 2013

model age

Spot 12-5

207Pb-235U

4.568

0.039

Zhou et al. 2013

model age

Spot 13

207Pb-235U

4.523

0.036

Zhou et al. 2013

model age

Spot 14-1

207Pb-235U

4.437

0.033

Zhou et al. 2013

model age

Spot 14-2

207Pb-235U

4.405

0.034

Zhou et al. 2013

model age

Spot 14-3

207Pb-235U

4.447

0.034

Zhou et al. 2013

model age

206Pb-238U

4.417

0.079

Zhou et al. 2013

model age

Spot 1-1

206Pb-238U

4.32

0.101

Zhou et al. 2013

model age

Spot 1-2

206Pb-238U

4.077

0.096

Zhou et al. 2013

model age

Spot 2-1

206Pb-238U

4.279

0.12

Zhou et al. 2013

model age

Spot 2-2

206Pb-238U

4.438

0.104

Zhou et al. 2013

model age

thirty-five point analyses of 14


zircon grains

thirty-five point analyses of 14


zircon grains

weighted average

weighted average

Spot 3

206Pb-238U

4.582

0.167

Zhou et al. 2013

model age

Spot 4

206Pb-238U

4.174

0.101

Zhou et al. 2013

model age

Spot 5-1

206Pb-238U

4.439

0.112

Zhou et al. 2013

model age

Spot 5-2

206Pb-238U

4.4

0.111

Zhou et al. 2013

model age

Spot 5-3

206

Pb-238U

4.41

0.104

Zhou et al. 2013

model
age

Spot 6

206

Pb-238U

5.196

0.119

Zhou et al. 2013

model
age

Spot 7-1

206

Pb-238U

4.301

0.101

Zhou et al. 2013

model
age

Spot 7-2

206

Pb-238U

4.417

0.1

Zhou et al. 2013

model
age

Spot 7-3

206

Pb-238U

4.165

0.1

Zhou et al. 2013

model
age

Spot 7-4

206

Pb-238U

4.476

0.101

Zhou et al. 2013

model
age

Spot 8

206

Pb-238U

4.857

0.112

Zhou et al. 2013

model
age

Spot 9-1

206

Pb-238U

4.495

0.105

Zhou et al. 2013

model
age

Spot 9-2

206

Pb-238U

4.923

0.114

Zhou et al. 2013

model
age

Spot 9-3

206

Pb-238U

4.267

0.101

Zhou et al. 2013

model
age

Spot 9-4

206

Pb-238U

4.852

0.108

Zhou et al. 2013

model
age

Spot 9-5

206Pb-238U

4.091

0.102

Zhou et al. 2013

model
age

Spot 9-6

206Pb-238U

4.616

0.112

Zhou et al. 2013

model
age

Spot 9-7

206Pb-238U

4.415

0.115

Zhou et al. 2013

model
age

Spot 10-1

206Pb-238U

4.584

0.105

Zhou et al. 2013

model
age

Spot 10-2

206Pb-238U

4.385

0.1

Zhou et al. 2013

model
age

Spot 10-3

206Pb-238U

4.435

0.103

Zhou et al. 2013

model
age

Spot 11

206Pb-238U

4.489

0.118

Zhou et al. 2013

model
age

Spot 12-1

206Pb-238U

4.271

0.099

Zhou et al. 2013

model
age

Spot 12-2

206Pb-238U

4.343

0.103

Zhou et al. 2013

model
age

Spot 12-3

206Pb-238U

4.206

0.101

Zhou et al. 2013

model
age

Spot 12-4

206Pb-238U

4.334

0.103

Zhou et al. 2013

model
age

Spot 12-5

206Pb-238U

4.607

0.11

Zhou et al. 2013

model
age

Spot 13

206Pb-238U

4.549

0.106

Zhou et al. 2013

model
age

Spot 14-1

206Pb-238U

4.346

0.1

Zhou et al. 2013

model
age

Spot 14-2

206Pb-238U

4.246

0.1

Zhou et al. 2013

model
age

Spot 14-3

206Pb-238U

4.288

0.1

Zhou et al. 2013

model
age

Al-Mg

4.5647

0.0004

Schiller, Baker, andmodel


Bizzarro 2010
age

Pu-Xe

4.521

0.02

Shukolyukov
andmodel
Begemann 1996a
age

Pu-Xe

4.507

0.016

Miura et al. 1998

relative to ADOR
average
samples
ADOR

of
seven
relative
to

model
age

Ibitira

K-Ar
whole-rock samples

Ar-Ar

3.2

Heymann,
Mazor,
and
Anders 1968 model age

4.49

Garrison
and
Bogard 1995 model age

using
15
stepwise
temperature extractions Ar-Ar

4.495

0.015

Bogard
and
Garrison 1995 plateau age

Sample at 400C

Ar-Ar

3.354

0.048

Bogard
and
Garrison 1995 extraction age

Sample at 500C

Ar-Ar

3.205

0.051

Bogard
and
Garrison 1995 extraction age

Sample at 600C

Ar-Ar

3.525

0.023

Bogard
and
Garrison 1995 extraction age

Sample at 700C

Ar-Ar

3.874

0.008

Bogard
and
Garrison 1995 extraction age

Sample at 775C

Ar-Ar

4.361

0.009

Bogard
and
Garrison 1995 extraction age

Sample at 825C

Ar-Ar

4.469

0.01

Bogard
and
Garrison 1995 extraction age

Sample at 875C

Ar-Ar

4.502

0.009

Bogard
and
Garrison 1995 extraction age

Sample at 930C

Ar-Ar

4.493

0.009

Bogard
and
Garrison 1995 extraction age

Sample at 975C

Ar-Ar

4.491

0.009

Bogard
and
Garrison 1995 extraction age

Sample at 1025C

Ar-Ar

4.509

0.008

Bogard
and
Garrison 1995 extraction age

Sample at 1100C

Ar-Ar

4.452

0.029

Bogard
and
Garrison 1995 extraction age

Sample at 1200C

Ar-Ar

4.491

0.098

Bogard
and
Garrison 1995 extraction age

Sample at 1300C

Ar-Ar

4.588

0.036

Bogard
and
Garrison 1995 extraction age

Sample at 1400C

Ar-Ar

4.48

0.103

Bogard
and
Garrison 1995 extraction age

Sample at 1550C

Ar-Ar

7.998

0.425

Bogard
and
Garrison 1995 extraction age

Ar-Ar

4.487

0.015

Yamaguchi
al. 2001

five extractions releasing


89% of Ar
Ar-Ar

4.487

0.016

Bogard
and
Garrison 2003 plateau age

4.4858

0.015

after
Bogard
Garrison (1995)

after
Bogard
Garrison (1995)

and

and Ar-Ar

et
plateau (model) age

Claydon,
plateau (model) age
Crowther, and
Gilmour 2012

207

Ibitira WR-1

Ibitira WR-2

pyroxene fraction
whole
rock
fraction
A017_7 residue
whole
rock
fraction
A017_7 wash-1
whole
rock
fraction
A017_7 wash-2
whole
rock
fraction
A039_5 residue
whole
rock
fraction
A039_5 wash-1
whole
rock
fraction
A039_5 wash-2
whole
rock
fraction
A047_3 residue
whole
rock
fraction
A047_3 wash-1
whole
rock
fraction
A047_3 wash-2
pyroxene-rich
fraction
GSC030_HF1 residue
pyroxene-rich
fraction
GSC030_HF1 wash-1
pyroxene-rich
fraction
GSC030_HF1 wash-2
pyroxene-rich
fraction
GSC030_HF2 residue
pyroxene-rich
fraction
GSC030_HF2 wash-1
pyroxene-rich
fraction
GSC030_HF2 wash-2
pyroxene-rich
fraction
A015_4 residue
pyroxene-rich
fraction
A015_4 wash-1
pyroxene-rich
fraction
A015_4 wash-2
pyroxene-rich
fraction
A015_5 residue
pyroxene-rich
fraction
A015_5 wash-1
pyroxene-rich
fraction
A015_5 wash-2
pyroxene-rich
fraction
A015_6 residue

207

Pb-206Pb

Pb-206Pb

207Pb-206Pb

4.55

4.554

4.556

0.01

Wasserburg et
al. 1977
model age

0.008

Chen
and
Wasserburg
1985
model age

0.006

Chen
and
Wasserburg
1985
model age

207Pb-206Pb

4.56

0.003

Manhs, Gpel,
and
Allgre
1987
model age

4.5558

0.0005

Iizuka
2013

et

Pb-Pb

4.5564

0.0011

Iizuka
2014

et

Pb-Pb

4.5543

0.0022

Iizuka
2014

et

Pb-Pb

4.584

0.0022

Iizuka
2014

et

Pb-Pb

4.556

0.0008

Iizuka
2014

et

Pb-Pb

4.5894

0.0022

Iizuka
2014

et

Pb-Pb

4.5881

0.0036

Iizuka
2014

et

Pb-Pb

4.5572

0.0013

Iizuka
2014

et

Pb-Pb

4.608

0.0024

Iizuka
2014

et

Pb-Pb

4.6043

0.0054

Iizuka
2014

et

Pb-Pb

4.5558

0.0007

Iizuka
2014

et

Pb-Pb

4.5533

0.0002

Iizuka
2014

et

Pb-Pb

4.556

0.0021

Iizuka
2014

et

Pb-Pb

4.5559

0.0007

Iizuka
2014

et

Pb-Pb

4.553

0.0002

Iizuka
2014

et

Pb-Pb

4.558

0.0024

Iizuka
2014

et

Pb-Pb

4.5558

0.0009

Iizuka
2014

et

Pb-Pb

4.5361

0.002

Iizuka
2014

et

Pb-Pb

4.6139

0.0549

Iizuka
2014

et

Pb-Pb

4.5506

0.0075

Iizuka
2014

et

Pb-Pb

4.5141

0.0021

Iizuka
2014

et

Pb-Pb

4.5709

0.0201

Iizuka
2014

et

Pb-Pb

4.552

0.001

Iizuka
2014

et

Pb-Pb

al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age

pyroxene-rich
fraction
A015_6 wash-1
pyroxene-rich
fraction
A015_6 wash-2
pyroxene-rich
fraction
A017_5 residue
pyroxene-rich
fraction
A017_5 wash-2
pyroxene-rich
fraction
A017_6 residue
pyroxene-rich
fraction
A017_6 wash-1
pyroxene-rich
fraction
A017_6 wash-2
pyroxene-rich
fraction
A039_2 residue
pyroxene-rich
fraction
A039_2 wash-1
pyroxene-rich
fraction
A039_2 wash-2
pyroxene-rich
fraction
A039_3 residue
pyroxene-rich
fraction
A039_3 wash-1
pyroxene-rich
fraction
A039_3 wash-2
pyroxene-rich
fraction
A047_1 residue
pyroxene-rich
fraction
A047_1 wash-1
pyroxene-rich
fraction
A047_1 wash-2
pyroxene-rich
fraction
A047_2 residue
pyroxene-rich
fraction
A047_2 wash-1
pyroxene-rich
fraction
A047_2 wash-2
plagioclase-rich fraction
A015_7 residue
plagioclase-rich fraction
A015_7 wash-1
plagioclase-rich fraction
A015_7 wash-2
plagioclase-rich fraction
A015_8 residue
plagioclase-rich fraction
A015_8 wash-1
plagioclase-rich fraction
A015_8 wash-2
plagioclase-rich fraction
A039_4 residue
plagioclase-rich fraction
A039_4 wash-1
plagioclase-rich fraction
A039_4 wash-2

4.5211

0.001

Iizuka
2014

et

Pb-Pb

4.5329

0.0242

Iizuka
2014

et

Pb-Pb

4.5581

0.0013

Iizuka
2014

et

Pb-Pb

4.5727

0.0025

Iizuka
2014

et

Pb-Pb

4.5546

0.0009

Iizuka
2014

et

Pb-Pb

4.552

0.001

Iizuka
2014

et

Pb-Pb

4.5265

0.0065

Iizuka
2014

et

Pb-Pb

4.5563

0.0007

Iizuka
2014

et

Pb-Pb

4.5634

0.0016

Iizuka
2014

et

Pb-Pb

4.539

0.0038

Iizuka
2014

et

Pb-Pb

4.5562

0.0008

Iizuka
2014

et

Pb-Pb

4.5822

0.0024

Iizuka
2014

et

Pb-Pb

4.5594

0.0045

Iizuka
2014

et

Pb-Pb

4.5565

0.0006

Iizuka
2014

et

Pb-Pb

4.5583

0.0014

Iizuka
2014

et

Pb-Pb

4.5815

0.0125

Iizuka
2014

et

Pb-Pb

4.5567

0.0007

Iizuka
2014

et

Pb-Pb

4.5632

0.002

Iizuka
2014

et

Pb-Pb

4.5843

0.0159

Iizuka
2014

et

Pb-Pb

4.5381

0.0052

Iizuka
2014

et

Pb-Pb

4.5232

0.001

Iizuka
2014

et

Pb-Pb

4.6072

0.0072

Iizuka
2014

et

Pb-Pb

4.5139

0.0038

Iizuka
2014

et

Pb-Pb

4.5456

0.0008

Iizuka
2014

et

Pb-Pb

4.6352

0.0032

Iizuka
2014

et

Pb-Pb

4.5636

0.0135

Iizuka
2014

et

Pb-Pb

4.5515

0.0007

Iizuka
2014

et

Pb-Pb

4.6311

0.0058

Iizuka
2014

et

Pb-Pb

al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age

plagioclase-rich fraction
A047_4 residue
plagioclase-rich fraction
A047_4 wash-1
plagioclase-rich fraction
A047_4 wash-2
plagioclase-rich fraction
A047_5 residue
plagioclase-rich fraction
A047_5 wash-1
plagioclase-rich fraction
A047_5 wash-2
weighted average of
residues of 8 pyroxenerich fractions and two
whole-rock fractions and
washes of two pyroxenerich and one plagioclaserich fractions (13 ages)
weighted average of
residues of one wholerock and six pyroxenerich fractions (seven
ages)

Ibitira WR-1

Ibitira WR-2

Ibitira WR-1

Ibitira WR-2

Ibitira WR-1

Ibitira WR-2

Single sample

4.5595

0.0108

Iizuka
2014

et

Pb-Pb

4.5606

0.0006

Iizuka
2014

et

Pb-Pb

4.6722

0.0123

Iizuka
2014

et

Pb-Pb

4.5528

0.0019

Iizuka
2014

et

Pb-Pb

4.5572

0.0005

Iizuka
2014

et

Pb-Pb

4.387

0.0275

Iizuka
2014

et

Pb-Pb

4.55635

0.00042

Iizuka
2014

et

Pb-Pb

4.55631

0.00028

Iizuka
2014

et

Pb-Pb

0.03

Chen
and
Wasserburg
1985
model age

0.013

Chen
and
Wasserburg
1985
model age

0.007

Chen
and
Wasserburg
1985
model age

0.005

Chen
and
Wasserburg
1985
model age

0.036

Chen
and
Wasserburg
1985
model age

0.022

Chen
and
Wasserburg
1985
model age

0.025

Shukolyukov
and Begemann
1996a
model age

206Pb-238U

4.629

206Pb-238U

4.567

207Pb-235U

4.577

207Pb-235U

4.559

208Pb-232Th

4.972

208Pb-232Th

4.658

Pu-Xe

4.581

al.
model age
al.
model age
al.
model age
al.
model age
al.
model age
al.
model age

al.
model age

al.
model age

Juvinas
whole rock samples

K-Ar

3.98

Heymann, Mazor,
and Anders 1968 model age

plagioclase sample

K-Ar

4.6

Hampel
1980

frB58

Pb-Pb

4.5676

0.0011

Manhs, Allgre,
and Provost 1984 model age

frB60

Pb-Pb

4.5676

0.0034

Manhs, Allgre,
and Provost 1984 model age

frB63

Pb-Pb

4.5648

0.001

Manhs, Allgre,
and Provost 1984 model age

frC64

Pb-Pb

4.5646

0.0011

Manhs, Allgre,
and Provost 1984 model age

frC65

Pb-Pb

4.5656

0.0008

Manhs, Allgre,
and Provost 1984 model age

et

al.
model age

frC66

Pb-Pb

4.5658

0.0007

Manhs, Allgre,
and Provost 1984 model age

frC67

Pb-Pb

4.5649

0.001

Manhs, Allgre,
and Provost 1984 model age

frD70

Pb-Pb

4.5534

0.0006

Manhs, Allgre,
and Provost 1984 model age

pla54

Pb-Pb

4.796

0.0075

Manhs, Allgre,
and Provost 1984 model age

pla57

Pb-Pb

4.7278

0.0049

Manhs, Allgre,
and Provost 1984 model age

pla62

Pb-Pb

4.821

0.0011

Manhs, Allgre,
and Provost 1984 model age

pla70

Pb-Pb

4.7979

0.0044

Manhs, Allgre,
and Provost 1984 model age

pyr62

Pb-Pb

4.5726

0.0038

Manhs, Allgre,
and Provost 1984 model age

pyr70

Pb-Pb

4.5768

0.0017

Manhs, Allgre,
and Provost 1984 model age

L1

Pb-Pb

4.634

0.003

Manhs, Allgre,
and Provost 1984 model age

L2

Pb-Pb

4.8131

0.0097

Manhs, Allgre,
and Provost 1984 model age

L3

Pb-Pb

4.8

0.0011

Manhs, Allgre,
and Provost 1984 model age

Res

Pb-Pb

4.953

0.0015

Manhs, Allgre,
and Provost 1984 model age

plaC

Pb-Pb

4.735

average age

Pb-Pb

4.527

0.024

two zircons Lee et al. 2009

207Pb-206Pb

4.549

0.006

clast
sample

weighted average of 20 model ages


determined on seven zircon grains
207Pb-206Pb

4.545

0.015

Zhou et al. 2013 model age

Spot 1-1

207Pb-206Pb

4.54

0.024

Zhou et al. 2013 model age

Spot 1-2

207Pb-206Pb

4.493

0.022

Zhou et al. 2013 model age

Spot 1-3

207Pb-206Pb

4.502

0.024

Zhou et al. 2013 model age

Spot 1-4

207Pb-206Pb

4.54

0.024

Zhou et al. 2013 model age

Spot 1-5

207Pb-206Pb

4.568

0.025

Zhou et al. 2013 model age

Spot 1-6

207Pb-206Pb

4.52

0.024

Zhou et al. 2013 model age

Spot 1-7

206Pb-238U

5.079

0.155

Zhou et al. 2013 model age

Spot 2-1

206Pb-238U

4.648

0.151

Zhou et al. 2013 model age

Spot 2-2

206Pb-238U

5.035

0.129

Zhou et al. 2013 model age

Spot 3-1

206Pb-238U

5.203

0.14

Zhou et al. 2013 model age

Spot 3-2

206Pb-238U

5.06

0.101

Zhou et al. 2013 model age

Spot 3-3

206Pb-238U

5.073

0.093

Zhou et al. 2013 model age

Spot 4-1

206Pb-238U

4.643

0.138

Zhou et al. 2013 model age

Spot 4-2

206Pb-238U

4.905

0.094

Zhou et al. 2013 model age

Spot 4-3

206Pb-238U

5.461

0.105

Zhou et al. 2013 model age

Spot 4-4

206Pb-238U

4.782

0.11

Zhou et al. 2013 model age

Spot 5-1

206Pb-238U

4.701

0.111

Zhou et al. 2013 model age

Manhs, Allgre,
and Provost 1984 model age
model age

Manhs
and
Allgre 1980
model age

Spot 5-2

206Pb-238U
206

Spot 6

Pb-238U

4.887

0.125

Zhou et al. 2013 model age

5.8

0.352

Zhou et al. 2013

model age

Pb- U

5.126

0.367

Zhou et al. 2013

model age

Al-Mg

4.5645

0.0004

Schiller, Baker, and Bizzarro


2010
model age

Pu-Xe

4.551

0.015

Pu-Xe

4.548

0.023

Miura et al. 1998

model age

average age of
three extractions
releasing 4578%
of 39Ar
Ar-Ar

4.48

0.007

Bogard and Garrison 2003

plateau age

sample at 350C

Ar-Ar

0.866

0.07

Bogard and Garrison 2003

extraction
age

sample at 400C

Ar-Ar

1.441

0.073

Bogard and Garrison 2003

extraction
age

sample at 500C

Ar-Ar

1.439

0.052

Bogard and Garrison 2003

extraction
age

sample at 600C

Ar-Ar

2.213

0.015

Bogard and Garrison 2003

extraction
age

206

Spot 7

238

Shukolyukov and Begemann


three samples 1996a
model age

Moama

sample
675C

at

sample
725C

at

sample
775C

at

sample
810C

at

sample
847C

at

sample
880C

at

sample
900C

at

sample
915C

at

sample
925C

at

sample
940C

at

sample
975C

at

sample
1025C

at

sample
1100C

at

sample
1200C

at

sample
1350C

at

sample
1600C

at

Moore County

Ar-Ar

2.865

0.017

Bogard and Garrison


2003
extraction age

Ar-Ar

3.743

0.01

Bogard and Garrison


2003
extraction age

Ar-Ar

4.167

0.01

Bogard and Garrison


2003
extraction age

Ar-Ar

4.345

0.011

Bogard and Garrison


2003
extraction age

Ar-Ar

4.396

0.012

Bogard and Garrison


2003
extraction age

Ar-Ar

4.426

0.012

Bogard and Garrison


2003
extraction age

Ar-Ar

4.473

0.012

Bogard and Garrison


2003
extraction age

Ar-Ar

4.484

0.012

Bogard and Garrison


2003
extraction age

Ar-Ar

4.483

0.011

Bogard and Garrison


2003
extraction age

Ar-Ar

4.447

0.009

Bogard and Garrison


2003
extraction age

Ar-Ar

4.363

0.011

Bogard and Garrison


2003
extraction age

Ar-Ar

4.181

0.012

Bogard and Garrison


2003
extraction age

Ar-Ar

4.217

0.018

Bogard and Garrison


2003
extraction age

Ar-Ar

4.133

0.042

Bogard and Garrison


2003
extraction age

Ar-Ar

4.294

0.107

Bogard and Garrison


2003
extraction age

Ar-Ar

4.209

0.188

Bogard and Garrison


2003
extraction age

K-Ar
whole-rock
samples

K-Ar

3.5

Heymann,
Mazor,
and Anders 1968
model age

4.46

Shukolyukov
and
Begemann 1996b
model age

45100%
extractions

Ar-Ar

4.25

0.03

Bogard and Garrison


2003
plateau age

350C

Ar-Ar

6.742

0.02

Bogard and Garrison


2003
step extraction age

425C

Ar-Ar

4.542

0.3

Bogard and Garrison


2003
step extraction age

500C

Ar-Ar

2.102

0.315

Bogard and Garrison


2003
step extraction age

550C

Ar-Ar

1.827

0.348

Bogard and Garrison


2003
step extraction age

625C

Ar-Ar

3.21

0.074

Bogard and Garrison


2003
step extraction age

700C

Ar-Ar

3.029

0.049

Bogard and Garrison


2003
step extraction age

750C

Ar-Ar

4.083

0.03

Bogard and Garrison


2003
step extraction age

800C

Ar-Ar

4.196

0.016

Bogard and Garrison


2003
step extraction age

830C

Ar-Ar

4.23

0.015

Bogard and Garrison


2003
step extraction age

852C

Ar-Ar

4.217

0.018

Bogard and Garrison


2003
step extraction age

875C

Ar-Ar

4.223

0.014

Bogard and Garrison


2003
step extraction age

895C

Ar-Ar

4.197

0.015

Bogard and Garrison


2003
step extraction age

920C

Ar-Ar

4.215

0.014

Bogard and Garrison


2003
step extraction age

950C

Ar-Ar

4.204

0.014

Bogard and Garrison


2003
step extraction age

975C

Ar-Ar

4.231

0.008

Bogard and Garrison


2003
step extraction age

1000C

Ar-Ar

4.25

0.008

Bogard and Garrison


2003
step extraction age

1025C

Ar-Ar

4.245

0.007

Bogard and Garrison


2003
step extraction age

1050C

Ar-Ar

4.218

0.008

Bogard and Garrison


2003
step extraction age

1100C

Ar-Ar

4.22

0.009

Bogard and Garrison


2003
step extraction age

1150C

Ar-Ar

4.189

0.01

Bogard and Garrison


2003
step extraction age

1155C

Ar-Ar

4.157

0.04

Bogard and Garrison


2003
step extraction age

1225C

Ar-Ar

4.117

0.03

Bogard and Garrison


2003
step extraction age

1325C

Ar-Ar

4.229

0.017

Bogard and Garrison


2003
step extraction age

1425C

Ar-Ar

4.267

0.008

Bogard and Garrison


2003
step extraction age

1500C

Ar-Ar

4.255

0.018

Bogard and Garrison


2003
step extraction age

1600C

0.61

Bogard and Garrison


2003
step extraction age

Ar-Ar

4.272

Pu-Xe

4.548

Shukolyukov
and
Begemann 1996a
model age

K-Ar

3.5

Heymann,
Mazor,
and Anders 1968
model age

Pasamonte
whole
samples

rock

high
temperature,
based on last
fraction
extracted
K-Ar

4.51

0.065

Podosek and Huneke


1973
model age

545C
extraction

K-Ar

3.595

0.062

Podosek and Huneke


1973
model age

630C
extraction

K-Ar

3.902

0.02

Podosek and Huneke


1973
model age

745C
extraction

K-Ar

4.066

0.005

Podosek and Huneke


1973
model age

850C
extraction

K-Ar

4.069

0.004

Podosek and Huneke


1973
model age

935C
extraction

K-Ar

4.132

0.006

Podosek and Huneke


1973
model age

1040C
extraction

K-Ar

4.119

0.006

Podosek and Huneke


1973
model age

1195C
extraction

K-Ar

4.364

0.009

Podosek and Huneke


1973
model age

1515C
extraction

K-Ar

4.511

0.015

Podosek and Huneke


1973
model age

bulk
plateau
age, based on
the first 80% Ar
extracted
Ar-Ar

4.15

0.3

Podosek and Huneke


1973
plateau age

low
temperature

Ar-Ar

0.05

Kunz et al. 1995

spectra model age

high
temperature,
75% gas loss

Ar-Ar

4.4

0.1

Kunz et al. 1995

spectra model age

sample 1, low
temperature
Ar-Ar

0.1

Kunz et al. 1995

spectrum
age

model

3.9

sample 1, high
temperature,
90% gas loss
Ar-Ar

0.1

Kunz et al. 1995

spectrum
age

model

4.4

sample 5, low
temperature
Ar-Ar

0.05

Kunz et al. 1995

spectrum
age

model

4.1

sample 5, high
temperature,
70% gas loss
Ar-Ar

0.1

Kunz et al. 1995

spectrum
age

model

4.4

300C

Ar-Ar

7.79

2.455

Kunz et al. 1995

extraction age

400C

Ar-Ar

6.836

0.202

Kunz et al. 1995

extraction age

470C

Ar-Ar

4.447

0.05

Kunz et al. 1995

extraction age

540C

Ar-Ar

4.065

0.034

Kunz et al. 1995

extraction age

600C

Ar-Ar

4.163

0.02

Kunz et al. 1995

extraction age

630C

Ar-Ar

4.241

0.022

Kunz et al. 1995

extraction age

670C

Ar-Ar

3.803

0.015

Kunz et al. 1995

extraction age

700C

Ar-Ar

3.898

0.013

Kunz et al. 1995

extraction age

730C

Ar-Ar

3.875

0.013

Kunz et al. 1995

extraction age

760C

Ar-Ar

3.921

0.014

Kunz et al. 1995

extraction age

790C

Ar-Ar

3.921

0.01

Kunz et al. 1995

extraction age

810C

Ar-Ar

3.938

0.009

Kunz et al. 1995

extraction age

830C

Ar-Ar

3.996

0.01

Kunz et al. 1995

extraction age

850C

Ar-Ar

3.989

0.014

Kunz et al. 1995

extraction age

880C

Ar-Ar

4.02

0.016

Kunz et al. 1995

extraction age

910C

Ar-Ar

4.024

0.011

Kunz et al. 1995

extraction age

940C

Ar-Ar

4.041

0.013

Kunz et al. 1995

extraction age

980C

Ar-Ar

4.08

0.016

Kunz et al. 1995

extraction age

1020C

Ar-Ar

4.162

0.026

Kunz et al. 1995

extraction age

1070C

Ar-Ar

4.184

0.019

Kunz et al. 1995

extraction age

1120C

Ar-Ar

4.21

0.027

Kunz et al. 1995

extraction age

1160C

Ar-Ar

4.314

0.048

Kunz et al. 1995

extraction age

1200C

Ar-Ar

4.325

0.041

Kunz et al. 1995

extraction age

1240C

Ar-Ar

4.438

0.069

Kunz et al. 1995

extraction age

1340C

Ar-Ar

4.348

0.029

Kunz et al. 1995

extraction age

1440C

Ar-Ar

4.397

0.015

Kunz et al. 1995

extraction age

1580C

Ar-Ar

4.443

0.072

Kunz et al. 1995

extraction age

300C

Ar-Ar

6.458

0.655

Kunz et al. 1995

extraction age

350C

Ar-Ar

5.442

0.189

Kunz et al. 1995

extraction age

400C

Ar-Ar

4.402

0.073

Kunz et al. 1995

extraction age

450C

Ar-Ar

4.189

0.032

Kunz et al. 1995

extraction age

500C

Ar-Ar

4.331

0.019

Kunz et al. 1995

extraction age

550C

Ar-Ar

4.487

0.013

Kunz et al. 1995

extraction age

600C

Ar-Ar

4.209

0.005

Kunz et al. 1995

extraction age

640C

Ar-Ar

4.1

0.004

Kunz et al. 1995

extraction age

680C

Ar-Ar

4.1

0.004

Kunz et al. 1995

extraction age

700C

Ar-Ar

4.105

0.005

Kunz et al. 1995

extraction age

720C

Ar-Ar

4.13

0.004

Kunz et al. 1995

extraction age

740C

Ar-Ar

4.14

0.005

Kunz et al. 1995

extraction age

760C

Ar-Ar

4.148

0.004

Kunz et al. 1995

extraction age

780C

Ar-Ar

4.176

0.006

Kunz et al. 1995

extraction age

800C

Ar-Ar

4.195

0.004

Kunz et al. 1995

extraction age

820C

Ar-Ar

4.203

0.006

Kunz et al. 1995

extraction age

840C

Ar-Ar

4.218

0.006

Kunz et al. 1995

extraction age

860C

Ar-Ar

4.24

0.005

Kunz et al. 1995

extraction age

890C

Ar-Ar

4.241

0.006

Kunz et al. 1995

extraction age

920C

Ar-Ar

4.217

0.006

Kunz et al. 1995

extraction age

950C

Ar-Ar

4.223

0.008

Kunz et al. 1995

extraction age

990C

Ar-Ar

4.305

0.009

Kunz et al. 1995

extraction age

1030C

Ar-Ar

4.357

0.011

Kunz et al. 1995

extraction age

1080C

Ar-Ar

4.381

0.014

Kunz et al. 1995

extraction age

1150C

Ar-Ar

4.436

0.006

Kunz et al. 1995

extraction age

1230C

Ar-Ar

4.483

0.007

Kunz et al. 1995

extraction age

1270C

Ar-Ar

4.567

0.066

Kunz et al. 1995

extraction age

1310C

Ar-Ar

5.407

0.341

Kunz et al. 1995

extraction age

1380C

Ar-Ar

4.666

0.222

Kunz et al. 1995

extraction age

300C

Ar-Ar

6.191

0.948

Kunz et al. 1995

extraction age

450C

Ar-Ar

6.099

0.169

Kunz et al. 1995

extraction age

530C

Ar-Ar

4.637

0.121

Kunz et al. 1995

extraction age

600C

Ar-Ar

4.477

0.077

Kunz et al. 1995

extraction age

670C

Ar-Ar

4.698

0.031

Kunz et al. 1995

extraction age

730C

Ar-Ar

4.007

0.027

Kunz et al. 1995

extraction age

790C

Ar-Ar

3.997

0.022

Kunz et al. 1995

extraction age

830C

Ar-Ar

4.076

0.021

Kunz et al. 1995

extraction age

880C

Ar-Ar

4.23

0.021

Kunz et al. 1995

extraction age

940C

Ar-Ar

4.267

0.022

Kunz et al. 1995

extraction age

1020C

Ar-Ar

4.221

0.016

Kunz et al. 1995

extraction age

1120C

Ar-Ar

4.255

0.023

Kunz et al. 1995

extraction age

1200C

Ar-Ar

4.287

0.083

Kunz et al. 1995

extraction age

1320C

Ar-Ar

4.437

0.033

Kunz et al. 1995

extraction age

1440C

Ar-Ar

4.739

0.385

Kunz et al. 1995

extraction age

Ar-Ar

4.245

0.024

Kunz et al. 1995

plateau age

sample 7, low
temperature,
48% width
Ar-Ar

0.1

Kunz et al. 1995

spectrum
age

model

sample 7, high
temperature,
35% gas loss
Ar-Ar

0.1

Kunz et al. 1995

spectrum
age

model

4.4

Revised
Podosek
Huneke
age

4.02

sample
plateau

and
1973
Ar-Ar
207
206

PbPb

207
206
207
206
207
206
207
206
207
206
207
206
207
206

206
207

and
plateau age

4.573

0.006

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.572

0.005

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.484

0.004

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.454

0.01

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.963

0.03

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.912

0.01

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.709

0.008

Unruh,
Nakamura,
and Tatsumoto 1977 model age

0.006

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.515

0.06

Unruh,
Nakamura,
and Tatsumoto 1977 model age

Pb- 4.454

0.019

Unruh,

PbPb

207

Quitt, Birck,
Allgre 2000

4.664

Nakamura,model age

206

Pb

207

PbPb 4.457

0.012

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.51

0.1

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.468

0.015

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.44

0.023

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.535

0.005

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.534

0.004

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.542

0.019

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.534

0.02

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.534

0.013

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.519

0.013

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.555

0.017

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbPb 4.52

0.004

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.874

0.038

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.814

0.038

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
5.694

0.083

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
5.597

0.092

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
3.38

0.35

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
5.336

0.7

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
0.296

0.005

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
0.277

0.003

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
0.103

0.009

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
0.377

0.014

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
0.376

0.016

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
2.953

0.3

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
3.29

0.085

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
3.215

0.107

Unruh,
Nakamura,
and Tatsumoto 1977 model age

206
207
206
207
206
207
206
207
206
207
206
207
206
207
206
207
206
207
206
207
206
207
206
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238
206
238

Pb- 4.903
U

and Tatsumoto 1977

0.041

Unruh,
Nakamura,model age
and Tatsumoto 1977

206
238
206
238

PbU
4.907

0.041

Unruh,
Nakamura,
and Tatsumoto 1977 model age

0.14

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.5

0.15

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
1.851

0.031

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
1.477

0.03

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
1.798

0.043

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
3.39

0.027

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.663

0.015

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.645

0.015

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.806

0.031

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.784

0.044

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.421

0.175

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
5.032

0.23

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
1.724

0.015

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
1.644

0.011

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
0.865

0.085

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
1.786

0.053

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
1.784

0.072

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
3.955

0.195

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.061

0.06

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.014

0.08

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.645

0.014

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.646

0.016

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.536

0.06

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
4.523

0.068

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
3.419

0.027

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU
3.161

0.036

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbU

206
238
206
238
206
238
206
238
206
238
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235
207
235

4.522

Pb- 3.396
U

0.036

Unruh,
Nakamura,model age
and Tatsumoto 1977

207
235
208
232
208
232
208
232
208
232

0.015

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 4.678

0.085

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 14.8

0.6

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 9.6

PbTh

208
232
208
232
208
232
208
232
208
232
208
232
208
232

two samples

PbU
4.136

Unruh,
Nakamura,
and Tatsumoto 1977 model age
0.03

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 0.57

0.2

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 4.922

0.5

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 4.16

0.25

Unruh,
Nakamura,
and Tatsumoto 1977 model age

0.295

PbTh 10.7

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 3.597

0.099

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 3.549

0.16

Unruh,
Nakamura,
and Tatsumoto 1977 model age

PbTh 8.627

0.35

Unruh,
Nakamura,
and Tatsumoto 1977 model age

0.019

Shukolyukov
and
Begemann 1996a
model age

Pu-Xe

4.576

K-Ar

2.7

Serra De Mag
whole
samples

rock

Heymann,
Mazor,
and Anders 1968
model age

average of four
extractions
having
the
same
age
within
uncertainties
Ar-Ar

3.386

0.007

Bogard and Garrison


2003
plateau age

350C

Ar-Ar

6.1

0.114

Bogard and Garrisonstep


extraction
2003
model age

450C

Ar-Ar

2.548

0.191

Bogard and Garrisonstep


extraction
2003
model age

525C

Ar-Ar

1.231

0.25

Bogard and Garrisonstep


extraction
2003
model age

600C

Ar-Ar

1.77

0.122

Bogard and Garrisonstep


extraction
2003
model age

650C

Ar-Ar

0.499

0.199

Bogard and Garrisonstep


extraction
2003
model age

700C

Ar-Ar

2.779

0.18

Bogard and Garrisonstep


extraction
2003
model age

775C

Ar-Ar

3.112

0.093

Bogard and Garrisonstep


extraction
2003
model age

825C

Ar-Ar

3.316

0.033

Bogard and Garrisonstep


extraction
2003
model age

875C

Ar-Ar

3.387

0.037

Bogard and Garrisonstep


extraction
2003
model age

925C

Ar-Ar

3.372

0.045

Bogard and Garrisonstep


extraction
2003
model age

975C

Ar-Ar

3.388

0.06

Bogard and Garrisonstep


extraction
2003
model age

1025C

Ar-Ar

3.395

0.041

Bogard and Garrisonstep


extraction
2003
model age

1075C

Ar-Ar

3.461

0.045

Bogard and Garrisonstep


extraction
2003
model age

1125C

Ar-Ar

3.511

0.037

Bogard and Garrisonstep


extraction
2003
model age

1175C

Ar-Ar

3.619

0.048

Bogard and Garrisonstep


extraction
2003
model age

1225C

Ar-Ar

3.528

0.103

Bogard and Garrisonstep


extraction
2003
model age

1275C

Ar-Ar

3.585

0.158

Bogard and Garrisonstep


extraction
2003
model age

1325C

Ar-Ar

3.689

0.193

Bogard and Garrisonstep


extraction
2003
model age

1400C

Ar-Ar

3.783

0.053

Bogard and Garrisonstep


extraction
2003
model age

1500C

Ar-Ar

3.91

0.028

Bogard and Garrisonstep


extraction
2003
model age

whole-rock
samples

K-Ar

3.8

545C

K-Ar

3.86

0.028

Podosek and Huneke


1973
step model age

630C

K-Ar

3.837

0.008

Podosek and Huneke


1973
step model age

745C

K-Ar

3.704

0.003

Podosek and Huneke


1973
step model age

850C

K-Ar

3.472

0.002

Podosek and Huneke


1973
step model age

935C

K-Ar

3.514

0.007

Podosek and Huneke


1973
step model age

1040C

K-Ar

3.64

0.003

Podosek and Huneke


1973
step model age

1195C

K-Ar

3.591

0.003

Podosek and Huneke


1973
step model age

K-Ar

3.923

0.006

Podosek and Huneke


1973
step model age

Stannern
Heymann,
Mazor,
and Anders 1968
model age

3.57

Ar-Ar

3.7

0.2

three samples,
low
temperature
Ar-Ar

3.3

0.3

Kunz et al. 1995

plateau age

three samples,
high
temperature,
80% gas loss
Ar-Ar

0.2

Kunz et al. 1995

plateau age

drill core 1,
basalt
clast,
300C
Ar-Ar

6.313

0.181

Kunz et al. 1995

step extraction age

350C

Ar-Ar

3.632

0.028

Kunz et al. 1995

step extraction age

400C

Ar-Ar

3.09

0.014

Kunz et al. 1995

step extraction age

450C

Ar-Ar

3.564

0.008

Kunz et al. 1995

step extraction age

500C

Ar-Ar

3.687

0.005

Kunz et al. 1995

step extraction age

1515C

mean of ages in literature

Shukolyukov
and
Begemann 1996a
model age

K-Ar

based on nine extractions atPodosek and Huneke


stepped temperatures
1973
plateau age

550C

Ar-Ar

3.679

0.005

Kunz et al. 1995

step extraction age

580C

Ar-Ar

3.636

0.004

Kunz et al. 1995

step extraction age

600C

Ar-Ar

3.523

0.005

Kunz et al. 1995

step extraction age

620C

Ar-Ar

3.48

0.007

Kunz et al. 1995

step extraction age

640C

Ar-Ar

3.392

0.005

Kunz et al. 1995

step extraction age

670C

Ar-Ar

3.408

0.005

Kunz et al. 1995

step extraction age

690C

Ar-Ar

3.471

0.006

Kunz et al. 1995

step extraction age

710C

Ar-Ar

3.511

0.005

Kunz et al. 1995

step extraction age

740C

Ar-Ar

3.543

0.005

Kunz et al. 1995

step extraction age

770C

Ar-Ar

3.58

0.006

Kunz et al. 1995

step extraction age

810C

Ar-Ar

3.597

0.006

Kunz et al. 1995

step extraction age

850C

Ar-Ar

3.595

0.006

Kunz et al. 1995

step extraction age

890C

Ar-Ar

3.57

0.006

Kunz et al. 1995

step extraction age

930C

Ar-Ar

3.537

0.006

Kunz et al. 1995

step extraction age

970C

Ar-Ar

3.478

0.007

Kunz et al. 1995

step extraction age

1020C

Ar-Ar

3.298

0.007

Kunz et al. 1995

step extraction age

1070C

Ar-Ar

3.474

0.012

Kunz et al. 1995

step extraction age

1120C

Ar-Ar

3.785

0.017

Kunz et al. 1995

step extraction age

1170C

Ar-Ar

4.003

0.013

Kunz et al. 1995

step extraction age

1220C

Ar-Ar

4.141

0.024

Kunz et al. 1995

step extraction age

1270C

Ar-Ar

4.229

0.026

Kunz et al. 1995

step extraction age

1330C

Ar-Ar

4.338

0.069

Kunz et al. 1995

step extraction age

1400C

Ar-Ar

4.247

0.049

Kunz et al. 1995

step extraction age

drill core 2,
recrystallized
classic matrix
300C
Ar-Ar

4.76

0.601

Kunz et al. 1995

step extraction age

350C

Ar-Ar

3.672

0.069

Kunz et al. 1995

step extraction age

400C

Ar-Ar

3.238

0.022

Kunz et al. 1995

step extraction age

450C

Ar-Ar

3.588

0.011

Kunz et al. 1995

step extraction age

500C

Ar-Ar

3.747

0.009

Kunz et al. 1995

step extraction age

540C

Ar-Ar

3.715

0.008

Kunz et al. 1995

step extraction age

580C

Ar-Ar

3.88

0.006

Kunz et al. 1995

step extraction age

600C

Ar-Ar

3.596

0.01

Kunz et al. 1995

step extraction age

620C

Ar-Ar

3.5

0.005

Kunz et al. 1995

step extraction age

640C

Ar-Ar

3.405

0.007

Kunz et al. 1995

step extraction age

670C

Ar-Ar

3.402

0.007

Kunz et al. 1995

step extraction age

690C

Ar-Ar

3.448

0.007

Kunz et al. 1995

step extraction age

710C

Ar-Ar

3.5

0.008

Kunz et al. 1995

step extraction age

740C

Ar-Ar

3.543

0.009

Kunz et al. 1995

step extraction age

770C

Ar-Ar

3.558

0.01

Kunz et al. 1995

step extraction age

810C

Ar-Ar

3.545

0.01

Kunz et al. 1995

step extraction age

850C

Ar-Ar

3.487

0.011

Kunz et al. 1995

step extraction age

890C

Ar-Ar

3.432

0.013

Kunz et al. 1995

step extraction age

930C

Ar-Ar

3.382

0.012

Kunz et al. 1995

step extraction age

970C

Ar-Ar

3.282

0.014

Kunz et al. 1995

step extraction age

1020C

Ar-Ar

3.095

0.014

Kunz et al. 1995

step extraction age

1070C

Ar-Ar

3.146

0.027

Kunz et al. 1995

step extraction age

1120C

Ar-Ar

3.492

0.05

Kunz et al. 1995

step extraction age

1170C

Ar-Ar

3.735

0.065

Kunz et al. 1995

step extraction age

1220C

Ar-Ar

4.001

0.114

Kunz et al. 1995

step extraction age

1270C

Ar-Ar

3.947

0.077

Kunz et al. 1995

step extraction age

1330C

Ar-Ar

4.122

0.126

Kunz et al. 1995

step extraction age

Ar-Ar

3.84

0.327

Kunz et al. 1995

step extraction age

PbPb

4.329

Manhs et al. 1975

model age

PbPb

4.13

Tera, Carlson,
Boctor 1987

PuXe

4.46

0.028

Shukolyukov
and
Begemann 1996a
model age

PuXe

4.434

0.013

Miura et al. 1998

model age

0.5

Miura et al. 1993

model age

84 matrix, 11
extractions
Ar-Ar

3.94

0.04

Takeda, Mori,
Bogard 1994

and
plateau age

84 clast,
extractions

Ar-Ar

3.98

0.03

Takeda, Mori,
Bogard 1994

and
plateau age

84
matrix,
temp. 300
Ar-Ar

3.76

0.05

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 400
Ar-Ar

3.95

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 500
Ar-Ar

4.01

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 600
Ar-Ar

3.96

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 700
Ar-Ar

3.97

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 775
Ar-Ar

3.95

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 850
Ar-Ar

3.88

0.04

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 950
Ar-Ar

3.91

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 1000
Ar-Ar

3.94

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 1050
Ar-Ar

3.97

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 1100
Ar-Ar

3.98

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 1150
Ar-Ar

3.98

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 1225
Ar-Ar

4.12

0.03

Takeda, Mori,
Bogard 1994

and
degassing age

84
matrix,
temp. 1350
Ar-Ar

4.35

0.03

Takeda, Mori,
Bogard 1994

and
degassing age

84

4.2

0.16

Takeda,

anddegassing age

1400C

and
model age

Yamato 75011
whole
sample

rock
K-Ar

11

matrix, Ar-Ar

Mori,

temp. 1500

Bogard 1994

84 clast, temp.
300
Ar-Ar

3.99

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
400
Ar-Ar

3.69

0.03

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
500
Ar-Ar

4.05

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
575
Ar-Ar

3.83

0.06

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
650
Ar-Ar

3.99

0.03

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
725
Ar-Ar

4.03

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
800
Ar-Ar

3.97

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
850
Ar-Ar

3.98

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
900
Ar-Ar

3.96

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
950
Ar-Ar

3.96

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
1000
Ar-Ar

3.99

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
1050
Ar-Ar

4.08

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
1100
Ar-Ar

4.15

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
1150
Ar-Ar

4.23

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
1225
Ar-Ar

4.36

0.03

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
1300
Ar-Ar

4.48

0.03

Takeda, Mori,
Bogard 1994

and
degassing age

84 clast, temp.
1500
Ar-Ar

4.5

0.02

Takeda, Mori,
Bogard 1994

and
degassing age

matrix,
low
temperature
Ar-Ar

0.04

Kunz et al. 1995

spectrum
age

model

3.91

matrix,
high
temperature,
90%
gas
extraction
Ar-Ar

0.1

Kunz et al. 1995

spectrum
age

model

4.4

clast,
low
temperature
Ar-Ar

0.03

Kunz et al. 1995

spectrum
age

model

3.98

clast,
high
temperature,
90%
gas
extraction
Ar-Ar

0.1

Kunz et al. 1995

spectrum
age

model

4.5

73
whole
BABI

Rb-Sr

4.28

0.24

Nyquist et al. 1986

model age

73
matrix,
<2.45
g/cu.
cm., BABI
Rb-Sr

4.53

0.07

Nyquist et al. 1986

model age

73
matrix,
2.452.85 g/cu.
cm., BABI
Rb-Sr

3.91

0.4

Nyquist et al. 1986

model age

73

4.71

0.19

Nyquist et al. 1986

model age

matrix,
rock,

matrix, Rb-Sr

2.853.3 g/cu.
cm., BABI
73 matrix, 3.3
g/cu. cm., BABI Rb-Sr

4.73

0.31

Nyquist et al. 1996

model age

84B
clast,
whole rock 1,
BABI
Rb-Sr

4.34

0.89

Nyquist et al. 1986

model age

84B
clast,
whole rock 3,
BABI
Rb-Sr

4.36

0.29

Nyquist et al. 1986

model age

84B
clast,
<2.45
g/cu.
cm., BABI
Rb-Sr

4.57

0.05

Nyquist et al. 1986

model age

84B
clast,
2.452.65 g/cu.
cm., BABI
Rb-Sr

4.52

0.15

Nyquist et al. 1986

model age

84B
clast,
2.652.85 g/cu.
cm., BABI
Rb-Sr

3.55

0.51

Nyquist et al. 1986

model age

84B clast, Plag,


BABI
Rb-Sr

3.73

0.54

Nyquist et al. 1986

model age

84B clast, Lt.


Px, BABI
Rb-Sr

4.11

0.11

Nyquist et al. 1986

model age

84B clast, Dk.


Px, BABI
Rb-Sr

4.51

0.14

Nyquist et al. 1986

model age

84B clast, most


Mag, BABI
Rb-Sr

4.66

0.12

Nyquist et al. 1986

model age

PbPb

4.551

0.008

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbPb

4.561

0.02

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbPb

4.552

0.01

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbPb

4.549

0.009

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbPb

4.543

0.013

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbU

5.04

0.078

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbU

4.816

0.079

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbU

4.548

0.047

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbU

4.842

0.073

Misawa, Yamaguchi,
and Kaiden 2005
model age

PbU

5.07

0.064

Misawa, Yamaguchi,
and Kaiden 2005
model age

73
matrix,
whole
rock,
chondrite
Sm-Nd 4.55

0.03

Nyquist et al. 1986

model age

73
matrix,
2.452.85 g/cu.
cm., chondrite
Sm-Nd 4.52

0.03

Nyquist et al. 1986

model age

73
matrix,
2.853.3 g/cu.
cm., chondrite
Sm-Nd 4.53

0.02

Nyquist et al. 1986

model age

0.02

Nyquist et al. 1996

model age

207

zircon 011-1a

206
207

zircon 011-1b

206
207

zircon 011-2a

206
207

zircon 011-2b

206
207

zircon 011-2c

206
206

zircon 011-1a

238
206

zircon 011-1b

238
206

zircon 011-2a

238
206

zircon 011-2b

238
206

zircon 011-2c

238

73 matrix, 3.3 Sm-Nd 4.51


g/cu.
cm.,

chondrite
84B
clast,
whole rock 1,
chondrite
Sm-Nd 4.55

0.02

Nyquist et al. 1986

model age

84B
clast,
whole rock 3,
chondrite
Sm-Nd 4.54

0.02

Nyquist et al. 1986

model age

84B
clast,
<2.45
g/cu.
cm., chondrite
Sm-Nd 4.51

0.04

Nyquist et al. 1986

model age

84B clast, Plag,


chondrite
Sm-Nd 4.56

0.04

Nyquist et al. 1986

model age

84B clast, Lt.


Px, chondrite
Sm-Nd 4.56

0.02

Nyquist et al. 1986

model age

84B clast, Dk.


Px, chondrite
Sm-Nd 4.53

0.02

Nyquist et al. 1986

model age

84B clast, most


Mag, chondrite Sm-Nd 4.58

0.02

Nyquist et al. 1986

model age

Fig. 6. Frequency versus radioisotope ages histogram diagram for the isochron ages for whole-rock samples and some or
all components of 12 eucrite achondrites, with color coding being used to show the ages obtained by the different
radioisotope dating methods.
An explanation for the agreement of ages
It has already been noted above that the Mn-Cr, Hf-W, Al-Mg, I-Xe, and Pu-Xe methods are calibrated directly or via
intermediaries against the Pb-Pb ages yielded by other meteorites, which invariably guarantees the ages obtained by these
methods will align with, or are adjusted to, the Pb-Pb ages that usually cluster around 4.554.57 Ga to date these eucrites.
Indeed, as Snelling (2014a, b) found, the ages of most of the 28 meteorites investigated so far have been established
directly by Pb-Pb ages or indirectly by Pb-Pb calibrated ages defining the clustering at 4.54.57 Ga.
Examination of the relevant literature reveals that the 87Rb decay rate has been determined by calibrating Rb-Sr dates
against Pb-Pb dates on the same rock samples (Nebel, Scherer, and Mezger 2011; Snelling 2014c). Similarly, Snelling
(2014d) has found that the 176Lu decay rate has also been determined by calibrating Lu-Hf dates against Pb-Pb dates on the
same mineral samples from the same rocks (Sderlund et al. 2004). In other words, the Rb-Sr and Lu-Hf dates were
adjusted to agree with the Pb-Pb ages, so it is hardly surprising when the adjusted decay constants are used to calculate
the Rb-Sr and Lu-Hf dates yielded by other rocks and minerals, and here by these meteorites, they often agree with the
meteorites 4.554.57 Ga Pb-Pb ages. It is also well known that the 40K decay constant has been calibrated against standard
samples of known Pb-Pb ages (Renne et al. 2010), so ultimately all K-Ar and Ar-Ar dates are adjusted accordingly. Even
the 147Sm decay constant has been calibrated against Pb-Pb ages on the same samples (Dickin 2005, pp. 7071; Lugmair
1974; Lugmair, Scheinin, and Marti 1975).
Therefore, ultimately all other radioisotopic dating methods have been effectively calibrated against Pb-Pb ages on the
same rocks and minerals, and in some cases on the same meteorites. These calibrations are simply done by determining
the Rb-Sr or Lu-Hf ages for specific rocks, meteorites, or minerals whose Pb-Pb ages have also been determined. Then it is
assumed that the Pb-Pb ages are the most accurate and true ages for those rocks, meteorites, or minerals. So if the Rb-Sr
or Lu-Hf ages disagree with the Pb-Pb ages, then the 87Rb and 176Lu half-life values are adjusted so that all the ages agree.
It is thus hardly surprising that often there is agreement between all the radioisotope dates yielded on the same rocks,
minerals and meteorites, except where it is interpreted that other factors have disturbed the radioisotope systems which thus
have yielded a plethora of different dates (see below). This ultimately means that all meteorite ages are dependent on the
reliability of determinations of the 238U and 235U decay constants, and the critical 238U/235U ratio. Yet discrepancies and
variations have now been found between the 238U/235U ratio in U-bearing earth-based (terrestrial) minerals and rocks and
the 238U/235U ratio in meteorites (Brennecka and Wadhwa 2012; Hiess et al. 2012). It also may be significant that the
estimated half-life of 238U is 4.468 Ga, which is almost identical to the claimed 4.554.57 Ga age of the earth, the asteroids,
and the meteorites derived from them.

Fig. 7. Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock samples and some or all
components of 12 eucrite achondrites, with color coding being used to show the ages obtained by the different radioisotope
dating methods.
Explanations for the scattering of ages
The usual explanation for such scattering of isochron and model ages is the thermal disturbance of the isotope systems
subsequent to formation of the parent body from which the meteorites came, affecting either the parent body or
subsequently the meteorites, or both (Bogard 2011; Lewis 1997; Trieloff et al. 2003). This must be especially the case where
in this instance the scattering of ages is rife in the K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, and U-Th-Pb systems. Confirmation of this
explanation would seem to be provided by the fact that these eucrites (basaltic achondrites) are invariably metamorphosed,
because it is well-documented that metamorphism resets these radioisotope systems (Faure and Mensing 2005; Snelling
2000).However, not all these eucrites are metamorphosed, at least not to the same extent. Pasamonte (fig.5a) and Serra de
Mag (fig. 5c) are regarded as least metamorphosed or unmetamorphosed, though in the latter the original igneous
pyroxenes have exsolved augite lamellae as a result of some metamorphism. Yet in both meteorites their radioisotope
systems have been greatly disturbed. For Pasamonte the Rb-Sr isochron ages are younger than the 4.554.57 Ga
clustering and the K-Ar, Ar-Ar, and U-Th-Pb model ages are widely scattered and are both younger and older. Similarly for
Serra de Mag the Sm-Nd and Pb-Pb isochron ages are younger than the 4.554.57 Ga clustering, whereas the Ar-Ar
model ages are widely scattered and invariably grossly younger.McSween et al. (2011), based on the work of Bogard and
Garrison (1995, 2003, 2009) and Bogard (2011), suggest another explanation for the lower Ar-Ar model ages for many
eucrites, such as for Cacheri, Caldera, Camel Donga, Ibitira, Juvinas, Moama, Moore County, Pasamonte, Serra de Mag,
Stannern, and Yamato75011 in this study. Brecciated eucrites evidence significant Ar degassing, which reset their Ar-Ar
ages to between 3.14.1 Ga. Since the timing of thermal and shock metamorphism is supposedly best determined from ArAr ages, Bogard and Garrison (2009) argued that these young Ar-Ar ages represent impact cratering events on the surface
of the parent asteroid 4-Vesta. They claimed that unbrecciated and unshocked eucrites have Ar-Ar ages near 4.5 Ga,
whereas brecciated eucrites yield younger Ar-Ar ages that cluster on a probability plot in clusters at ~3.5, 3.8, 3.9, and 4.0
Ga (fig. 8). These they concluded were impact events that brecciated the eucrites and reset their Ar-Ar ages. However, in
this study Ibitira, Moore County, Pasamonte, and Serra de Mag are all unbrecciated, as evidenced in thin sections (fig. 5),
yet they yield many Ar-Ar ages younger than 4.5 Ga, some around 4.5 Ga, and quite a few well above 4.5 Ga.
Fig. 8 Probability plot of 40Ar-39Ar ages of unbrecciated and
brecciated eucrites, based on the work of Bogard and Garrison
(1995, 2003, 2009), after Bogard (2011) and McSween et al.
(2011). The peaks are interpreted as representing major impact
events on their parent asteroid 4-Vesta.Nevertheless, thermal
disturbances tend to reset the radioisotope systems so that the
resultant ages are lower than their original ages (Lewis 1997),
whereas for all these eucrites the radioisotope ages are both
younger and older than the believed formation age of the
parent body, determined as 4.554.57 Ga from the clustering
of all the radioisotope systems at that age. So perhaps there
might be an additional explanation for the scattering of both the
isochron and model ages for all these eucrites (basaltic
achondrites).
Internal layering in parent 4-Vesta
NASAs Dawn spacecraft mission to 4-Vesta investigated the
asteroids surface during 20112012 with a Visible and Infrared
(VIR) Spectrometer for mineral identification and mapping, and
a Gamma Ray and Neutron Detector (GRaND) for
geochemical analysis and mapping (McSween et al. 2014). As
a consequence new evidence was provided to strengthen considerably the connection of the HED meteorites to 4-Vesta
(McSween et al. 2013, 2014). For example, high-resolution visible-near infrared spectra indicated pyroxene compositions
and abundances like those of the HED meteorites, the mapped petrologic complexity (unlike the homogeneity of smaller
asteroids) was consistent with the range of rock compositions among the HED meteorites, and geochemical measurements
of the regolith were similar to the compositions of HED meteorites (and unlike those of other achondrite types). These
results have further strengthened uniformitarians attempts to explain the natural formation of 4-Vesta in the context of all
other bodies in the solar system, and how it was shaped or modified by subsequent processes to produce its current internal
structure and its surface features.Regarded as a leftover planetary building block, with an average diameter of 510 km (317
mi) and a mean density of 3456 kg/m3 (216 lbs/ft3), 4-Vesta was once called the smallest terrestrial planet (McSween et al.
2014). It is believed by uniformitarians to be an intact example of a large, differentiated protoplanet, like those claimed to be
the building blocks of terrestrial planet accretion from the solar nebula, and based on analyses of the HED meteorites is
postulated to have a basaltic upper crust (as represented by the noncumulate basaltic eucrites) that overlies a cumulate
eucrite mid-crust layer and then a diogenite lower crust, with a depleted peridotitic (harzburgite) mantle beneath and then an
iron core (Mandler and Elkins-Tanton 2013; Zuber et al. 2011). Petrological analyses of eucrites coupled with
thermalevolution modeling has led to the proposal of two possible mechanisms of silicate-metal differentiation leading to the
formation of the basaltic achondrites of Vestas crustthe earlier model of equilibrium partial melting (Stolper 1977), or the
later now popular model of crystallization of residual liquids (eucrites) and cumulates (diogenites) from a postulated cooling
magma ocean (Mandler and Elkins-Tanton 2013; Righter and Drake 1997). For a body the size of 4-Vesta it is postulated
that subsequent eruptions directly to the surface via dikes would be mechanically very difficult, so instead, large magma
chambers would likely have formed in the subsurface and flows of basaltic eucrites could have erupted episodically from
these chambers. Geochemical trends in basaltic eucrites (the main group trend and the Stannern trend) would seem to
require multiple magmas and complex processes within this asteroids crust (Barrat et al. 2007).Thus a complex history for
4-Vesta has been elucidated by uniformitarians using the range of ages obtained for the HED meteorites yielded by the
various radioisotope dating techniques. They even postulate that the differentiation of the crust and mantle occurred close to
4.56 Ga within 13 Myr of the formation of the solar system, as defined by the Pb-Pb isochron age of the CAIs in chondrites,
along with core formation by silicate-metal fractionation at the same time or soon thereafter (McSween et al. 2011, 2014).
The earliest magmatic activity supposedly occurred at ~5 Myr after formation of the solar system, and then persisted
perhaps for as much as ~50150 Myr. Then in common with other bodies in the solar system, in the period from 4.1 Ga to
3.5 Ga often called the late heavy bombardment 4-Vesta was evidently struck by careening, massive bolides, as deduced
from the reset Ar-Ar ages (fig. 8).

A creation perspective
Asteroid 4-Vesta was a completely formed entity, with an iron core, ultramafic mantle, and basaltic crust, all the necessary
silicate-metal fractionation and crust-mantle differentiation happening exceedingly rapidly within hours, and thus not
requiring the millions of years postulated by uniformitarians.Given the general consensus that the asteroids consist of
leftover material from the formation of the solar system, Snelling (2014a, b) proposed that the accepted coincident 4.55
4.57 Ga ages for the earth and many meteorites could be due to the earth and the parent asteroids having been created
from the same primordial material.The simplest unifying assumption would therefore be that all such primordial material may
have had the same created isotopic endowment. This assumption seems to be borne out by the earth apparently having the
same time-integrated Pb isotopic endowment and thus being the same Pb-Pb age as the meteorites plotted on the
geochron (Patterson 1956). The earths current Pb isotopic endowment was represented on that geochron by the Pb
isotopic composition of a modern oceanic sediment sample, which would appear to contain the time-integrated Pb isotopic
endowment from the earths beginning which was then processed through the earths subsequent rock cycle (Tyler 1990).
However, while the possibility that the created initial ratios of parent to daughter elements were different for the earth
compared to those created for other solar system objects could be considered, that possibility seems unwarranted all the
solar system objects (planets, moons, and asteroids)werw created from the same primordial material Nevertheless, if there
were created differences in the initial Pb isotopic ratios, then it should not have been possible to plot the meteorites and the
earth on the same Pb-Pb geochron, or meteorites on the same Pb-Pb, U-Pb, Rb-Sr, Sm-Nd, Lu-Hf, and Re-Os isochrons
(Snelling in prep.).It would also seem reasonable to propose that some of all the isotopeswere created at the beginning in
the primordial material, including isotopes that subsequently also formed by radioisotope decay as daughter isotopes from
parent isotopes, regardless of when radioisotope decay started. Even the conventional scientific community has assumed
the initial material of the solar system had the primeval Pb isotopic ratios as measured in the troilite (iron sulfide) in the
Canyon Diablo iron meteorite (Faure and Mensing 2005). Thus 4-Vesta had been formed, all the necessary silicate-metal
fractionation and crust-mantle differentiation of that primordial material to produce its internal layering had happened
exceedingly rapidly within hours. That silicate-metal fractionation and crust-mantle differentiation may have also resulted in
some redistribution or mixing of parent and daughter atoms, the latter having been originally created rather than derived via
radioactive decay. These processes may explain some of the scattering in the radioisotope ages for these eucrite
meteorites, especially those older than the 4.554.57 Ga clustering.At what point in time radioactive decay began is unclear
from Scripture, and is still a matter of debate among creationists. The RATE project considered the possibility of a large
amount of accelerated decay occurring during the Creation Week, as radioisotope decay was not regarded as decay in the
sense of deterioration of matter (Vardiman, Snelling, and Chaffin 2005). It is instead a transmutation process, by which one
element is changed into another. The daughter element is certainly not inferior to the parent element..In the context of the
decay evident today due to the operation of the second law of thermodynamics, Anderson (2103) contended that there is no
real biblical evidence to suggest that the second law was inoperable prior to the curse, and so argued that rather the second
law was in effect from the beginning of creation. He thus also suggested that the tendency toward entropy implicit in the
second law was never of a kind that conflicted with Gods declaration that the creation was very good, or that eventuated in
the death of any sentient creature. On the other hand, it could be argued that radioisotope decay is more than the operation
of the second law of thermodynamics; the additional outcome being the radiation produced which is harmful to lifes
biological and chemical makeup. Indeed, if billions of years of accelerated radioisotope decay had occurred mainly during
the early part of the Creation Week, as considered a possibility by the RATE team (Vardiman, Snelling, and Chaffin 2005),
the enormous burst of radiation would surely have been detrimental to all life on the earth, for example, the plants of Day
Three. It is for this reason that many creationists are not comfortable with postulating that accelerated radioisotope decay
happened during the Creation Week, so maybe there was no radioisotope decay at all until it was started as part of the
curse.The lack of any evidence of a pattern of isochron ages in these 12 basaltic achondrite meteorites (eucrites), as well as
in the 16 chondrite meteorites previously studied (Snelling 2014a, b), that matches the pattern found during the RATE
project (Snelling 2005c; Vardiman, Snelling, and Chaffin 2005) would therefore strongly suggest that all these meteorites
and their parent asteroids have not experienced any episode of accelerated radioisotope decay, either at the time of the
creation of the primordial material on Day One or of their formation on Day Four, or since. This could then be taken to infer
that no accelerated radioisotope decay occurred anywhere in the solar system during the Creation Week, including on the
earth during Days OneThree. Such a conclusion is based on the assumption that the mechanism of small changes to the
binding forces in the nuclei of the parent radioisotopes proposed as the cause of a past episode of accelerated radioisotope
decay (Vardiman, Snelling, and Chaffin 2005) would thus have affected every atom making up the earth, and by logical
extension every atom of the universe at the same time.
Where to from here?
As concluded by Snelling (2014a, b) from his studies of 16 chondrite meteorites, based on the assumptions made the 4.55
4.57 Ga radioisotope ages for the Bereba, Cacheri, Camel Donga, Ibitira, Juvinas, Pasamonte, Serra de Mag, and
Yamato 75011 basaltic achondrite meteorites (eucrites) obtained primarily by Pb-Pb radioisotope isochron and model age
dating of various constituent minerals and fractions (for example, Iizuka et al. 2014; Manhs, Allgre, and Provost 1984;
Misawa, Yamaguchi, and Kaiden 2005; Tera, Carlson, and Boctor 1997; Zhou et al. 2013) are likely not their true real-time
ages. The assumptions on which the radioisotope dating methods are based are simply unprovable, and in the light of the
possibility of an inherited primordial geochemical signature, subsequent resetting of radioisotope clocks due to impact
cratering of asteroids, and the evidence in earth rocks for past accelerated radioisotope decay, mixing of isotopes and
resetting of radioisotope clocks, these assumptions are unreasonable.However, we still need to develop a coherent and
comprehensive explanation of what these radioisotope compositions in both meteorites and earth rocks really represent and
mean within our young-age creation-Flood framework for earth and universe history. We have some possible clues already,
as discussed here, and a clearer picture may yet emerge from continued investigations now in progress. Examination of the
radioisotope dating data for the remaining groups of achondrite meteorites is necessary to complete our understanding of
what the meteorite radioisotope ages might mean. If only the earth was affected by accelerated radioisotope decay during
the Flood, then it is also necessary to examine the radioisotope dating data for lunar rocks to determine whether the moon
was affected by that aspect of the Flood catastrophe or not. Additionally, the radioisotope dating data for many more earth
rocks from all levels of the geologic record need to be collated and examined. If accelerated radioisotope decay only
occurred during the Flood, then it might be expected that the radioisotope ages of pre-Flood (mostly Precambrian) strata
determined by the different methods would be noticeably discordant (Snelling 2005c), whereas the radioisotope ages of
the strata formed during the Flood (mostly Phanerozoic) would be mostly concordant. This difference might be expected due
to the pre-Flood rocks having already been formed and their radioisotope clocks started before the onset of the
accelerated radioisotope decay during the Flood, when their radioisotope clocks would have been speeded up by different
amounts according to the atomic weights of the parent radioisotopes. In contrast, the Flood rocks would have had their
radioisotopes reset when those rocks formed and so their radioisotope clocks started only during the accelerated

radioisotope decay of the Flood event. It may take the collation and examination of the huge radioisotope dating data sets of
as many different earth rocks as possible from all levels of the geologic record to enable any firm conclusions to be
made.Whatever the radioisotope dating data for the earths rocks may reveal, it is already well-established that there are so
many problems with the radioisotope dating methods which render them totally unreliable in providing time markers for the
different stages in the earths history. Indeed, the investigations of determinations of the decay constants of each of the
parent radioisotopes needs to be completed to provide further documentation of the uncertainties in that key time
assumption. Therefore, even though most of these eucrites (basaltic achondrite meteorites) yield a consistent Pb-Pb
isochron and model age of 4.554.57 Ga that cannot be their true real-time age, which according to the creation paradigm is
only about 6000 real-time years.
Conclusions
After decades of numerous careful radioisotope dating investigations of basaltic achondrite meteorites (eucrites) their Pb-Pb
isochron and model age of 4.554.57 Ga has been well established. This date for these eucrites is supported for many of
them by a strong clustering of their Pb-Pb isochron and model ages in the 4.554.57 Ga range, as well as being confirmed
by both isochron and model age results via the U-Pb method, and to a lesser extent, by the Rb-Sr, Lu-Hf, and Sm-Nd
methods. The Hf-W, Mn-Cr, Al-Mg, I-Xe, and Pu-Xe methods are all calibrated against the Pb-Pb isochron method, so their
results are not objectively independent. Thus the Pb-Pb isochron dating method stands supreme in the conventional
evolutionary uniformitarian community as the ultimate, most precise tool for determining the ages of the basaltic achondrite
meteorites.There are no discernible patterns in the isochron and model ages for these eucrites, apart from considerable
scatter of the Rb-Sr, Sm-Nd, and some Pb-Pb isochron ages, and the considerable scatter of the K-Ar, Ar-Ar, Rb-Sr, U-Pb,
Th-Pb, and even some Pb-Pb model ages. These basaltic achondrite ages do not follow the systematic pattern found in
Precambrian rock units during the RATE project. The -decay isochron ages are not always older than the -decay isochron
ages for particular eucrites, and among the -decayers the isochron ages are not always older according to the increasing
heaviness of the atomic weights of the parent radioisotopes, or the increasing lengths of their half-lives. Thus there appears
to be no consistent evidence in these basaltic achondrite meteorites similar to the evidence found in earth rocks of past
accelerated radioisotope decay.Any explanation for the 4.554.57 Ga age for these basaltic achondrite meteorites needs to
consider their origin. These meteorites are regarded as fragments of the asteroid 4-Vesta, debris derived from impact
cratering of its surface and collisions with other asteroids. Even in the naturalistic paradigm the asteroids, and thus the
meteorites, are regarded as primordial material left over from the formation of the solar system.Todays measured
radioisotope compositions of these eucrites may reflect a geochemical signature of that primordial material, which included
atoms of all elemental isotopes. Therefore if some, or perhaps most, of the daughter isotopes were thus inherited by these
basaltic achondrite meteorites when they were formed from that primordial material, and the parent isotopes in the
meteorites have only been subjected to some subsequent radioisotope decay (and none at accelerated rates), then the
4.554.57 Ga Pb-Pb isochron age for these eucrites cannot be their true real-time age, which according to the creation
paradigm is only about 6000 real-time years.However, these conclusions and suggested explanations as discussed are still
only tentative, their confirmation or adjustment awaiting the examination of more radioisotope dating data from meteorites in
remaining groups. Furthermore, further extensive studies of the radioisotope dating of lunar rocks and rocks from all levels
of the earths geologic record are required to attempt to systematize what proportions of the isotopes in each radioisotope
dating system measured today are due to inheritance from the primordial material, past accelerated radioisotope decay,
and mixing, additions, and subtractions in the earths mantle and crust through earth history.
Radioisotope Dating of Meteorites: IV. The Primitive and Other Achondrites
by Dr. Andrew A. Snelling on May 6, 2015
Abstract
Meteorites date the earth with a 4.55 0.07 Ga Pb-Pb isochron called the geochron. They appear to consistently yield
4.554.57 Ga radioisotope ages, adding to the uniformitarians confidence in the radioisotope dating methods. Among the
achondrites, meteorites not containing chondrules, are the primitive achondrites, angrites, aubrites, mesosiderites, and
irons. Many radioisotope dating studies in the last six decades have used the K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, U-Th-Pb, Lu-Hf,
Re-Os, Mn-Cr, Hf-W, Al-Mg, I-Xe, and Pu-Xe methods to yield an abundance of isochron and model ages for these
achondrites from whole-rock samples, and mineral and other fractions. Such age data for 10 of these achondrites were
tabulated and plotted on frequency versus age histogram diagrams. They strongly cluster in many of these achondrites at
4.554.57 Ga, dominated by Pb-Pb, U-Pb, and Pb-Pb calibrated isochron and model ages, testimony to the Pb-Pb
techniques supremacy as the uniformitarians ultimate dating tool, which they consider very reliable. These ages are
sometimes confirmed by Ar-Ar, Rb-Sr, Lu-Hf, Re-Os, and Sm-Nd isochron and/or model ages, but agreement could be due
to calibration with the Pb-Pb system. There is also scattering of many K-Ar, Ar-Ar, Rb-Sr, Re-Os, Sm-Nd, and a few U-Pb,
Pb-Pb, Th-Pb ages, in most cases likely due to thermal disturbances resulting from impact cratering of the parent asteroids.
No pattern was found in these meteorites isochron ages similar to the systematic patterns of isochron ages found in
Precambrian rock units during the RATE project, so there is no evidence of past accelerated radioisotope decay having
occurred in these achondrites, and therefore on their parent asteroids. This is not as expected, yet it is the same for all
meteorites so far studied. Thus it is argued that accelerated radioisotope decay must have only occurred on the earth, and
only the 500600 million years worth we have physical evidence for during the Flood. Otherwise, due to their 4.554.57 Ga
ages these achondrites and their parent asteroids are regarded as originally representing primordial material that was
made from which the non-earth portion were created of the solar system.Thus todays measured radioisotope compositions
of these achondrites could reflect a geochemical signature of that primordial material, which included atoms of all
elemental isotopes. So if most of the measured daughter isotopes were already in these achondrites when they were
formed on their parent asteroids, then their 4.554.57 Ga ages obtained by Pb-Pb, U-Pb, and Pb-Pb-calibrated isochron
and model age dating are likely not their true real-time ages, which according to the creation paradigm is only about 6000
real-time years. It is anticipated that further investigation of radioisotope ages data for groups of meteorites, and martian
and lunar meteorites, for lunar rocks, and for rocks from every level in the earths geologic record, should enable these
interim ideas to be confirmed or modified.
Shop NowKeywords: meteorites, classification, primitive achondrites, Acapulco, Caddo County IAB iron, angrites, Angra
dos Reis, DOrbigny, aubrites, Norton County, Shallowater, mesosiderites, Estherville, IIE irons, Colomera, Kodakanal,
Weekeroo Station, radioisotope dating, K-Ar, Ar-Ar, Rb-Sr, Sm-Nd, U-Th-Pb, Re-Os, Lu-Hf, Mn-Cr, Hf-W, Al-Mg, I-Xe, Pu-Xe,
isochron ages, model ages, discordant radioisotope ages, scattering of radioisotope ages, accelerated radioactive decay,
thermal disturbance, resetting, primordial material, geochemical signature, mixing, inheritance, Day Three Upheaval, the
Flood
Introduction

In 1956 Claire Patterson at the California Institute of Technology in Pasadena reported a Pb-Pb isochron age of
4.550.07Ga for three stony and two iron meteorites, which since then has been declared the age of the earth (Patterson
1956). Adding weight to that claim is the fact that many meteorites appear to consistently date to around the same age
(Dalrymple 1991, 2004), thus bolstering the evolutionary communitys confidence that they have successfully dated the age
of the earth and the solar system at around 4.56Ga. These apparent successes have also strengthened their case for the
supposed reliability of the increasingly sophisticated radioisotope dating methods.Creationists have commented little on the
radioisotope dating of meteorites, apart from acknowledging the use of Pattersons geochron to establish the age of the
earth, and that many meteorites give a similar old age. Morris (2007) did focus on the Allende carbonaceous chondrite as an
example of a well-studied meteorite analyzed by many radioisotope dating methods, but he only discussed the radioisotope
dating results from one, older (Tatsumoto, Unruh, and Desborough 1976) paper.In order to rectify this lack of engagement
by the creationist community with the meteorite radioisotope dating data, Snelling (2014a) obtained as much radioisotope
dating data as possible for the Allende CV3 carbonaceous chondrite meteorite (due to its claimed status as the most studied
meteorite), displayed the data, and attempted to analyze them. He found that both isochron and model ages for the total
rock, separated components, or combinations of these strongly clustered around a Pb-Pb age of 4.564.57Ga, the earliest
(Tatsumoto, Unruh, and Desborough 1976) and the latest (Amelin et al. 2010) determined Pb-Pb isochron ages at
4.5530.004Ga and 4.567180.0002Ga respectively being essentially the same. Apart from scatter of the U-Pb, Th-Pb, RbSr, and Ar-Ar ages, no systematic pattern was found in the Allende isochron and model ages similar to the systematic
pattern of isochron ages found in Precambrian rock units during the RATE project that was interpreted as produced by an
episode of past accelerated radioisotope decay (Snelling 2005c; Vardiman, Snelling, and Chaffin 2005).Snelling (2014b)
grouped together all the radioisotope ages obtained for ten ordinary (H, L, and LL) and five enstatite (E) chondrites and
similarly displayed the data. They generally clustered, strongly in the Richardton (H5), St. Marguerite (H4), Bardwell (L5),
Bjurbole (L4), and St. Sverin (LL6) ordinary chondrite meteorites, at 4.554.57 Ga, dominated by Pb-Pb and U-Pb isochron
and model ages, but confirmed by Ar-Ar, Rb-Sr, Re-Os, and Sm-Nd isochron ages. There was also scatter of the U-Pb, ThPb, Rb-Sr, and Ar-Ar model ages, in some cases possibly due to thermal disturbance. Again, no pattern was found in these
meteorites isochron ages indicative of past accelerated radioisotope decay.Snelling (2014c) subsequently compiled all the
radioisotope ages for 12 eucrite (basaltic) achondrites. The data for many of these meteorites again strongly clustered at
4.554.57Ga, dominated by Pb-Pb and U-Pb isochron and model ages but confirmed by Rb-Sr, Lu-Hf, and Sm-Nd isochron
ages. There was also scatter of the U-Pb, Pb-Pb, Th-Pb, Rb-Sr, K-Ar, and Ar-Ar model ages, in most cases likely due to
thermal disturbances resulting from metamorphism or impact cratering of the parent asteroid, identified as 4-Vesta. Again,
no pattern was found in these meteorites isochron ages similar to the systematic patterns of isochron ages found in
Precambrian rock units during the RATE project, so there was no evidence of past accelerated radioisotope decay having
occurred in these eucrites, and therefore on their parent asteroid.Snelling (2014a,b,c) then sought to discuss the possible
significance of this clustering in terms of various potential creationist models for the history of radioisotopes and their decay.
He favored the idea that asteroids and the meteorites derived from them are residual primordial material from the
formation of the solar system Thus he argued that todays measured radioisotope compositions of all these chondrites and
eucrites may reflect a geochemical signature of that primordial material, which included atoms of all elemental isotopes. So
if some of the daughter isotopes were already in these chondrites and eucrites when they were formed, then the 4.55
4.57Ga ages for them obtained by Pb-Pb and U-Pb isochron and model age dating are likely not their true real-time ages,
which according to the creation paradigm is only about 6000 real-time years. However, Snelling (2014a,b,c) suggested that
drawing final conclusions from the radioisotope dating data for just these 16 chondrite and 12 eucrite meteorites was still
premature, and recommended further studies of more meteorites from still other classification groups. This present
contribution is therefore designed to further document the radioisotope dating data for more meteorites, the primitive
achondrites and the other achondrites, the latter encompassing the angrites, aubrites, mesosiderites (stony-irons), and
irons, so as to continue the discussion of the potential significance of these data.
The Classification of Primitive and Other Achondrite Meteorites
The most recent classification scheme for the meteorites is that of Weisberg, McCoy, and Krot (2006), which is reproduced
in Fig. 1. Based on their bulk compositions and textures, Krot et al. (2005) divided meteorites into two major categories,
chondrites (meteorites containing chondrules) and achondrites (meteorites not containing chondrules or non-chondritic
meteorites). They further subdivided the achondrites into primitive achondrites and igneously differentiated achondrites.
However, Weisberg, McCoy, and Krot (2006) simply subdivided all meteorites into three categorieschondrites, primitive
achondrites, and achondrites (fig. 1).

Fig. 1. The classification system for meteorites (after Weisberg, McCoy, and Krot 2006).
The non-chondritic meteorites or achondrites contain virtually none of the components found in chondrites. It is
conventionally claimed that they were derived from chondritic materials by planetary melting, and that fractionation caused
their bulk compositions to deviate to various degrees from chondritic materials (Krot et al. 2005). The degrees of melting that
these rocks experienced are highly variable, and thus, these meteorites have been divided into the two major categories
primitive and differentiated. However, there is no clear cut boundary between these categories.The differentiated nonchondritic meteorites, or achondrites (fig. 1), are conventionally regarded as having been derived from parent bodies that
experienced large-scale partial melting, isotopic homogenization (ureilites are the only exception), and subsequent
differentiation. Based on abundance of Fe-Ni metal, these meteorites are commonly divided into three typesachondrites,
stony-irons, and irons. Each of these types contains several meteorite groups and ungrouped members (fig. 1). According to
uniformitarians, several groups of achondrites and iron meteorites are likely to be genetically related and were possibly
derived from single asteroids or planetary bodies.
The Primitive Achondrites

Several groups of stony (silicate-rich) meteorites that have essentially bulk chondritic compositions but achondritic textures
apparently formed by low degrees of melting have been collectively referred to as primitive achondrites (Krot et al. 2005;
Mittlefehldt 2005). Such partial melting would have not occurred during passage through the earths atmosphere, because
there is only sufficient time during passage for the outer surfaces of meteorites to be heated momentarily. Thus, these have
only been partially melted on their parent asteroids and therefore not completely differentiated, so they are regarded by
uniformitarians as chondritic meteorites that began their conversion to achondrites but the process was aborted before
completion (Norton 2002). They are therefore intermediate forms between the chondrites and achondrites. These include
the acapulcoites and lodranites in the acapulcoites-lodranite clan, the winonaites and silicate-inclusion-bearing IAB and
IIICD irons in the winonaites-IAB-IIICD silicate inclusions clan, and the brachinites and ureilites (fig. 1), although the latter
two groups are sometimes included in the differentiated achondrites.
Acapulco
This type specimen for the acapulcoites fell at El Quemado near Acapulco, Mexico, in August 1976 (Norton 2002; Palme et
al. 1981). A single, partially crust-covered massive stone weighing 1.9kg (4.2lb) was recovered from a crater 30cm (11.8in)
in diameter (Palme et al. 1981; Zipfel et al. 1995). At the time it was a confusing meteorite since it has a bulk composition
comparable to average H chondrites but a highly crystallized achondritic texture and unique mineral compositions
intermediate between those of E and H chondrites, with some similarities to those of silicate inclusions in some iron
meteorites. A second meteorite, which had fallen in 1868 a few miles east of Lodran in Punjab, Pakistan, was found to be
very similar in all characteristics, so these two anomalous stony meteorites remained unique until others of their kind began
to be recovered in Antarctica. Today, about 20 of these meteorites are known, most from Antarctica with a few from the
Sahara, plus the Monument Draw meteorite found in Texas in 1985 (Norton 2002; Zipfel et al. 1995).Acapulco is a finegrained equigranular rock that consists of a tight assemblage of orthopyroxene, olivine, and plagioclase,associated with
comparatively large amounts of metallic Fe-Ni alloys, and minor amounts of diopside (clinopyroxene), troilite [FeS], chromite
[FeCr2O4], and the phosphates apatite and whitlockite (Palme et al. 1981). The olivine with a fayalite content of 1012.5%
(Fa1012.5) and the orthopyroxene with a ferrosilite content of 1113% (Fs 1113) are the two main minerals and occur in equal
amounts. They are thus both Mg-rich, but their Fe/(Fe+Mg) ratio is sufficiently high (oxidized Fe) to place Acapulcos
mineralogy between the H and E chondrites. The plagioclase has an oligoclase composition, averaging An 14Ab82Or4, while
the clinopyroxene is emerald-green chromian diopside (En54Fs5Wo41). The phosphates occur either in irregular patches or as
small grains, the latter often being associated with the sulfide or metal phases. Their modal abundance of 1.6wt% is higher
than that of ordinary (O) equilibrated chondrites (0.6wt%). Also in contrast to the O chondrites whitlockite [Ca 9(MgFe)
(PO4)6PO3OH] constitutes only 10% of the phosphates, the remaining 90% being chlorapatite [Ca10(PO4)6(OH,Cl)2].
The high amount of the Fe-Ni metal alloys kamacite and taenite (~20 wt%) and their heterogeneous distribution in this
meteorite convinced some early investigators that this was some kind of stony-iron meteorite (Norton 2002). The degree of
textural equilibrium observed in Acapulco is unusually high compared with type 6 and 7 O chondrites (see figs 5 and 6 in
Snelling 2014b), indicating thermal metamorphism at a high temperature and/or of longer duration, or an igneous origin
(Zipfel et al. 1995). While the peak metamorphic temperature is unknown, it was sufficiently high to melt the Fe-Ni alloys and
troilite. Oxygen isotopic compositions cluster between the terrestrial and carbonaceous chondrite fractionation line, showing
its lack of homogeneity, so differentiation and igneous processes were apparently insufficient to homogenize it (Norton
2002).
Caddo County IAB Iron
The ~16 kg (~35 lb) Caddo County IAB iron meteorite was found in Oklahoma in 1987 (Takeda et al. 2000). Petrologic and
chemical studies showed that there were silicate inclusions in this meteorite, including small basaltic inclusions and even
larger basaltic/gabbroic segregations up to several centimeters (an inch or so) wide. These segregations are not clasts like
those within a breccia, as they have transitional margins located mainly at silicate/metal alloy boundaries only a few
centimeters (an inch or so) from the fine-grained ultramafic silicates. This is similar to what occurs in winonaites, which is
why the IAB iron meteorites like Caddo County are classified with the winonaites in the winonaites-IAB-IIICD silicates clan.
Even the metal compositions of IAB and IIICD iron meteorites are similar enough to warrant the inclusion of these
meteorites in this clam within the primitive achondrites (Mittlefehldt 2005).The winonaites-IAB-IIICD silicate inclusions clan
contains members with primitive chondritic compositions, basalt-depleted lithologies, and lithologies containing mafic
segregations (Benedix et al. 2000). Modal plagioclase contents are like those of O chondrites, and some winonaites contain
relict chondrules. Texturally winonaites show internal heterogeneity and substantial differences between members
(Mittlefehldt 2005). The textures are typically metamorphic. Winonaites contain olivine, orthopyroxene, clinopyroxene,
plagioclase, troilite, Fe-Ni alloys, chromite, apatite, graphite, K-feldspar, daubreelite [FeCr 2S4], schreibersite [(Fe,Ni)3P], and
alabandite [MnS]. The silicate and other inclusions in IAB iron meteorites are heterogeneous and have been subdivided into
five types(i) angular, chondritic silicate; (ii) non-chondritic silicate-rich; (iii) sulfide-rich; (iv) rounded, often graphite-rich;
and (v) phosphate-bearing (Benedix et al. 2000). The graphite-rich inclusions only rarely contain silicates, and some sulfiderich inclusions contain silicates that are apparently angular chondritic silicates. The mineralogy of these silicate inclusions in
IAB iron meteorites are diverse, likely due to reaction with the host Fe-Ni metal alloys (Mittlefehldt 2005). They contain
olivine, orthopyroxene, clinopyroxene, plagioclase, troilite, Fe-Ni alloys, chromite, daubreelite, graphite, and the phosphates
apatite, whitlockite, and rarely, brianite [Na2CaMg(PO4)2].The coarse-grained gabbroic inclusions in the Caddo County IAB
iron meteorite consists of chromian diopside up to 2 1.3 mm (~0.08 0.05 in) in size set between plagioclase crystals
more than 3 mm (0.12 in) in length (Takeda et al. 2000). The modal abundances are plagioclase (59%) and diopside (28%),
with less abundant small, rounded grains of orthopyroxene (5%) and olivine (8%) (fig. 2). The coarse-grained gabbroic
material is not a clast included in the Fe-Ni metal alloys, or in the finer-grained, recrystallized chondritic material. Rather, it is
within the silicate masses, with transitional boundaries with the surrounding ultramafic silicates. In the transitional region
between the two lithologies medium-sized orthopyroxene and olivine grains are dominant. The olivine with Fa 2.53.3 is strongly
magnesian, as is the fairly uniform orthopyroxene with En 9094Fs5
7Wo12.5 (Benedix et al. 2000; Takeda et al. 2000). The plagioclase
with An1618Ab7981Or3 is within the range typical of primitive
achondrites. Diopside (clinopyroxene) crystals display some slight
zoning in CaO at the expense of MgO (En 5361Fs2.5Wo3744) and in Cr
(0.51.2 wt% Cr2O3).
Fig. 2. Photomicrograph of a silicate inclusion in the primitive
achondrite the Caddo County IAB iron meteorite in transmitted light
with partially crossed polars. This inclusion is a basaltic clast
composed of plagioclase (pl, gray), clinopyroxene (cpx, whitish) and
olivine (ol, pink) (after Krot et al. 2005).

Uniformitarians suggest it is likely that both the IAB irons and the winonaites came from the same parent body (Norton
2002). Their textures and compositions suggest that they are a mix of chondrite and metal breccia. Since iron and silicates
are immiscible and tend to separate in a melt, finding silicates associated with irons would not be expected. In fact, the
relative rarity of silicate inclusions in iron meteorites would seem to be a strong argument for iron meteorites formed by
magmatic processes within differentiated asteroids resulting in metallic core formation. However, these silicated irons, such
as the IAB iron meteorites, do not seem to have been associated with magmatic processes that led to differentiation
because their trace elements show only weak trends or slopes that are inconsistent with fractional crystallization. Yet the
silicates were heated sufficiently to recrystallize them or in some cases melt them (hence the metamorphic textures), but the
metal was not completely melted. Thus there is no consensus yet on how the metal phase was formedimpact melting,
fractional melting with melt segregation, batch melting with melt segregation, and fractional crystallization are all suggested
mechanisms (Mittlefehldt 2005). But all these suggested processes have serious difficulties, as they are not supported by
the observed textures and compositions of these meteorites, or experimentally derived solid metal/liquid metal partition
coefficients. The most comprehensive model attempting to explain the origin of these meteorites proposes the breakup and
reassembly of a hot, partially differentiated parent body (Benedix et al. 2000). The breakup and reassembly event is
supposed to have effectively frozen the evolution of the parent body, and mixed different materials together.
The Other Achondrites
The other achondrites are sometimes referred to as the differentiated achondrites because they would appear to represent
the products of classical (as understood here on the earth) igneous processes acting on the silicate-oxide system of
asteroidal bodiespartial to complete melting, and magmatic crystallization (Mittlefehldt 2005). Iron meteorites represent
the complementary metal-sulfide system products of this process. These other achondrites therefore include the so-called
asteroidal meteorites the HED or Vesta clan (howardites, eucrites, and diogenites), the angrites, and the aubrites; the socalled stony-ironsthe mesosiderites and pallasites; and all the other groups of iron meteorites (fig. 1). Snelling (2014c) has
already discussed the HED clan meteorites, particularly the eucrites (basaltic achondrites). The martian (SNC) and lunar
meteorites are also achondrites.
AngritesAngra dos Reis (ADOR) and DOrbigny
The angrite group currently consists of 13 meteorites, the best known of which are Angra dos Reis (ADOR), LEW (Lewis
Cliffs) 86010, Asuka 881371, DOrbigny, and Sahara 99555 (Mittlefehldt 2005). These meteorites are linked by identical
oxygen-isotopic compositions, similar unusual mineralogy, and several distinctive geochemical characteristics. Although
some are petrologically anomalous, the preponderance of distinctive characteristics indicates that they plausibly originated
on a common parent body. The angrites thus appear to be medium- to coarse-grained mafic igneous rocks of basaltic
composition (though critically silica undersaturated and highly alkali-depleted) from the crust of a differentiated asteroid (Krot
et al. 2005; Mittlefehldt 2005). They are believed by uniformitarians to have formed as partial melts of primitive source
materials under relatively oxidizing conditions. As the angrites do not seem to form a single fractionation sequence, several
parent melts are required, although many could have followed a similar partial melting/crystallization path. The unusual
pyroxene compositions in angrites (see below) appear familiar to the Ca-Al-Ti inclusions in CV chondrites (carbonaceous,
Vigarano-like chondrites) such as Allende (Snelling 2014a). If there was a chondritic precursor for these angrites, the closest
would likely be the CV chondrites.The first recognized and most famous angrite is the Angra dos Reis (ADOR), a single
stone weighing about 1.5 kg (3.3 lb) which fell into Angra dos Reis Bay in the State of Rio de Janeiro, Brazil, in January
1869 (Norton 2002). It remained a unique meteorite for over a century. Then over three consecutive years three more
angrites were found and recognized from the Antarctic ice-fieldsLEW 86010 (found in 1986), LEW 87051 (found in 1987),
and Asuka 881371 (found in 1988). Subsequently the remarkable specimen Sahara 99555, a single stone weighing 2.71 kg
(5.97 lb), was found in May 1999 in the Libyan Sahara Desert. And recently (in 2001) the new angrite DOrbigny, a 16.55 kg
(36.49 lb) stone, became available for study, though it had been found back in July 1979 by a farmer in a corn field near
Buenos Aires, Argentina (Grossman and Zipfel 2001; Kurat et al. 2004).The Angra dos Reis (ADOR) meteorite is an
ultramafic igneous rock, 95% composed by an unusual variety of clinopyroxene with a Ca-Al-Ti-rich and very low Fe content
commonly called fassaite [Ca(Mg,Fe,Ti,Al)(Si,Al)2O6], but formerly named by the International Mineralogical Association as
aluminian-ferrain-diopside (Mittlefehldt 2005; Norton 2002). This pyroxene in ADOR is homogeneous with a composition of
En33Fs12Wo55. Accessory minerals include olivine and the Ca-rich olivine kirschsteinite [Ca(Mg,Fe)SiO 4] with uniform
compositions of Fo53 and Fo38respectively. Only trace plagioclase (An >99) is sometimes present. Aluminous spinel, troilite,
whitlockite, ulvspinel [Fe2TiO4]-magnetite [Fe3O4] solid solution, and Ni-rich metal are common minor or trace phases.
ADOR has an equilibrated texture, with groundmass olivine and pyroxene occurring as small xenomorphic, equidimensional
grains, and larger poikilitic pyroxene grains enclosing smaller grains.DOrbigny when found was mostly covered with dark
gray-brown fusion crust, and had a somewhat unusual shape (Kurat et al. 2004) (fig. 3). This meteorite had a shape like a
loaf of bread, with front and back shields curved in a semi-parallel way and intergrown on one side with the opposite side
opening like a clam. The compact rock macroscopically had a medium-grained micro-gabbroic texture with abundant
scattered vesicles or druses sometimes lined with glasses (Grossman and Zipfel 2001; Kurat et al. 2004; Mikouchi and
MacKay 2001) (fig. 4). Its texture is a quenched igneous texture ophitic to subophitic, common graphic intergrowths of
olivine and plagioclase, and some are porphyritic with coarse subhedral to euhedral olivine grains up to 1 cm (0.4 in) in
diameter (fig. 5). The modal abundance is plagioclase 39.4%, fassaite 27.7%, Mg-rich olivine 19.4%, kirschsteinite
[CaFeSiO4] 11.9%, spinel (mostly ulvspinel) 0.6%, troilite 0.5%, and Ca silico-phosphate 0.5% (Kurat et al. 2004; Mikouchi
and McKay 2001). Trace phases include the Fe-Ni metal taenite, Ni-pyrrhotite and pentlandite. The fassaite and olivine
show extensive chemical zoning, whereas the plagioclase is homogeneous at An100. The fassaite has more magnesian
cores at En63, but rims are zoned down to En0,. Its Al2O3 and TiO2 contents are high and variable. The TiO 2 content
increases from 1.4 wt% in the cores to 5 wt% in the rims, whereas Al 2O3 shows complex variation with En content, usually
being homogeneous at 8 wt% in the cores, but dropping to 6 wt% at the outer rims. Olivine phenocrysts have cores with
Fo63 but are zoned, the CaO content increasing as Fo decreases from 0.8 wt% in the cores to ~8 wt% at Fo 10, at which point
the olivine becomes an intergrowth of sub-calcic kirschsteinite and Ca-rich fayalite (Fe-rich olivine). Granoblastic olivine
grains that are interpreted to be xenocrystic are much more Mg-rich at Fo90.

Fig. 3. The DOrbigny angrite meteorite as it was found. (A)


Front shield with elongated regmaglypts (small, well-defined
indentations or pits on the surface produced by selective
erosion during passage through the atmosphere) pointing
radially from the stagnation point (center left) and open
round vugs of variable sizes. Scars are from partially
removed fusion crust gouged by the plow. Specimens for
first identification and bulk analysis were taken from the
compact lithology of the front shield at upper right (visible
cut). (B) Side view at lower part in (A) showing the edge at
which the front shield and the back side pan, both consisting
of compact rock, are intergrown. Note the open round vugs
and the porous lithology to the right and left of the pan (large
irregularly shaped open spaces). Pan and shield are
covered by fusion crust, edges are worn and without fusion
crust. (C) Inclined view at back side pan from upper side in
(A), opposite side from (B), depicting the pan and the porous lithology situated between the compact shield and pan
lithologies. Note the abundant irregularly shaped open spaces, some of which are partially filled by caliche (white). Length of
the meteorite was 34 cm (13.4 in) (after Kurat et al. 2004).

Fig. 4. Small rock chips of the DOrbigny angrite meteorite (after Mikouchi and McKay 2001). (a) Note fine-grained lithology
at the right end of chip. (b) Note pale green olivine megacryst.

Fig. 5. Anorthite (white)-olivine (brown) intergrowth in transmitted light. (a) Both phases are intimately intergrown, olivine is
in places replaced by augite (dark brown). (b) Same as (a) but under uncrossed polarizers. Olivine has mostly the same
optical orientation (consists of a single crystal) and anorthite forms many twinned individual grains (after Kurat et al. 2004).
AubritesNorton County and Shallowater
About 8% of all known achondrites are aubrites, named after the fall of the type specimen near Aubres, France, on
September 14, 1836 (Keil 2010; Norton 2002). The aubrites (or enstatite achondrites) are highly reduced enstatite
pyroxenites, their silicate minerals being essentially FeO free (Krot et al. 2005; Mittlefehldt 2005). They are nearly all
monomict breccias, with very coarse fragments of enstatite many millimeters across, embedded in a matrix of crushed and
comminuted enstatite. They consist of 7598 vol. % FeO-free enstatite (En>99), with variable lesser amounts of albitic
plagioclase (Ab8895), and virtually FeO-free, high-Ca diopside and Mg-olivine (forsterite). Aubrites have only small amounts
of iron averaging <2 wt% Si-bearing Fe-Ni metal, with the remainder as troilite. Other phases include a host of accessory
unusual sulfides in which elements that are normally lithophile are chalcophile, such as Ti in heidite [FeTi2S4], Na in
caswellsilverite [NaCrS2], and Ca in oldhamite [CaS]. Clasts of both igneous and impact-melt origin are common, and the
precursors to the breccias were mostly coarse-grained (probably plutonic) orthopyroxenites, that is, very coarse enstatite
grains exhibiting remnant original igneous textures.Aubrites share nearly identical unusual silicate mineralogy and oxygenisotopic composition, and a similar highly reduced nature with enstatite (E) chondrites. This suggests to uniformitarians that
they may be closely related to enstatite chondrite-like parent bodies (Krot et al. 2005; Mittlefehldt 2005; Norton 2002).
However, even though the aubrites perhaps formed from the partial melting of a single body, they are not thought to have
formed on the same parent bodies as the E chondrites. Yet to uniformitarians the aubrites are clearly of igneous origin in
which large enstatite crystals formed by extensive fractional crystallization of an Mg-rich ultramafic magma, although some
suggest that while the aubrite parent body was melted sufficiently to segregate metal and sulfide into a core, large-scale
silicate fractionation did not occur. Nevertheless, since the aubrites now are nearly all breccias, their parent body must have
melted, recrystallized, and then fragmented by collision, later to be reassembled into an enstatite rubble pile (asteroid).The
Norton County aubrite fell on September 14, 1948 in Norton County, Kansas, and when recovered the stone weighed 998 kg
(2200 lb) (Keil 2010). It is the largest known and thus probably the best-studied aubrite, largely because of the ample
material available for research (fig. 6). It is a fragmental impact breccia consisting of coarse and angular enstatite crystal
fragments, embedded into a clastic matrix, mostly made of comminuted enstatite (fig. 7), together with a host of mineral and
lithic fragments of igneous and impact-melt origin (Keil 2010). The precursor materials were most likely plutonic igneous
rocks (fig. 8), which formed (in uniformitarian thinking) through extensive melting and igneous differentiation of the aubrite
parent body, a process responsible for the formation of the observed diverse lithologies. These include dunites (represented
by forsterite crystals), plutonic orthopyroxenites (represented by most enstatite crystals in the matrix), plutonic pyroxenites
(the pyroxenitic clasts; fig. 8), plagioclase-silica rocks (like the feldspathic clasts), and impact-melt breccias (the microporphyritic clasts and the diopside-plagioclase-silica-clast). The most noteworthy are the pyroxenitic clasts, which occur in
large, crystalline masses up to 8 cm (3 in) in size in many hand specimens and are clearly of plutonic origin. They consist of
xenomorphic-granular and tightly intergrown, coarse enstatite (En99Wo1), with minor primary diopside (En55Wo45), and
accessory kamacite, taenite, titanoan troilite, oldhamite [CaS], and caswellsilverite (fig. 8). Xenomorphic plagioclase (Ab 7888)
occurs in the interstices between enstatite grains and forms myrmekitic intergrowths with silica (probably tridymite).
Casanova, Keil, and Newsom (1993) studied the mineralogy and trace element composition of metal and metal nodules, up
to 1.5 cm (0.6 in) in size, in the Norton County aubrite
and concluded that the metal did not undergo fractional
crystallization in a core but, rather, is a fraction of the
metallic Fe-Ni which, during partial melting of an
enstatite chondrite-like precursor lithology, was not
completely segregated from the silicates. A remarkable,
oldhamite-dominated igneous lithology consists of clasts
containing oldhamite [CaS] single crystals up to 2 cm
(0.8 in) in size, with inclusions of ferromagnesian
alabandite [MnS], troilite, daubreelite [FeCr2S4],
caswellsilverite [NaCrS2], and metallic Fe-Ni, which are
usually in intimate contact with a silicate portion
consisting of enstatite, forsterite, and/or plagioclase and
testify to the formation of the oldhamite from a melt
(Wheelock et al. 1994).
Fig. 6. The Norton County aubrite meteorite (after
Norton 2002). This is a monomict breccia with coarse
clasts of enstatite and a comminuted matrix of the same
mineral. Fe-Ni metal is present as rusty brown grains on
the edges of the specimen. The face measures 32 mm (1.26 in) across.
Fig. 7. Photomicrograph of the brecciated texture of the
Norton County aubrite meteorite, which is typical for the
common aubrite fragmental impact breccias, consisting
mostly of fragments of enstatite and enstatite-dominated
lithic clasts of igneous parentage. Transmitted light, polars
partially crossed. Longest horizontal dimension equals 6
cm (2.4 in) (after Keil 2010).
Fig. 8. Photomicrograph of the plutonic, xenomorphicgranular texture of a pyroxenitic clast from the Norton
County aubrite, dominated by large enstatite crystals with
diopside exsolution and minor diopside with enstatite
exsolution. Transmitted light, polars partially crossed.
Longest horizontal dimension is 3.5 cm (1.37 in) (after Keil
2010).

The Shallowater aubrite was found in Lubbock County, Texas in 1936


and weighed in at 4.6 kg (10 lb) (Keil 2010; Keil et al. 1989). It is a nonbrecciated, coarse-grained orthopyroxenite that consists of large, up to
4.5 cm (1.8 in), subhedral poikilitic orthoenstatite crystals (up to 80 vol.
%) which enclose smaller orthoenstatite crystals of identical composition
(En>99) and which also contain as inclusions and in the interstices
xenoliths of an assemblage of low-Ca clinoenstatite (1 vol.%), forsterite
(Mg-olivine) (2.9 vol.%), plagioclase (Ab8088An017Or311) (2.5 vol.%),
metallic Fe-Ni (3.3 vol.%, which is more than five times more than in
other aubrites), troilite (2.9 vol.%), schreibersite [(Fe,Ni) 3P] (0.4 vol.%),
weathered opaques (8 vol.%), and traces of niningerite (a magnesiumiron-manganese sulfide) and oldhamite [CaS] (Keil et al. 1989).
Regarded as definitely of igneous origin, the rock is deduced to have
experienced a complex, three-stage cooling history, which prompted
Keil et al. (1989) to suggest that its distinct parent body formed by the
reassembly of the debris from the collision of a partly or totally molten
enstatite-rich aubrite-like body with a solid, enstatite-chondrite-like body
(represented by the xenoliths), followed by rapid cooling and crystallization. They also suggested that this was not the main
aubrite parent body (asteroid) from which other aubrites came, but rather that Shallowater represents a sample from a
second different aubrite parent body.
MesosideritesEstherville
Mesosiderites are stony-iron polymict breccias composed of angular fragments of different mineral compositions, an unusual
mixture of silicates mostly from basaltic, gabbroic and orthopyroxenite sources with Fe-Ni metal and troilite (Mittlefehldt
1990, 2005; Norton 2002). The silicate fragments and Fe-Ni metal are in nearly equal proportions, which is why these
meteorites were originally classified as stony-irons. The Fe-Ni metal may be uniformly distributed within a silicate matrix with
large silicate fragments enclosed by the Fe-Ni metal, or the Fe-Ni metal may be found in clumps or aggregates surrounded
by silicates, as in the case of the Estherville mesosiderite (fig. 9). Superficially they appear distinct from the basaltic (HED)
achondrites because of their large Fe- Ni metal content, but mineral, chemical, and oxygen isotopic connections are
recognized between the basaltic (HED) achondrites and the mesosiderites. It is because the silicate fragments are of crustal
rocks from a differentiated body with both the rock fragments and silicate minerals very similar to those in the HED suite
basaltic achondrite meteorites that both the mesosiderites and HED achondrites were the result of impact mixing on different
parent asteroidal bodies.
Fig. 9. The mesosiderite meteorite Estherville (after Norton
2002). This slab is about 10 cm (about 4 in) across at its
longest dimension. The Fe-Ni metal is fragmental and in
aggregates scattered through the dark silicate matrix.The
silicates consist of mineral and lithic clasts set in a finegrained fragmental to impact-melt matrix (Mittlefehldt 2005).
The most common lithic clasts are basalts, gabbros, and
orthopyroxenites, while dunites are minor and anorthosites
are rare. The most common mineral clasts are centimetersized orthopyroxene and olivine fragments, while millimetersized plagioclase fragments are less common. Olivine clasts
are typically single-crystal fragments, varying in composition
from Fo92 to Fo58. Mesosiderites also contain fine-grained
olivine whose compositions may have been altered by
metamorphic equilibration. The pyroxene clasts are
dominated by the low-Ca orthopyroxene hypersthene with
smaller amounts of low-Ca pigeonite. Compositional ranges
are from about Fs20 to Fs40. Plagioclase grains are calcic, almost pure anorthite. Basaltic and gabbroic clasts are composed
of ferroan pigeonite and calcic plagioclase, with minor to accessory silica, whitlockite, augite, chromite, and ilmenite. Fe-Ni
metal and troilite are common, but these have been added during brecciation. The matrix varies from cataclastic texture with
highly angular mineral fragments to igneous-textured. The Fe-Ni metal portion of mesosiderites has a uniform composition
unlike iron meteorites, with a Ni content of between 7 and 10%.The fall of the Estherville mesosiderite in Iowa, USA
occurred on May 10, 1879 (Powell 1971). After a brilliant fireball had been seen, a shower of several large masses and
many small fragments fell, totaling 320 kg (710 lb). Esthervilles Fe-Ni metal distribution is patchy surrounded by silicates so
that the Fe-Ni metal clumps and aggregates which total 56 wt% of this meteorite appear isolated in planar sections (fig. 9).
Its texture is characterized by brecciated, angular silicate lithic fragments and individual silicate crystals. The major silicate
mineral constituents (> 10 vol.%) are calcic plagioclase and orthopyroxene, while pigeonite and olivine are minor
constituents (210 vol.%), and accessories (< 2 vol.%) include augite, tridymite, apatite, whitlockite, schreibersite, chromite,
and ilmenite (Powell 1971). Lamellae of rutile occur within the ilmenite, and troilite is also an accessory phase as small
irregular blebs (small, usually rounded inclusions) interstitial to the silicate grains within the lithic fragments.
IIE IronsColomera, Kodakanal, and Weekeroo Station
Of the meteorites seen to fall, only about 4% are irons (Norton 2002). Yet some 40% of all finds are irons, because they are
stronger and more easily pass through the earths atmosphere with less fragmenting than stony meteorites so they tend to
be larger and far easier to recognize and find than stony meteorites. Even though they are subject to rusting, being
composed of Fe-Ni alloys, they are still much more resistant to weathering than stony meteorites. The cut face of an iron
meteorite is usually featureless, its apparent uniformity only broken by occasional inclusions. However, if the surface is
etched with dilute nitric acid a unique texture appears. Called Widmansttten structure, it is the result of the intergrowth of
lamellae or plates of two Fe-Ni alloys, the low-Ni kamacite and high-Ni taenite. Iron meteorites that display this
Widmansttten structure are called octahedrites, since the intergrown kamacite/taenite plates are arranged parallel to the
eight equilateral triangular faces of a back-to-back pyramidal structure called an octahedral dipyramid or simply an
octahedron.Irons are regarded by uniformitarians as differentiated meteorites that represent the cores of parent asteroidal
bodies. In their view, precursor refractory minerals first condensed out of the solar nebula. After a brief period of heating,
chondrules formed and crystallized. Then the original chondritic parent body formed through accretion and agglomeration of
these chondrules and other solid condensates such as Fe-Ni metal grains. After building up to perhaps 100200 km (about

60120 mi) in diameter, the chondritic body supposedly melted and differentiated to produce a layered differentiated parent
body with a metallic core surrounded by a largely silicate mantle and crust. Subsequently, the crust and mantle rocks were
stripped away by multiple impacts to expose and eventually denude the metallic core. Finally, the Fe-Ni metallic core which
may have measured several kilometers (a few miles) across would have been fragmented.There are two classifications of
iron meteorites based on either structural or chemical criteria. Earlier classifications were based on the structural
appearance of the cut and etched surfaces of irons, especially the texture of the Widmansttten structure, that is, on how
wide or narrow are the kamacite plates (Norton 2002). The width depends on the bulk Ni content, so the structural
classification is based on band width of the kamacite lamellae and Ni content (table 1). This is primarily an octahedrite
classification with seven sub-groups ranging from coarsest to finest kamacite plate widths as their Ni content increases. Two
other iron meteorite groups define the low and high end of the Ni content range, the hexahedrites and ataxites respectively.
However, there are a few unique iron meteorites that do not conform to any of these classification groups, as they display
unusual structures and/or chemistry, such as a random-appearing polycrystalline texture, troilite nodules, and/or silicate
inclusions. Furthermore, the textural groups in this structural classification give the impression they are distinct from each
other. Yet they are not well-defined groups but grade smoothly from one texture into another. The Ni content boundaries for
each textural group are also somewhat arbitrary.
Table 1. Structural classification of iron meteorites according to their internal structure (after Norton 2002).
The structural parameters in that structural
Group
Symbol Band Width (mm) Nickel (%)
classification have been combined with several
chemical parameters such as nickel and trace
Hexahedrite H
>50
4.56.5
element contents to produce a more definite
Octahedrites O
classification with meaningful distinct genetic groups
that could represent different parent bodies (table 2).
Coarsest
0gg
3.350
6.57.2
There are certain trace elements such as gallium
(Ga), germanium (Ge), and iridium (Ir) that like Ni are
Coarse
Og
1.33.3
6.57.2
siderophile (or iron-loving), so they are used to subdivide the iron meteorites into distinct chemical
Medium
Orn
0.51.3
7.410.3
groups. Experiments have shown that because Ni
tends to accumulate and concentrate in the liquid
Fine
Of
0.20.5
7.812.7
phase, then the first solid Fe-Ni alloy accumulating,
Finest
Off
<0.2
7.812.7
presumably at the developing core of a differentiated
asteroid, is relatively low in Ni. As the crystallization
Plessitic
Opl
<0.2
Kamacite spindles
process continues both the melt and crystallizing
solid metal become richer in Ni, although the metal is
Ataxite
D
No structure
>16.0
still less rich in Ni than the melt. This fractional
crystallization results in metal alloys with various Ni contents. These three siderophile trace elements have different affinities
for Fe as a solid versus Fe as a liquid, so fractional crystallization affects them much more strongly. For example, Ir prefers
to combine with the first crystallizing Fe metal which is low in Ni. With increasing Ni concentration in the melt, the amount of
Ir in the crystallizing metal decreases since most of it is partitioned into the first metal to crystallize.
Table 2. Structural and chemical relationships in iron meteorites (after Norton 2002).
Chemical Groups Frequency (%)

Band Width (mm)

Ni (wt%)

Structural Groups

IA

17.0

1.03

6.48.7

Ogg, Og, Om

IB

1.7

0.011.0

8.725

Om D

IC

2.1

<3

6.16.8

Anom, Og

IIA

8.1

>50

5.35.7

IIB

2.7

515

5.76.4

Ogg

IIC

1.4

0.060.07

9.311.5

Opl

IID

2.7

0.40.8

9.611.3

Om, Of

IIE

2.5

0.72

7.59.7

Ogg Om (Anom)

IIF

1.0

0.050.21

8.410

Of D

IIIA

24.8

0.91.3

7.19.3

Om

IIIB

7.5

0.61.3

8.410.5

Om

IIIC

1.4

0.23

1013

Ogg Off

IIID

1.0

0.010.05

1623

Off D

IIIE

1.7

1.31.6

8.29.0

Ogg

IIIF

1.0

0.51.5

6.87.8

Ogg Of

IVA

8.3

0.250.45

7.49.4

Of

IVB

2.3

0.0060.03

1626

Therefore, when Ir abundances in the crystallized Fe-Ni metals in iron meteorites are plotted against their Ni contents (which
have resulted from melts with different initial bulk Ni contents) distinct narrow fields appear, often with negative slopes (fig.
10). These small but distinct fields, made up of five or more meteorites, initially allowed most iron meteorites to be classified
into one of 17 chemical groups (table 2). A few of these groups were subsequently found to be related to each other. For
example, IIIA and IIIB irons both have very narrow structural and Ni content variations, and both are in the Om structural
group, so they are now usually combined as IIIAB irons. Other chemical groups have been similarly paired, thus reducing

the number of distinct groups to 13. About 15% of all known iron meteorites do not fall into one of these groups and remain
anomalous. A detailed description of the individual iron meteorite groups is found in Scott and Wasson (1975).
Fig. 10. A log-log plot of Ni (nickel) abundance versus Ir (iridium)
abundance in iron meteorites. This allows for resolution of different
compositional fields so that the meteorites are divided into differently
labelled groups that can be distinguished from each other based on their
Ir concentrations (after Scott and Wasson 1975).
Because the purpose of this study was to focus on individual meteorites
dated by more than one radioisotope method, the only group of iron
meteorites that contained representatives which met that criterion was
the IIE irons. They are a small group with very diverse characteristics in
terms of metal textures and the mineralogy of their silicate inclusions,
which are primarily what has been radioisotope dated. The chemical
composition of their Fe-Ni metals is very restricted (figs. 10 and 11) and
inconsistent with fractional crystallization (Scott and Wasson 1975;
Wasson and Wang 1986). Their silicate inclusions range from angular
metamorphosed chondrites (containing chondrules) to highly
differentiated silicates in small feldspar-rich globules (Bogard, Garrison,
and McCoy 2000).
Fig. 11. A log-log plot of the abundance of Ni (nickel) versus Ge
(germanium) in iron meteorites. Thirteen fields are distinguished, each
representing a different chemical group and a postulated different parent
body (after Scott and Wasson 1975).
The IIE irons Colomera, Kodakanal, and Weekeroo Station (fig. 12) all
contain globular differentiated silicate inclusions (Bogard, Garrison,
and McCoy 2000). They are rounded to elongated and typically can
reach 1 cm (0.4 in) and comprise about 10 vol. % of the bulk of each
meteorite. Inclusions in Weekeroo Station are dominantly
orthopyroxene (24 vol. %), clinopyroxene (16 vol. %) and plagioclase
(59 vol. %) in a ratio of 1:1:2, whereas Colomera and Kodakanal
inclusions contain major clinopyroxene (28 and 21 vol. % respectively)
and plagioclase (67 and 73 vol. % respectively) in a ratio of ~1:2 or 1:3
with only minor orthopyroxene (both 3 vol. %). Colomera, Kodakanal,
and Weekeroo Station contain coarse-grained (up to 5 mm or 0.2 in
grain size) gabbroic inclusions, partially to wholly cryptocrystalline
(consisting of crystals that are too small to be distinguishable under a
microscope) inclusions, and glassy inclusions. The ratio of these types
can differ significantly. Most inclusions in Weekeroo Station are at least
partially cryptocrystalline, and are typified by a corona structure in
which large (up to several millimeters) corroded augite (Fs 17.66) is
rimmed by orthopyroxene (bronzite, Fs22.2) and surrounded by a finegrained radiating structure of acicular feldspar, tridymite, and glass
(Bogard, Garrison, and McCoy 2000; Bunch, Keil, and Olsen 1970).
Colomera also contains similar cryptocrystalline, plus gabbroic
inclusions consisting of 15 mm (0.030.19 in) grains of feldspar (both
plagioclase and K-feldspar), augite (Fs812Wo4045), and orthopyroxene
(Fs1923), along with millimeter-sized chromite grains, plus accessory
rutile, schreibersite, troilite, and whitlockite. Glassy inclusions are found in both Colomera and Kodakanal. A variety of shock
features are evident in the silicates of these three IIE iron meteorites, including deformation twins, planar fractures, and
undulatory extinction.
Fig. 12. A polished and etched slab of the coarse octahedrite iron IIE silicate-bearing meteorite Weekeroo Station. The
silicate inclusions consist of plagioclase, orthopyroxene, and clinopyroxene (after Krot et al. 2005).
The Radioisotope Dating of the Primitive and Other Achondrite Meteorites
To thoroughly investigate the radioisotope dating
of the primitive and other achondrites all the
relevant literature was searched. The objective
was to find primitive and other achondrites that
have been dated by more than one radioisotope
method, and a convenient place to start was
Dalrymple (1991, 2004), who compiled lists of
such data. The ten primitive and other achondrite
meteorites that were found to have been dated
multiple times by more than one radioisotope
methodthe primitive achondrites Acapulco and
Caddo County IAB iron, the angrites Angra dos
Reis (ADOR) and DOrbigny, the aubrites Norton County and Shallowater, the mesosiderite Estherville, and the IIE irons
Colomera, Kodakanal, and Weekeroo Station thus became the focus of this study. When papers containing radioisotope
dating results for these achondrites were found, the reference lists were also scanned to find further relevant papers. In this
way a comprehensive set of papers, articles, and abstracts on radioisotope dating of these achondrite meteorites was
collected. While it cannot be claimed that all the papers, articles, and abstracts which have ever been published containing
radioisotope dating results for these achondrites have thus been obtained, the cross-checking undertaken between these
publications does indicate the data set obtained is very comprehensive.All the radioisotope dating results of these ten

primitive and other achondrites were then compiled and tabulated. For ease of viewing and comparing the radioisotope
dating data, the isochron and model ages for some or all components of each of these ten achondrites were tabulated
separately for each groupthe isochron ages and the model ages for the primitive achondrites in Tables 3 and 4
respectively, the isochron ages and the model ages for the angrites in Tables 5 and 6 respectively, the isochron ages and
the model ages for the aubrites in Tables 7 and 8 respectively, the isochron ages and the model ages for the mesosiderite
Estherville in Tables 9 and 10 respectively, and the isochron ages and the model ages for the IIE irons in Tables 11 and 12
respectively. The data in these tables were then plotted on frequency versus age histogram diagrams, with the same color
coding being used to show the ages obtained by the different radioisotope dating methodsthe isochron and the model
ages for whole-rock samples and some or all components of each of these ten achondritesthe primitive achondrites (figs.
13 and 14 respectively), the angrites (figs. 15 and 16 respectively), the aubrites (figs. 17 and 18 respectively), the
mesosiderite Estherville (figs. 19 and 20 respectively), and the IIE irons (figs. 21 and 22 respectively).
Discussion
In contrast to the Allende CV3 carbonaceous chondrite meteorite (Snelling 2014a), there have been fewer radioisotope ages
obtained for these five groups of primitive and other achondrite meteorites, because except for the angrite DOrbigny fewer
radioisotope methods have been used on them, and fewer radioisotope determinations have been undertaken. Yet the
outcome is similar to that found for the ordinary and enstatite chondrites (Snelling 2014b) and the eucrites (basaltic
achondrites) (Snelling 2014c). There is no consistent pattern evident of the -decay mode K-Ar (and Ar-Ar), Rb-Sr, Lu-Hf,
and Re-Os isochron ages increasing in that order according to the parents atomic weights or their decay rates (half-lives),
or of the -decay mode Sm-Nd isochron ages always being younger than the U-Th-Pb isochron ages according to the
parents atomic weights or their decay rates (half-lives). Such a pattern would be potentially indicative of a past episode of
accelerated radioisotope decay, as suggested by Snelling (2005c) and Vardiman, Snelling, and Chaffin (2005) from their
radioisotope investigations of earth rocks and minerals. Thus it could be concluded from these data that no accelerated
radioisotope decay event has occurred on the asteroids which parented these achondrite meteorites.
Table 3. Isochron ages for whole-rock samples and some or all components of the two primitive achondrites Acapulco and
Caddo County IAB iron, with the details and literature sources.
Sample

Method

Date

Error +/-

thirty-three analyses
of two aliquots of
plagioclases
Ar-Ar

4.507

phosphates

4.557

Note

Source

Type

0.018

Renne 2000

isochron
age

0.002

Gpel,
Manhs,
Allgre 1992 abstract

ACAPULCO (acapulcoite)

sixteen samples

sixteen samples

sixteen samples
three phosphate and
three
multigrain
fractions

Pb-Pb

207

207

Pb-206Pb 4.5565

Pb-206Pb 4.5565

0.0013

phosphatesexternal
normalization (EN)

Amelin
2005;
Wadhwa, and
2006

Amelin,
Lugmairisochron
age

0.001

Amelin
2005;
phosphatesdouble spikeWadhwa, and
(DS)
2006

Amelin,
Lugmairisochron
age

Amelin
2005;
Wadhwa, and
2006

Amelin,
Lugmairisochron
age

207

Pb-206Pb 4.55652

0.00078

207

Pb-206Pb 4.5551

0.0013

three phosphate and


three
multigrain
fractions
U-Pb
four
minerals
+
leachate + residue
(fines)
Sm-Nd

4.555

andisochron
age

0.0032

phosphatesEN+DS

Amelin and Pravdivtsevaisochron


2005
age
concordia-constrained
linear 3D isochron

Amelin and Pravdivtsevaisochron


2005
age

4.605

0.032

Prinzhofer,
Papanastassiou,
Wasserburg 1992

Mn-Cr

4.555

0.0012

Zipfel et al. 1995

Mn-Cr

4.551

0.0012

Lugmair and Shukolyukovisochron


1998
age

phosphates

I-Xe

4.557

0.002

Brazzle et al. 1999

isochron
age

feldspars

I-Xe

4.562

0.003

Brazzle et al. 1999

isochron
age

fifteen
extractions
from one sample
Ar-Ar

4.504

0.012

Bogard, Garrison,
Takeda 2005

andisochron
age

twenty-five
extractions from one
sample
Ar-Ar

4.51

0.008

Bogard
Garrison,
Takeda 2005

andisochron
age

andisochron
age
isochron
age

CADDO COUNTY (IAB iron)

twenty
extractions
from one sample
Ar-Ar
mineral
separates
plus whole rock
Rb-Sr

4.483

4.57

Bogard
Garrison,
Takeda 2005

0.012

andisochron
age

0.23

plagioclase,
diopside,
impure diopside, and whole
rock
Liu et al. 2002a

isochron
age

Liu et al. 2002a separates


plus two each more plag
and diop.
Liu et al. 2002b

isochron
age

seven
mineral
separates
(plag.,
diop.) plus whole rock Rb-Sr

4.52

0.03

Plag
and
cpx
separates
(silicate
inclusion)
Sm-Nd

4.41

0.03

Stewart, Papanastassiou,isochron
and Wasserburg 1993
age

three samples of cpx


and plag.
Sm-Nd

4.53

0.02

Stewart, Papanastassiou,isochron
and Wasserburg 1996
age

mineral
separates
plus whole rock
Sm-Nd

4.5

0.04

eight extractions from


one sample
I-Xe

4.5579

0.0001

plagioclase,
diopside,
impure diopside, and whole
rock
Liu et al. 2002a
Bogard, Garrison,
Takeda 2005

isochron
age
andisochron
age

Table 4. Model ages for whole-rock samples and some or all components of the two primitive achondrites Acapulco and
Caddo County IAB iron, with the details and literature sources.
Sample

Method

Date

Error +/- Note

Source

Type

Ar-Ar

4.5

0.01

Bogard et al. 1993

plateau age

Ar-Ar

4.503

0.011

McCoy et al. 1996

step
age

Ar-Ar

4.51

0.02

McCoy et al. 1996

plateau age

Ar-Ar

4.514

0.016

Pellas et al. 1997

plateau age

Ar-Ar

4.554

Renne 2000

model age

Ar-Ar

4.509

0.016

Renne 2000

step heating
plateau age

Ar-Ar

4.501

0.031

Renne 2000

step heating
plateau age

Pellas et al. 1997 age, revised Ar-Ar

4.502

0.005

Trieloff,
Jessberger,
and Fini 2001
plateau age

average of McCoy et al.


(1996), Pellas et al. (1997),
and Renne (2000) revised
Ar-Ar

4.504

0.003

Trieloff,
Jessberger,
and Fini 2001
plateau age

updated decay constants

Ar-Ar

4.518

0.01

Bogard 2011

model age

K-Ar

4.7

0.3

Palme et al. 1981

model age

Pb-Pb

4.5554

0.0052

Zhou et al. 2012

model age

weighted mean of five oldest


of twelve samples
U-Th/He

4.538

0.032

Min et al. 2003

model age

weighted mean of three oldest


of twelve samples
U-Th/He

4.576

0.03

Min et al. 2003

model age

merrilite

Pu-Xe

4.41

0.016

Pellas et al. 1997

model age

apatite

Pu-Xe

4.4

0.016

Pellas et al. 1997

model age

Ar-Ar

4.52

0.005

average

Takeda et al. 2000

plateau age

fifteen extractions from one


sample
Ar-Ar

4.508

0.013

(or 4.528)

Bogard et al. 2005

plateau age

twenty-five extractions
one sample

4.506

0.01

(or 4.527)

Bogard et al. 2005

plateau age

ACAPULCO
six extractions

two plagioclases

apatite
grains
measurements)

heating

(73
weighted mean

CADDO COUNTY (IAB iron)


mean value
extractions

of

seventeen

from
Ar-Ar

twenty extractions from one


sample
Ar-Ar

4.489

Ar-Ar

4.528

Ar-Ar

4.527

0.023

(or 4.487)

Bogard et al. 2005

plateau age

0.013

using
latest
decay
constants on Bogard et
al. 2005 ages
Vogel and Renne 2008 plateau age

0.01

using
latest
decay
constants on Bogard et
al. 2005 ages
Vogel and Renne 2008 plateau age
using
latest
decay
constants on Bogard et
al. 2005 ages
Vogel and Renne 2008 plateau age

Ar-Ar

4.487

0.023

Ar-Ar

4.472

0.02

Vogel and Renne 2008 plateau age

Ar-Ar

4.542

0.02

Vogel and Renne 2008 plateau age

Ar-Ar

4.552

0.02

Vogel and Renne 2008 plateau age

Ar-Ar

4.552

0.02

Vogel and Renne 2008 plateau age

Ar-Ar

4.532

0.02

Vogel and Renne 2008 plateau age

Ar-Ar

4.562

0.02

Vogel and Renne 2008 plateau age

mineral separates

Rb-Sr

4.55

Liu et al. 2002a

model age

mineral separates

Rb-Sr

4.59

Liu et al. 2002a

model age

4.561

Markowski et al. 2006


in Vogel and Renne
2008
model age

plagioclase separates
silicate inclusion

from

Hf-W

0.002

Table 5. Isochron ages for whole-rock samples and some or all components of the two angrite achondrites Angra dos Reis
and DOrbigny, with the details and literature sources.
Sample

Method

Reading

Err +/-

six total meteorite, a pyroxene, and


two whitlockite samples
Rb-Sr

4.47

0.24

two whole rock samples plotted with


one DOrbigny whole rock sample
Rb-Sr

4.56

Note

Source

Type

Wasserburg et al. 1977

isochron
age

ANGRA DOS REIS (ADOR)

three samples plotted


Severin phosphates

with

Hans,
Kleine,
Bourdon 2010

andisochron
age

Pb-Pb

4.5566

0.0002

Kleine et al. 2012

isochron
age

207

Pb-204Pb

4.558

0.006

Tera and Carlson 1999

isochron
age

207

Pb-206Pb

4.555

0.005

Tatsumoto, Knight, andisochron


Allgre 1973
age

207

Pb-206Pb

4.551

0.004

Chen
1981

4.5586

0.0006

Zartman, Jagoutz,
Bowring 2006

St.

isochron
with
several
other
meteorites and Canyon Diablo
two samples on isochron wih 11
Alende and two St Severin samples

pyroxene and plagioclase separates -207Pb-206Pb

and

Wasserburgisochron
age
andisochron
age

pyroxene fractions 1R-3R + 5R

207

Pb-206Pb

4.5571

0.0036

Amelin 2008

isochron
age

all five pyroxene fractions

207

Pb-206Pb

4.55766

0.00013

Amelin 2008

isochron
age

pyroxene separate plotted with four


mined separates from another
meteorite

207

Pb/206PbPb/206Pb

4.5578

0.00042

isochron
Lugmair and Galer 1992 age

pyroxene fractions 1R-3R + 5R + 5R


wash
U-Pb

4.5576

0.00015

Amelin 2008

isochron
age

all five pyroxene fractions + 4R and


5R washes
U-Pb

4.5571

0.0005

Amelin 2008

isochron
age

three phosphate and four pyroxene


samples
Sm-Nd

4.55

0.04

isochron
Lugmair and Marti 1977 age

four total rock, four whitlockite, and Sm-Nd

4.564

0.037

Jacobsen

204

andisochron

one pyroxene analeses


whole rock, pyroxene, and mixed
fractions
Hf-W

Wasserburg 1984

age

4.556

0.0009

Kleine et al. 2012

isochron
age

plagioclase separates, relative to


Allende CAIs
Al-Mg

4.5603

0.0004

Nyquist et al. 2003

isochron
age

whole rock, olivene, pyroxene and


plagioclase separates (relative to
LEW 86010)
Al-Mg

4.5627

0.0011

Spivak-Birndorf, Wadhwa,isochron
and Janney 2005
age

whole rock, olivene, pyroxene and


plagioclase (2) separates, relative to
E60 Pb-Pb
Al-Mg

4.56242

0.00029

Spivak-Birndorf, Wadhwa,isochron
and Janney 2009
age

whole rock, olivene, pyroxene and


plagioclase (2) separates, relative to
CAI Pb-Pb
Al-Mg

4.5625

0.0003

Schiller,
Baker,
Bizzarro 2010

andisochron
age

Spivak-Birndorf,
Wadhwa,
and
Janney (2009) data relative to
Allende CAI Pb-Pb
Al-Mg

4.5628

0.0004

Schiller,
Baker
Bizzarro 2010

andisochron
age

six separates (whole rock


pyroxene); relative to CAIs

Hf-W

4.5624

0.0013

Markowski et al. 2007

relative to E60 CAI Pb-Pb using


Markowski et al (2007) data
Hf-W

4.5631

0.0008

Spivak-Birndorf, Wadhwa,isochron
and Janney 2009
age

Markowski et al. (2007) data relative


to Allende CAI Pb-Pb
Hf-W

4.5637

0.001

Schiller,
Baker,
Bizzarro 2010

relative to Brennecka and Wadhwa


(2011) Pb-Pb age: whole rock (2);
pyroxene (3); and olivene separates
(1)
Hf-W

4.5634

0.0003

Kleine et al. 2012

isochron
age

recalibration of Kleine et al. (2012)


value relative to Connelly et al
(2012) Pb-Pb age of CAIs
Hf-W

4.5628

0.0006

Kruijer et al. 2014

isochron
age

whole rock, pyroxene (2) fractions,


relative to LEW 86010
Mn-Cr

4.5616

0.0005

Nyquist et al. 2003

isochron
age

whole rock, chromite and two glass


fractions
Mn-Cr

4.5622

0.0006

Glavin, Jagoutz,
Lugmair 2003

andisochron
age

eleven
fractionstotal
rock,
silicates, spinel (2), olivine (4),
pyroxene and glass (2) relative to
LEW 86010
Mn-Cr

4.5629

0.0006

Glavin et al. 2004

whole rock, olivine, pyroxene and


plagioclase separates (relative to
LEW 86010)
Mn-Cr

4.5625

0.0005

Spivak-Birndorf, Wadhwa,isochron
and Janney 2005
age

relative to LEW 86010

Mn-Cr

4.5622

0.0005

Sugiura, Miyazaki, andisochron


Yanai 2005
age

whole rock

Mn-Cr

4.5632

0.0006

Shukolyukov and Lugmairisochron


2007
age

relative to LEW 86010 Pb-Pb and


Glavin et al. (2004) data
Mn-Cr

4.5637

0.0004

Spivak-Birndorf, Wadhwa,isochron
and Janney 2009
age

relative to LEW 86010 Pb-Pb and


Nyquist et al. (2003) and Sugiura,
Miyazaki, and Yanai (2005) data
Mn-Cr

4.5629

0.0007

Spivak-Birndorf, Wadhwa,isochron
and Janney 2009
age

Sugiura, Miyazaki, and Yanai (2005)


data relative to Allende CAI Pb-Pb
Mn-Cr

4.5638

0.0007

Schiller,
Baker,
Bizzarro 2010

andisochron
age

Glavin et al. (2004) data relative to


Allende CAI Pb-Pb
Mn-Cr

4.5646

0.0003

Schiller,
Baker,
Bizzarro 2010

andisochron
age

four-step dissolution of whole rock


(minus drusy pyroxenes)
Pb-Pb

4.563

0.0025

Zartman, Jagoutz,
Bowring 2006

andisochron
age

pyroxene (5) and whole rock (3) Pb-Pb

4.56463

0.00028

Amelin 2007

DORBIGNY

and

isochron
age

andisochron
age

isochron
age

isochron

separates

age

pyroxene (4) and whole rock (5)


separates
Pb-Pb

4.56424

0.00029

Amelin 2008

isochron
age

pyroxene (4) and whole rock (6)


separates
Pb-Pb

4.56453

0.0002

Amelin 2008

isochron
age

plagioclase separate plotted with


eleven other samples, relative to
LEW 86010
Rb-Sr

4.56

Nyquist et al. 2003

isochron
age

one whole rock sample plotted with


two ADOR whole rock samples
Rb-Sr

4.56

Hans,
Kleine,
Bourdon 2010

whole rock sample plotted with


whole rock samples of four other
angrites
Lu-Hf

4.576

0.049

Amelin, Wimpenny, andisochron


Yin 2011
age

four mineral fractions and two whole


rock samples
Lu-Hf

4.52

0.097

Sanborn, Carlson,
Wadhwa 2012

whole rock with mafic minerals (2)

4.6

0.07

Nyquist et al. 2003

pyroxene and plagioclase separates Sm-Nd

3.08

0.05

Tonui,
Ngo,
andisochron
Papanastassiou 2003
age

whole
rock,
pyroxene
plagioclase fractions

4.507

0.089

Sanborn Carlson,
Wadhwa 2011

pyroxene (5) and whole rock (3)


separates (3D-linear)
U-Pb

4.565

0.0023

Amelin 2007

isochron
age

pyroxene (4) and whole rock (5)


separates
U-Pb

4.56464

0.00027

Amelin 2008

isochron
age

Sm-Nd

and
Sm-Nd

andisochron
age

andisochron
age
isochron
age

andisochron
age

Table 6. Model ages for whole-rock samples and some or all components of the two angrite achondrites Angra dos Reis and
DOrbingy, with the details and literature sources.
Sample

Method

Reading

Error +/-

Note

Source

Type

ANGRA DOS REIS (ADOR)


207

Pb-206Pb 4.546

0.001

Wasserburg
1977

et

total meteorite

al.model
age

207

Pb-206Pb 4.543

0.001

Wasserburg
1977

et

whitlockite-A

al.model
age

207

Pb-206Pb 4.544

0.001

Wasserburg
1977

et

whitlockite-B

al.model
age

207

Pb-206Pb 4.54

Wasserburg
1977

et

whitlockite-C

al.model
age

whitlockite

207

Pb-206Pb 4.553

0.008

Chen and Wasserburgmodel


1981
age

whole rock

207

Pb-206Pb 4.551

0.004

Chen and Wasserburgmodel


1981
age

pyroxene separate

207

Pb-206Pb 4.5578

0.00042

Lugmair
1992

207

Pb-206Pb 4.55765

0.00013

Amelin 2008

model
age

207

Pb-206Pb 4.55768

0.00013

Amelin 2008

model
age

fraction 1R, Px

207

Pb-206Pb 4.5577

0.0003

Amelin 2008

model
age

fraction 1R, Px

207

Pb-206Pb 4.5577

0.0003

Amelin 2008

model
age

fraction 2R, Px

207

Pb-206Pb 4.5577

0.0003

Amelin 2008

model
age

fraction 2R, Px

207

Pb-206Pb 4.5577

0.0003

Amelin 2008

model
age

fraction 3R, Px

207

Pb-206Pb 4.5575

0.0003

Amelin 2008

model

five
pyroxene
fractionsweighted
average, primordial Pb
five
pyroxene
fractionsweighted
average,204Pb analytical

and

Galermodel
age

age
fraction 3R, Px

207

Pb-206Pb 4.5575

0.0003

Amelin 2008

model
age

fraction 4R, Px

207

Pb-206Pb 4.5567

0.0003

Amelin 2008

model
age

fraction 4R, Px

207

Pb-206Pb 4.5587

0.0003

Amelin 2008

model
age

fraction 5R, Px

207

Pb-206Pb 4.5578

0.0003

Amelin 2008

model
age

fraction 5R, Px

207

Pb-206Pb 4.5578

0.0003

Amelin 2008

model
age

4W3, Px Wash-3

207

Pb-206Pb 4.5565

Amelin 2008

model
age

5W3, Px Wash-3

207

Pb-206Pb 4.5588

Amelin 2008

model
age

using 2012 U isotope compositions

Pb-Pb

4.5566

0.00026

Brennecka
Wadhwa 2012

206

Pb-238U

4.63

0.07

Wasserburg
1977

et

total meteorite

al.model
age

206

Pb-238U

4.61

0.07

Wasserburg
1977

et

whitlockite-A

al.model
age

206

Pb-238U

4.546

0.038

Wasserburg
1977

et

whitlockite-B

al.model
age

206

Pb-238U

4.5

Wasserburg
1977

et

whitlockite-C

al.model
age

whitlockite

206

Pb-238U

4.5

0.2

Chen and Wasserburgmodel


1981
age

whole rock

206

Pb-238U

4.55

0.06

Chen and Wasserburgmodel


1981
age

pyroxene separate

206

Pb-238U

4.586

0.015

Lugmair
1992

fraction 1R, Px

206

Pb-238U

4.5547

Amelin 2008

model
age

fraction 2R, Px

206

Pb-238U

4.553

Amelin 2008

model
age

fraction 3R, Px

206

Pb-238U

4.556

Amelin 2008

model
age

fraction 4R, Px

206

Pb-238U

4.5685

Amelin 2008

model
age

fraction 5R, Px

206

Pb-238U

4.3984

Amelin 2008

model
age

4W3, Px Wash-3

206

Pb-238U

5.5349

Amelin 2008

model
age

5W3, Px Wash-3

206

Pb-238U

5.6581

Amelin 2008

model
age

208

Pb-232Th 4.53

0.1

Wasserburg
1977

et

total meteorite

al.model
age

208

Pb-232Th 4.6

0.11

Wasserburg
1977

et

whitlockite-A

al.model
age

208

Pb-232Th 4.54

0.053

Wasserburg
1977

et

whitlockite-B

al.model
age

208

Pb-232Th 4.54

Wasserburg
1977

et

whitlockite-C

al.model
age

pyroxene separate

235

U-207Pb

4.566

0.0047

Lugmair
1992

andmodel
age

and

and

Galermodel
age

Galermodel
age

DORBIGNY
whole rock

Al-Mg

4.56601

0.00025

Baker et al. 2005

model
age

relative to CAI Pb-Pb

Al-Mg

4.56375

0.0004

Schiller, Baker,
Bizzarro 2010

andmodel
age

relative to DOrbigny Pb-Pb

Al-Mg

4.5656

0.0004

Schiller, Baker,
Bizzarro 2010

andmodel
age

whole rock

Pb-Pb

4.51808

0.00009

Baker et al. 2005

model
age

hand-picked druse pyroxene fractions


subjected to leaches of differing
severity
Pb-Pb

4.5632

0.0011

Zartman, Jagoutz, andmodel


Bowring 2006
age

hand-picked druse pyroxene fractions


subjected to leaches of differing
severity
Pb-Pb

4.5643

0.0008

Zartman, Jagoutz, andmodel


Bowring 2006
age

hand-picked druse pyroxene fractions


subjected to leaches of differing
severity
Pb-Pb

4.5638

0.0008

Zartman, Jagoutz, andmodel


Bowring 2006
age

hand-picked druse pyroxene fractions


subjected to leaches of differing
severity
Pb-Pb

4.5694

0.0028

Zartman, Jagoutz, andmodel


Bowring 2006
age

hand-picked druse pyroxene fractions


subjected to leaches of differing
severity
Pb-Pb

4.5585

0.0033

Zartman, Jagoutz, andmodel


Bowring 2006
age

weighted mean of the first three handpicked druse pyroxene fractions


(above) with two other designated as
aberrant
Pb-Pb

4.5639

0.0006

Zartman, Jagoutz, andmodel


Bowring 2006
age

4.563

0.001

Spivak-Birndorf,
Wadhwa, and Janneymodel
2005
age

pyroxene (5) and whole rock (3)


separates (weighted average)
Pb-Pb

4.56448

0.00024

Amelin 2007

model
age

pyroxene (5) and whole rock (8)


separates (weighted average assuming
primordial Pb)
Pb-Pb

4.56442

0.00012

Amelin 2008

model
age

pyroxene (5) and whole rock (8)


separates (weighted average assuming
analytical Pb)
Pb-Pb

4.56453

0.00019

Amelin 2008

model
age

recalculated Amelin (2008) weighted


average (primordial Pb) due to
new 238U-235U ratio value
Pb-Pb

4.5638

0.0004

model
Brennecka et al. 2010 age

Pb-Pb

4.56336

0.00034

Bouvier and Wadhwamodel


2010
age

Pb-Pb

4.56334

0.0003

Brennecka
Wadhwa 2011

andmodel
age

Using 2012 U isotope compositions

Pb-Pb

4.56337

0.00025

Brennecka
Wadhwa 2012

andmodel
age

druse pyroxene fraction

207

Pb-206Pb 4.559

0.0011

Jagoutz et al. 2002

model
age

groundmass pyroxene fraction

207

Pb-206Pb 4.548

0.0011

Jagoutz et al. 2002

model
age

weighted mean

207

Pb-206Pb 4.557

0.0015

Jagoutz et al. 2002

model
age

matrix pyroxene fraction

207

Pb-206Pb 4.549

0.002

Jagoutz et al. 2003

model
age

matrix pyroxene fraction

207

Pb-206Pb 4.557

0.002

Jagoutz et al. 2003

model
age

Pb-Pb

corrected for
determination

new 238U-235U

ratio

druse pyroxene fraction

207

Pb-206Pb 4.5554

0.0019

Jagoutz et al. 2003

model
age

druse pyroxene fraction

207

Pb-206Pb 4.556

0.004

Jagourtz et al. 2003

model
age

druse pyroxene fraction

207

Pb-206Pb 4.557

0.001

Jagourtz et al. 2003

model
age

12R Px

207

Pb-206Pb 4.5666

0.0014

assuming
primordial Pb

Amelin 2008

model
age

13R Px

207

Pb-206Pb 4.5649

0.0006

assuming
primordial Pb

Amelin 2008

model
age

15 R WR (fragment)

207

Pb-206Pb 4.5646

0.0004

assuming
primordial Pb

Amelin 2008

model
age

16R WR (fines)

207

Pb-206Pb 4.5621

0.0004

assuming
primordial Pb

Amelin 2008

model
age

17R Px

207

Pb-206Pb 4.5647

0.0004

assuming
primordial Pb

Amelin 2008

model
age

18R Px

207

Pb-206Pb 4.5642

0.0003

assuming
primordial Pb

Amelin 2008

model
age

19R WR minus Px

207

Pb-206Pb 4.5645

0.0003

assuming
primordial Pb

Amelin 2008

model
age

20R WR (rel. fine)

207

Pb-206Pb 4.5644

0.0003

assuming
primordial Pb

Amelin 2008

model
age

22R WR

207

Pb-206Pb 4.57

0.0015

assuming
primordial Pb

Amelin 2008

model
age

23R WR

207

Pb-206Pb 4.5636

0.0013

assuming
primordial Pb

Amelin 2008

model
age

21W3 Px

207

Pb-206Pb 4.5644

0.0004

assuming
primordial Pb

Amelin 2008

model
age

22W3 WR

207

Pb-206Pb 4.5643

0.0003

assuming
primordial Pb

Amelin 2008

model
age

22W3 WR

207

Pb-206Pb 4.5644

0.0004

assuming
primordial Pb

Amelin 2008

model
age

12R Px

207

Pb-206Pb 4.5673

0.0014

assuming
analytical Pb

Amelin 2008

model
age

13R Px

207

Pb-206Pb 4.5649

0.0006

assuming
analytical Pb

Amelin 2008

model
age

15 R WR (fragment)

207

Pb-206Pb 4.5648

0.0004

assuming
analytical Pb

Amelin 2008

model
age

16R WR (fines)

207

Pb-206Pb 4.5642

0.0004

assuming
analytical Pb

Amelin 2008

model
age

17R Px

207

Pb-206Pb 4.5648

0.0004

assuming
analytical Pb

Amelin 2008

model
age

18R Px

207

Pb-206Pb 4.5642

0.0003

assuming
analytical Pb

Amelin 2008

model
age

19R WR minus Px

207

Pb-206Pb 4.5645

0.0003

assuming
analytical Pb

Amelin 2008

model
age

20R WR (rel. fine)

207

Pb-206Pb 4.5646

0.0003

assuming
analytical Pb

Amelin 2008

model
age

22R WR

207

Pb-206Pb 4.5741

0.0015

assuming
analytical Pb

Amelin 2008

model
age

23R WR

207

Pb-206Pb 4.5642

0.0013

assuming
analytical Pb

Amelin 2008

model
age

21W3 Px

207

Pb-206Pb 4.5645

0.0004

assuming
analytical Pb

Amelin 2008

model
age

22W3 WR

207

Pb-206Pb 4.5643

0.0003

assuming
analytical Pb

Amelin 2008

model
age

assuming
analytical Pb

Amelin 2008

model
age

4.4916

assuming
primordial Pb

Amelin 2008

model
age

Pb-238U

4.6133

assuming
primordial Pb

Amelin 2008

model
age

206

Pb-238U

4.8416

assuming
primordial Pb

Amelin 2008

model
age

16R WR (fines)

206

Pb-238U

4.8929

assuming
primordial Pb

Amelin 2008

model
age

17R Px

206

Pb-238U

4.6946

assuming
primordial Pb

Amelin 2008

model
age

18R Px

206

Pb-238U

4.9719

assuming
primordial Pb

Amelin 2008

model
age

19R WR minus Px

206

Pb-238U

5.0804

assuming
primordial Pb

Amelin 2008

model
age

20R WR (rel. fine)

206

Pb-238U

4.9574

assuming
primordial Pb

Amelin 2008

model
age

22R WR

206

Pb-238U

3.6722

assuming
primordial Pb

Amelin 2008

model
age

23R WR

206

Pb-238U

2.8337

assuming
primordial Pb

Amelin 2008

model
age

21W3 Px

206

Pb-238U

5.471

assuming
primordial Pb

Amelin 2008

model
age

22W3 WR

206

Pb-238U

5.2906

assuming
primordial Pb

Amelin 2008

model
age

22W3 WR

206

Pb-238U

5.5678

assuming
primordial Pb

Amelin 2008

model
age

22W3 WR

207

Pb-206Pb 4.5644

12R Px

206

Pb-238U

13R Px

206

15 R WR (fragment)

0.0004

Table 7. Isochron ages for whole-rock samples and some or all components of the two aubrite achondrites Norton County
and Shallowater, with the details and literature sources.
Sample

Method

Reading

Error +/- Note

Source

Type

eight fractionsthree breccias and four splits,


three enstatites
Rb-Sr

4.7

0.1

Bogard et al. 1967a

isochron age

samples of feldspar (2), pyroxenes (4), matrix


(2), and whole rock (1)
Rb-Sr

4.48

0.04

Minster
1976

isochron age

whole rock and feldspar (2) samples plotted


with eight samples from four other meteorites Rb-Sr

4.516

0.029

Minster, Rickard, and


Allgre 1979
isochron age

revised decay constant

Rb-Sr

4.44

0.04

Bogard, Dixon,
Garrison 2010

Pb-Pb

4.55

0.003

Huey
1973

Ar-Ar

4.55

0.05

McCoy et al. 1995

NORTON COUNTY

and

and

Allgre

and
isochron age

Kohman
isochron age

SHALLOWATER
isochron age

Table 8. Model ages for whole-rock samples and some or all components of the two aubrite achondrites Norton County and
Shallowater, with the details and literature sources.
Sample

Method

Reading

Error +/- Note

Source

Type

K-Ar

4.4

0.65

Geiss and Hess 1958

model age

5.09

Kirsten,
Krankowsky,
Zhringer 1963

and

K-Ar

4.68

Kirsten,
Krankowsky,
Zhringer 1963

and

K-Ar
K-Ar

4.5

Bogard et al. 1967a

NORTON COUNTY

523.3X breccia IV

model age
model age
model age

4965F breccia, split A

K-Ar

4.2

Bogard et al. 1967a

model age

4965F breccia, split B

K-Ar

4.3

Bogard et al. 1967a

model age

4965F breccia, enstatite

K-Ar

4.5

Bogard et al. 1967a

model age

Ar-Ar

4.454

extractions releasing up to
87% of Ar from sample NCOkada
Ar-Ar

4.676

Bogard, Dixon, and Garrisonaverage spectrum


2010
age

summed across all extractions


from sample NC-Okada
Ar-Ar

4.198

Bogard, Dixon, and Garrison


2010
total age

samples NC-15961,
celsius extraction

1300
0.018

Bogard, Dixon, and Garrisonmaximum


2010
spectrum age

Pb-Pb

4.55

0.003

Huey and Kohman 1973

model age

Ar-Ar

4.53

0.05

McCoy et al. 1995

total age

five extractions releasing 6


45% Ar
Ar-Ar

4.539

0.003

McCoy et al. 1995

plateau age

Ar-Ar

4.535

0.02

Bogard 2011

plateau age

K-Ar

4.53

0.03

McCoy et al. 1995

model age

I-Xe

4.566

0.002

Brazzle et al. 1999

absolute age

I-Xe

4.5633

0.0004

Gilmour et al. 2006

closure age

I-Xe

4.5623

0.0004

Gilmour et al. 2009; Hohenberg


and Pravdivtseva 2008
closure age

SHALLOWATER

Table 9. Isochron ages for whole-rock samples and some or all components of the mesosiderite achondrite Estherville, with
the details and literature sources.
Sample

Method

Reading

Error +/- Note

Source

Type

Ar-Ar

3.55

0.096

Murthy, Alexander, and Saito 1978 isochron age

Rb-Sr

4.411

0.088

Rb-Sr

4.26

0.1

Murthy, Coscio, and Sabelin 1977 isochron age

Rb-Sr

4.24

0.03

Murthy, Alexander, and Saito 1978 isochron age

Rb-Sr

4.542

0.203

Sets A and C

Re-Os

4.25

0.5

one sample fits with 21 other


meteorites
Luck and Allgre 1983

Re-Os

4.6

0.05

fits to this isochron

Shen,
Papanastassiou,
Wasserburg 1998

Pb-Pb

4.556

0.035

Set A

Brouxel and Tatsumoto 1990

isochron age

Pb-Pb

4.553

0.035

Set Bfourteen analyses

Brouxel and Tatsumoto 1991

isochron age

Sm-Nd

4.533

0.094

Sets A and C

Brouxel and Tatsumoto 1991

isochron age

ESTHERVILLE

nine
separates

eleven
separates

one sample fits with five


other meteorites
Cumming 1969

isochron age

Brouxel and Tatsumoto 1991

isochron age
isochron age
and
isochron age

Table 10. Model ages for whole-rock samples and some or all components of the mesosiderite achondrite Estherville, with
the details and literature sources.
Sample Method

Reading

Error +/- Note

Source

Type

Bogard et al. 1990

plateau age

ESTHERVILLE
Ar-Ar

3.62

Ar-Ar

3.94

0.1

two samples averaged with 18 otherBogard


meteorite samples
1998

U-Pb

4.571

0.018

Set A

Brouxel and Tatsumoto


1990
concordia age

U-Pb

4.569

0.061

Set Bfourteen analyses

Brouxel and Tatsumoto


1991
concordia age

and

Garrison
plateau age

Table 11. Isochron ages for whole-rock samples and some or all components of the three iron IIE achondrites Colomera,
Kodakanal, and Weekeroo Station, with the details and literature sources.
Sample

Method

Reading

Error +/- Note

Source

Type

sixteen inclusions (feldspar, diopside, glass)

Rb-Sr

4.61

0.04

Sanz, Burnett, and


Wasserburg 1970
isochron age

fifteen inclusions only

Rb-Sr

4.64

0.14

Sanz, Burnett, and


Wasserburg 1970
isochron age

twenty inclusions (feldspar, diopside, glass)

Rb-Sr

4.64

0.04

Sanz, Burnett, and


Wasserburg 1970
isochron age

revised decay constant

Rb-Sr

4.51

0.04

Bogard, Garrison, and


McCoy 2000
isochron age

revised decay constant

Rb-Sr

4.59

0.127

Snyder et al. 2001

revised decay constant

Rb-Sr

3.73

0.1

Gpel, Manhs, and


Allgre 1985
isochron age

revised decay constant

Rb-Sr

3.81

0.1

Gpel, Manhs, and


Allgre 1985
isochron age

eleven density and mineral separates of


Burnett and Wasserburg 1967
Rb-Sr

3.743

0.044

Snyder et al. 2001

four samples plotted; five samples from three


other meteorites
Re-Os

4.624

0.017

Birck and Allgre 1998 isochron age

207

Pb-206Pb 3.676

0.003

Gpel, Manhs, and


Allgre 1985
isochron age

207

Pb-206Pb 3.675

0.003

Gpel, Manhs, and


Allgre 1985
isochron age

238

U-206Pb

3.68

0.023

Gpel, Manhs, and


Allgre 1985
isochron age

238

U-206Pb

3.679

0.005

Gpel, Manhs, and


Allgre 1985
isochron age

238

U-206Pb

3.684

0.022

Gpel, Manhs, and


Allgre 1985
isochron age

235

U-207Pb

3.677

0.009

Gpel, Manhs, and


Allgre 1985
isochron age

235

U-207Pb

3.678

0.009

Gpel, Manhs, and


Allgre 1985
isochron age

Rb-Sr

4.37

0.2

Burnett
and
Wasserburg 1967
isochron age

Rb-Sr

4.39

0.07

Evensen et al. 1979

Rb-Sr

4.28

0.23

Bogard, Garrison, and


McCoy 2000
isochron age

Rb-Sr

4.64

0.04

Snyder et al. 2001

isochron age

Rb-Sr

4.35

0.07

Snyder et al. 2001

isochron age

Rb-Sr

4.33

0.23

Snyder et al. 2001

isochron age

Sm-Nd

0.7

Snyder et al. 2001

isochron age

Hf-W

4.555

Schulz et al. 2012

isochron age

COLOMERA

isochron age

KODAKANAL

one clinopyroxene and two alkali-rich glass


samples extracted from the meteorite (and
leached, as well as the leachates)
one clinopyroxene and two alkali-rich glass
samples extracted from the meteorite (and
leached, as well as the leachates)
one clinopyroxene and two alkali-rich glass
samples extracted from the meteorite (and
leached, as well as the leachates)
one clinopyroxene and two alkali-rich glass
samples extracted from the meteorite (and
leached, as well as the leachates)
one clinopyroxene and two alkali-rich glass
samples extracted from the meteorite (and
leached, as well as the leachates)
one clinopyroxene and two alkali-rich glass
samples extracted from the meteorite (and
leached, as well as the leachates)
one clinopyroxene and two alkali-rich glass
samples extracted from the meteorite (and
leached, as well as the leachates)

isochron age

WEEKEROO STATION
nine fractions plotted

isochron age

Table 12. Model ages for whole-rock samples and some or all components of the three iron IIE achondrites Colomera,
Kodakanal, and Weekeroo Station, with the details and literature sources.

Sample

Method

Reading

Error +/- Note

Source

Type

ten extractions releasing 2572%


Ar from feldspar sample
Ar-Ar

4.47

0.01

Bogard, Garrison, and McCoy


2000
plateau age

all extractions above 25% Ar


release from feldspar
Ar-Ar

4.48

0.01

Bogard, Garrison, and McCoyaverage spectrum


2000
age

seven extractions above 75% Ar


release from feldspar
Ar-Ar

4.5

0.02

Bogard, Garrison, and McCoyaverage spectrum


2000
age

Ar-Ar

4.451

0.017

Bogard, Garrison, and McCoystep


2000
age

extraction

feldspar sample 400C

Ar-Ar

3.194

0.025

Bogard, Garrison, and McCoystep


2000
age

extraction

500C

Ar-Ar

3.499

0.013

Bogard, Garrison, and McCoystep


2000
age

extraction

550C

Ar-Ar

3.852

0.007

Bogard, Garrison, and McCoystep


2000
age

extraction

600C

Ar-Ar

4.021

0.006

Bogard, Garrison, and McCoystep


2000
age

extraction

625C

Ar-Ar

4.11

0.004

Bogard, Garrison, and McCoystep


2000
age

extraction

650C

Ar-Ar

4.142

0.002

Bogard, Garrison, and McCoystep


2000
age

extraction

675C

Ar-Ar

4.323

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

705C

Ar-Ar

4.387

0.002

Bogard, Garrison, and McCoystep


2000
age

extraction

725C

Ar-Ar

4.403

0.002

Bogard, Garrison, and McCoystep


2000
age

extraction

750C

Ar-Ar

4.437

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

775C

Ar-Ar

4.456

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

800C

Ar-Ar

4.464

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

815C

Ar-Ar

4.465

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

830C

Ar-Ar

4.465

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

850C

Ar-Ar

4.469

0.002

Bogard, Garrison, and McCoystep


2000
age

extraction

870C

Ar-Ar

4.473

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

890C

Ar-Ar

4.479

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

915C

Ar-Ar

4.48

0.002

Bogard, Garrison, and McCoystep


2000
age

extraction

940C

Ar-Ar

4.473

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

965C

Ar-Ar

4.465

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

1000C

Ar-Ar

4.468

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

1050C

Ar-Ar

4.482

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

1100C
1125C

Ar-Ar

4.488

0.001

Bogard, Garrison, and McCoystep

extraction

COLOMERA

2000

age

Ar-Ar

4.499

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

1150C

Ar-Ar

4.501

0.002

Bogard, Garrison, and McCoystep


2000
age

extraction

1175C

Ar-Ar

4.507

0.002

Bogard, Garrison, and McCoystep


2000
age

extraction

1200C

Ar-Ar

4.51

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

1250C

Ar-Ar

4.507

0.001

Bogard, Garrison, and McCoystep


2000
age

extraction

1300C

Ar-Ar

4.485

0.009

Bogard, Garrison, and McCoystep


2000
age

extraction

1350C

Ar-Ar

4.552

0.015

Bogard, Garrison, and McCoystep


2000
age

extraction

1450C

Ar-Ar

5.164

0.052

Bogard, Garrison, and McCoystep


2000
age

extraction

1550C
five analyses of one sample

Rb-Sr

4.7

Burnett and Wasserburg 1967 model age

six analyses of another sample

Rb-Sr

3.8

Burnett and Wasserburg 1967 model age

four analyses of another sample

Rb-Sr

3.1

Burnett and Wasserburg 1967 model age

two samples

K-Ar

3.53

0.1

Bogard et al 1967b

K-Ar

2.47

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (Me943G)

K-Ar

3.09

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (P1159-04)

K-Ar

3.38

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (P1159-04)

K-Ar

3.32

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (Me574B)

K-Ar

3.27

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (Me574B)

K-Ar

3.59

0.1

Bogard,
Burnett,
Wasserburg 1969

and

feldspar (Me943-1)

Rb-Sr

3.83

0.1

Bogard,
Burnett,
Wasserburg 1969

and

mixed phases (Me574A)

Rb-Sr

3.74

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (Me574B)

Rb-Sr

3.8

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (P1159-04)

Rb-Sr

3.82

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (P1159-04) split A

Rb-Sr

3.97

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (P1159-04) split B

Rb-Sr

3.76

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (Me943G)

Rb-Sr

3.69

0.1

Bogard,
Burnett,
Wasserburg 1969

and

glass (Me943G)

average of seven determinations


on four fractions
K-Ar

4.49

0.1

Bogard et al. 1967b

model age

Weekeroo H (oligoclase) p<2.96,


105-177
K-Ar

4.05

Bogard et al. 1967b

model age

KODAKANAL
average
age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age
model age

WEEKEROO STATION

model

K-Ar

4.52

Bogard et al. 1967b

model age

K-Ar

4.56

Bogard et al. 1967b

model age

Weekeroo H (oligoclase) p<2.96,


62-50
K-Ar

4.54

Bogard et al. 1967b

model age

Weekeroo N-B2 <150

K-Ar

4.37

Bogard et al. 1967b

model age

K-Ar

4.48

Bogard et al. 1967b

model age

K-Ar

4.36

Bogard et al. 1967b

model age

K-Ar

4.51

Bogard et al. 1967b

model age

K-Ar

4.54

0.03

Niemeyer 1980

model age

K-Ar

4.3

0.1

Bogard, Garrison, and McCoy


2000
model age

Ar-Ar

4.54

0.03

Niemeyer 1980

Ar-Ar

4.49

0.03

Bogard, Garrison, and McCoy


2000
plateau age

I-Xe

4.555

Brazzle et al 1999; Niemeyer


1980
model age

Re-Os

7.1

Niemeyer and Esser 1991;


Snyder et al. 2001
model age

Weekeroo HC <150

eight extractions

eight extractions

plateau age

The Primitive AchondritesAcapulco and Caddo County IAB Iron


The Pb-Pb and U-Pb isochron ages for Acapulco all agree on a 4.554.57 Ga age for this meteorite, with matching support
from Mn-Cr and I-Xe isochron ages (fig. 13). One Sm-Nd isochron age is older, and one Ar-Ar isochron age is younger.
Similarly, nine Ar-Ar model ages are younger by a similar amount from the one Pb-Pb model 4.554.57 Ga age (fig. 14).
However, one Ar-Ar-model age agrees with that one Pb-Pb model age, while another Ar-Ar model age is substantially older.
Additionally, two Pu-Xe model ages are younger than that one Pb-Pb model 4.55 4.57 Ga age, while two U-Th/He model
ages straddle either side of it.

Fig. 13. Frequency versus radioisotope ages histogram diagram for the isochron ages for whole-rock samples and some or
all components of the two primitive achondrites Acapulco and Caddo County IAB iron, with color coding being used to show
the ages obtained by the different radioisotope dating methods.

Fig. 14. Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock samples and some or all
components of the two primitive achondrites Acapulco and Caddo County IAB iron, with color coding being used to show the
ages obtained by the different radioisotope dating methods.
The silicates in the Caddo County IAB iron meteorite yield a main grouping of two Ar-Ar, one Rb-Sr and one Sm-Nd isochron
ages at a 4.504.525 Ga age, slightly lower than the minor grouping of one Rb-Sr and one I-Xe isochron ages at 4.554.57
Ga (fig. 13). Two other Sm-Nd and one Ar-Ar younger isochron ages are scattered below these groupings. On the other
hand, the main grouping of model ages is at 4.554.57 Ga and consists of three Ar-Ar, one Rb-Sr and one Hf-W model ages
(fig. 14). One Rb-Sr model age is older, but ten scattered Ar-Ar model ages are younger.
The AngritesAngra dos Reis and DOrbigny
Both of these angrites are strongly dated at 4.554.57 Ga, particularly by Pb-Pb isochron and model ages, but supported by
U-Pb, Sm-Nd, Rb-Sr, and Hf-W isochron ages and U-Pb model ages (Angra dos Reis), and U-Pb, Rb-Sr, Lu-Hf, Sm-Nd, MnCr, Hf-W, and Al-Mg isochron ages and Al-Mg model ages (DOrbigny) (figs. 15 and 16). There is minimal scatter of isochron
ages, with just one younger Rb-Sr isochron age for Angra dos Reis, and one older Sm-Nd isochron age and one very much
younger Sm-Nd isochron age for DOrbigny (fig. 15). There are older U-Pb and Th-Pb model ages scattered above the
strong 4.554.57 Ga age peak for Angra dos Reis, and younger Pb-Pb, U-Pb, and Th-Pb model ages scattered below it (fig.
16). On the other hand, there are numerous U-Pb model ages widely scattered either side of the strong 4.554.57 Ga age
peak for DOrbigny, and just two slightly younger Pb-Pb model ages (fig. 16).

Fig. 15. Frequency versus radioisotope ages histogram diagram for the isochron ages for whole-rock samples and some or
all components of the two angrite achondrites Angra dos Reis and DOrbigny, with color coding being used to show the ages
obtained by the different radioisotope dating methods.

Fig. 16. Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock samples and some or all
components of the angrite achondrite Angra dos Reis, with color coding being used to show the ages obtained by the
different radioisotope dating methods.

Fig. 16 (cont). Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock samples and
some or all components of the angrite achondrite DOrbigny, with color coding being used to show the ages obtained by the
different radioisotope dating methods.
The AubritesNorton County and Shallowater
While there is one Pb-Pb isochron age of 4.55 Ga for the Norton County aubrite, there are three younger Rb-Sr isochron
ages scattered below it (table 7 and fig. 17). There is only one Ar-Ar isochron age of 4.55 Ga for the Shallowater aubrite
(table 7 and fig. 17). Similarly, there is only one Pb-Pb model age of 4.55 Ga for Norton County, yet there are ten scattered
K-Ar and Ar-Ar model ages, three older and seven younger (table 8 and fig. 18). On the other hand, there is a close
grouping of model ages for Shallowater, four Ar-Ar model ages grouped at 4.534.539 Ga and three I-Xe model ages
grouped at 4.5624.566 Ga (table 8 and fig. 18).

Fig. 17. Frequency versus radioisotope ages histogram diagram for the isochron ages for whole-rock samples and some or
all components of the two aubrite achondrites Norton County and Shallowater, with color coding being used to show the
ages obtained by the different radioisotope dating methods.

Fig. 18. Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock samples and some or all
components of the two aubrite achondrites Norton County and Shallowater, with color coding being used to show the ages
obtained by the different radioisotope dating methods.
The MesosideritesEstherville
The silicates in the Estherville mesosiderite yield two Pb-Pb isochron ages grouped at 4.554.56 Ga, but their Ar-Ar, Rb-Sr,
and Sm-Nd isochron ages are younger, the Ar-Ar and some Rb-Sr isochron ages being particularly younger (table 9 and fig.
19). For comparison, the two Re-Os isochron ages yielded by the Fe-Ni metal component of this meteorite are scattered
either side of the 4.554.56 Ga Pb-Pb isochron ages for the silicates, one being much younger, similar to two much younger
Rb-Sr isochron ages. The only two U-Pb model ages are at 4.57 Ga, whereas the only two Ar-Ar model ages are very much
younger (table 10 and fig. 20).

Fig. 19. Frequency versus radioisotope ages histogram diagram for the isochron ages for whole-rock samples and some or
all components of the mesosiderite achondrite Estherville, with color coding being used to show the ages obtained by the
different radioisotope dating methods.

Fig. 20. Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock samples and some or all
components of the mesosiderite achondrite Estherville, with color coding being used to show the ages obtained by the
different radioisotope dating methods.
The IIE IronsColomera, Kodakanal and Weekeroo Station
The silicates in the Colomera and Weekeroo Station IIE irons yield a few results at the 4.554.57 Ga target age indicated
by the strong peaks and main groupings of isochron and model ages for many other meteorites (tables 11 and 12).
However, four of the five Rb-Sr isochron ages for Colomera are older than the target age, and the other is younger (fig.
21). For Weekeroo Station only the sole Hf-W isochron age matches the target age, whereas five of the six Rb-Sr isochron
ages are much younger, and the other is older, plus the one Sm-Nd isochron age is extremely young (fig. 21). There is one
Ar-Ar model age for Colomera that matches the target age, whereas many other Ar-Ar model ages are younger, and some
very much younger (fig. 22). Additionally, Colomera yields one much older Rb-Sr model age and two very much younger RbSr model ages. Weekeroo Station yields a K-Ar model age that matches the target age, supported by an I-Xe model age,
whereas apart from one much older K-Ar model age, all its other K-Ar model ages are either younger or much younger
(table 12 and fig. 22). Interestingly, the Fe-Ni metal in Weekeroo Station yields a much older Re-Os model age.

Fig. 21. Frequency versus radioisotope ages histogram diagram for the isochron ages for whole-rock samples and some or
all components of the three iron IIE achondrites Colomera, Kodakanal, and Weekeroo Station, with color coding being used
to show the ages obtained by the different radioisotope dating methods.

Fig. 22. Frequency versus radioisotope ages histogram diagram for the model ages for whole-rock samples and some or all
components of the three iron IIE achondrites Colomera, Kodakanal, and Weekeroo Station, with color coding being used to
show the ages obtained by the different radioisotope dating methods.
In stark contrast, the silicates in the Kodakanal IIE iron meteorite yield both isochron and model ages that indicate the
radioisotope systems in its silicates have been greatly disturbed, as all of this meteorites isochron and model ages are very
much younger than the 4.554.57 Ga target age (note the adjusted scales of radiometric ages for this meteorite in figs 21
and 22). Interestingly, all the U-Pb and Pb-Pb isochron ages peak at 3.6753.7 Ga, and the Rb- Sr isochron ages are
slightly older, whereas its Fe-Ni metal component yields a sole Re-Os isochron 4.62 Ga age, which might be closer to the
target age if this isochron age was recalculated using the latest Re decay constant (Birck and Allgre 1998) (table 11 and
fig. 21). The K-Ar and Rb-Sr model ages in Kodakanals silicates are all scattered between 2.74 Ga and 3.97 Ga, but the
Rb-Sr model ages are all older than the K-Ar model ages (table 12 and fig. 22).
Agreement of Ages
As noted previously by Snelling (2014a, b, c) for chondrite and eucrite meteorites, there is generally strong agreement
among most of the Pb-Pb isochron and model ages for most of these meteorites in these five different groups for a common
4.554.57 Ga age. Occasionally, where used, some U-Pb, Rb-Sr, Ar-Ar, Sm-Nd, Lu-Hf, and maybe Re-Os, isochron ages,
and some U-Pb, Rb-Sr, Ar-Ar, and K-Ar model ages match this common 4.554.57 Ga age. This is to be expected, since the
Rb, Lu, Re, Sm, and K decay constants are all ultimately calibrated against the U-Pb method and thus the U decay
constants (Dickin 2005, pp. 7071; Lugmair, Scheinin, and Marti 1975; Nebel, Scherer, and Mezger 2011; Renne et al.
2010; Selby et al. 2007; Sderlund et al. 2004). This ultimately means that all meteorite ages are dependent on the reliability
of determinations of the 238U and 235U decay constants, and the critical 238U/235U ratio. However, such calibrations are
dependent on assuming the crucial 238U/235U ratio is constant in all rocks and minerals, but recently significant variations in
this ratio have been measured in both terrestrial minerals and meteorites (Brennecka and Wadhwa 2012; Hiess et al. 2012).
It also may be significant that the estimated half-life of 238U is 4.468 Ga, which is almost identical to the claimed 4.554.57
Ga age of the earth, the asteroids, and the meteorites derived from them.
Additionally, whenever I-Xe, Hf-W, Mn-Cr, and Al-Mg isochron ages, and I-Xe, Hf-W, and Al-Mg model ages were
determined they always matched the target 4.554.57 Ga age defined by the Pb-Pb isochron and model ages. This is
because all these methods are calibrated against the 4.554.57 Ga Pb-Pb isochron and model ages of other meteorites and
their components. The I-Xe method is calibrated against the 4.557 Ga Pb-Pb model age of phosphate separates from the
Acapulco meteorite and normalized against the I-Xe determinations of the Shallowater aubrite (Bogard Garrison, and
Takeda 2005; Brazzle et al. 1999; Gpel, Manhes, and Allgre 1994). The Hf-W method is calibrated by an Hf-W isochron
obtained on the H5 chondrite St. Marguerite being calibrated against the Pb-Pb model ages of whole-rock fragments and
phosphates from St. Marguerite (Gpel, Manhes, and Allgre 1994; Kleine et al. 2002, 2004, 2005; Markowski et al. 2006),
and also by calibrating the Hf-W isochron obtained on the angrite DOrbigny against its Pb-Pb isochron age corrected for U
isotopic variations (Amelin 2008; Brennecka and Wadhwa 2011; Kleine et al. 2012). The Mn-Cr method is calibrated by an
Mn-Cr isochron obtained on the H5 chondrite St. Marguerite being calibrated its Pb-Pb model age and the Pb-Pb isochron
age of the angrite LEW 86010, or calibrated by the Mn-Cr isochrons of measured meteorites being calibrated against the
Mn-Cr isochron for the angrite LEW 86010, which is in turn calibrated against its Pb-Pb isochron age (Gpel, Manhes, and
Allgre 1994; Lugmair and Galer 1992; Lugmair and Shukolyukov 1998; Polnau and Lugmair 2001). The Al-Mg method is
calibrated against the Al-Mg age of the Ca-Al inclusions (CAIs) in the CV chondrite Efremovka E60 calibrated against their
Pb-Pb isochron age or is calibrated by the Al-Mg isochron for the angrite DOrbigny calibrated against either the Pb-Pb
isochron age of the CAIs in the Allende CV chondrite or the Pb-Pb isochron age of DOrbigny (Amelin et al. 2002; Amelin
2008; Jacobsen et al. 2008; Schiller, Baker, and Bizzarro 2010; Spivak-Birndorf, Wadhwa, and Janney 2009).
Scattering of Ages
To explain the scattering of ages either side of the 4.554.57 Ga target age appeal is usually made to the effects of thermal
disturbance of the radioisotope systems, either during the supposed cooling of the parent asteroid body from its formation in
the aftermath of the supposed solar nebula, or during subsequent fragmentation and reassembly of the parent asteroid,
and/or as a result of impact cratering of the parent asteroids (Bogard 2011; Bogard, Dixon, and Garrison 2010; Bogard et al.
1990, 1993; Bogard and Garrison 1998; Bogard, Garrison, and McCoy 2000; Bogard, Garrison, and Takeda 2005; Lewis
1997). This must be especially the case where among the meteorites studied here the scattering of ages is common in the
K-Ar, Ar-Ar, I-Xe, Pu-Xe, Rb-Sr, Sm-Nd, Re-Os, U-Th/He, and U-Th-Pb systems. Ar and He being gases are readily mobile
within and between minerals, and metamorphism is known to reset all these radioisotope systems (Faure and Mensing
2005; Snelling 2000).Thermal disturbances tend to reset the radioisotope systems so that the resultant ages are usually
lower than their original ages (Lewis 1997), which is very evident for the Kodakanal iron IIE meteorite. On the other hand,
for many of these meteorites the radioisotope ages are both younger and older than the believed formation age of the
parent bodies, determined as 4.554.57 Ga from the clustering of all the radioisotope systems at that age. Birck and Allgre
(1998) noted that the older Re-Os age for the Kodakanal iron IIE meteorite could be reconciled by simply adjusting the 187Re
decay constant to thus force the older Re-Os age to agree with the expected 4.554.57 Ga age. Indeed, Snelling (2014d, e,
2015) has documented how decay constants have repeatedly been adjusted so as to bring Rb-Sr, Lu-Hf, and Re-Os ages
into agreement with Pb-Pb and U-Pb ages, including for meteorites. Thus some of the scattering of ages of these meteorites
could be due to the decay constants used to calculate the respective ages. Nevertheless, many of the much older ages in
the U-Th-Pb isotopic system, and perhaps even the few in the K-Ar, Ar-Ar, and Rb-Sr isotopic systems, could be due to
parent radioisotope migration out of, or daughter isotope migration into, minerals and meteorites during thermal
disturbances.
Ga clustering.
At what point in time radioactive decay began is unclear from Scripture, and is still a matter of debate among creationists.
The RATE project considered the possibility of a large amount of accelerated decay occurring during the Creation Week, as
radioisotope decay was not regarded as decay in the sense of deterioration of matter (Vardiman, Snelling, and Chaffin
2005). It is instead a transmutation process, by which one element is changed into another. The daughter element is
certainly not inferior to the parent elementIn the context of the decay evident today due to the operation of the second law of
thermodynamics, Anderson (2103) contended that there is no real biblical evidence to suggest that the second law was
inoperable prior to the curse, and so argued that rather the second law was in effect from the beginning of creation. He thus
also suggested that the tendency toward entropy implicit in the second law was never of a kind that conflicted with Gods
declaration that the creation was very good, or that eventuated in the death of any sentient creature. On the other hand, it
could be argued that radioisotope decay is more than the operation of the second law of thermodynamics; the additional
outcome being the radiation produced which is harmful to lifes biological and chemical makeup. Indeed, if billions of years
of accelerated radioisotope decay had occurred mainly during the early part of the Creation Week, as considered a
possibility by the RATE team (Vardiman, Snelling, and Chaffin 2005), the enormous burst of radiation would surely have

been detrimental to all life on the earth, for example, the plants of Day Three. It is for this reason that many creationists are
not comfortable with postulating that accelerated radioisotope decay happened during the Creation Week, so maybe there
was no radioisotope decay at all until it was started as part of the curse.The lack of any evidence of a pattern of isochron
ages in these ten primitive and other achondrite meteorites, as well as in the 28 chondrite and basaltic achondrite (eucrite)
meteorites previously studied (Snelling 2014a, b, c), that matches the pattern found during the RATE project (Snelling
2005c; Vardiman, Snelling, and Chaffin 2005) would therefore strongly suggest that all these meteorites and their parent
asteroids have not experienced any episode of accelerated radioisotope decay, either at the time of the creation of the
primordial material on Day One or of their formation on Day Four, or since. This could then be taken to infer that no
accelerated radioisotope decay occurred anywhere in the solar system during the Creation Week, including on the earth
during Days OneThree. Such a conclusion is based on the assumption that the mechanism of small changes to the binding
forces in the nuclei of the parent radioisotopes proposed as the cause of a past episode of accelerated radioisotope decay
(Vardiman, Snelling, and Chaffin 2005) would thus have affected every atom making up the earth, and by logical extension
every atom of the universe at the same time.But why is there evidence of an episode of accelerated radioisotope decay in
the earth, but not in these parent asteroids and their meteorite fragments? The answer would seem to be that the
accelerated radioisotope decay only occurred during the catastrophic global Flood event on the earth and that the parent
asteroids were not similarly affected. However, if the earths atoms were affected by accelerated radioisotope decay during
the Flood, then surely every other atom in the universe would have been similarly affected..Another reason for arguing that
accelerated radioisotope decay occurred in earths rocks only during the Flood is that the earths rocks contain the physical
evidence of only 500600 million years worth of radioisotope decay (as calculated using todays measured decay rates),
which equates to the same time period postulated by uniformitarians during which the geologic record of the Flood
accumulated (Snelling 2009). This physical evidence of radioisotope decay in earths rocks is provided by radiohalos and
fission tracks (Snelling 2005a, b). Each fully-formed U radiohalo, no matter where in the geologic record it occurs or the
supposed age of its host rock, only represents up to 100 million years worth of U decay, so even if it is in a Precambrian
(pre- Flood) granite it still only records up to 100 million or so years worth of U decay that occurred during the Flood, at the
same time as new granites containing U radiohalos were forming in plutons which had intruded into fossil-bearing (and
therefore Flood-deposited) sedimentary strata. On the other hand, up to 600 or so million years worth of fission tracks are
found in zircon grains matching the radioisotope ages of those same zircon grains at the base of the strata record of the
Flood (Snelling 2005b).This still does not fully explain why there is such a spread of radioisotope ages from 4.03 Ga to the
present in the earths rocks, or why the radioisotope ages of the oldest earth rock (the Acasta Gneiss, Canada) (Bowring,
Williams, and Compston 1989; Stern and Bleeker 1998) and the oldest mineral in an earth rock (a zircon grain in the Jack
Hills sandstone, Western Australia) (Valley et al. 2014; Wilde at al. 2001) are 4.03 Ga and 4.4 Ga respectively rather than
4.56 Ga, the supposed age of the earth. The answer might be that the earths rocks, subsequent to their creation first
suffered from the processes of mixing of isotopes and resetting of radioisotope clocks in the mantle and crust during the
creation Great Upheaval when the dry landwere formed, and then further suffered from mixing of isotopes and resetting of
radioisotope clocks in the mantle and crust as a consequence of the catastrophic plate tectonics during the Flood, as well
as the concurrent accelerated radioisotope decay during the Flood. These mixing processes would not just have affected
isochron ages in that the mixing and adding or removing of parent and/ or daughter isotopes produces mixing lines that are
then interpreted as isochrons, but the same mixing, adding and/or subtracting of parent and/or daughter isotopes would
have reset the radioisotope model ages. In these ways the radioisotope clocks would have been reset during both the
Great Upheaval and the Flood, just as they are when pre-Flood crustal rocks (with older radioisotope ages) were melted to
form granite magmas, which when they crystallized had their radioisotope clocks reset to record the younger granite
formation ages.By comparison, the Day Three Great Upheaval which the Scriptures describe as occurring on the earth
could not have affected other bodies in the solar system, including the asteroids which were the source of the meteorites in
this study, because these other bodies were not created until Day Four. Thus we can be dogmatic that this event was earthspecific, as it was designed to produce the continental crust and the dry land on the earth in readiness for the subsequent
creation of plants, birds, animals, and man. In any case, so far our observations of the surfaces of asteroids do not indicate
any continental crust on them akin to that which formed on the earth on Day Three. However, when they were formed on
Day Four their formation could have included incredibly rapid silicate-metal fractionation and mantle-crust differentiation with
accompanying redistribution of the previously created parent and (what uniformitarians now interpret as) daughter isotopes.
The scatter in the radioisotope ages of the 38 meteorites studied thus far, the ten in this study and the 28 in the previous
studies, would then be due to processes subsequent to the initial creation of the parent asteroids from the primordial
material on Day Four of the Creation Week which have reset the radioisotope clocks at various times. Such processes
would include the initial fractionation and differentiation, heating due to impact cratering, impact disintegration, and recoalescing of the asteroids, space weathering, and heating on passage of the meteorites through the earths atmosphere
(Bogard and Garrison 2009; Cloutis, Binzel, and Gaffey 2014; Michel 2014; Norton 2002, ).
Where to From Here?
As concluded by Snelling (2014a, b, c) from his studies of 16 chondrite meteorites and 12 basaltic achondrites (eucrites),
based on the assumptions made the 4.554.57 Ga radioisotope ages for the Acapulco, Caddo County, Angra dos Reis,
DOrbigny, Norton County, Shallowater, Estherville, and Weekeroo Station achondrites obtained primarily by Pb-Pb and UPb radioisotope isochron and model age dating of various constituent minerals and fractions, but also by radioisotope
methods directly calibrated against Pb-Pb meteorite ages (for example, Amelin 2007, 2008; Amelin and Pravdivtseva 2005;
Brazzle et al. 1999; Brouxel and Tatsumoto 1990, 1991; Gilmour et al. 2006; Huey and Kohman 1973; Kleine et al. 2012;
Markowski et al. 2006; Schulz et al. 2012; Zhou et al. 2012) are likely not their true real-time ages. The assumptions on
which the radioisotope dating methods are based are simply unprovable, and in the light of the possibility of an inherited
primordial geochemical signature, subsequent resetting of radioisotope clocks due to impact cratering of asteroids, and the
evidence in earth rocks for past accelerated radioisotope decay, mixing of isotopes and resetting of radioisotope clocks,
these assumptions are unreasonable.However, we still need to develop a coherent and comprehensive explanation of what
these radioisotope compositions in both meteorites and earth rocks really represent and mean within our young-age
creation-Flood framework for earth and universe history. We have some possible clues already, as discussed here, and a
clearer picture may yet emerge from continued investigations now in progress. Examination of the radioisotope dating data
for groups of meteorites is necessary to complete our understanding of what the meteorite radioisotope ages might mean.
If only the earth was affected by accelerated radioisotope decay during the Flood, then it is also necessary to examine the
radioisotope dating data for martian meteorites, and lunar meteorites and rocks to determine whether Mars and the Moon
were affected by that aspect of the Flood catastrophe or not. Additionally, the radioisotope dating data for many more earth
rocks from all levels of the geologic record need to be collated and examined. If accelerated radioisotope decay only
occurred during the Flood, then it might be expected that the radioisotope ages of pre-Flood (mostly Precambrian) strata
determined by the different methods would be noticeably discordant (Snelling 2005c), whereas the radioisotope ages of

the strata formed during the Flood (mostly Phanerozoic) would be mostly concordant (Snelling 2005b). This difference might
be expected due to the pre-Flood rocks having already been formed and their radioisotope clocks started before the onset
of the accelerated radioisotope decay during the Flood, when their radioisotope clocks would have been speeded up by
different amounts according to the atomic weights of the parent radioisotopes for the full duration of the Flood event. In
contrast, the Flood rocks would have had their radioisotopes mixed and then reset when those rocks formed. Thus their
radioisotope clocks only started at different times during the accelerated radioisotope decay of the Flood event, and
consequently Flood rocks experienced less accelerated radioisotope decay than pre-Flood rocks. It may take the collation
and examination of the huge radioisotope dating data sets of as many different earth rocks as possible from all levels of the
geologic record to enable any firm conclusions to be made.Whatever the radioisotope dating data for the earths rocks may
reveal, it is already well-established that there are so many problems with the radioisotope dating methods which render
them totally unreliable in providing time markers for the different stages in the earths history. Indeed, the investigations of
determinations of the decay constants of each of the parent radioisotopes needs to be completed to provide further
documentation of the numerous uncertainties in that key time assumption. Therefore, even though most of these primitive
and other achondrites yield a consistent Pb-Pb, U-Pb, and Pb-Pb-calibrated isochron and model age of 4.554.57 Ga, that
cannot be their true real-time age, which according to the creation paradigm is only about 6000 real-time years.
Conclusions
After six decades of numerous careful radioisotope dating investigations of these achondrite meteorites their Pb-Pb, U-Pb,
and Pb-Pb-calibrated isochron and model age of 4.554.57 Ga has been well established. This date for these achondrites is
supported for most of them by a strong clustering of their Pb-Pb, U-Pb, and Pb-Pb-calibrated isochron and model ages in
the 4.554.57 Ga range, as well as sometimes being confirmed by both isochron and model age results via the K-Ar, Ar-Ar,
Rb-Sr, Lu-Hf, Re-Os, and Sm-Nd methods. The Hf-W, Mn-Cr, Al-Mg, and I-Xe methods are all calibrated against the Pb-Pb
method, so their results are not objectively independent. Thus the Pb-Pb dating method stands supreme in the conventional
evolutionary uniformitarian community as the ultimate, most precise tool for determining the ages of these achondrite and
other meteorites.There are no discernible patterns in the isochron and model ages for these achondrites, apart from
scattering of many K-Ar, Ar-Ar, Rb-Sr, Re-Os, and Sm-Nd isochron ages and a few Pb-Pb and U-Pb isochron ages, and the
considerable scattering of many K-Ar, Ar-Ar, and Rb-Sr model ages and some Pb-Pb, U-Pb, and Th-Pb model ages. The
radioisotope ages for these achondrites do not follow the systematic pattern found in Precambrian rock units during the
RATE project. The -decay isochron ages are not always older than the -decay isochron ages for particular achondrites,
and among the -decayers the isochron ages are not always older according to the increasing heaviness of the atomic
weights of the parent radioisotopes, or the increasing lengths of their half-lives. Thus there appears to be no consistent
evidence in these achondrite meteorites similar to that found in earth rocks of past accelerated radioisotope decay.
Therefore it could be concluded from these data that no accelerated radioisotope decay event has occurred on the asteroids
which parented these achondrite meteorites.Any explanation for the 4.554.57 Ga age for these achondrite meteorites
needs to consider their origin. These meteorites are regarded as fragments of asteroids, debris derived from impact
cratering of their surfaces and collisions with other asteroids. Even in the naturalistic paradigm the asteroids, and thus the
meteorites, are regarded as primordial material left over from the formation of the solar system. Todays measured
radioisotope compositions of these achondrites may reflect a geochemical signature of that primordial material, which
included atoms of all elemental isotopes. Therefore if some, or perhaps most, of the daughter isotopes measured today in
these achondrite meteorites were thus inherited by them when they were formed from that primordial material, and the
parent isotopes in these meteorites have only been subjected to some subsequent radioisotope decay (and none at
accelerated rates), then the 4.554.57 Ga Pb-Pb, U-Pb, and Pb-Pb-calibrated isochron and model age for these
achondrites cannot be their true real-time age, which according to the creation paradigm is only about 6000 real-time
years.However, these conclusions and suggested explanations as discussed are still somewhat tentative, their confirmation
or adjustment awaiting the examination of more radioisotope dating data for groups of meteorites, and martian and lunar
meteorites. Furthermore, further extensive studies of the radioisotope dating of lunar rocks and rocks from all levels of the
earths geologic record are required to attempt to systematize what proportions of the isotopes in each radioisotope dating
system measured today are due to inheritance from the primordial material, to accelerated radioisotope decay during the
Flood, and mixing, additions and subtractions in the earths mantle and crust through earth history, particularly during the
Day Three Upheaval and then subsequently during the Flood.
Acknowledgments
The invaluable help of my research assistant Lee Anderson Jr. in compiling these radioisotope dating data into the tables
and then plotting the data in the color coded age versus frequency histogram diagrams is acknowledged. Also, our
production assistant Laurel Hemmings is especially thanked for her painstaking work in preparing these papers for
publication.

Vous aimerez peut-être aussi