Vous êtes sur la page 1sur 36

Paleoreef Maps: Evaluation of a Comprehensive Database

on Phanerozoic Reefs1
Wolfgang Kiessling,2,3 Erik Flgel,2 and Jan Golonka4

ABSTRACT
To get a better understanding of controls on reef
development through time, we created a comprehensive database on Phanerozoic reefs. The
database currently comprises 2470 reefs and contains information about geographic position/paleoposition, age, reef type, dimensions, environmental
setting, paleontological and petrographical features, and reservoir quality of each buildup.
Reef data were analyzed in two qualitatively different ways. The first type of analysis was by visualization of paleogeographic reef distribution maps. Five
examples (Late Devonian, Early Permian, Late
Triassic, Late Jurassic, middle Miocene) are presented
to show the potential of paleoreef maps for paleogeographic and paleoclimatological reconstructions.
The second type of analysis was a numerical processing of coded reef characteristics to realize major
trends in reef evolution and properties of reef carbonates. The analysis of paleolatitudinal reef distributions through time shows pronounced asymmetries in some time slices, probably related to
climatic asymmetries rather than controlled by plate
tectonic evolution alone. The dominance of particular reef builders through time suggests that there
are seven cycles of Phanerozoic reef development.
First curves for the Phanerozoic distribution of bioerosion in reefs, bathymetric setting, and debris
Copyright 1999. The American Association of Petroleum Geologists. All
rights reserved.
1Manuscript received December 8, 1997; revised manuscript received
February 3, 1999; final acceptance March 3, 1999.
2 Institut fr Palontologie, Loewenichstrae 28, D-91054 Erlangen,
Germany.
3Current address: Museum fr Naturkunde, Invalidenstr. 43, D-10115
Berlin, Germany; e-mail: wolfgang.kiessling@rz.hu-berlin.de
4Jagiellonian University, Institute of Geological Sciences, 30-063 Krakow,
Oleandry 2a, Poland.
This study was supported by the German Research Foundation (Projects
Fl 42/75, Fl 42/80-1) and was partly embedded in the priority program on
controls on biogenic sedimentation: reef evolution. Fruitful discussions with
R. Koch (Erlangen), R. Leinfelder (Stuttgart), and B. Senowbari-Daryan
(Erlangen) are gratefully acknowledged. D. Ford (Dallas) is thanked for
editorial remarks. D. Jovanovic (Beograd), M. Link (Erlangen), R. Scasso
(Buenos Aires), B. Senowbari-Daryan (Erlangen), and T. Steuber (Erlangen)
provided important unpublished data. J. Collins, P. Playford, and J. Wilson
are thanked for their reviews. The remarks of J. Collins were especially useful
to improve the manuscript.

1552

potential of reefs are presented. The observed pattern in the temporal and spatial distribution of reefs
with reservoir quality may assist in hydrocarbon
exploration. Statistical tests on the dependencies of
reefal reservoir quality suggest that large size, high
debris potential, low paleolatitude, high amount of
marine aragonite cement, and a platform/shelf edge
setting favor reservoir quality. Reefal reservoirs are
significantly enhanced in times of high evaporite
sedimentation, elevated burial of organic carbon,
low oceanic crust production, low atmospheric
CO 2 content, and cool paleoclimate, as well as
when they are present in aragonite oceans.
INTRODUCTION
Modern coral reefs are the most complex and taxonomically diverse marine ecosystems (Paulay,
1997; Hatcher, 1997). Biotic composition, sedimentary development, and diagenetic history of recent
reefs reflect only a small part of the long-lasting history of reefs that started about 2 b.y. ago with stromatolitic buildups and continued with a wealth of
different organic buildups in the Phanerozoic.
Current and past changes of climatic and oceanographic conditions are recorded by reef biota and
reef sediments. Because reefs also are considered
important as alkalinity and climate-controlling carbonate factories (Berger, 1982; Hubbard, 1997), as
tracers of fossil benthic communities (Kauffman and
Fagerstrom, 1993), and as hydrocarbon reservoirs
(Greenlee and Lehmann, 1993), it is worthwhile to
learn more about this fascinating ecosystem.
Intense research on fossil reefs has accumulated a
tremendous amount of data during the last 150 yr
(Flgel and Flgel-Kahler, 1992), and several papers
review Phanerozoic evolution of reefs (Newell,
1971; Heckel, 1974; Wilson, 1975; James, 1983;
Sheehan, 1985; Fagerstrom, 1987; Copper, 1988,
1989; Talent, 1988; Flgel and Flgel-Kahler, 1992;
James and Bourque, 1992; Kauffman and Fagerstrom, 1993; Wood, 1993, 1995). These papers
focus on important biological and geological
aspects of reef evolution; however, many open
questions still exist regarding the driving forces of
AAPG Bulletin, V. 83, No. 10 (October 1999), P. 15521587.

Kiessling et al.

reef development and evolution mainly due to


three shortcomings: (1) many comprehensive reef
studies are limited to one time slice of variable
extent, (2) investigation of reef features through
time are treated mostly qualitatively and semiquantitatively rather than strictly quantitatively, and (3)
the interpretation of fossil reef distributions is
biased by using simplified and commonly obsolete
paleogeographic maps and oversimplified or exaggerated distribution plots.
The most comprehensive database on ancient
reefs was presented by Flgel and Flgel-Kahler
(1992); however, although more than 2000 references are cited in their papers, it is not much
more than a literature survey and is hampered by
previous points two and three. In this study, we
use a comprehensive computer database of
Phanerozoic reefs together with new paleogeographic maps to analyze the variation of reef features and reef localities through time. The database
was applied in two independent ways. The first
application was a visual evaluation of paleogeographic reef distribution maps. The concentration
of reef settings in particular areas, the latitudinal
distribution and boundaries of reef occurrences,
and the relation of paleogeography and reefs were
directly observed. In addition, selected reef features could be highlighted by defining different
colors or setting filter functions. Five paleogeographic maps (Upper Devonian, Lower Permian,
Upper Triassic, Upper Jurassic, middle Miocene)
are featured in this paper for a discussion of their
potential in reef studies.
The second application was a numerical analysis
of coded reef features. Interpretation is based on
the examination of diagrams and statistical analyses. Although this method produces less straightforward results, it allows a much more profound discussion than the maps alone. Secular trends in reef
evolution are best studied by using quantitative
data. We limit our discussion to eight selected reef
features: abundance, size, paleoposition, biotic
composition, bioerosion, bathymetry, debris potential, and reservoir potential.

1553

ing either detailed descriptions of single reefs or


dealing with reef distribution and characteristics
on a regional scale. Currently, 1700 references are
considered in the database. Reef characteristics
were transformed into numerical codes to allow an
objective and statistically meaningful analysis of the
data (Table 1).
To permit a comparison of reef abundance over
time, a minimum distance for a reef to be counted
separately within the same time slice was chosen.
Owing to the scale of paleogeographic maps, this
distance was set to 20 km; however, if more closely
spaced reefs were of different ages, from a different
paleogeographic setting, or of significantly different composition, they were included in the
database. Reef attributes are assigned only if adequate descriptions are available, otherwise the
database fields remained blank (missing value in
statistical analyses).
The database structure contains seven main
headings: (1) reef identification and references, (2)
age, (3) present-day position and paleoposition, (4)
general features, (5) environment, (6) paleontological features, and (7) petrographical features and
reservoir potential.
Reef Identification and References
The definite identification of each reef in the
database is possible by using a specific identification number for each reef. This number is necessary to link the main database with related data
sets (e.g., references, paleopositions). A trivial
name for each reef is also provided to permit a
quick assignment of its location. This field contains a short name of the reef or of the locality and
the name of the country or United States state
where the reef is situated. Each reef is assigned at
least one reference but may be assigned to up to
six references. References are coded by four-digit
numbers and are linked to the reference database
containing the whole reference.
Age

DATABASE STRUCTURE
The database contains outcropping reefs, as well
as reefs known only from the subsurface (drilling
and seismic exploration). Modern-type reefs and
reef mounds, mud mounds, and major biostromes
were considered, but are described separately.
Most reef data used were extracted from published references (cf. Flgel and Flgel-Kahler,
1992), but we also included personal communications and unpublished reports. The literature analysis was focused on comprehensive papers provid-

Under this heading, several fields are included.


Three fields are reserved for the stratigraphic age of
the reef, including system, series, and stage.
Another field contains the chronostratigraphic age
in millions of years (Ma), according to the slightly
modified geological time scale of Gradstein and Ogg
(1996). The most important field in this category is
the time slice. A total of thirty-two time slices were
used, encompassing the time between the
PrecambrianCambrian boundary and the late
MiocenePliocene. The six megasequences of Sloss

1554

Paleoreef Maps

Table 1. Two Examples of Reefs as Coded in the Paleoreef Database*

Category
1
2

3
4

5
6

Field
Reef name
Reef number
References
Age/System
Age/Series
Age/Stage
Age/time slice
Age/m.y.
Age/reliability
Location/today
Plate
Location/paleo
Reef type
Size/thickness
Size/width
Size/extension
Environment
Bathymetry
Biota main
Biota detailed

Frasnian Stromatoporoid Mound

Guild
Diversity

Couvin, Belgium
1560
839, 960, 976
Devonian
Upper
Middle Frasnian
9
371
3 = exact and reliable age assignment
50.0500N, 4.4667E
315 = Avalonia
20.0639S, 4.7912E
2 = reef mound
3 = 100 m500 m
4 = >500 m
40 km
1a = intraplatform
2 = below fair-weather wave base
4 = stromatoporoids
9 = stromatoporoids, corals
(bryozoans, algae/microbes)
3 = binder dominated
2 = moderate

Bioerosion/macro
Bioerosion/micro
Micrite
Sparite
Reservoir potential
Debris potential

Absent
Absent
3 = abundant
2 = moderately abundant
No
2 = moderate

Miocene Coral Reef


Cap Blanc, Mallorca, Spain
679
645, 670, 1194
Tertiary
Upper
TortonianMessinian
32
7
3 = exact and reliable age assignment
39.4167N, 2.7833E
320 = Balearic Islands
39.3528N, 1.9269E
1 = reef
2 = 10 m100 m
2 = 20 m100 m
20 km
2a = platform margin
1 = above fair-weather wave base
1 = scleractinian corals
1 = corals, red algae (hydrozoans,
foraminifera, microbes, sponges)
1 = framework dominated
1 = low (strongly dominated
by Porites)
Present
Present
2 = moderately abundant
1 = few
Yes (high porosity)
2 = moderate

*Remarks in parentheses refer to an additional memo field. Numbers in the first column indicate categories as titled in text. References: 645 = Pomar
(1991); 670 = Esteban (1979); 839 = Lecompte (1958); 960 = Lecompte (1970); 1194 = Pomar et al. (1996).

(1963) form the major frame for our time slices


because they can be correlated intercontinentally
and have been shown to be linked to global tectonics
(Sloss, 1972; Soares et al., 1978) and high-order
eustatic sea level f luctuations (e.g., Haq et al.,
1988). The definition of time slices corresponds to
supersequences as defined by Mobil and partially
defined by Golonka and Ford (1997a, b) and
Golonka et al. (1997a, b), but owing to limited data
not all Phanerozoic supersequences are separated.
The boundaries of the time slices are defined by
second- to third-order eustatic sea level minima.
Thus, the time slices represent time intervals that
may embrace several stages or transit system
boundaries but also may cut stages. The philosophy
behind this approach is a more natural subdivision
of the geological record not biased by regional differences and not a priori influenced by biological
evolution. The time intervals represented by each
time slice are listed in Table 2. The last field within
the category age estimates the reliability of age
determination given in the literature.

Present-Day Position and Paleoposition


Present-day coordinates are represented by two
numbers referring to latitude and longitude
(Greenwich coordinates). Southern and western
coordinates are negative numbers. The correct
assignment of a plate number is crucial for calculating the paleogeographic reef position. The plates
are coded in the database by three-digit numbers.
The calculated paleopositions also are represented
by two fields: paleolatitude and paleolongitude.
The method used for paleogeographic reconstructions is described in a following section.
General Features
Reef Type
Four reef types were distinguished: (1) true
reefs, where the organisms form a rigid framework,
(2) reef mounds, where matrix/cement and organisms are about equally important and the buildup is

Kiessling et al.

1555

essentially matrix-supported, (3) mud mounds,


where organisms are minor constituents of the
buildup consisting predominantly of carbonate
mud, and (4) biostromes, where dense growth of
reef-building organisms is evident, but there is no
depositional relief. Because the difference between
small biostromes and fossiliferous beds is hazy and
small biostromes were too numerous to be counted, only major biostromes within a time slice were
recorded in the database.
Size
The vertical and lateral extensions of a reef body
were registered. Because most of the data provided
in publications are not ver y accurate in this
respect, a rather coarse interval classification was
used: (1) less than 10 m thick and less than 20
wide, (2) 10100 m thick and 20100 m wide, (3)
100500 m thick or wide, (4) more than 500 m
thick or wide. Reef complexes containing stacked
smaller buildups were described as one single reef
at a given locality. If several reefs were described
aligned in a continuous belt, the length of the reef
belt in kilometers was recorded.
Environment
We have separated four principal depositional
settings for reefs, each subdivided into several environments (Figure 1).
(1) Shelf or platform: (1a) within shallow carbonate platform, (1b) intraplatform sag, (1c) epeiric sea, (1d) coastal, transitional, marginal marine,
and (1e) open-marine shelf.
(2) Platform or shelf margin: (2a) platform margin bordering shallow basins, (2b) platform or shelf
margin bordering deep basins, and (2c) atoll structure, seamounts.
(3) Slope or ramp: (3a) upper slope or inner
ramp and (3b) lower slope or outer ramp.
(4) Basin: (4a) moderately deep basin (above
photic zone, <200 m) and (4b) deep basin.
For some reefs it is difficult to assign the correct
depositional environment, e.g., the difference
between epeiric sea and open-marine shelf is not
always precisely defined. The same is true for epeiric sea, open-marine shelves, and ramps [see
Burchette and Wright (1992) for a comprehensive
discussion]. Ramps were only counted as such if
they were bordering the open ocean.
Bathymetry: The determination of paleowater
depth is quite reliable if two intervals are considered: 1 = above fair-weather wave base; 2 = below
fair-weather wave base. A finer bathymetric classification is not feasible in most cases owing to lack of
specific data.

Figure 1Schematic position of the different reef environment settings used in the database. SL = sea level, 1a
= shallow carbonate platform, 1b = intraplatform sag, 1c
= epeiric sea, 1d = coastal, transitional, marginal
marine, 1e = open-marine shelf, 2a = platform margin
bordering shallow basins, 2b = platform/shelf margin
bordering deep basins, 2c = atoll structure, seamounts,
3a = upper slope or inner ramp, 3b = lower slope or
outer ramp, 4a = moderately deep basin (<200 m), 4b =
deep basin.

Paleontological Features
Biotic Composition
The biotic composition of reefs in the database
refers only to reef-building organisms. Reef-dwelling
or destructive organisms were not considered. Reefbuilders were defined as sessile organisms having
the potential to contribute significantly to buildup
formation by constructing, baffling, or binding.
Two fields in the database refer to the biotic composition of reef builders. The quantitatively dominant reef
building group is listed in the first field. In reefs

1556

Paleoreef Maps

with a pronounced horizontal and vertical zonation,


the dominant reef-builder in the reef core or climax
stage was listed, even if the group was not dominant
in the whole buildup. In the second field, all important reef-builders in a buildup were considered. Sixtyfive combinations of reef-building groups have been
distinguished so far. The combinations are coded by
two-digit numbers.
Dominant Guild
Following the guild model of Fagerstrom (1987,
1991), the dominant guild of reef builders in each reef
was determined: The constructor guild, the baffler
guild, and the binder/encruster guild were separated.
Some fossil groups can be always assigned to the
same guild (e.g., stromatolites = binder guild), but
most groups are classified into different guilds
depending on growth form (Fagerstrom, 1987, 1988).
Diversity
The diversity field also exclusively refers to reef
builders. Because the quality of published data is
quite heterogeneous, a rather coarse interval classification of diversity was applied: 1 = low diversity
(less than five species or one species strongly predominant); 2 = moderate diversity, and 3 = high
diversity (more than 25 species).
Bioerosion
We distinguished macroborings and microborings. Because the study of bioerosion in fossil reefs
is still in its infancy (Vogel, 1993), we decided not
to consider more than binary (presence/absence)
data. Evidence provided in the original text or a figure showing boring traces was necessary for a positive indication in the bioerosion fields.
Petrographical Features and Reservoir
Quality
Petrography
The amount of micrite and sparite (synsedimentary or early diagenetic marine carbonate cement)
was listed in two separate fields. They were quantified with respect to the proportion of biota, micrite,
and sparite using three intervals. We also noted if
the reefs were heavily dolomitized because dolomitization reduces the reliability of paleontological
data and provides information on reservoir quality.

one can evaluate whether a reef has a reservoir potential; therefore, a binary field was included in the
database saying whether a reef may have a reservoir
potential. In the subsurface, only reefs with tested
reservoir quality are included, whereas outcropping
reefs need to have preserved high porosity (>3%) and
at least moderate thickness (>10 m) to be classified as
having reservoir quality. Outcropping equivalents of
subsurface reservoirs were not included if porosity
had been destroyed by surface diagenesis. Seals and
source rocks were not considered. The database
lumps productive and nonproductive reefal reservoirs.
Debris Potential
Many ancient reefs consist predominantly of
debris formed by reef organisms and reworking of
lithified reef rocks (Zankl, 1977; Hubbard et al.,
1990). Because the significance and amount of
debris in reefs are thought to vary considerably
through time, we tried to quantify the production of
debris in reefs. Owing to the limited information on
the absolute amount of debris produced by a reef,
we quantified the relative debris production. Again,
poor data did not allow us to separate more than
three intervals: 1 = low, 2 = moderate, and 3 = high
debris production. Low debris production is supposed for reefs with a high proportion of
autochthonous reef carbonates or reefs lacking forereef debris. Many reefs with low debris production
are from deeper water environments, but reefs in
protected lagoonal environments and buildups dominated by certain fossil groups (e.g., microbes) also
are unlikely to produce high amounts of debris. The
absolute amount of debris production is not relevant
for this field. A 200-m-thick reef can be classified as
highly debris producing as can be a reef of less than
10 m thickness if they both consist almost exclusively of rudstones and reworked boundstones. Thus,
the values in this field reflect the potential of a reef
to produce debris rather than the volume of debris
produced. The actual debris production of a reef can
be calculated with the aid of reef dimensions.
Additional Information
If a detailed study was available, additional information from that study was included under remarks (thickness of reef in meters, species names, average porosity, etc.) to allow a later refinement of the database.
EVALUATION OF DATA COMPLETENESS AND
RELIABILITY OF INTERPRETATIONS

Reservoir Quality
Reservoir quality of reefs from published data is usually difficult to quantify precisely. In many cases, however,

Currently, 2470 Phanerozoic reefs are in the


database assigned to 32 time slices as defined in

Kiessling et al.

Table 2. The number of reefs in a single time slice


ranges between 14 and 245 (74 on average). This
tremendous range in number can be attributed to
two possible reasons: either it reflects a real difference between times of reef crises and peaks in reef
development or it reflects a bias of unequal reef
data availability.
Before starting to draw conclusions from the statistical analysis of reef characteristics, we should
try to find out which one of these possibilities better applies to our database. Figure 2 depicts the
number of reefs in each time slice and the average
number of reefs per million years. How real is the
trend depicted by Figure 2? As an example, the
abundance of Early Cambrian reefs is certainly
underrepresented in the database. The abundant
Siberian reefs of this time (Rowland and Gangloff,
1988) are not all included in the database because
informative descriptions are limited to rather few
reefs. Despite this, decreased reef numbers in the
Middle Cambrian to Early Ordovician represent a
real trend caused by the disappearance of archaeocyathid-microbe communities in the Middle
Cambrian (James and Debrenne, 1980) and the
absence of reef-building corals and stromatoporoids. The radiation of the latter groups in the
MiddleLate Ordovician is reflected in an increasing reef abundance in the Late Ordovician and
Silurian (see also Copper, 1994). In the Devonian,
the increasing amount of reefs toward the
FrasnianFamennian boundary appears to be real.
The prominent peak of GivetianFrasnian reefs
probably is somewhat exaggerated as a result of
intense exploration activities in Canada and Russia
related to their economic importance. Although
the peak may be too high in relation to the surrounding time slices, it likely ref lects an actual
peak in the number of reefs. The reduced amount
of buildups after the early Famennian reef crisis
also is likely to represent reality; however, the low
abundance of Late Carboniferous to Artinskian
reefs is suggested to be artificial because most of
the reefs are situated in Russia, where data access
is difficult and the actual number of reefs may be
much higher. The high number of reefs in the
Triassic is related to the intense research in the
Alpine-Mediterranean area during the last 25 yr
(Flgel, 1981) and hence reflects the real situation.
The drop of reef numbers in the Early Jurassic
appears to be real (TriassicJurassic reef crisis).
The abundance of Early Jurassic reefs would be
even lower if the common Lithiotis bioherms
were not considered. The bloom of Late Jurassic
reefs is real (Leinfelder, 1994), as is the decrease in
the number of reefs in the Cretaceous; however,
the abundance of reefs in the Cretaceous would
increase if all minor rudist-biostromes were included. Because the decision of including or omitting a

1557

rudist association is subjective, the trend shown


within the Cretaceous is somewhat artificial. The
low amount of reefs in the Paleogene is thought to
reflect the actual situation, as does the increase of
reefs in the Neogene. The Neogene peak is much
more pronounced when normalized reef numbers
are examined, since Cenozoic time slices are considerably shorter than Paleozoic and Mesozoic
time slices (Table 2).
In summar y, the amount of reef data in the
database largely reflects the real situation in most
cases, but an occasional bias is present, related to
unequal quantity and quality of published data and
some subjective categorization. Our graphs of reef
distribution (Figure 2; see also Figure 13) are in
good agreement with the reef abundance curve of
Talent (1988). Talents approach has the advantage
of possessing higher stratigraphic resolution
(stages) and better ref lecting Phanerozoic reef
crises; however, our charts provide more quantitative information about what Talent (1988) called
episodes of reef construction. The Silurian,
Frasnian, Late Jurassic, and Miocene peaks in reef
abundance are obviously more important than indicated in Talents curve.
Information on ancient reefs is rather heterogeneous because type and quality of information
varies depending on scientific scope and scale of
study (description of one particular reef vs. regional surveys or reviews); hence, the reliability of
attributes in the database as the basis for statistical
analyses is another important question. All essential
information and an at least fair stratigraphic assignment could be found in the references for about
48% of the reefs in the database. An average of 22%
of reefs in the database lack detailed information
on paleontological, environmental, or petrographical attributes, and 30% of the reefs in the database
lack an indication of reef size or dominant biota
(Figure 3). Certain reef characteristics (e.g., bioerosion or biotic diversity) mostly refer to European
and North American reefs because these attributes
are rarely described in other areas; however, adequate descriptions of some reefs can be found in all
regions and time slices. Assuming that other reefs
in the same area with the same biotic composition
and age do not differ notably in main features, the
negative influence of heterogeneous knowledge
may be insignificant.
Similar arguments can be given for the reefs not
represented in the database. Even the best known
time slices are incomplete regarding the amount of
reefs in the database due to burial, erosion, or lack
of investigation in particular areas. Although this
fact certainly biases the interpretation potential of
global reef distribution, its effect on the statistical
evaluation is relatively low, as indicated by a comparison of statistical analyses at different stages of

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32

Cambrian
Cambrian
CambrianOrdovician
Ordovician
Ordovician
Silurian
Silurian
SilurianDevonian
Devonian
Devonian
DevonianCarboniferous
Carboniferous
Carboniferous
CarboniferousPermian
Permian
Permian
Triassic
TriassicJurassic
Jurassic
Jurassic
Jurassic
JurassicCretaceous
Cretaceous
Cretaceous
Cretaceous
Cretaceous
Paleogene
Paleogene
Paleogene
PaleogeneNeogene
Neogene
Neogene

Period

Shortcut for
Time Slice
Early Cambrian
Middle Cambrian
Tremadocian
Arenigian
Caradocian
Llandoverian
Wenlockian
Lochkovian
EmsianEifelian
GivetianFrasnian
Tournaisian
ViseanSerpukhovian
MoscovianKasimovian
Asselian
Artinskian
Guadalupian
Ladinian
Norian
Pliensbachian
BajocianBathonian
Oxfordian-Kimmeridgian
Berriasian
Barremian
Albian
Turonian
Campanian
Ypresian
Lutetian
Rupelian
Aquitanian
Serravallian
Messinian

*Time slices are defined by supersequences.

Time
Slice
No.

Table 2. Definition of Time Slices Used in the Reef Database*

Nemakit-Daldynian
Toyonian
Franconian
Arenigian
Darriwilian
Rhuddanian
Wenlockian
Middle Pridolian
Late Pragian
Givetian
Middle Famennian
Middle Visean
Bashkirian
Gzhelian
Sakmarian
Roadian
Induan
Late Carnian
Late Hettangian
Middle Aalenian
Late Bathonian
Late Tithonian
Late Valanginian
Late Aptian
Late Cenomanian
Middle Campanian
Thanetian
Lutetian
Priabonian
Chattian
Burdigalian
Tortonian

Lower Boundary
520
502
488
472
452
435
425
412
396
368
348
328
302
287
277
255
232
218
195
169
152
140
126
105
90
76
53
45
33
22
14
6

Age of Plate Tectonic


Reconstruction (Ma)

Toyonian
Dresbachian
Tremadocian
Llanvirnian
Ashgillian
Telychian
Early Pridolian
Middle Pragian
Eifelian
Early Famennian
Early Visean
Serpukhovian
Kasimovian
Asselian
Kungurian
Changhsingian
Early Carnian
Middle Hettangian
Early Aalenian
Middle Bathonian
Middle Tithonian
Early Valanginian
Early Aptian
Middle Cenomanian
Early Campanian
Selandian
Ypresian
Bartonian
Rupelian
Aquitanian
Serravallian
Pliocene

Upper Boundary
33
14
15
21
21
15
10
16
22
20
22
15
27
11
17
20
24
21
24
12
20
12
18
23
13
23
9
12
9
8
10
9

Duration of
Time Slice (m.y.)

1558
Paleoreef Maps

Kiessling et al.

20

300

Absolute number of reefs


250

1559

18

Number per m.y.

No. of Reefs

14
200
12

150

10

Normalized No. of Reefs

16

100
6

4
50
2

Ea
M rly C
id
dl am
e
C bri
Tr am an
em br
ad ian
o
Ar cia
C eni n
ar gi
Ll ado an
an c
do ia
W ve n
en ria
L lo n
Em oc ckia
h
G sia kov n
ive n
i
/
tia Ei an
n/ fel
Vi
ia
F
s
M ean To ras n
os /S ur nia
co e na n
vi rpu is
an k ia
/K ho n
as vi
im an
ov
As ian
se
A
li
G rtin an
ua sk
da ia
lu n
La pia
di n
ni
an
Ba Pli
N
jo ens or
ci
i
an ba an
/B ch
Ki ath ian
m o
m ni
er an
i
Be dgia
rri n
Ba as
rre ian
m
ia
Al n
b
Tu ian
C ro
am ni
pa an
Yp nia
re n
s
Lu ian
te
R tia
u n
Aq pel
ui ian
Se ta
rra nia
v n
M alli
es an
si
ni
an

Figure 2Number of reefs in the database. The absolute number (bars) and the number of reefs per million years
(line) in each time slice are indicated. Abundance peaks are likely to represent real times of reef prosperity.

data input. Because the reef database will always


remain incomplete, the stability of calculated
means of reef attributes with an increasing amount
of data is a crucial factor. In Figure 4, three stages
in database development (September 1996,
October 1997, December 1998) are compared
with respect to the percentage of major reefs
(thickness >100 m) in each time slice for the
LochkovianMessinian. The 1996 curve takes into
account 1460 reefs, the 1997 curve refers to 1816
reefs, and the new curve considers 2042 reefs.
Although up to 160% of data was added to the time
slices, the percentage of major reefs changed
insignificantly in most of the time slices. Even the
relatively strong deviations in some slices
(Asselian, Aquitanian) do not change the overall
pattern. Stronger deviations are noted when a random selection of reef data is studied. We compare
the curve based on the whole database (2470
reefs), with the curve based on a random selection
of 1000 reefs (Figure 5). The two curves differ in
details, yet the major trend is still apparent in the
random selection. We conclude that further
progress in database development will modify

details of the results, but the major patterns and


trends are already visible.
The stability of trends uncovered by the database
proves its value as a predictive tool. A rough estimate of attributes will be possible on reefs in unexplored areas.
PALEOGEOGRAPHIC RECONSTRUCTIONS
All reefs were assigned to time slices and
plate numbers, a necessary step for extracting
evolutionary trends and calculating paleopositions. The plate tectonic model describes the
relative motions between approximately 300
plates and terranes. This model was constructed using P LATES and PALEOMAP software, which
integrates computer graphics and data management technology with a highly structured and
quantitative description of tectonic relationships. The heart of this program is the rotation
file, which is constantly updated as new paleomagnetic data become available. Hot-spot volcanoes serve as reference points for the calculation

Good

Fair

Poor

Figure 3Completeness of information in the database. Good = information is present for all fields of a data set and stratigraphic assignment is at
least fairly exact; fair = detailed information on paleontology, paleoenvironment, or petrography is missing; poor = information on reef dimensions,
paleontology, petrography, or paleoenvironment is lacking.

n n n an an an an an an an an an an an an an an an an ian an an an an an an an an an an an an
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
ia ia ia
i
i
i
i
i
i
i
i
i
br br oc nig oc ver ock kov ifel sn ais hov ov sel nsk lup din or ch hon idg ias em Alb ron an res tet pel tan vall sin
i
a
d
d
a
r
N
p
u
r
r
m
m
l
i
s
s
n
t
m
u
a
a
e
E
o
r
p
r
u
a
r
t
h
i
k
a
a a a Ar
n c n/ /F ur u
A Ar ad L
T am Y L
s
R qu err Me
sb Ba me Be Ba
C C m
ar nd e
A S
u
C Lla W Lo sia tian To erp /Ka
C
en an/ im
i
ly dle Tre
l
G
r
K
P ci
/S ian
Em ive
Ea Mid
n
jo
a ov
G
Ba
se sc
i
o
V
M

0%

50%

100%

Reefs (%)

1560
Paleoreef Maps

Kiessling et al.

1561

60

50

% Major Reefs 96
% Major Reefs 97
% Major Reefs 98

40

30

20

10

Em Loc
G sia hko
ive n v
tia /Ei ian
n/ fel
Vi
Fr ia
se
M an To asn n
os /S u
i
co er rna an
vi pu is
an k ia
/K ho n
as via
im n
ov
As ian
Ar sel
G tins ian
ua k
da ian
lu
La pia
di n
ni
Ba P
a
jo lien No n
ci s r
an ba ia
/B ch n
Ki ath ian
m on
m
er ian
id
Be gia
n
r
Be rias
rre ian
m
ia
Al n
Tu bia
C ron n
am ia
pa n
Yp nia
re n
s
Lu ian
te
R tia
up n
Aq el
u ia
Se ita n
rra nia
v n
M allia
es n
si
ni
an

Figure 4Comparison of three stages in database development. The percentage of major reefs (thicker than 100 m)
in each time slice is indicated for three stages in database development (September 1996, October 1997, December
1998). Note the stability of the overall trend despite a substantial increase of reef data.

of paleolongitudes (Golonka and Bocharova,


1997). Magnetic data have been used to define
paleolatitudinal position of continents and rotation of plates. For Jurassic and younger a ge
slices, the magnetic anomalies have been used to
define the evolution of the oceans. Ophiolites
and deep-water sediments mark paleo-oceans,
which were subducted and included into fold
belts.
The plate tectonic reconstruction maps have
been combined with paleoenvironmental and
lithological data, enabling the generation of
detailed paleogeographic maps. Thirty-two maps
were constructed that depict the changing configuration of mountains, land, shallow seas, and deep

ocean basins for distinctive time intervals encompassing a time span from the Early Cambrian to
the late MiocenePliocene. Generally, the individual maps illustrate the conditions present during
the maximum marine transgressions of a higher
frequency cyclicity. Relative sea level cyclicity
(Haq et al., 1988; Ross and Ross, 1988; Greenlee
and Lehmann, 1993), chronostratigraphy (Gradstein
and Ogg, 1996), and regional unconformities provide the basis to partition the higher frequency
depositional cycles into subdivisions ranging
from 8 to 33 Ma. The calculated paleolatitudes
and paleolongitudes were loaded into the database and used to generate the paleoreef maps discussed in this paper (Figures 611).

1562

Paleoreef Maps

Figure 5Percentage
of major reefs in each
time slice as calculated
from the database.
(a) Diagram based on
a random selection
of 1000 reefs.
(b) Diagram based on
all 2470 data. Although
the two graphs exhibit
differences in detail,
the overall trend is
the same. Large
amounts of thick reef
complexes are noted
from the Devonian to
the Tournaisian, from
the Sakmarian to the
Norian, and in the
Cenozoic.

Figure 6Global distribution of GivetianFrasnian reefs. Ocean surface currents were derived from the plate tectonic configuration. 1 = Mountains, 2 = land,
3 = shelf, 4 = deep water, 5 = predicted upwelling zones, slightly modified from Golonka et al. (1994), 6 = reefs thicker than 100 m, 7 = reefs between 10 and
100 m thickness, 8 = reefs thinner than 10 m, 9 = reefs without thickness data. Note the close association of many reefs with predicted upwelling sites and the
wide latitudinal distribution of reefs.

1564

Paleoreef Maps

DISCUSSION OF PALEOREEF MAPS


We selected five examples to exhibit the potential of paleoreef maps for interpreting Phanerozoic
reef distributions. We briefly describe the maps
with respect to the latitudinal extent of the reef
belt in relation to paleoclimate, regional reef concentrations, proposed upwelling, and selected reef
attributes. More detailed interpretations of the
maps are hampered by the lack of sophisticated
paleo-oceanographic models. In the Late Triassic
maps, we additionally exemplify our paleogeographic reconstructions in detail. The evolution of
particular features in a wider context is referred to
in the subsequent section.
Late MiddleLate Devonian
The global proliferation of reefs during this time
(Late MiddleLate Devonian (Givetian-Frasnian,
slice 9; Figure 6) (Moore, 1989; Copper, 1994) is
well represented in the database. A total of 245
reefs were recorded. The diversity of reef types is
relatively low. Reef mounds with stromatoporoids
and tabulate/rugosan corals as the main constituents are dominant. Reefs are approximately
equally abundant on both hemispheres. The
extraordinarily warm climate of this time (Dickins,
1993) is reflected by reefs occurring in abnormally
high latitudes. The belt of reef development
extends from around 50N (Amur region and Sette
Daban Range, east Siberia) (Bolshakova et al.,
1994) to 42S. The southernmost buildups are
Givetian mud mounds of Algeria (Wendt et al.,
1997). The southernmost reef mounds dominated
by stromatoporoids or corals are situated in Morocco (Cattaneo et al., 1993) and in central Afghanistan (Mistaen, 1985).
There are six centers of reef development: (1)
Siberian platform, Russia, (2) Timan-Pechora
Province, Russia, (3) Alberta, Canada, (4) central
Europe, (5) Morocco, and (6) south China; however, from Figures 3 and 6 it becomes obvious that
although there are large numbers of reefs, there is
not a proportionally large number of reefs with
good data. This discrepancy is partly attributed to
the fact that many Devonian reefs are located in
Russia, where data access is difficult, and because
many GivetianFrasnian reefs occurring in the subsurface of Canada and Russia are known only
through seismic exploration and drilling.
The abundance of thick reef complexes is outstanding (Figures 5, 6). More than 30% of all recorded reefs are producing reservoirs or may be regarded as having reservoir quality. The major sites of
hydrocarbon production are Alberta, Volga-Urals,
and Timan-Pechora.

Many of the GivetianFrasnian reefs are located


at or close to sites with more than 70% probability
of coastal upwelling (Golonka et al., 1994). Reefs
along the former northeastern margin of the
Siberian craton are situated in a belt with predicted
upwelling in the Northern Hemisphere Summer.
Coastal upwelling also is predicted along the western coast of Australia, where the famous Canning
basin barrier reef developed (Playford, 1980), in
the Moravian karst region, and along the former
western margin of the south China block. We conclude that GivetianFrasnian reefs could thrive in
areas in or close to coastal upwelling. This position
may reflect the ability of Devonian reef builders to
cope with higher nutrient levels than modern reefs
(Hallock and Schlager, 1986; Wood, 1993, 1995).
An alternative explanation could be the barred
basin model of Whalen (1995) that emphasized the
role of tectonic barriers in deflecting cool eastern
boundary currents and preventing high nutrient
concentrations in the barred basins. Although this
model may apply for south China, Australia, and
North America, the direct association of reefs and
upwelling in the Pechora-Urals and Moravian karst
regions confirms that GivetianFrasnian buildups
could grow in water with high nutrient concentration owing to an autecology of reef builders different from builders of modern reefs. Considering the
great reservoir potential of Frasnian reefs, it is not
surprising that their distribution is closely linked
with source rock distribution (compare Klemme
and Ulmishek, 1991).
Early Permian (Asselian, Slice 13)
We recorded a total of 74 reefs in the Early
Permian (Asselian, slice 13; Figure 7) time slice. Most
of them are reef mounds and mud mounds rather
than true reefs. Algae (mostly phylloid algae and
Palaeoaplysina) dominate most of the buildups.
More than one-half of the reefs have a reservoir
potential or are producing reservoirs (Figure 15).
More than 30% of the buildups were situated below
the fair-weather wave base (Figure 14b).
The main locations of reef growth were (1) Urals
and Timan-Pechora Province, Russia, (2) Barents
Sea north of Norway, and (3) Texas and New
Mexico. Buildups with reservoir quality occur especially in Russia.
The Asselian time slice is special in that most of
the reefs are situated in the Northern Hemisphere,
although most of the marine shelves are located in
the Southern Hemisphere. Reef mounds and mud
mounds are present to latitudes greater than 40N
on the Northern Hemisphere, but do not extend
beyond 15S in the Southern Hemisphere. This pronounced asymmetry in reef distribution (Figure 12)

Figure 7Global distribution of Asselian reefs. Ocean surface currents were derived from the plate tectonic configuration. 1 = Mountains, 2 = land, 3 = shelf,
4 = deep water, 5 = predicted upwelling zones, slightly modified from Golonka et al. (1994), 6 = reefs with reservoir quality, 7 = reefs without or unknown
reservoir quality. Continental ice sheets are not indicated.

Figure 8Global distribution of NorianRhaetian reefs. Ocean surface currents were derived from the plate tectonic configuration. 1 = Mountains, 2 = land, 3 =
shelf, 4 = deep water, 5 = predicted upwelling zones, slightly modified from Golonka et al. (1994), 6 = reefs with high debris potential, 7 = reefs with moderate
debris potential, 8 = reefs with low debris potential, 9 = reefs without data indicating debris potential. The majority of reefs are situated in Mediterranean
Tethys. See Figure 9 for a magnification of the western Tethys.

Kiessling et al.

1567

Figure 9Distribution of Tethyan reefs in the NorianRhaetian time slice with reef numbers. 1 = Nondepositional
landmasses (orange color indicates topographic highs), 2 = terrestrial depositional environment (undifferentiated),
3 = fluvial-lacustrine depositional environment, 4 = coastal, transitional, marginal-marine environment, 5 =
shallow-sea, shelf environment, 6 = slope environment, 7 = deep ocean basins with sediments, 8 = deep ocean
basins with little or no sediments (primarily oceanic crust), 9 = sandstone/siltstone, 10 = shale, clay, mudstone, 11 =
carbonates, 12 = evaporites, 13 = mixed sandstone/shale, 14 = mixed carbonate/shale, 15 = dolomites, 16 = red beds,
17 = intrashelf, intraplatform reefs, 18 = platform or shelf margin reefs, 19 = slope, ramp, and basin reefs, 20 = reefs
with no data on paleoenvironment, 21 = reef number in the database, 22 = oceanic spreading center and transform
faults, 23 = subduction zone. Ad = Adria, Ib = Iberia, Ki = Kirsehir and Sakariya, Lu = Lut, Mo = Moesia, Pi = Pindos
ocean, Si = Sicily, SM = Serbo-Macedonian terrane, Ti = Tisa, Ta = Taurus.

is likely related to climatic asymmetry during the


Gondwana glaciation. According to Dickins (1993),
the Asselian was the time of maximum Southern
Hemisphere glaciation, whereas there is no evidence for ice in the Northern Hemisphere; however, the different composition of high-latitude

mounds in the Northern Hemisphere points to a


pronounced climatic zonation even in the ice-free
part of the Early Permian. The reef-builder Palaeoaplysina, regarded herein as an enigmatic alga
(Watkins and Wilson, 1989), exemplifies this difference. Palaeoaplysina was suggested to represent a

1568

Paleoreef Maps

cool-water dweller (Davies and Nassichuk, 1973),


but new data provide evidence for Palaeoaplysina
favoring moderate temperatures (Ritter and Morris,
1997). Mounds with Palaeoaplysina are found
north of 25N paleolatitude in the Urals, Svalbard,
and along the northern margin of North America,
but extended down to 15N at the western margin
of North America. The low-latitude occurrences are
likely related to relatively cool eastern boundary
currents along the Panthalassa margin.
Considering the high reservoir potential of
Asselian reefs (Figure 7), it is amazing that they usually are far from modeled coastal upwelling areas;
however, many reefs are located close to mountain
areas (central Pangean mountain range, protoUrals). Continental weathering may have led to
high nutrient concentrations in those settings. Reef
distribution is closely linked with source rock distribution (Klemme and Ulmishek, 1991) in this
interval.
Late Triassic (NorianRhaetian, Slice 18)
The Norian supersequence corresponds with a
low first-order sea level stand at a time of high continental emergence. The supercontinent Pangea
began its breakup. Early greenhouse conditions
prevailed (Frakes et al., 1992). The MiddleLate
Triassic was characterized by a pronounced monsoonal climate (Mutti and Weissert, 1995). There is
no evidence of significant continental glaciation.
A total of 117 reefs were recorded in the Late
Triassic (NorianRhaetian, slice 17; Figures 8 and 9)
time slice. The diversity of reef types is moderate.
Most of the NorianRhaetian reefs and reef mounds
are characterized by a predominance of scleractinian corals or coralline sponges (Flgel, 1981, 1994).
The Norian reefs of the Mediterranean Tethys are
characterized by a high debris potential (Figure 8)
and a high amount of platform/shelf margin reefs.
Compared to the ScythianCarnian time slice
(Figure 12, slice 16), there is a significant spatial
expansion of reefs. This is true for the advance
toward higher latitudes, as well as the augmentation along the western margin of North America.
The northern boundary of reef growth is demarcated by occurrences in the Northern Caucasus, in
Sichote Alin, and in Japan at around 40N. The
southernmost occurrences of reefs were located at
38S (Chile and Papua New Guinea). Nearly all
shelf areas between 40N and 40S encompassed
reefs or at least reef mounds, but there appears to
be a gap at the equatorial Southern Hemisphere. As
calculated from the database, the diversity of reef
builders obviously is much more affected by paleogeographic setting (comparison of eastern paleoPacific and Tethys) than by latitudes.

Based on data of Kristan-Tollmann (KristanTollmann and Tollmann, 1981, 1982), Stanley (1988,
1994) developed a Late Triassic reef distribution
map. Although Stanleys map showed a pattern similar to that of our map, Stanleys map uses an obsolete
paleogeography and places much emphasis on reefs
located on Panthalassa terranes. Although we agree
that many Late Triassic reefs were situated on
seamounts and terranes, which later accreted at
North America, our reconstructions suggest that
they were located much closer to the American margin than is proposed by Stanley (1988).
Reefs appear close to predicted upwelling sites
of Golonka et al. (1994) along the western margin
of North America and South America and along the
former northern margin of Australia. The predicted
upwelling for North America may not be real
because east Pacific terranes that could have
deflected eastern boundary currents were not considered in the model of Golonka et al. (1994); however, reefs in South America and along the
Australian margin are likely to be associated with
upwelling. Reefal carbonates in Peru are mostly
limited to small-scale coral-sponge biostromes
(Stanley, 1994), but there are true reefs in Timor
(Vinassa de Regny, 1915) and offshore northwest
Australia (Rhl et al., 1991; Colwell et al., 1994).
Figure 9 provides an example of a regional paleogeographic map combined with paleoreef positions, allowing a more detailed evaluation. In the
western Tethys, Late Paleozoic and Triassic rifting
and sea-floor spreading resulted in several separated carbonate platforms (Dercourt et al., 1993;
Philip et al., 1996; Golonka and Gahagen, 1997).
The large amount of western Tethyan reefs was distributed on a large carbonate platform that existed
during most of the Mesozoic, spreading from
Apulia through the Ionian and Hellenide terranes to
the Taurus zone (Dercourt et al., 1993). This zone
was connected with the AlpineInner Carpathian
area, which forms the marginal carbonate platform
of Europe. The AlpineInner Carpathian area contains abundant reefs. The narrow branch of neoTethys recorded in the deep-water sediments of
Sicily (Kozur, 1990; Catalano et al., 1991), Lago
Negro (Marsella et al., 1993), and Crete (Kozur and
Krahl, 1987), as well as in the Mammonia ophiolite
complex in Cyprus (Robertson and Woodcock,
1979; Morris, 1996), separated the Adria-Taurus
platform from the African continent. The incipient
Pindos ocean separated the Pelagonian, Sakhariya,
and Kirsehir carbonate platforms from the IonianTaurus platform (Robertson et al., 1991, 1996;
Stampfli et al., 1991). The Lut (Iran) carbonate platform with numerous reefs belongs to the Cimmerian
continent, separated from Gondwana by the main
neo-Tethyan oceanic branch (Sengr, 1984; Sengr
et al., 1984; Ricou, 1996).

Kiessling et al.

The western Tethys reefs exhibit distinct patterns (Figure 9; Table 3) in terms of tectonicsedimentary patterns, the frequency of the reefs,
the types of principal reef builders, and the distribution of reef biota. Reefs formed (1) on marginal
carbonate platforms close to the Eurasian continent
(northwestern Caucasus, western Carpathians, and
northern calcareous Alps), (2) on various separated
carbonate platforms facing deeper water basins
(southern Alps, Apennines, Sicily, Greece, southern
Turkey), and (3) in restricted shelf environments
(southern Spain).
Reefs are concentrated in an inner shelf zone parallel to the northwestern coast of the Tethys and
extending a length of several hundred kilometers,
and within a region consisting of separated platforms
and comprising an area of about 1.8 106 km2.
NorianRhaetian reefs include coralline sponge
reefs, coralline spongecoral reefs, coral reefs, dasycladacean algal reefs, serpulid-carbonate cement
reefs, and microbial reefs (Table 3). The last three
types, known from southern Spain, the Apennines,
and the southern Alps, indicate particular environmental conditions excluding the regular reef
builders. Coralline sponge reefs and coral reefs
occur in close association, but coral reefs seem to
be more common in the inner shelf zone.
The comparison of the taxonomic composition
of the coral and coralline sponge faunas indicates
relations between the reefs situated in the northwestern Tethys (northern Alps) and those connected with the separated platforms. Reefs on separated platforms, however, yield a high number of
endemic taxa. In contrast, there are significant
paleontological differences considering the biotic
composition of reefs formed on the Cimmerian
continent (Lut block, Iran).
Platform margin reefs (e.g., Dachstein-type reefs;
northern and southern Alps, Sicily) are distinctly
thicker than intraplatform reefs that are usually less
than 50 m thick (Table 3).
Late Jurassic (OxfordianKimmeridgian)
There are currently 164 reefs considered in the
Late Jurassic (OxfordianKimmeridgian; Figure 10)
time slice. The Late Jurassic is marked by high
diversity of reef types (Leinfelder, 1994; Leinfelder
et al., 1996), including thrombolitic mounds, siliceous sponge mounds, and coral reefs. Reefs of all
types are found in Europe, where reefs are most
widespread in this time slice. Reefs outside Europe
are mostly coral-dominated reefs and biostromes.
There is seismic evidence for a major reef trend
along the eastern shelf of North America, but most
reefs in this trend are poorly known. Rare wells indicate small-scale stromatolite-Tubiphytes bioherms

1569

and thrombolitic mounds rather than coral reefs in


the OxfordianKimmeridgian (Ellis et al., 1985;
Pratt and Jansa, 1989).
In contrast to the Late Jurassic reef distribution
map of Leinfelder (1994), we do not indicate any
reefs in western and northwestern North America.
According to Beauvais (1992), there are only scattered coral occurrences that can hardly be called
reefs. With the exception of this discrepancy, our
distribution plots are similar.
Bioherms can be found in all latitudes between
45S (Neuquen basin, Argentina) (Legarreta, 1991)
and 52N (Japan). The paleogeographic data for the
Japanese reefs, however, are rather doubtful and
the reefs may have originated in much lower latitudes. Thus, the northern limit of reef distribution
is marked by occurrences in Britain, Poland, and
Uzbekistan at around 40N. The high paleolatitude
(54S) reefs of Patagonia, Argentina, reported by
Ramos (1978) are assigned to the Tithonian
Berriasian time slice (R. Scasso, 1997, personal
communication).
Scleractinian-dominated reefs could thrive in
higher latitudes in the Southern Hemisphere than
in the Northern Hemisphere. Although this is in
agreement with an asymmetrical temperature distribution on both hemispheres (Kiessling and
Scasso, 1996), it is puzzling that Argentinean reefs
could exist in a setting that should have been influenced by cool eastern boundary currents. This
paradox is best explained by the barred basin
model of Whalen (1995), that is, the proto-Andean
arc may have acted to deflect eastern boundary
currents.
On a global scale, warm surface water (more
than 18C in winter) can be assumed between
40N and 55S paleolatitude. This statement is supported by other paleoclimatic data (Vakhrameev,
1991; Ditchfield et al., 1994; Hallam, 1994;
Ditchfield, 1997). We agree with Leinfelder (1994)
that the Upper Jurassic reef distribution reflects a
maritime greenhouse effect.
Late Jurassic upwelling sites were suggested
through paleoclimatic modeling by Parrish and
Curtis (1982), Ross et al. (1992), Golonka et al.
(1994), and Price et al. (1995). Reefs generally
are rare in predicted upwelling areas; however,
some reefal areas are closely associated with the
predicted upwelling centers of Golonka et al.
(1994) that include intrashelf coral biostromes in
Saudi Arabia (El-Asaad, 1991), diverse reef types
in the Caucasus and Crimea (Kuznetsov, 1993;
Michailova, 1968), and coral reefs in Amu-Darya
(Fortunatova et al., 1986). The actual source rock
distribution (Klemme and Ulmishek, 1991) is
coincident with some reef occurrences in the
Gulf of Mexico, in the Amu-Darya region, and in
Arabia.

1570

Paleoreef Maps

Table 3. List of Upper Triassic Reefs Evaluated in the Mediterranean Tethys*


Number
1
4
5
6
7
9
11
12
13
16
17
19
21
25
26
28
32
40
44
45
72
75
76
77
80
83
84
92
93
94
101
102
106
119
120
137
138
150
155
169
172
179
208
212
235
246
268
343
359
396
474
1494
1720
1721
2136
2156
2191
2192
2336
2342
2343
2344
2348
2353

Name
Tilkideligi Tepe, Turkey
Aksu, Turkey
Adnet, Salzburg, Austria
Rtelwand, Austria
Rio Blanco, Spain
Grimming, Austria
Argolis, Greece
Feichtenstein, Austria
Gruber, Austria
Wilde Kirche, Austria
Hochschwab, Austria
Pokljuka, Slovenia
Hohe Wand, Austria
Mala Fatra, Slovakia
Panormide, Sicily, Italy
Gosaukamm, Austria
Gesuse, Austria
Steinplatte, Austria
Hochknig, Austria
Luda-Kamcia, Bulgaria
Begunjscica, Slovenia
Marawand, Delijan, Iran
Lakaftari, Esfahan, Iran
Tabas, Iran
Hoher Gll, Bavaria, Germany
Waliabad, Iran
Monte Genuardo, Sicily, Italy
Dereky, Turkey
Triglav, Slovenia
Vascau, Romania
Gela, Sicily
Rhtikon, Switzerland
Lattari Mountains, Italy
Topuk, Turkey
Rahatalana Yayla, Turkey
Karakorum, Kashmir
Northern Kaukasus, Russia
Yesilova, Turkey
Cyprus
Zdial Plateau, Slovakia
Liptovska Osada, Slovakia
N-Calabria, Italy
Val Adrara, Italy
Korfu, Greece
Monte Lieggio, Salerno, Italy
Madonie, Sicily
Naybandan, Iran
Kocagedik, Turkey
Durmitor, Montenegro
Pico de la Carne, Spain
Epidauros, Greece
Vapa, Zlatibor, Serbia
Artavaggio, southern Alps, Italy
Vello, southern Alps, Italy
Bobrovcek, High Tatra, Slovakia
Albenza, Southern Alps
Meimeh, Delijan, Iran
Kuhbanan, Kerman, Iran
Simferopol, Ukraine
Monte Cetona, Tuscany, Italy
Ponte Arverino, Umbria, Italy
Monti Simbruini, Apennines, Italy
Salzbrunnen, Bagherabad, Iran
Mahallat, Iran

Stage
Norian
Norian
Rhaetian
Rhaetian
Rhaetian
Upper Norian
NorianRhaetian
Rhaetian
Rhaetian
Rhaetian
Upper Norian
Upper Norian
Upper Norian
Rhaetian
Upper Norian
Upper Norian
Upper Norian
Rhaetian
Upper Norian
Norian
Rhaetian
NorianRhaetian
Middle Norian
NorianRhaetian
Upper Norian
Rhaetian
NorianRhaetian
Norian
Upper Norian
Norian
Norian
Rhaetian
Rhaetian
Norian
Upper Norian
Lower Norian
Norian
Upper Norian
CarnianNorian
Norian
Norian
Norian
Rhaetian
Rhaetian
Norian
Upper Norian
Norian
CarnianNorian
Norian
Norian
Norian
Norian
Upper Norian
Upper Norian
Rhaetian
Rhaetian
NorianRhaetian
NorianRhaetian
Norian-Rhaetian
Rhaetian
Rhaetian
Norian-Rhaetian
Rhaetian
Norian

*Upper Triassic reefs are time slice 18; see also Figure 9.
**Thickness data: 1 = <10 m; 2 = 10-100 m; 3 = 100-500 m; 4 = >500 m.

Type
Reef
Reef
Reef
Reef
Reef mound
Reef
Reef
Reef
Reef mound
Reef
Reef
Reef
Reef
Biostrome
Reef
Reef
Reef
Biostrome
Reef

Reef
Reef
Reef
Reef mound
Reef
Reef mound
Reef
Reef
Reef
Reef mound
Reef
Reef mound
Mud mound
Reef
Reef
Reef
Reef
Reef
Reef
Reef
Reef
Reef
Biostrome
Reef
Mud mound
Reef
Biostrome

Reef
Biostrome
Reef
Reef mound
Reef
Reef
Biostrome
Reef
Reef
Biostrome
Reef
Mud mound
Mud mound
Reef
Reef
Reef mound

Thickness**
2
2
3
3
2
3
3
3
3
2
1
3
2
1
3
3
3
1
3

2
2
2
2
4
2
3
2
3
2
3
1
1
2
1
3
2
3

2
3

1
1
1
3
2
3
3
2
2
2
1
1
1
1
2
1
2

1
2
1
2

Dominant Reef
Builder
Coralline sponges
Coralline sponges
Corals
Corals
Serpulids
Coralline sponges
Coralline sponges
Corals
Corals
Corals
Corals
Corals
Coralline sponges
Corals
Coralline sponges
Coralline sponges
Coralline sponges
Corals
Corals

Coralline sponges
Coralline sponges
Corals
Corals
Corals
Coralline sponges
Coralline sponges
Coralline sponges
Corals
Corals
Algae
Corals
Microbes
Coralline sponges
Corals
Corals
Coralline sponges
Coralline sponges
Corals
Corals
Corals
Corals
Corals
Corals
Serpulids
Coralline sponges
Coralline sponges
Corals

Algae
Coralline sponges
Corals
Serpulids
Serpulids
Corals
Corals
Coralline sponges
Corals
Corals
Serpulids
Microbes
Serpulids
Corals
Corals

Figure 10Global distribution of OxfordianKimmeridgian reefs. Ocean surface currents were derived from the plate tectonic configuration. 1 = Mountains,
2 = land, 3 = shelf, 4 = deep water, 5 = predicted upwelling zones, slightly modified from Golonka et al. (1994), 6 = intrashelf/intraplatform reefs, 7 = platform
or shelf margin reefs, 8 = slope, ramp, and basin reefs, 9 = reefs without data indicating paleoenvironment.

1572

Paleoreef Maps

Middle Miocene (BurdigalianSerravallian)

EVALUATION OF PALEOGEOGRAPHICAL REEF


DISTRIBUTION THROUGH TIME

During the early to middle Miocene (Burdigalian


Serravallian; Figure 11), reefs achieved an almost
modern-type global distribution pattern, but
Miocene reefs in the Mediterranean area and the
para-Tethys extended beyond the modern coral
reef distribution. Framework-dominated coral-algal
buildups represent the great majority of reefs.
Although the map depicts 146 reefs, it is still considered incomplete. Reef data probably are missing
in the present-day Indian Ocean and some Pacific
atolls owing to poor knowledge of subsurface geology; however, the proposed distribution of
earlymiddle Miocene coral reefs of Franseen et al.
(1996) appears to be exaggerated. Judging from the
strong similarity with the distribution map of
Miocene shallow-water carbonates (Sun and
Esteban, 1994), Franseen et al. (1996) obviously
lumped reefal and platform carbonates in their plot.
According to Jordan et al. (1990), Miocene reefs
existed in latitudes between 27S and 48N. Our
map, however, shows a much wider extent of reef
distribution (42S to 50N). Fear y and James
(1995) provided seismic evidence for the existence of a nearly 500-km-long middle Miocene barrier reef along the southern margin of Australia;
however, little is known about the composition of
this reef trend. The northern limit of reef occurrences is marked by algal-vermetid reef mounds in
Poland (Pisera, 1985). Other reefs of the northern
para-Tethys are composed mostly of bryozoans,
coralline algae, serpulids, and sessile foraminifera
(Pisera, 1996). Reefs with corals playing a significant role in reef construction are found up to 47N
(Dullo, 1983).
Sun and Esteban (1994) recognized two end
members of paleoclimatic and depositional setting
of Miocene reefs: (1) humid, oceanic tropicalsubtropical settings and (2) arid, land-locked temperatesubtropical settings. Reefs in southeast Asia and the
western Pacific belong in the first category. These
reefs are mostly diverse coral-algal reefs, and many of
them form important hydrocarbon reservoirs.
Another important center of reef development was
situated in the Caribbean, which can be described as
a transitional setting (Sun and Esteban, 1994). The
Mediterranean area and incipient Red Sea form the
third center of middle Miocene reef development.
These reefs can be assigned to the arid, land-locked
depositional setting.
Considering the modern aspect of most middle
Miocene reefs, it is not surprising that almost all of
them were outside the upwelling centers modeled
by Parrish and Curtis (1982) and Golonka et al.
(1994). Actual source rock occurrences (Klemme
and Ulmishek, 1991) are virtually never accompanied by coral-algal reefs.

The reef distribution on the paleogeographic


maps can be compared with modern reef distributions to evaluate the differences. There are two
major deviations: (1) variable paleolatitudinal concentrations of reef growth and (2) variable amounts
of reefs on the eastern margins of the oceans.
The paleolatitudinal distribution of reefs was different during most of the Phanerozoic from the
almost symmetrical pattern of Holocene zooxanthellate coral reefs that are restricted to a belt
between 33N and 31S. The average paleolatitudinal site of reefs (mean of all reef paleolatitudes
in one time slice) changed significantly (Figure
12). Throughout the Phanerozoic a principal shift
from the Souther n Hemisphere toward the
Northern Hemisphere occurred. This trend follows the general northward drift of continents
during the Phanerozoic. The average paleolatitude
of Cambrian reefs was at 18S. Interestingly, there
were very few reefs in the Cambrian equatorial
region (10S to 10N). In the latest Cambrian and
Early Ordovician, a strong shift toward the equator
is evident. With the exception of small mounds in
the San Juan Province of Argentina (Armella, 1994),
all reefs were situated in low paleolatitudes. In
the MiddleLate Ordovician, the average paleolatitude of reefs returned to higher southern latitudes
(around 15S) before another shift toward the
equator occurred in the Silurian. After a return to
relatively high southern paleolatitudes in the
Lochkovian, the average position of reefs shifted
finally toward the Equator during the Devonian
and Early Carboniferous. The Frasnian time slice
(Figure 12, slice 10) represents one of the rare
periods in the Phanerozoic when the reefs were
almost symmetrically arranged on both hemispheres. This is especially remarkable because
most of the continental shelves were still situated
in the Southern Hemisphere at that time. The
mean paleolatitude of reefs shifted to the
Northern Hemisphere in the Moscovian and never
returned back to the Southern Hemisphere. This
major shift to the north was completed in the
Asselian (Early Permian), when most of the reefs
were concentrated in the interval from 30 to
40N (Figure 12, slice 14). After a southward shift
during the Middle and Late Permian, a stepwise
movement of main reef localities to higher northern latitudes can be observed. In the Oxfordian
Kimmeridgian time slice, the average paleolatitude of reefs was 25N. After a backstepping
toward the equator, the mean paleolatitude
reached its highest Phanerozoic value in the latest
Cretaceous (nearly 30N). The average paleolatitude of reefs remained at relatively high northern

Figure 11Global distribution of Serravallian (middle Miocene) reefs. Ocean surface currents were derived from the plate tectonic configuration. 1 = Mountains, 2 = land, 3 = shelf, 4 = deep water, 5 = predicted upwelling zones, slightly modified from Golonka et al. (1994), 6 = true reefs, 7 = reef mounds, 8 = mud
mounds/banks, 9 = biostromes, 10 = reefs without data indicating type. Continental ice sheets are not indicated.

Figure 12Average paleolatitudinal sites of reefs from the Early Cambrian to the late MiocenePliocene. The concentration of reefs is largely related to the distribution of large shelf areas, but oceanographic asymmetries also are important. There is a concentration of Late Carboniferous and Early Permian reefs on
the Northern Hemisphere, although the majority of shelf areas was situated in the Southern Hemisphere (Figure 7). This asymmetry is likely related to the
Gondwana glaciation.

Kiessling et al.

latitudes in the Tertiary, reaching 25N in Ypresian,


but only 11N in Lutetian.
Flgel (1994) analyzed the changing paleolatitudinal distribution of Pangean (Carboniferous
Jurassic) reefs. His pattern exhibits the same trend,
but owing to the limited stratigraphic resolution
(stratigraphic systems), he was not able to observe
the MoscovianArtinskian Northern Hemisphere
excursion.
We mentioned that the general trend of changing reef distribution on both hemispheres is strongly related to the distribution of continents and shelf
areas. The more intense geologic research in the
Northern Hemisphere biases the concentration of
reefs in the database, thereby influencing the pattern depicted in Figure 12; however, rapid changes
in the paleolatitudinal distribution are hardly
explained by plate tectonic configuration or heterogeneous distribution of data. Asymmetries of climate and surface current properties also likely had
a major impact on the pattern, and is especially
true for the late Paleozoic Gondwana glaciation
that caused a jump of the reef locations toward the
Northern Hemisphere.
The paleogeographic sites of major reef development also indicate significant oceanographic
changes and reflect the changing ecology of reef
builders. Before the Middle Triassic, reefs were not
concentrated at the western margins of large
oceans. Instead, numerous reefs developed on the
eastern margin of Panthalassa and along relatively
shallow intracontinental seaways. A longitudinal
distribution similar to that of the Holocene was not
achieved before the Oligocene.
PALEOGEOGRAPHIC ASPECTS
Based on the knowledge of the pattern discussed, we now discuss the applicability of reefs
for paleogeographic reconstructions; however, this
approach must be taken with care because we
move into the dangerous grounds of circular reasoning by trying to both derive a better paleogeography and evaluate the spatial reef distribution
through time from the database. The evaluation of
paleolatitudes by using a uniform approach must
consider different paleoclimatic conditions and the
possibility of cool-water adapted reefs. Even today,
the main reef belt is situated in the tropics and subtropics, but particular reef types (e.g., Lophelia
mounds) are found up to the polar circle (Teichert,
1958; Henrich et al., 1996); hence the classification
of reef attributes is crucial for using reefs as paleolatitudinal indicators.
In some time slices there is a distinct difference in
the composition of high-latitude reefs compared
with low-latitude reefs (e.g., Asselian, Figure 7). In

1575

these time slices, reefs of a particular composition


are likely to indicate water temperature and thus
may give a hint to paleolatitude. In other time slices
(e.g., Norian, OxfordianKimmeridgian), reefs at or
beyond the boundary of modern reef distribution do
not show pronounced compositional differences
compared with lower latitude reefs in the same
paleoenvironment. In these cases it is very difficult
to use reefs as paleolatitudinal indicators. The longitudinal setting (western vs. eastern margin of
oceans) commonly had a more pronounced impact
on reef composition and biodiversity than the latitude. This fact additionally restricts the application
of reefs for paleolatitudinal reconstructions, but at
the same time enhances the possibilities for terrane
reconstructions (Belasky and Runnegar, 1993).
Similar restrictions apply for the paleoclimatic interpretation of reef distribution maps. Icehouse time
slices do not exhibit a more equatorial concentration of reefs than greenhouse time slices.
From the global pattern depicted in our reef distribution maps (Figures 611) and associated discussions, we are able to argue that the plate tectonic
reconstruction of some areas needs to be revised.
This is true for the Japanese archipelago in the
TriassicJurassic and for parts of far eastern Russia
(e.g., Koryak and Sikhote Alin) in the Paleozoic. A
better knowledge of Devonian reefs in Russia would
help to judge the validity of the high paleolatitudes
assigned to many Russian microcontinents (Figure 6).
VARIATION OF SELECTED REEF ATTRIBUTES
THROUGH TIME
The analysis of secular variation in reef features
was done by plotting diagrams and conducting
statistical tests. Owing to the rather coarse classification intervals of many reef features, only distinct
differences are reflected in the diagrams. In this
paper, we limit the discussion to major biotic
composition, bioerosion, bathymetr y, debris
potential, and reservoir potential of Phanerozoic
reefs.
If the dominant reef-building group in each reef
is plotted on a diagram (Figure 13), the typical biotic composition of Phanerozoic reefs is relatively
well represented. Similar to James and Bourque
(1992), we recognize more reef cycles than had
been previously (e.g., James, 1983) suggested.
Following is a list of reef attributes through time.
Seven reef cycles are evident.
(1) CambrianEarly Ordovician: Reef mounds
and microbial mounds, usually dominated by
microbes/stromatolites. The climax of reef development was reached in the Early Cambrian through
the significant contribution of archaeocyathids to
reef building.

1576

Paleoreef Maps

Figure 13Dominant
Phanerozoic reef types
and reef builders.
The curves to the left
indicate the cumulative
number of reefs and reef
mounds and the number
of mud mounds and
biostromes in each time
slice. Horizontal bars on
the right depict the
cumulative number of
reefs in which a particular
fossil group is dominant.
Others refers to
brachiopods,
pelmatozoans, and
foraminifera. Seven
Phanerozoic cycles of
reef building are evident
(large numbers).
Major mass extinctions
are demarcated by
starred lines.

(2) Middle OrdovicianLate Devonian: Reef


mounds and reefs, usually dominated by stromatoporoids and tabulate or rugose corals. Bryozoans are
additionally important in the Ordovician and Silurian.
The climax of reef development was reached in the

GivetianFrasnian. In comparison with the graphs of


James (1983) and James and Bourque (1992), we note
a significant decrease of reef abundance in the Early
Devonian (compare also Talent, 1988). This cycle was
terminated by the FrasnianFamennian reef crisis.

Kiessling et al.

(3) Latest DevonianEarly Permian: Reef mounds


and mounds, usually dominated by algae, microbes,
and bryozoans. There is no distinct climax of reef
development in this cycle.
(4) Middle PermianMiddle Triassic: Reef
mounds, mounds, and reefs, usually dominated by
pharetronid (calcareous) sponges, microbes, or
corals. Although this cycle is interrupted by the
PermianTriassic extinction, middle and Late
Permian reefs do not differ significantly from
Middle Triassic reefs with respect to their high-rank
taxonomic composition and reef attributes (verified by cluster analysis).
(5) Late Triassicearliest Cretaceous: Reefs, reef
mounds, mounds, and biostromes usually dominated by scleractinian corals. Again, this cycle is interrupted by a major extinction (TriassicJurassic
boundary). The Late Triassic is transitional between
cycles 4 and 5, but cluster analysis showed that it is
more closely related to cycle 5. The Early Jurassic is
exceptional owing to the great number of bivalve
bioconstructions. The climax of reef development
was reached in the Late Jurassic.
(6) Cretaceous: Biostromes and reef mounds usually dominated by rudist bivalves or scleractinian
corals. An indistinct climax was reached in the late
Early Cretaceous. This cycle was terminated by the
CretaceousTertiary extinction event, which led to
the total extinction of rudists.
(7) Tertiary: Reefs, usually dominated by scleractinian corals, with significant contribution by
coralline red algae. The climax of reef development
was reached in the middle Miocene.
Reefs with evidence of bioerosion are underrepresented in the database because bioerosion is
rarely described or mentioned in publications, even
though it may be present; however, the observed
pattern (Figure 14a) may give some hint to actual
variations in the intensity of bioerosion. We emphasize bioerosion because it is thought to be an important factor in reef evolution (Vogel, 1993), and the
intensity of bioerosion may indicate past nutrient
levels (Hallock, 1988). The oldest reefs with evidence of bioerosion are known from the Early
Cambrian (James et al., 1977; Gandin and Debrenne,
1984). During most of the Paleozoic, few reefs
(<10%) showed evidence of bioerosion. The most
striking increase in bioerosion is evident after the
PermianTriassic extinction. This expansion of bioerosion is likely to reflect the real situation and can
be explained by the Mesozoic diversification of
predators [Mesozoic marine revolution of Vermeij
(1977)]. The abundance of reefs with evidence of
bioerosion commonly is in marked contrast to the
diversity of macroborings, e.g., macroborings greatly
diversified in the Devonian (Kobluk et al., 1978),
but fewer than 10% of the reefs recorded in the
database show evidence of bioerosion; however,

1577

this observation is biased by generally few descriptions of bioerosion in reefs.


The percentage of reefs formed below fairweather wave base is relatively high between the
Devonian and the Triassic (Figure 14b). The most
prominent peak is in the Tournaisianearly Visean
time slice, where 65% of the reefs grew in deepwater environments. The Permian is another period with a considerable number of deeper water
reefs (up to 31%). In all other time slices 325% of
the reefal structures evidently formed below wave
base. Note that after the FrasnianFamennian reef
crisis most of the subsequent reefs grew in deeper
water, whereas after the TriassicJurassic crisis considerably fewer reefs formed below wave base.
The average debris potential of reefs varies considerably (Figure 14c), but there is a significant
increase through the Phanerozoic (r = 0.77, <
0.001). After some debris potential of EarlyMiddle
Cambrian reefs, there was a sharp decline until the
Middle Ordovician. During the Silurian and
Devonian, the mean debris potential achieved its
highest Paleozoic values. The entire Carboniferous
and most of the Permian are characterized by relatively low debris potential. Late PermianLate
Triassic reefs had about the same mean potential
to produce debris as the SilurianDevonian reefs.
After the latest Triassic reef crisis, the average
debris potential was low (predominance of
Lithiotis mounds). During the Jurassic until the
earliest Cretaceous, debris potential increased
rapidly and continuously. An opposing trend is
obser ved during the Cretaceous and early
Paleogene. A significant increase in debris potential is noted in the Lutetian, followed by a decline
in the OligoceneMiocene.
The relative amount of debris produced by a reef
depends on different parameters, such as water
energy, topography, mineralogy, architecture of
framework, and bioerosion. We used the database
to explore the relationships among the different
reef features and debris potential. A simplified correlation matrix of normalized values is depicted in
Table 4. This table indicates that a significant correlation exists between debris potential and nearly all
quantitative parameters of the database. High
debris potential is more likely to occur in reefs
with great thickness, high diversity of reef builders,
little micrite content, evidence of bioerosion, and
in shallow water and low latitude settings.
REEFS AS RESERVOIRS
Carbonates and especially reefs and reefal sediments form important hydrocarbon reservoirs
(Roehl and Choquette, 1985; Greenlee and
Lehmann, 1993; Kuznetsov, 1997). Roehl and

1578

Paleoreef Maps

Figure 14Comparison of bioerosion, bathymetric setting, and debris production of reefs through time. (a) Percentage of reefs with evidence of bioerosion in each time slice; (b) percentage of reefs grown below fair-weather
wave base in each time slice; (c) average debris potential of reefs in each time slice (1 = low; 3 = high). Debris potential tends to increase during times of few deep-water reefs and many reefs showing evidence of bioerosion.

Kiessling et al.

Choquette (1985) stated that 61% of the recoverable oil in giant fields is from carbonate reservoirs.
Paleogeographic plots showing sites of reservoir
reefs compared with reefs without reservoir quality
(Figure 7) can assist to detect general patterns of
reservoir distribution and may provide a tool in
hydrocarbon exploration.
The reservoir potential of reefs varies extremely
through the Phanerozoic (Figure 15). Our data suggest that there are four periods containing reefal
reservoirs. Abundant reefs with reservoir quality are
recognized from the Silurian to the Late Permian, in
the Late Jurassic, in the middle Cretaceous, and in
the Miocene. The peak in absolute reservoirs is evident in the GivetianFrasnian (more than 80 reefs),
but the relative amount of reservoirs is highest in
the Asselian (more than 50% of reefs).
The overall pattern is similar to the results of
Greenlee and Lehmann (1993), but there are differences regarding the quantity of reefal reservoirs
through time. Although reservoir quality as defined
in this paper does not necessarily imply productive
hydrocarbon reservoirs, we think that our results
better reflect the importance of particular times for
hydrocarbon exploration in reefs. Three arguments
substantiate this statement.
(1) Greenlee and Lehmann (1993) mostly referred to Exxon data and particularly excluded
buildups from the former Soviet Union and China,
whereas our database takes into account reefs from
all over the world.
(2) Greenlee and Lehmann (1993) exclusively
referred to subsurface data of producing reservoirs,
whereas our study includes outcrop data and data
from reefs with reservoir potential but lacking hydrocarbon accumulation (e.g., Capitan reef trend).
(3) The diagram in Figure 15 is based on 2470
reef data, whereas Greenlee and Lehmann (1993)
referred to 29 reefal reservoirs.
Despite all of these methodological differences,
we have an unequal temporal distribution of reefal
reservoirs similar to that of Greenlee and Lehmann
(1993). For instance, we recognize reservoir quality
(productive and nonproductive) for nearly 40% of
the Permian reefs, whereas there are almost no reefal
reservoirs in the Triassic. The interpretation of such
tremendous differences is difficult. We agree with
Greenlee and Lehmann (1993) that times of reefal
reservoirs correspond to times of extensive source
deposition, but we see no relation with high-order
eustatic sea level fluctuations. We note the secular
differences in reefal reservoirs, although all of our
time slices are defined by second-order sea level. A
quantitative test of the percentage of reservoirs and
the mean first-order sea level in time slices (Vail et al.,
1977; Hallam, 1984) did not produce any significant
correlations. Third-order sea level fluctuations certainly can have an important effect on the reservoir

1579

Table 4. Correlation Matrix (Based on Spearman-Rho)


of Debris Potential and Reservoir Quality with
Quantitative Reef Attributes*

Size
Diversity
Micrite
Sparite
Debris Potential
Water Depth
Absolute Paleolatitude
Bioerosion

Debris
Potential

Reservoir
Quality

+
+

+
+
+
+

*+ and indicate slope of significant (<0.01) correlations.

quality of reefs (Sun and Esteban, 1994; Sun and


Wright, 1998), but this does not explain the strong
variations between supersequences.
Klemme and Ulmishek (1991) noted six major
source rock intervals: Silurian, Late Devonian
Tournaisian, PennsylvanianEarly Permian, Late
Jurassic, AptianTuronian, and OligoceneMiocene;
hence, the Phanerozoic deposition of source rocks
correlates stratigraphically with the development of
reefal reservoirs. Although the correlation is not perfect concerning absolute amounts, it is sufficient to
presume a link between the driving forces for source
rock deposition and reefal reservoir development.
This link does not imply that abundant source rock
deposition directly enhances the reservoir quality of
reefs. More likely physicochemical parameters
(paleoclimate, sea water chemistry, plate tectonics)
may favor source rock deposition and reefal reservoir
development simultaneously.
Productive reefs occur in icehouse (Moscovian
Asselian, Miocene) and in greenhouse (Silurian
Devonian, middle Cretaceous) climatic cycles, and in
calcite-dominated, as well as in aragonite-dominated,
reefs. Additionally, reefal reservoirs occur in all types
of reefs: biostromal complexes (middle Cretaceous),
mud mounds (e.g., MoscovianAsselian), reef
mounds (e.g., Frasnian), and framework-dominated
coral reefs (Miocene).
The areal or temporal vicinity of porous reefs to
petroleum source rocks is another crucial point in
defining productive reservoirs. Due to the likely
absence of photosymbiosis before the Late
Triassic (Stanley and Swart, 1995), reefs were not
adapted to oligotrophic settings (Wood, 1993,
1995). Although it was suggested that some
Paleozoic tabulate corals and stromatoporoids harbored symbionts (Wood, 1995), the observed pattern of reef distribution (Figure 6) does not support
a generally nutrient-limited setting for stromatoporoid-dominated reefs. Asselian bioherms do not
flourish in areas with proposed upwelling (Figure 7),

1580

Paleoreef Maps

90

Reefs With Reservoir Quality (No. and %)

80

70

60

50

40

30

20

10

Ea
r
M ly C
id
dl am
e
C bri
Tr am an
em b
ad rian
o
Ar cian
en
C
a igia
Ll rad n
an oc
do ia
n
W ver
en ia
n
L loc
Em oc kia
hk n
G sia ov
ive n
i
tia /Ei an
n/ fel
Vi
F
ra ian
s
M ean To sn
os /S u
ia
co e rna n
vi rpu is
an k ia
/K ho n
as vi
im an
ov
i
As an
s
e
Ar li
t a
G ins n
ua ki
da an
lu
p
La ian
di
ni
an
Ba P
jo lien No
ci s ria
an b
/B ac n
a h
Ki tho ian
m
m nia
n
er
id
Be gia
rr n
Ba iasi
rre an
m
ia
Al n
b
Tu ia
n
C ron
am ia
n
pa
Yp nian
re
Lu sian
te
t
R ian
up
Aq e
u lia
Se itan n
rra ian
va
M llia
es n
si
ni
an

Figure 15Amount (white bars) and percentage (black bars) of reefs with reservoir potential in each time slice. Note
the pronounced differences between time slices and the almost total lack of reefal reservoirs from the Triassic to
Middle Jurassic.

Table 5. Quantitative Comparison of Reservoir Quality with Qualitative Reef Attributes*

Good
Environment

Shelf/platform margin

Reef Type

Reef mound

Dominant Reef Builder

Algae
Tubiphytes

Dominant Guild

Constructor
Binder

Moderate
Intrashelf/intraplatform
Slope/ramp
Basin
True reef
Mud mound
Microbes
Coralline sponges
Scleractinians
Pelecypods
Bryozoans
Baffler

Poor

Biostrome
Siliceous sponges
Tabulates/rugosa
Worms
Others

*Based on one-way ANOVA models. Mean reservoir potential is significantly enhanced in shelf/platform margin environments, in reef mounds, in algal or
Tubiphytes buildups, and in buildups dominated by the constructor or binder guild.

but their vicinity to prominent mountain ranges


may indicate elevated nutrient levels through continental erosion. Thus, the ability of many Paleozoic

reefs and mounds to flourish in nutrient-rich surface water may enhance their reservoir potential in
comparison to MesozoicCenozoic reefs.

Kiessling et al.

A statistical analysis of the database permits a


more quantitative approach to the problem. In a
first step we produced a correlation matrix of all
standardized quantitative reef features (Table 4).
This resulted in a significant positive correlation
between reservoir quality and reef size (thickness
and extension), debris potential, and the amount of
marine cement. The negative correlation of bioerosion and reservoir potential is an artifact related to
the facts that buildup reservoirs are usually poorly
described paleontologically and that reservoirs are
common in Paleozoic reefs with generally little evidence of bioerosion. The correlation of reservoir
quality and debris potential is logical considering
that many reefal reservoirs are located in reefderived debris rather than in the autochthonous
reef rock (Alsharhan, 1985; Crevello et al., 1985).
The positive correlation of reefal reservoir quality
with the amount of marine cement is contrary to
what one would expect and may be explained by
the usually unstable aragonitic mineralogy of early
diagenetic sparite, especially in algal-rich reefs
(Mazzullo and Cys, 1979; James et al., 1988) that
generally form good reservoir rocks. The similarity
in the abundance of biocementstones (Webb,
1996) and reefal reservoirs in the Paleozoic hints to
another possible interpretation: Grotzinger and
Knoll (1995) proposed that reduced shelf space for
carbonate precipitation, increased calcium flux to
the oceans, and deep basinal anoxia should favor
massive carbonate precipitation; hence, global
oceanographic peculiarities in particular time intervals simultaneously may favor both reefal reservoir
quality and precipitation of marine cement.
The relation of qualitative (environment, reef
type, dominant biota, dominant guild) and quantitative reef attributes was analyzed with one-way
ANOVA (analysis of variance) models (Table 5).
Shelf/platform margin reefs contain significantly
more reservoirs than do reefs from all other environments. The differences in reservoir abundance
among intrashelf/intraplatform, slope/ramp, and
basinal reefs are insignificant. Reef mounds are
more likely to form reservoirs on average than true
reefs, mud mounds, or biostromes. The number of
reservoirs in true reefs and mud mounds does not
differ significantly, but biostromes have the significantly lowest reservoir potential of all reef types
separated in the database. Algal- or Tubiphytesdominated buildups are more likely to form reservoirs than are reefs predominated by microbes,
coralline sponges, scleractinian corals, pelecypods,
or bryozoans. Bryozoans form reservoirs more
commonly than reefs predominated by siliceous
sponges, tabulate/rugose corals, worms, or other
reef builders. Reefs predominated by the constructor and binder guilds bear significantly more reservoirs than buildups with prevailing bafflers.

1581

The combination of statistical methods permits


the definition of an ideal Phanerozoic reefal reservoir: a large reef mound at the shelf/platform margin, dominated by constructors or binders and
Tubiphytes or algae containing abundant marine
cement, producing large quantities of debris, and
situated in a low paleolatitude. Not all of these
attributes will be realized together in one particular
reef (e.g., algal and Tubiphytes reefs are virtually
never dominated by the constructor guild), but the
more of these features that are realized in a buildup,
the more likely it represents a prospective target. In
summary, the qualitative statement of Burchette and
Wright (1992), who emphasized the organic and
sedimentary facies as a major control of buildup
reservoir potential, can be confirmed by our data.
In addition, other physicochemical parameters in
the earth system need to be analyzed as well. This
can be done by a correlation matrix of mean reef
attributes and published earth system parameters for
each time slice. We have tested the correlations of
the percentage of reservoirs in the time slices with
eustatic sea level curves (see preceding sections),
global carbonate and evaporite sedimentation
(Kazmierczak et al., 1985; Bluth and Kump, 1991),
oceanic crust production (Gaffin, 1987), atmospheric CO2 concentrations and global paleotemperature
(Frakes, 1979; Berner, 1994), atmospheric oxygen
concentrations (Berner and Canfield, 1989), Mg/Ca
ratio in seawater (Hardie, 1996; Stanley and Hardie,
1998), calcite vs. aragonite percentage in ooids and
in skeletal organisms (Mackenzie and Agegian, 1989;
Stanley and Hardie, 1998), the strontium isotope
curve of Burke et al. (1982), and the carbonate 13C
curve of Holser (1992). All published curves were
averaged to fit our time slice definition. The percentage of reefal reservoirs is significantly enhanced in
time slices with large reef numbers, a high percentage of deeper water mounds, a high mean sparite
content in reefs, high global evaporite sedimentation, high atmospheric oxygen and low carbon dioxide concentration, low global mean temperatures,
low oceanic crust production, high 13C in carbonate, low percentage of calcite ooids, and elevated
Mg/Ca (aragonite ocean). Mean reservoir quality in
time slices is not correlated with global carbonate
sedimentation and 87Sr/86Sr in seawater.
The strongest correlations ([r] > 0.6, < 0.01)
are noted with global temperature (Berner, 1994)
(calculated from the CO2 curve), atmospheric oxygen concentration, 13C, and evaporite sedimentation area (Figure 16). Considering the qualitative
statement that reefal reservoirs occur in warm and
in cool paleoclimatic modes, the strong negative
correlation of reefal reservoir quality and calculated paleotemperature (Berner, 1994) is surprising.
The correlation is weaker (but still significant) with
paleoclimatic conditions derived directly from the

1582

Paleoreef Maps

geologic record (Frakes, 1979). One might conclude that cool global paleoclimate favors reservoir
development in reefs; however, reservoir quality is
enhanced in lower absolute paleolatitudes (Table
4), hence the relation of paleoclimate and reefal
reservoir quality remains controversial. Because the
long-term 13C curve and the oxygen concentration curve can be taken as measures of net organic
carbon burial (Berner and Canfield, 1989; Holser et
al., 1996), one may presume an actual relation
between organic carbon burial and reefal reservoir
development. This agrees with the previously discussed stratigraphic correlation of reservoir and
source rock formation on a supersequence level.
Evaporite sedimentation area (Bluth and Kump,
1991) is positively correlated with both 13C and
the percentage of reefal reservoirs (Figure 16),
hence a connection among these three measures
may be presumed but is difficult to explain. Degens
and Paluska (1979) proposed a model to explain
the relation between evaporite sedimentation and
source rock generation. They emphasized a pickling effect of saline brines from evaporite deposits
that prevented rapid organic matter degradation. A
model more directly linking evaporites and secondary porosity generation in reefal carbonates
was suggested by Sun (1992). He observed a significant leaching of Miocene skeletal aragonite by
hypersaline brines and stated that leaching by
hypersaline brines may be at least as important as
leaching by meteoric water in the geologic record.
Although the global applicability of this model is
disputable, it would fit excellently to our data and
may explain the strong difference in reservoir quality between the Guadalupian and the Ladinian time
slices, which are similar in other reef attributes.
In summary, there is not a solitary explanation
for the observed pattern depicted in Figure 15.
The database suggests that Phanerozoic reefal
reservoir distribution is closely related to biotic
evolution and geologic parameters that affected
paleo-oceanographic water mass properties on a
global scale. We suspect a direct control of evaporite sedimentation and global paleoclimate,
whereas largely unknown factors simultaneously
control global reservoir quality, global organic carbon burial, precipitation of marine cement in
reefs, and the percentage of deeper water reefs.
Low-order (first- and second-order) sea level fluctuations do not play a significant role in defining
productive reservoirs.

paleogeographic reconstructions. The patterns


depicted in paleogeographic reef maps (Figures
611), the overall patterns observed after numerical
processing of the database (Figures 2, 5, 1216), and
statistical results (Tables 4, 5), are likely to represent
reality, although the database is, and always will
remain, incomplete. Some bias is caused by incomplete data, poor knowledge of reefs, heterogeneous
distribution of information in time and space, and
the occasional subjective treatment of data.
In comparison with published reef distribution
maps, our paleoreef maps are less sketchy, allow
the application of different filters, and permit the
distinction of particular reef types. These possibilities considerably enhance the potential use of reefs
for paleoclimatic and paleogeographic studies. The
paleogeographic distribution pattern of reefs with
reservoir quality (Figure 7) can provide a tool for
hydrocarbon exploration.
The diagrams in Figure 14 present the first proposal of secular variations of bioerosion, debris
potential, and bathymetry in and of reefs through
time. These figures suggest an overall increase of
bioerosion and debris potential through time,
whereas there appears to be a decrease of deeper
water reefs from the Carboniferous to the Tertiary.
Our curves for reef abundance, biotic composition, paleolatitudinal distribution, and reservoir
potential show similarities with published results,
but also exhibit significant deviations. The variations
in reef abundance (Figure 2) and biotic composition
(Figure 13) on a high-ranked taxonomic level can be
grouped into seven cycles of reef development. The
cycles may cross times of reef crises and mass
extinctions (OrdovicianSilurian, PermianTriassic),
suggesting that the reef ecosystem is more stable
than commonly thought. The paleolatitudinal concentration of reefs (Figure 12) is characterized by an
overall Phanerozoic northward drift. This roughly
follows the drift of continental plates, but climatic
asymmetries affected the global pattern considerably, especially during the late Paleozoic Gondwana
glaciation. The reservoir potential of Phanerozoic
reefs (Figure 15) exhibits enormous changes
through time. These changes are likely to be more
controlled by biotic and paleo-oceanographic evolution rather than by sea level fluctuations. Time slices
with elevated organic carbon burial, raised evaporite
sedimentation area, lower paleotemperatures, and a
large percentage of aragonite ooids are characterized
by significantly enhanced reefal reservoir quality on
a global scale (Figure 16).

CONCLUSIONS
OUTLOOK
The paleoreef maps database offers new opportunities for understanding the reef ecosystem, for
analyzing past climates, and for plate tectonic and

The current state of the database allows only statistical tests based on rather long time intervals. To

Kiessling et al.

% Reefal reservoirs

50
4

13 C
40
Total evaporites (105 km2)

30
Surface air temperature (C)

13
C

Reef Reservoirs (%) Evaporite Depositional Area (C)

60

1583

1
20
0
10
-1

-2

Ea
r
M ly C
id
dl am
e
C bri
Tr am an
em br
ad ian
Ar ocia
e n
C nig
ar
ia
a
Ll do n
an c
do ian
W ve
en ria
n
L loc
Em oc kia
hk n
G sia o
ive n via
tia /Ei n
n/ fel
Vi
Fr ia
s
a n
M ean To sn
os /S ur ia
co e na n
vi rpu isi
an k an
/K ho
as via
im n
o
As vian
s
A eli
G rtin an
ua sk
da ian
lu
p
La ian
di
ni
a
Ba Pli
N n
jo ens or
ci
an ba ian
/B ch
Ki ath ian
m o
m ni
er an
i
Be dgia
rri n
Ba as
rre ian
m
ia
Al n
Tu bian
C ron
am ia
pa n
Yp nian
re
s
Lu ian
te
R tian
u
Aq pe
ui lian
Se tan
rra ian
va
M llia
es n
si
ni
an

Figure 16Correlation of reefal reservoir percentage in time slices with selected global physicochemical variables.
Global surface air temperature (Berner, 1994) is inversely correlated with reservoir frequency, whereas the global
area of evaporite sedimentation (Bluth and Kump, 1991) and the global 13C in carbonates ( PDB) (Holser, 1992)
are positively correlated.

evaluate reef attributes on a finer stratigraphic level,


more data are required and international cooperation
is essential. The authors are currently editing a publication on Phanerozoic reef patterns to which several
reef specialists are contributing. With this major project, we expect that shortcomings and biases in the
database will be significantly reduced; nevertheless,
we conclude this paper with a call for data. We are
willing to offer parts of the database in exchange for
new data. All paleoreef maps discussed in this paper
can be downloaded from our internet home page
(http://www.geol.uni-erlangen.de/pal/palreef.htm).
REFERENCES CITED
Alsharhan, A. S., 1985, Depositional environment, reservoir
units evolution, and hydrocarbon habitat of Shuaiba
Formation, Lower Cretaceous, Abu Dhabi, United Arab
Emirates: AAPG Bulletin, v. 69, p. 899912.
Armella, C., 1994, Thrombolitic-stromatolitic cycles of the
CambroOrdovician boundary sequence, Precordillera
Oriantal basin, western Argentina, in J. Bertrand-Sarfati and
C. Monty, eds., Phanerozoic stromatolites II: Dordrecht,

Kluwer Academic Publishers, p. 421441.


Beauvais, L., 1992, Corals of the circum-Pacific region, in G. E. G.
Westermann, ed., The Jurassic of the circum-Pacific:
Cambridge, Cambridge University Press, p. 324327.
Belasky, P., and B. Runnegar, 1993, Biogeographic constraints for
tectonic reconstructions of the Pacific region: Geology, v. 21,
p. 979982.
Berger, W. H., 1982, Increase of carbon dioxide in the
atmosphere during deglaciation: the coral reef hypothesis:
Naturwissenschaften, v. 69, p. 8788.
Berner, R. A., 1994, Geocarb II: a revised model of atmospheric
CO 2 over Phanerozoic time: American Journal of Science,
v. 294, p. 5691.
Berner, R. A., and D. Canfield, 1989, A new model for
atmospheric oxygen over Phanerozoic time: American Journal
of Science, v. 289, p. 333361.
Bluth, G. J. S., and L. R. Kump, 1991, Phanerozoic paleogeology:
American Journal of Science, v. 291, p. 284308.
Bolshakova, L. N., M. P. Geller, R. V. Goryunova, V. N. Dubatolov,
A. B. Ivanovskiy, V. N. Kosmyn, V. A. Luchinina, B. S. Sokolov,
and Yu. I. Tesakov, 1994, Paleozoic coral reefs in Russia:
Stratigrafiya Geologicheskaya Korrelyatsiya, v. 2/1, p. 4654.
Burchette, T. P., and V. P. Wright, 1992, Carbonate ramp
depositional systems: Sedimentary Geology, v. 79, p. 357.
Burke, W. H., R. E. Denison, E. A. Hetherington, R. B. Koepenik, H. F.
Nelson, and J. B. Otto, 1982, Variations of seawater 87Sr/86Sr
through Phanerozoic time: Geology, v. 10, p. 516519.

1584

Paleoreef Maps

Catalano, R., P. Di Stefano, and H. Kozur, 1991, Permian


circumpacific deep-water faunas from the western Tethys
(Sicily, Italy): new evidence for the position of the Permian
Tethys: Palaeogeography, Palaeoclimatology, Palaeoecology,
v. 87, p. 75108.
Cattaneo, G., A. Tahiri, M. Zahraoui, and D. Vachard, 1993, La
sdimentation rcifale du Givtien dans la Meseta marocaine
nord-occidentale: Comptes Rendus de lAcademie des Sciences
Serie II, v. 317, p. 7380.
Colwell, J. B., U. Rhl, U. von Rad, and E. Kristan-Tollmann, 1994,
Mesozoic sedimentary and volcaniclastic rocks dredged from the
northern Exmouth plateau and Rowley terrace, offshore
northwest Australia: Australian Geological Survey Organisation
Journal of Australian Geology and Geophysics, v. 15, p. 1142.
Copper, P., 1988, Ecological succession in Phanerozoic reef
ecosystems: is it real?: Palaios, v. 3, p. 136152.
Copper, P., 1989, Enigmas in Phanerozoic reef development:
Association of Australasian Palaeontologists Memoir, v. 8, 371 p.
Copper, P., 1994, Ancient reef ecosystem expansion and collapse:
Coral Reefs, v. 13, p. 311.
Crevello, P. D., P. M. Harris, D. L. Stoudt, and L. R. Baria, 1985,
Porosity evolution and burial diagenesis in a Jurassic reefdebris reservoir, Smackover Formation, Hico Knowles field,
Louisiana, in P. O. Roehl and P. W. Choquette, eds., Carbonate
petroleum reservoirs: New York, Springer-Verlag, p. 387406.
Davies, G. R., and W. W. Nassichuk, 1973, The hydrozoan?
Palaeoaplysina from the upper Paleozoic of Ellesmere Island,
Arctic Canada: Journal of Paleontology, v. 47, p. 251265.
Degens, E. T., and A. Paluska, 1979, Hypersaline solutions interact
with organic detritus to produce oil: Nature, v. 281,
p. 666668.
Dercourt, I., L. E. Ricou, and B. Vrielnyck, eds., 1993, Atlas of
Tethys paleoenvironmental maps: Paris, Gauthier-Villars, 307 p.
Dickins, J. M., 1993, Climate of the Late Devonian to Triassic:
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 100,
p. 8994.
Ditchfield, P. W., 1997, High northern palaeolatitude Jurassic
Cretaceous palaeotemperature variation: new data from Kong
Karls Land, Svalbard: Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 130, p. 163175.
Ditchfield, P. W., J. D. Marshall, and D. Pirrie, 1994, High latitude
palaeotemperature variation: new data from the Tithonian to
Eocene of James Ross Island, Antarctica: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 107, p. 79101.
Dullo, W.-C., 1983, Fossildiagenese im mioznen Leitha-Kalk der
Paratethys von sterreich: Ein Beispiel fr Faunenverschiebungen durch Diageneseunterschiede: Facies, v. 8, p. 1112.
El-Asaad, G. M. A., 1991, Oxfordian hermatypic corals from
central Saudi Arabia: Gobios, v. 24, p. 267287.
Ellis, P. M., P. D. Crevello, and L. S. Eliuk, 1985, Upper Jurassic and
Lower Cretaceous deep-water buildups, Abenaki Formation,
Nova Scotia shelf: SEPM Core Workshop, v. 6, p. 212248.
Esteban, M., 1979, Significance of the upper Miocene coral reefs
of the western Mediterranean: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 29, p. 169188.
Fagerstrom, J. A., 1987, The evolution of reef communities: New
York, John Wiley, 600 p.
Fagerstrom, J. A., 1988, A structural model for reef communities:
Palaios, v. 3, p. 217220.
Fagerstrom, J. A., 1991, Reef-building guilds and a checklist for
determining guild membership: Coral Reefs, v. 10, p. 4752.
Feary, D. A., and N. P. James, 1995, Cenozoic biogenic mounds
and buried Miocene(?) barrier reef on a predominantly coolwater carbonate continental marginEucla basin, western
Great Australian Bight: Geology, v. 23, p. 427430.
Flgel, E., 1981, Paleoecology and facies of Upper Triassic reefs in
the northern Calcareous Alps, in D. F. Toomey, ed., European
fossil reef models: SEPM Special Publication 30, p. 291359.
Flgel, E., 1994, Pangean shelf carbonates: controls and paleoclimatic
significance of Permian and Triassic reefs: Geological Society
of America Special Paper, v. 228, p. 247266.

Flgel, E., and E. Flgel-Kahler, 1992, Phanerozoic reef evolution:


basic questions and data base: Facies, v. 26, p. 167278.
Fortunatova, N. K., I. G. Micheev, A. G. Ibragimov, V. P.
Farbirovich, and A. G. Shvez-Teneta-Gurii, 1986, Metodi
vydeleniya rifovych fazii v verchnejurskich karbonatych
otlozheniyach Yzhnogo Uzbekistana, in B. S. Sokolov, ed.,
Fanerosoiskie rifi i koralli SSSR: Moskow, Nauka, p. 149161.
Frakes, L. A., 1979, Climates through geologic time: Amsterdam,
Elsevier, 310 p.
Frakes, L. A., J. E. Francis, and J. I. Syktus, 1992, Climate modes of
the Phanerozoic: the history of the earths climate over the
past 600 million yr: Cambridge, Cambridge University Press,
274 p.
Franseen, E. K., M. Esteban, W. C. Ward, and J.-M. Rouchy, 1996,
Introduction, in E. K. Franseen, M. Esteban, W. C. Ward, and
J.-M. Rouchy, eds., Models for carbonate stratigraphy from
Miocene reef complexes of Mediterranean regions: SEPM
Concepts in Sedimentology and Paleontology, v. 5, p. vix.
Gaffin, S., 1987, Ridge volume dependence on seafloor generation
rate and inversion using long term sea level change: American
Journal of Science, v. 287, p. 596611.
Gandin, A., and F. Debrenne, 1984, Lower Cambrian bioconstructions in southwestern Sardinia (Italy): Geobios,
Memoir Special, v. 8, p. 231240.
Golonka, J., and N. Y. Bocharova, 1997, Hot spots activity and the
break-up of Pangea: Gaea Heidelbergensis, v. 3, p. 143.
Golonka, J., and D. Ford, 1997a, Absaroka (Pennsylvanian
Triassic) paleoenvironment and lithofacies (abs.): Canadian
Society of Petroleum GeologistsSEPM Joint Convention,
Program with Abstracts, p. 109.
Golonka, J., and D. Ford, 1997b, Sauk and Tippecanoe
(CambrianSilurian) paleoenvironment and lithofacies (abs.):
Canadian Society of Petroleum GeologistsSEPM Joint
Convention, Program with Abstracts, p. 110.
Golonka, J., and L. Gahagan, 1997, Tectonic model of the
Mediterranean terranes (abs.): AAPG Bulletin, v. 81, p. 1378.
Golonka, J., M. I. Ross, and C. R. Scotese, 1994, Phanerozoic
paleogeographic and paleoclimatic modeling maps: Canadian
Society of Petroleum Geologists Memoir, v. 17, p. 147.
Golonka, J., D. Ford, and J. Bednarczyk, 1997a, Kaskaskia
(DevonianMississippian) paleoenvironment and lithofacies
(abs.): Canadian Society of Petroleum GeologistsSEPM Joint
Convention, Program with Abstracts, p. 110.
Golonka, J., D. Ford, M. Edrich, R. Pauken, J. Wildharber, and N. Y.
Bocharova, 1997b, Zuni (JurassicCretaceous) paleoenvironment
and lithofacies (abs.): Canadian Society of Petroleum
GeologistsSEPM Joint Convention, Program with Abstracts, p. 111.
Gradstein, F. M., and J. G. Ogg, 1996, Geological time scale for
the Phanerozoic: Episodes, v. 19, p. 34.
Greenlee, S. M., and P. J. Lehmann, 1993, Stratigraphic framework
of productive carbonate buildups, in R. G. Loucks and J. F.
Sarg, eds., Carbonate sequence stratigraphy: recent
developments and applications: AAPG Memoir 57, p. 4362.
Grotzinger, J. P., and A. H. Knoll, 1995, Anomalous carbonate
precipitates: is the Precambrian the key to the Permian?:
Palaios, v. 10, p. 578596.
Hallam, A., 1984, Pre-Quaternary changes of sea level: Annual
Reviews of Earth and Planetary Science, v. 12, p. 205243.
Hallam, A., 1994, Jurassic climates as inferred from the sedimentary
and fossil record, in J. R. L. Allen, B. J. Hoskins, B. W. Sellwood,
R. A. Spicer, and P. J. Valdes eds., Palaeoclimates and their
modelling: London, Chapman and Hall, p. 7988.
Hallock, P., 1988, The role of nutrient availability in bioerosion:
consequences to carbonate buildups: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 63, p. 275291.
Hallock, P., and W. Schlager, 1986, Nutrient excess and the
demise of coral reefs and carbonate platforms: Palaios, v. 1,
p. 389398.
Haq, B. U., J. Hardenbol, and P. R. Vail, 1988, Mesozoic and
Cenozoic chronostratigraphy and cycles of sea-level change, in
C. K. Wilgus, B. S. Hastings, C. G. St. C. Kendall, H. W.

Kiessling et al.

Posamentier, C. A. Ross, and J. C. van Wagoner, eds., Sea-level


changesan integrated approach: SEPM Special Publication 42,
p. 71108.
Hardie, L. A., 1996, Secular variation in seawater chemistry: an
explanation for the coupled secular variation in the
mineralogies of marine limestones and potash evaporites over
the past 600 m.y.: Geology, v. 24, p. 279283.
Hatcher, B. G., 1997, Coral reef ecosystems: how much greater is
the whole than the sum of the parts?: Coral Reefs, v. 16,
p. S77S91.
Heckel, P. H., 1974, Carbonate buildups in the geologic record: a
review, in L. F. Laporte, ed., Reefs in time and space: SEPM
Special Publication 18, p. 90154.
Henrich, R., A. Freiwald, A. Wehrmann, P. Schfer, C. Samtleben,
and H. Zankl, 1996, Nordic cold-water carbonates: occurrences
and controls, in J. Reitner, F. Neuweiler, and F. Gunkel, eds.,
Global and regional controls on biogenic sedimentation: Gttinger
Arbeiten zur Geologie und Palontologie, Sonderband 2,
p. 3552.
Holser, W. T., 1992, Stable isotope geochemistry of sulfate and
chloride rocks: Lecture Notes in Earth Sciences, v. 43,
p. 153176.
Holser, W. T., M. Magaritz, and R. L. Ripperdan, 1996, Global
isotopic events, in O. H. Walliser, ed., Global events and event
stratigraphy: Berlin, Springer-Verlag, p. 6388.
Hubbard, D. K., 1997, Reefs as dynamic systems, in C. Birkeland,
ed., Life and death of coral reefs: London, Chapman and Hall,
p. 4367.
Hubbard, D. K., A. I. Miller, and D. Scaturo, 1990, Production and
cycling of calcium carbonate in a shelf-edge reef system (U.S.
Virgin Islands): applications to the nature of reef systems in
the fossil record: Journal of Sedimentary Petrology, v. 60,
p. 335360.
James, N. P., 1983, Reef environment, in P. A. Scholle, D. G.
Bebout, and C. H. Moore, eds., Carbonate depositional
environments: AAPG Memoir 33, p. 345440.
James, N. P., and P.-A. Bourque, 1992, Reefs and mounds, in R. G.
Walker and N. P. James, eds., Facies models: response to sea
level change: St. Johns, Geological Association of Canada,
p. 323347.
James, N. P., and F. Debrenne, 1980, Lower Cambrian bioherms:
pioneer reefs of the Phanerozoic: Acta Palaeontologica
Polonica, v. 25, p. 655668.
James, N. P., D. R. Kobluk, and S. G. Pemberton, 1977, The oldest
macroborers: Lower Cambrian of Labrador: Science, v. 197,
p. 980983.
James, N. P., J. L. Wray, and R. N. Ginsburg, 1988, Calcification of
encrusting aragonitic algae Peyssonneliaceae: implications for
the origin of late Paleozoic reefs and cements: Journal of
Sedimentary Petrology, v. 58, p. 291303.
Jordan, C. F., Jr., M. W. Colgan, S. H. Frost, D. Bosence, and
M. Esteban, 1990, An overview of Miocene reefs: AAPG
Bulletin, v. 76, p. 688689.
Kauffman, E. G., and J. A. Fagerstrom, 1993, The Phanerozoic
evolution of reef diversity, in R. E. Ricklefs and D. Schluter,
eds., Species diversity in ecological communities: Chicago,
University of Chicago Press, p. 315329.
Kazmierczak, J., V. Ittekkot, and E. T. Degens, 1985, Biocalcification through time: environmental challenge and
cellular response: Palontologische Zeitschrift, v. 59, p. 1533.
Kiessling, W., and R. Scasso, 1996, Ecological perspectives of Late
Jurassic radiolarian faunas from the Antarctic Peninsula, in
A. C. Riccardi, ed., Advances in Jurassic research: Zurich,
Transtec, v. 1/2, p. 317326.
Klemme, H. D., and G. F. Ulmishek, 1991, Effective petroleum source
rocks of the world: stratigraphic distribution and controlling
depositional factors: AAPG Bulletin, v. 75, p. 18091851.
Kobluk, D. R., N. P. James, and S. G. Pemberton, 1978, Initial
diversification of macroboring ichnofossils and exploitation of
the macroboring niche in the lower Paleozoic: Paleobiology,
v. 4, p. 163170.

1585

Kozur, H., 1990, Deep-water Permian in Sicily and its possible


connection with the Himalaya-Tibet region: Fifth HimalayaTibet-Karakorum Workshop, p. 27.
Kozur, H., and J. Krahl, 1987, Erster Nachweis von Radiolarien in
tethyalen Perm Europas: Neues Jahrbuch fr Geologie und
Palontologie, Abhandlungen, v. 174, p. 357372.
Kristan-Tollman, E., and A. Tollmann, 1981, Die Stellung der Tethys in
der Trias und die Herkunft ihrer Fauna: Mitteilungen der
sterreichischen Geologischen Gesellschaft, v. 74/75, p. 129135.
Kristan-Tollman, E., and A. Tollmann, 1982, Die Entwicklung der
Tethystrias und Herkunft ihrer Fauna: Geologische Rundschau,
v. 71, p. 9871019.
Kuznetsov, V. G., 1993, Late JurassicEarly Cretaceous carbonate
platform in the northern Caucasus and Precaucasus, in J. A. T.
Simo, R. W. Scott, and J.-P. Masse, eds., Cretaceous carbonate
platforms: AAPG Memoir 56, p. 455463.
Kuznetsov, V. G., 1997, Oil and gas in reef reservoirs in the
former USSR: Petroleum Geoscience, v. 3, p. 6571.
Lecompte, M., 1958, Les recifs paleozoiques Belgique: Geologische Rundschau, v. 47, p. 384401.
Lecompte, M., 1970, Die Riffe im Devon der Ardennen und ihre
Bildungsbedingungen: Geologica et Paleontologica, v. 4,
p. 2571.
Legarreta, L., 1991, Evolution of a CallovianOxfordian carbonate
margin in the Neuqun basin of west-central Argentina: facies,
architecture, depositional sequences and global sea-level
changes: Sedimentary Geology, v. 70, p. 209240.
Leinfelder, R. R., 1994, Distribution of Jurassic reef types: a mirror
of structural and environmental changes during breakup of
Pangea: Canadian Society of Petroleum Geologists Memoir,
v. 17, p. 677700.
Leinfelder, R. R., W. Werner, M. Nose, D. U. Schmid, M. Krautter,
M. Laternser, M. Takacs, and D. Hartmann, 1996, Paleoecology,
growth parameters and dynamics of coral, sponge and
microbolite reefs from the Late Jurassic, in J. Reitner,
F. Neuweiler, and F. Gunkel, eds., Global and regional controls
on biogenic sedimentation: Gttinger Arbeiten zur Geologie
und Palontologie, Sonderband 2, p. 227248.
Marsella, E., H. Kozur, and B. DArgenio, 1993, Monte Facito
Formation (Scythianmiddle Carnian); a deposit of the
ancestral Lagonegro basin in southern Apennines: Bolletino de
Sevisio Geologico Italia, v. 119, p. 225248.
Mazzullo, S. J., and J. M. Cys, 1979, Marine aragonite sea floor
growths and cements in Permian phylloid algal mounds
Sacramento Mountains New Mexico USA: Journal of
Sedimentary Petrology, v. 49, p. 917936.
Michailova, M. V., 1968, Biogermnie massivy v verchnejurskich
otlozheniyach gornogo Krima i Severnogo Kavkaza,
Iskopaemie rifi i metodika ich izuchenya. Trudy tretei
paleoekologolitologicheskoi sessii: Moscow, Uraliskii Filial
Akademii Nauk SSSR, p. 196209.
Mistaen, B., 1985, Phnomnes rcifaux dans le Dvonien
dAfghanistan (Montagnes Centrales): Societe Gologie du
Nord, v. 11, 381 p.
Moore, P. F., 1989, Devonian reefs in Canada and some adjacent
areas, in H. H. J. Geldsetzer, N. P. James, and G. E. Tebbutt,
eds., ReefsCanada and adjacent areas: Canadian Society of
Petroleum Geologists Memoir, v. 13, p. 367390.
Morris, A., 1996, A review of palaeomagnetic research in the
Troodos ophiolite, Cyprus, in A. Morris and D. H. Tarling,
eds., Palaeomagnetism and tectonics of the Mediterranean
region: Geological Society of London Special Publication
no. 105, p. 311324.
Mutti, M., and H. Weissert, 1995, Triassic monsoonal climate and
its signature in LadinianCarnian carbonate platforms
(southern Alps, Italy): Journal of Sedimentary Research, v. B65,
p. 357367.
Newell, N. D., 1971, An outline history of tropical organic reefs:
American Museum Novitates, v. 2465, p. 137.
Parrish, J. T., and R. L. Curtis, 1982, Atmospheric circulation,
upwelling, and organic-rich rocks in the Mesozoic and

1586

Paleoreef Maps

Cenozoic eras: Palaeogeography, Palaeoclimatology,


Palaeoecology, v. 40, p. 3166.
Paulay, G., 1997, Diversity and distribution of reef organisms, in
C. Birkeland, ed., Life and death of coral reefs: London,
Chapman and Hall, p. 298353.
Philip, J., J.-P. Masse, and G. Camoin, 1996, Tethyan carbonate
platforms, in A. E. M. Nairn, L.-E. Ricou, B. Vrielnyck, and
J. Dercourt, eds., The oceans basins and margin, v. 8, the
Tethys Ocean: New York, Plenum Press, p. 239266.
Pisera, A., 1985, Paleoecology and lithogenesis of the middle
Miocene (Badenian) algal-vermetid reefs from the Roztocze Hills,
south-eastern Poland: Acta Geologica Polonica, v. 35, p. 89155.
Pisera, A., 1996, Miocene reefs of the Paratethys: a review, in
E. K. Franseen, M. Esteban, W. C. Ward, and J.-M. Rouchy,
eds., Models for carbonate stratigraphy from Miocene reef
complexes of Mediterranean regions: SEPM, Concepts in
Sedimentology and Paleontology, v. 5, p. 97104.
Playford, P. E., 1980, Devonian Great Barrier Reef of Canning
basin, Western Australia: AAPG Bulletin, v. 64, p. 814840.
Pomar, L., 1991, Reef geometries, erosion surfaces and highfrequency sea-level changes, upper Miocene reef complex,
Mallorca, Spain: Sedimentology, v. 38, p. 243269.
Pomar, L., W. C. Ward, and D. G. Green, 1996, Upper Miocene
reef complex of the Llucmajor area, Mallorca, Spain, in E. K.
Franseen, M. Esteban, W. C. Ward, and J.-M. Rouchy, eds.,
Models for carbonate stratigraphy from Miocene reef
complexes of Mediterranean regions: SEPM, Concepts in
Sedimentology and Paleontology, v. 5, p. 191225.
Pratt, B. R., and L. F. Jansa, 1989, Upper Jurassic shallow water
reefs of offshore Nova Scotia, in H. H. J. Geldsetzer, N. P.
James, and G. E. Tebbutt, eds., ReefsCanada and adjacent
areas: Canadian Society of Petroleum Geologists Memoir, v. 13,
p. 741747.
Price, G. D., B. W. Sellwood, and P. J. Valdes, 1995,
Sedimentological evaluation of general circulation model
simulations for the greenhouse Earth: Cretaceous and Jurassic
case studies: Sedimentary Geology, v. 100, p. 159180.
Ramos, V. A., 1978, Los arrecifes de la Formacion Cotidiano
(Jurasico superior) en la Cordillera Patagonia y su significado
paleoclimatico: Ameghiniana, v. 15, p. 97111.
Ricou, L.-E., 1996, The plate tectonic history of the past Tethys
Ocean, in A. E. M. Nairn, L.-E. Ricou, B. Vrielnyck, and
J. Dercourt, eds., The oceans basins and margin, v. 8, the
Tethys Ocean: New York, Plenum Press, p. 970.
Ritter, S. M., and Morris, T. H., 1997, Oldest and lowest latitudinal
occurrence of Palaeoaplysina: Middle Pennsylvanian Ely
Limestone, Burbank Hills, Utah: Palaios, v. 12, p. 397401.
Robertson, A. H. F., and N. H. Woodcock, 1979, Mamonia
complex, southwest Cyprus: evolution and emplacement of a
Mesozoic continental margin: Geological Society of America
Bulletin, v. 90, p. 651665.
Robertson, A. H. F., P. D. Clift, P. Degnan, and G. Jones, 1991,
Paleogeographic and paleotectonic evolution of eastern
Mediterranean neotethys: Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 87, p. 289344.
Robertson, A. H. F., J. E. Dixon, S. Brown, A. Collins, A. Morris,
E. A. Pickett, I. Sharp, and T. Ustaomer, 1996, Alternative
tectonic models for the late Palaeozoicearly Tertiary
development of Tethys in the eastern Mediterranean region, in
A. Morris and D. H. Tarling, eds., Palaeomagnetism and
tectonics of the Mediterranean region: Geological Society of
London Special Publication 105, p. 239263.
Roehl, P. O., and P. W. Choquette, 1985, Introduction, in P. O.
Roehl and P. W. Choquette, eds., Carbonate petroleum
reservoirs: New York, Springer-Verlag, p. 115.
Rhl, U., T. Dumont, U. von Rad, R. Martini, and L. Zaninetti,
1991, Upper Triassic Tethyan carbonates off northwest
Australia (Wombat plateau, ODP Leg 122): Facies, v. 25,
p. 211252.
Ross, C. A., and J. R. P. Ross, 1988, Late Paleozoic transgressiveregressive deposition, in C. K. Wilgus, B. S. Hastings, C. G.

St. C. Kendall, H. W. Posamentier, C. A. Ross, and J. C.


van Wagoner, eds., Sea-level changesan integrated
approach: SEPM Special Publication no. 42, p. 227247.
Ross, C. A., G. T. Moore, and D. N. Hayashida, 1992, Late Jurassic
paleoclimate simulationpaleoecological implications for
ammonoid provinciality: Palaios, v. 7, p. 487507.
Rowland, S. M., and R. A. Gangloff, 1988, Structure and
paleoecology of Lower Cambrian reefs: Palaios, v. 3, p. 111135.
Sengr, A. M. C., 1984, The Cimmeride orogenic system and the
tectonics of Eurasia: Geological Society of America Special
Paper 195, 82 p.
Sengr, A. M. C., Y. Yilmaz, and O. Sungurlu, 1984, Tectonics of
the Mediterranean Cimmerides: nature and evolution of the
western termination of paleo-Tethys, in J. E. Dixon and A. H.
F. Robertson, eds., The geological evolution of the eastern
Mediterranean: Oxford, Blackwell, p. 77112.
Sheehan, P. M., 1985, Reefs are not so differentthey follow the
evolutionary pattern of level-bottom communities: Geology,
v. 13, p. 4649.
Sloss, L. L., 1963, Sequences in the cratonic interior of North
America: Geological Society of America Bulletin, v. 74,
p. 93113.
Sloss, L. L., 1972, Synchrony of Phanerozoic sedimentary-tectonic
events of the North American craton and the Russian platform:
24th International Geological Congress, Section 6, p. 2432.
Soares, P. C., P. M. B. Landim, and V. J. Fulfaro, 1978, Tectonic
cycles and sedimentary sequences in the Brazilian
intracratonic basins: Geological Society of America Bulletin,
v. 89, p. 181191.
Stampfli, G., J. Marcoux, and A. Baud, 1991, Tethyan margins in
space and time: Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 87, p. 373409.
Stanley, G. D., Jr., 1988, The history of early Mesozoic reef
communities: a three-step process: Palaios, v. 3, p. 170183.
Stanley, G. D., Jr., 1994, Late Paleozoic and early Mesozoic reefbuilding organisms and paleogeography: the TethyanNorth
American connection: Courier Forschungs-Institut Senckenberg, v. 172, p. 6975.
Stanley, G. D., Jr., and P. K. Swart, 1995, Evolution of the coralzooxanthellae symbiosis during the Triassic: a geochemical
approach: Paleobiology, v. 21, p. 179199.
Stanley, S. M., and L. A. Hardie, 1998, Secular oscillations in the
carbonate mineralogy of reef-building and sediment-producing
organisms driven by tectonically forced shifts in seawater
chemistry: Palaeogeography, Palaeoclimatology, Palaeoecology,
v. 144, p. 319.
Sun, S. Q., 1992, Skeletal aragonite dissolution from hypersaline
seawater: a hypothesis: Sedimentary Geology, v. 77, p. 249257.
Sun, S. Q., and M. Esteban, 1994, Paleoclimatic controls on
sedimentation, diagenesis, and reservoir quality: lessons from
Miocene carbonates: AAPG Bulletin, v. 78, p. 519543.
Sun, S. Q., and V. P. Wright, 1998, Controls on reservoir quality of
an Upper Jurassic reef mound in the Palmers Wood field area,
Weald basin, southern England: AAPG Bulletin, v. 82,
p. 497515.
Talent, J. A., 1988, Organic reef-building: episodes of extinction
and symbiosis?: Senckenbergiana lethaea, v. 69, p. 315368.
Teichert, C., 1958, Cold- and deep-water coral banks: AAPG
Bulletin, v. 42, p. 10641082.
Vail, P. R., R. M. J. Mitchum, R. G. Todd, J. M. Widmier, S. I.
Thompson, J. B. Sangree, J. N. Bubb, and W. G. Hatlelid, 1977,
Seismic stratigraphy and global changes of sea-level, in C. E.
Payton, ed., Seismic stratigraphyapplications to hydrocarbon
exploration: AAPG Memoir 26, p. 49212.
Vakhrameev, V. A., 1991, Jurassic and Cretaceous floras and
climates of the Earth: Cambridge, Cambridge University Press,
318 p.
Vermeij, G. J., 1977, The Mesozoic marine revolution: evidence
from snails, predators and grazers: Paleobiology, v. 3,
p. 245258.
Vinassa de Regny, P., 1915, Triadische Algen, Spongien,

Kiessling et al.

Anthozoen und Bryozoen aus Timor: Palaeontologie von


Timor, v. 4/8, p. 75117.
Vogel, K., 1993, Bioeroders in fossil reefs: Facies, v. 28, p. 109114.
Watkins, R., and E. C. Wilson, 1989, Paleoecologic and
biogeographic significance of the biostromal Palaeoaplysina
in the lower McCloud Limestone, eastern Klamath Mou:
Palaios, v. 4, p. 181192.
Webb, G. E., 1996, Was Phanerozoic reef history controlled by
the distribution of nonenzymatically secreted reef carbonates
(microbial carbonate and biologically induced cement)?:
Sedimentology, v. 43, p. 947971.
Wendt, J., Z. Belka, B. Kaufmann, R. Kostrewa, and J. Hayer,
1997, The worlds most spectacular carbonate mud mounds

1587

(Middle Devonian, Algerian Sahara): Journal of Sedimentary


Research, section A, v. 67, p. 424436.
Whalen, M. T., 1995, Barred basins: a model for eastern ocean
basin carbonate platforms: Geology, v. 23, p. 625628.
Wilson, J. L., 1975, Carbonate facies in geologic history: Berlin,
Springer-Verlag, 471 p.
Wood, R., 1993, Nutrients, predation and the history of reefbuilding: Palaios, v. 8, p. 526543.
Wood, R., 1995, The changing biology of reef-building: Palaios,
v. 10, p. 517529.
Zankl, H., 1977, Quantitative aspects of carbonate production in a
Triassic reef complex: Proceedings of the Third International
Coral Reef Symposium, v. 2, p. 379382.

ABOUT THE AUTHORS


Wolfgang Kiessling
Wolfgang Kiessling is currently a
postdoctoral fellow at Humboldt
University in Berlin. He has been a
research associate at the University of
Erlangen, where he received his
Ph.D. in 1995. Previously, he worked
as consultant for Shell Philippines
Exploration. His major research interests are ecosystem evolution, marine
paleoecology, micropaleontology,
and carbonate sedimentology.
Erik Flgel
Erik Flgel is full professor of
paleontology and head of the Institute
of Paleontology at the ErlangenNrnberg University. He is editor of
the international journal Facies. He
has worked on microfacies analysis
and depositional models of carbonate rocks, calcareous algae, paleoecology of PermianJurassic reefs in
the Alpine-Mediterranean region,
and in the field of archeometry. His
current research interests are the evolution of Phanerozoic
reefs and the integration of microfacies data into applied
facies concepts.

Jan Golonka
Jan Golonka is a senior lecturer at
the Jagiellonian University in Krakow,
Poland. He received his M. Sc. degree
(1967) from the University of Mining
and Metallurgy, and Ph.D. (1978)
from the Geological Institute, Poland.
From 1967 to 1981 he was a university lecturer and research geologist in
the Geological Institute. He worked
for Mobil in Dallas from 1982 to 1999.
His research interests focus on paleogeography, plate tectonics, paleoclimatology, and global and
regional geology.

Vous aimerez peut-être aussi