Vous êtes sur la page 1sur 32

Economic Geology

Vol. 96, 2001, pp. 17431774

The Evolution of a Porphyry Cu-Au Deposit, Based on LA-ICP-MS Analysis of


Fluid Inclusions: Bajo de la Alumbrera, Argentina
THOMAS ULRICH,,* DETLEF GNTHER,**
ETH Zrich, Isotope Geochemistry and Mineral Resources, Department of Earth Sciences, ETH Zentrum NO, 8092 Zrich, Switzerland
AND

CHRISTOPH A. HEINRICH

ETH Zrich, Isotope Geochemistry and Mineral Resources, Department of Earth Sciences, ETH Zentrum NO, 8092 Zrich, Switzerland,
and Faculty of Mathematics and Natural Sciences, University of Zrich, Ramistrasse 17, 8006 Zrich, Switzerland

Abstract
The chemical and physical evolution of magmatic to hydrothermal processes in the porphyry Cu-Au deposit
of Bajo de la Alumbrera (northwestern Argentina) has been reconstructed with a quantitative fluid inclusion
study. Fluid inclusion petrography, microthermometry, and single inclusion microanalysis by Excimer laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-MS) are combined to determine the evolution
of pressure, temperature, and ore metal concentrations (including Cu and Au) in the fluids. Complementary
hydrogen and oxygen isotope analyses are used to further constrain the water sources in the evolving system.
The combined data provide a new level of insight into the mechanisms of metal sourcing and ore mineral precipitation in a porphyry-style magmatic-hydrothermal system.
Based on previously reported observations of the igneous geology, alteration geochemistry, and veining history of the subvolcanic porphyries at Alumbrera, the distribution of fluid inclusion types in space and time is
documented. Six major inclusion types are distinguished. The highest temperature brine inclusions (up to
750C; P >1 kbar) are mainly recorded in barren quartz magnetite veins in the core of the alteration system.
These polyphase brine inclusions (halite sylvite + multiple opaque and transparent daughter crystals) are interpreted as the most primitive magmatic fluid recorded at the level of the deposit. They are of moderately
high salinity (5060 wt % NaCl equiv) dominated by NaCl, KCl, and FeCl2, and contain on average 0.33 wt
percent Cu and 0.55 ppm Au. Upon cooling and decompression, these saline liquids exsolve a vapor phase,
which is preferentially enriched in Cu relative to its main salt components but probably plays a minor role in
the formation of this particular deposit because of the inferred small mass fraction of vapor. Cooling and decompression from the highest initial P-T conditions down to about 450C causes magnetite K silicate alteration but no saturation in Au or Cu sulfides, as recorded by continually high ore metal concentrations in the
fluid inclusions. Coprecipitation of Cu and Au as chalcopyrite and native gold ( some early bornite) occurs
over a narrow range of decreasing fluid temperature. With cooling from ~400 to 305C, the Cu concentration
in the brine drops by about one order of magnitude to less than ~0.07 wt percent, without a proportional decrease in major salt components. Ore mineral precipitation extracts ~85 percent of the Cu and Au from the
fluid. It is associated with potassic alteration, as shown by a concomitant decrease in the K/Na ratio of the cooling magmatic brine and by an increase in its Ba and Sr concentrations (elements which are probably liberated
in the destruction of calcic igneous minerals).
The fluid chemical data demonstrate that the metal ratios in this and probably many other porphyry-style
ore deposits are primarily controlled by the magmatic source of the ore brines. On the other hand, the final hypogene ore grade of the deposit is controlled by the efficiency of ore mineral precipitation. At Alumbrera,
metal extraction is governed by the efficiency of cooling a high flux of magmatic fluid within a small rock volume. Dilution of residual magmatic fluids, as recorded by aqueous fluid inclusions of decreasing salinities and
temperatures below 295C, follows after the main stage of copper introduction and is associated with feldspardestructive (phyllic) alteration. Geometric relationships, fluid analyses, and stable isotope data together indicate that phyllic alteration results from postmineralization hydrothermal activity involving minor mixing between meteoric water, residual brine, and a waning input of magmatic vapor.

Introduction
HYDROTHERMAL fluid transport associated with shallow
crustal magmatism is a key process contributing to the formation of major metal resources. Among the magmatic-hydrothermal ore deposits, porphyry Cu-Au-Mo deposits are
the economically most important today and have been the
Corresponding

author: e-mail, thomas.ulrich@anu.edu.au


*Present address: Geology Department, The National University of Australia, 0200 Canberra, Australia.
**Present address: Laboratory for Inorganic Chemistry, ETH Zentrum
CAB, 8092 Zrich, Switzerland.
0361-0128/01/3206/1743-32 $6.00

subject of intensive exploration and research over the last


decades. The formation of porphyry-style and other magmatic-hydrothermal orebodies requires a subtle interplay between large-scale tectonics, magma generation, and emplacement in the upper crust, advective flow of hydrothermal fluid
of diverse provenance, and complex chemical reactions between fluids, precipitating ore minerals, and host rocks undergoing hydrothermal alteration. The complexity of these
processes demands an integrated approach of geologic, geochemical, and experimental investigations (cf. Hedenquist
and Richards, 1998).

1743

1744

ULRICH ET AL.

The key role of hydrothermal fluids in the formation of ore


deposits associated with shallow intrusions was recognized
very early (e.g., Lindgren, 1905), and numerous studies on
fluid inclusions have demonstrated many common features
among porphyry copper deposits worldwide (e.g., Roedder,
1971; Nash, 1976; Chivas and Wilkins, 1977; Etminan, 1977;
Eastoe, 1978; Wilson et al., 1980; Weisbrod, 1981; Preece and
Beane, 1982; Reynolds and Beane, 1985; Quan et al., 1987;
John, 1989; Bodnar and Cline, 1991; Cline and Bodnar, 1994;
Beane and Bodnar, 1995). High-temperature fluid inclusions
with very high salinity are generally associated with low-density vapor inclusions and indicate a magmatic contribution to
the metal-transporting fluids (Bodnar, 1995). Stable isotope
data of igneous phases, alteration minerals, and fluid inclusions have shown that magmatic as well as meteoric waters
are involved in porphyry copper systems (e.g., Taylor, 1974;
Batchelder, 1977; Solomon, 1983; Bowman et al., 1987; Dilles
et al., 1992; Oreskes and Einaudi, 1992; Rye, 1993; Sheets et
al., 1996; Hezarkhani and Williams-Jones, 1998). Fluids from
both sources may be essential for ore formation (Henley and
McNabb, 1978), but the present weight of geologic, experimental, and fluid chemical evidence indicates that the metals
are introduced dominantly by magmatic fluids (Candela and
Holland, 1986; Heinrich et al., 1992; Bodnar, 1995; Cline,
1995; Shepherd and Chenery, 1995; Hedenquist and
Richards, 1998; Ulrich et al., 1999). However, the sources and
the geochemical role especially of low- to intermediate-salinity fluids of variable density remain ambiguous (Shinohara
and Hedenquist, 1997; Heinrich et al., 1999).
Quantitative information about the major and trace element composition of the metal-transporting hydrothermal
fluids has been a limiting factor in reconstructing the magmatic-hydrothermal enrichment process in time and space.
Petrographic observations of ore mineral daughter crystals
precipitated inside fluid inclusions (Roedder, 1984, see refer-

ences therein) and more recent microanalytical data (Diamond et al., 1990; Boiron et al., 1991; Heinrich et al., 1992;
Mavrogenes et al., 1995; Damman et al., 1996) show that
magmatic fluid inclusions can contain very high ore metal
concentrations. With the recent development of laser ablation-inductively coupled plasma-mass spectrometry (LAICP-MS; Gnther et al., 1998; Audtat et al., 2000a, b), a single inclusion analytical technique of sufficient sensitivity,
quantitative accuracy, and routine applicability has become
available for extensive studies of texturally complex samples
(Audtat et al., 1998). In this paper we report the first study
of a world-class magmatic-hydrothermal ore deposit, where
LA-ICP-MS analyses of some 300 individual fluid inclusions
in 27 samples are integrated with conventional fluid inclusion
techniques and other geochemical data within a good geologic framework, as reported in Ulrich and Heinrich (2001).
The aim of this study is to quantify the behavior of ore metals
and other elements involved in hydrothermal mass transfer
by high-temperature brines, vaporlike fluids, and low-salinity
aqueous solutions. By combination with stable isotope data
we determine the sources of fluids and metals independently
and define the mass transfer and the dominant causes for
metal precipitation in the orebody.
Geologic Setting
The Cu-Au deposit of Bajo de la Alumbrera in northwestern Argentina (lat 2730' S, long 6660' W) is hosted by several high K dacite porphyries that intruded into andesitic
flows and pyroclastic rocks of the Faralln Negro Volcanic
Complex. From crosscutting relationships and petrographic
observations, several porphyries can be distinguished (Table
1) that form a composite intrusive complex (Ulrich and Heinrich, 2001; J. M. Proffett, writ. commun.). The Ar/Ar dates of
biotite from the earliest mineralized P2 Porphyry (7.10

TABLE 1. Relative Timing Relationships between the Porphyry Intrusions at Bajo de la Alumbrera
Porphyry

Timing relationships

Alteration/mineralization

Los Amarillos Porphyry

Cut by Early P3 and Northwest Porphyries

Intensely feldspar destructive; barren to weakly mineralized

NE Porphyry

Cut by Early P3; no other timing relationships

Moderate potassic, chlorite epidote; weakly to moderately


mineralized

P2 Porphyry

Cut by Early P3 Porphyry and all younger porphyries;


associated with purple quartz veins

Quartz-magnetite, strong potassic; strongly mineralized

Early P3 Porphyry

Cuts veins of P2 Porphyry; is cut by Late P3 and


younger porphyries

Partly quartz-magnetite, strong potassic; moderately to


strongly mineralized

Quartz-Eye Porphyry

Cut by Late P3; contains most of the purple quartz veins

Intensely feldspar destructive; weakly to moderately


mineralized

Late P3 Porphyries
(Campamiento and
North Porphyries)

Cuts Early P3 Porphyry; is intruded by later porphyries

Weak-moderate potassic, weak feldspar destructive; weakly


mineralized

Northwest Porphyry dikes

Intrudes P2 and Early P3 Porphyries

Weak potassic; barren

Post-Mineral Porphyry

Cuts Late P3 Porphyries

Commonly unaltered, slightly feldspar destructive; barren

Summary of the geologic relationships among the different porphyry intrusions; for a complete description of the geology see Ulrich and Heinrich (2001)
and J.M. Proffett (unpub. data)
0361-0128/98/000/000-00 $6.00

1744

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

0.13) and a postore dike (6.83 0.07 Ma) constrain the late
Miocene age of mineralization (Sasso and Clark, 1998). The
P2 and the Early P3 Porphyries are the two distinct intrusions
hosting the bulk of the economic copper-gold mineralization,
which was introduced by at least two separate magmatic-hydrothermal fluid pulses into a dense network of quartz magnetite-chalcopyrite veinlets. The slightly later Quartz-Eye
Porphyry is overprinted by a stockwork of characteristic pink
quartz veins that are associated with intense feldspar-destructive alteration in outcrops. These veins contain only minor
bornite and chalcopyrite. At least four subsequent stages of
minor intrusion are weakly altered and essentially unmineralized, and all have much lower vein densities than the P2, the
Early P3, and the Quartz Eye Porphyries (Ulrich and Heinrich, 2001; J. M. Proffett, writ. commun.; Table 1).
The alteration sequence is characterized by an early quartzmagnetite assemblage in the center of the deposit. This alteration is concentrated on the P2 Porphyry, but a second
weaker phase of this alteration also affects parts of the Early
P3 Porphyry. The adjacent potassic alteration (K feldspar +
biotite magnetite) is most intensely developed in the mineralized porphyries but affects in a less intense form (secondary biotite) some of the later intrusions, such as the
Northwest Porphyry dikes. This alteration is surrounded by a
propylitic zone at the fringe of the deposit, overprinting
mainly the andesitic volcanic rocks and some porphyry dikes.
Concentric zones of quartz-magnetite, potassic, and propylitic alteration of the mineralized stock and all later intrusions
are overprinted by feldspar-destructive alteration, which advanced from the top of the deposit downward along steep
crosscutting structures and lobes. Mineralization is dominated by chalcopyrite and intimately intergrown gold. Ore
mineral precipitation is associated with potassic alteration and
occurs as disseminations in the porphyries and proximal andesitic volcanic rocks and as texturally late phases in quartz
magnetite veinlets. Economic Cu and Au concentrations
occur in a high-grade ore shell above and around a barren
core of highly quartz-magnetite-altered but sulfide-poor
porphyries.
Sample Material for Fluid Inclusion and Isotope Studies
Samples with veins and phenocrysts for petrographic, fluid
inclusion, and stable isotope studies were taken from original
surface outcrops and the developing mine, but most samples
were collected from diamond drill core (Appendices 1 and 2).
The selection of samples was based on core logging and surface mapping of the main alteration zones (quartz-magnetite,
potassic, propylitic, and feldspar-destructive alteration) and
the distribution of high Cu-Au grades. The vertical sampling
extends to a depth of 550 m below the premining land
surface.
Vein densities and crosscutting relationships were most
clearly observed in the P2, Early P3, and the Quartz-Eye Porphyries. Fluid inclusion and isotope data were derived from
samples of the major vein and alteration types in these intrusions. The Early P3 Porphyry is intimately associated with the
second major pulse of mineralization and potassic and quartzmagnetite alteration. It provided the preferred and some of
the best fluid inclusion material. Exceptionally well preserved
fluid inclusions were also found in a 5- to 8-cm-thick vuggy
0361-0128/98/000/000-00 $6.00

1745

vein of coarse-grained purple quartz from a block of QuartzEye Porphyry in the southeastern part of the open pit (sample BLA97). Euhedral crystals up to 1 cm contain inclusion
assemblages in a particularly clear time sequence. Biotite inclusions indicate that this quartz crystallized during potassic
alteration. Pyrite and possibly anhydrite (now gypsum) were
deposited later in the vugs, associated with feldspar-destructive overprint of the wall rock.
Fluid inclusion assemblages in the small quartz veinlets related to feldspar-destructive and chlorite-epidote alteration
were inadequate for analysis. Suitable inclusions probably representing the feldspar-destructive stage were observed on secondary trails in strongly feldspar destructively overprinted
samples, even though the quartz host in these samples belongs
to veins of the potassic stage. Secondary trails are very common in these samples and cut trails of primary or pseudosecondary inclusions preserved from the potassic veining stage.
For a reconnaissance stable isotope study, selected minerals were separated from veins and whole-rock samples to obtain an enriched mineral fraction of 90 percent purity or
more. Fluid inclusion samples were also used for isotopic
analysis, as far as possible. Quartz phenocryst separates of the
least altered intrusion of the Northwest Porphyry were analyzed to obtain an isotopic signature representative for the
magmas. Coexisting quartz and magnetite from the quartzmagnetite alteration zone were used to estimate both the
18O composition and the temperature conditions of the earliest alteration stage at Alumbrera. To estimate the oxygen
isotope composition of fluids associated with the main alteration zones, we analyzed the quartz veins related to potassic
alteration in three samples of the Early P3 and Quartz-Eye
Porphyries and quartz separates from late veinlets and matrix
quartz associated with feldspar-destructive alteration. Secondary biotite was analyzed for 18O in three wall-rock andesites with intense potassic alteration, and two albite samples were analyzed to represent the chlorite-epidote zone.
Hydrogen isotopes were determined on biotite of potassically altered rocks, on chlorite of the chlorite-epidote zone,
and sericite and/or clay fractions from samples with feldspardestructive overprint. Hydrogen isotopes of fluid inclusions
were analyzed after thermal extraction of inclusion water
from veins related to the quartz-magnetite, the potassic, and
the feldspar-destructive alteration.
To obtain an estimate of the local meteoric water at the
time of hydrothermal activity, we analyzed quartz and fluid
inclusion extracts from the post-sulfide stage of an epithermal
vein of the Faralln Negro gold mine, located 15 km northwest from Alumbrera along the same structural trend
(Malvicini and Llambias, 1963).
Petrography and Distribution of Fluid Inclusions
Fluid inclusion types
Petrography and all subsequent single inclusion measurements strictly concentrated on fluid inclusion assemblages,
i.e., closely associated groups or trails of inclusions with visually identical phase ratios and similar shape (Goldstein and
Reynolds, 1994). Six types of fluid inclusions were distinguished according to the number, the nature, and the volume
proportions of phases present at room temperature (Fig. 1).

1745

1746

ULRICH ET AL.

polyphase brine

opaque-bearing brine

brine
halite

hematite
halite

halite

20m

20m

A
aqueous

20m

B
high-density

20m

20m

lowdensity

chalcopyrite

20m

Vein quartz

Low-density
vapor inclusions

High-density
vapor inclusions
Quartz phenocryst
50m

20m

Polyphase
brine inclusions

Low-density
vapor inclusions

Low-density
vapor inclusions

Polyphase
brine inclusions

20m

50m

FIG. 1. Photomicrographs of inclusion types and relative timing among assemblages. A. Polyphase brine inclusion with
hematite and three unidentified daughter phases. B. Opaque-bearing brine inclusion characterized by one or two opaque
daughter crystals. C. Simple brine inclusions containing only a halite daughter crystal, which has a similar or slightly larger
size than the vapor bubble. D. Two-phase aqueous inclusion (325 m) with a larger or similar proportion of liquid than
vapor. E. High-density vapor inclusions, commonly with a chalcopyrite crystal (1545 m). F. Low-density vapor inclusions
containing more than 80 percent vapor. G. A low-density vapor trail is cut by a similar trail with small high-density inclusions
(sample BLA97). H. Fluid inclusion relationship in a phenocryst that is crosscut by a quartz veinlet (stippled line shows the
contact). The small polyphase and opaque-bearing brine inclusions (I) occur exclusively in the phenocryst (see also Fig. 2A).
J. and K. Trails of small polyphase brine inclusions are cut by later low-density vapor trails.
0361-0128/98/000/000-00 $6.00

1746

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

Three inclusion types are halite saturated, generally have


rounded isometric or negative crystal shapes, and are typically
20 to 30 m in size. The polyphase brine inclusions contain a
dense packing of halite and up to five other transparent and
opaque daughter phases (Fig. 1A). Sylvite crystals, always
smaller than the halite crystal, were only identified with certainty in a small number of inclusions, as the only isotropic
crystal other than halite. Metallic phases including hematite
(red, platy) and chalcopyrite (opaque, triangular) are invariably present in polyphase brine inclusions. The size of measured inclusions ranges from 10 to 50 m. Some of these
highly saline inclusions show minor variations in occurrence
and nature of daughter crystals even among clearly coeval inclusions in an assemblage, probably due to non-nucleation of
internal precipitates (non-nucleation; Roedder, 1984). The
opaque-bearing brine inclusions have only a halite crystal plus
one or two opaque phases, one of which can be hematite. Inclusion size ranges from 5 to 40 m (Fig. 1B). The simple
brine inclusions are between 5 and 35 m large and contain a
halite cube as the only daughter crystal (Fig. 1C).
Aqueous inclusions have a high filling ratio (avg 80% liquid
and 20% vapor) and occur as irregular-shaped to rounded inclusions (Fig. 1D) in texturally late quartz crystals or as secondary inclusions in earlier quartz. They rarely are larger than
15 m.
Vapor-rich inclusions were divided, according to their filling ratio, into high-density vapor inclusions (6090% vapor;
Fig. 1E) and low-density vapor inclusions (generally >90%
vapor; Fig 1F). Both types commonly reach 25 m in size and
have slightly rounded to perfect negative crystal shapes. Liquid occupying less than 20 percent is generally invisible due
to internal reflection of the inclusion walls, except in particularly large inclusions near protrusions. High-density vapor inclusions often contain a small chalcopyrite crystal (Fig. 1E).
No daughter crystals have been observed in low-density
inclusions.
Relative timing of fluid entrapment
The sequence of fluid entrapment was inferred from the
distribution of inclusion assemblages in veins of different paragenetic stages (Ulrich and Heinrich, 2001), from petrographic observations of crosscutting trails of inclusions along
individual fractures (Fig. 1G, J, K), or rare primary inclusions
or pseudosecondary trails that can be attributed to specific
growth zones in euhedral vein crystals (Fig. 1H, I; see Roedder, 1984; Goldstein and Reynolds, 1994). However, the commonly random distribution and superimposition of numerous
inclusions in most samples made the determination of relative
timing between inclusion assemblages very difficult.
Veins related to quartz-magnetite and potassic alteration
phases of the P2 and Early P3 Porphyries have indistinguishable inclusion populations, and we infer (but cannot unambiguously prove) that the fluid history of the two main mineralizing pulses was similar. Rare crosscutting relationships
between inclusion trails indicate that the polyphase brine inclusions are earlier than the other brine inclusion types, and
that trails containing only polyphase brine inclusions are earlier than trails containing simultaneously trapped polyphase
and vapor inclusions. No direct crosscutting relationships involving simple brine inclusions were observed, but these are
0361-0128/98/000/000-00 $6.00

1747

considered later than the polyphase and opaque-bearing


brine inclusions, because of their generally lower homogenization temperature in individual samples.
Vapor inclusions occur as trails and clusters and low-, as
well as high-density, types were found together with
polyphase brine inclusions in boiling assemblages. These are
single trails containing vapor- as well as liquid-rich inclusions,
trapped from a two-phase fluid in one microfracture. The former coexistence of brine and vapor is unequivocally demonstrated by such boiling trails, but polyphase brine and vapor
inclusions are more commonly randomly distributed in large
numbers in the core of quartz crystals. The possible coeval
trapping of such assemblages is difficult to prove. Some assemblages of high-density vapor inclusions in phenocrysts
and veins show early and probably primary character, whereas
low-density vapor inclusions generally occur on sharp secondary trails cutting through entire grains, or in paragenetically late veins, indicating a later stage of entrapment.
Some samples from the Quartz-Eye Porphyry have preserved a particularly clear record of fluid entrapment, and inclusion assemblages from sample BLA97 (Fig. 2B) have
therefore been singled out in the subsequent data presentation. Quartz phenocrysts contain a generation of small (10
m) polyphase brine inclusions with one or more transparent
daughter crystals and a prominent red hematite flake. The
same type of inclusion was also found together with rare highdensity vapor inclusions (Fig. 1E) in a primary growth zone in
the core of the euhedral quartz crystal in the vuggy vein. The
crystal records a complex fluid history starting with primary
or early pseudosecondary polyphase brine inclusions without
coexisting vapor, followed by later polyphase brine inclusions
with high-density vapor inclusions. Subsequent pseudosecondary polyphase brine plus high- and low-density vapor inclusions are associated with biotite precipitation, before the
crystal is rimmed by a thin final growth zone containing irregular aqueous and low-density vapor inclusions. The latter
fluids slightly predate or are coeval with the precipitation of
pyrite in the residual vug and probably with feldspar-destructive alteration of the wall rock.
Aqueous inclusions are generally late relative to all other
liquid-rich inclusions and to all phases of quartz magnetite
stockwork veining. They were found in late anhydrite veins
and as secondary inclusions in earlier quartz veins occurring
in samples with an intense feldspar-destructive overprint.
Here they commonly occur together with low-density vapor
inclusions but never in clear boiling assemblages.
Deposit-scale distribution of fluid inclusions
The distribution of the different inclusion types on the
scale of the deposit was studied in an attempt to delineate the
relationships between fluids, alteration, and ore deposition.
The relative abundance of inclusions was visually estimated
for each type of inclusion in a total of 38 polished sections
(Table 2). The schematic section (Fig. 3) shows the ore-grade
distribution and alteration zones in section 47, with additional
sample and alteration data projected from neighboring
sections (see Ulrich and Heinrich, 2001; J.M. Proffett, writ.
commun.).
Halite-saturated inclusions generally exceed 50 percent of
the inclusions in samples from the central part of the deposit.

1747

1748

ULRICH ET AL.

Sample BLA 1/97 A, II

Sample 45-61.3 S

magnetite
pyrite

phenocryst

high-density
vapor inclusions

vein quartz

polyphase brine and


high-density vapor
inclusions

high-density
vapor inclusions

high-density
vapor inclusions
polyphase and
high-density
vapor inclusions

aqueous
inclusions
polyphase brine
inclusions and
low-density
vapor inclusions

polyphase
brine inclusions

heterogeneous
trapping of vapor-rich
and polyphase
brine inclusions

secondary polyphase
brine inclusions
low-density
vapor inclusions and
polyphase brine inclusions
1 mm

growth zone with irregular


small aqueous and
vapor-rich inclusions

primary? polyphase and


opaque-bearing brine
inclusions with hematite

low-and high-density
vapor inclusions

high-density
high-density
vapor inclusions vapor inclusions and opaquebearing brine inclusions

biotite
low-density
vapor inclusions
small polyphase and
opaque-bearing brine
inclusions,some
high-density
vapor inclusions
opaque-bearing
brine inclusions

primary and pseudosecondary? polyphase


brine inclusions

1 mm

FIG. 2. A. Sketch of a phenocryst cut by a veinlet (sample 45-61.3 S). Small, hematite-bearing inclusions (smallest dots)
are restricted to the phenocryst, whereas later generations of polyphase and opaque-bearing brine inclusions also occur in
the vein. B. Simplified drawing of crystals from sample BLA97 showing the relative timing relationships of saline and vapor
inclusions. The polyphase and opaque-bearing brine inclusions occur in the core, whereas the vapor inclusions are on trails
and commonly along the outer zone of the crystal.

The samples with the largest numbers of polyphase and


opaque-bearing brine inclusions are associated with the lowgrade core and the quartz-magnetite, as well as the potassic,
alteration. Simple brine inclusions are not very abundant and
show no association with the alteration zones nor any correlation with ore grades.
Toward the fringe of the deposit, polyphase and opaquebearing brine inclusions diminish and vapor-rich inclusions,
invariably associated with aqueous inclusions, become predominant. This zoning is correlated with the feldspar-destructive overprint of some of these samples. Aqueous inclusions throughout the deposit are secondary and clearly belong
to the incipient feldspar-destructive overprint in many samples. Petrography and abundance distribution of the different
inclusion types possibly indicate a trend of decreasing salinity
from the center to the periphery of the deposit and from early
to late fluid stages in any one sample. The salinity and temperature distribution is further quantified in the following
section on the basis of microthermometric measurements.
Microthermometry
Experimental approach and evaluation of measurements
Heating and freezing experiments were conducted on a
Linkam THMS600 stage for homogenization temperature up
to 600C and on a Linkam TS1500 stage for temperatures
0361-0128/98/000/000-00 $6.00

above 600C. The precision for freezing runs is about 0.1C


and for heating runs 2C at 500C. The uncertainty for hightemperature measurements with the Linkam TS1500 stage
may be up to 20C. Apparent salinities are reported in weight
percent NaCl equivalent, based on the halite solubility equation for halite-saturated inclusions and on the final ice melting for halite-undersaturated inclusions (Bodnar and Vityk,
1994). Minor CO2 detected by laser Raman microprobe in
some of the high-density vapor inclusions may cause errors in
the calculation of salinities, but clathrates were generally absent or minor in abundance at the temperature of final ice
melting (see Table 3 for abbreviations).
Microthermometric results reported in Tables 4, 5, 6, and all
figures strictly refer to fluid inclusion assemblages. The salinity
within a single assemblage is quite uniform (typically 3 wt %
NaCl equiv) but varies greatly between different assemblages.
Freezing measurements and heating experiments up to the
temperature of halite dissolution were carried out before LAICP-MS analysis to obtain the apparent salinities needed for
internal standardization of elemental analyses (see below).
Final homogenization above 600C led to stretching and decrepitation of some inclusions. Limited high-temperature
homogenization data were therefore measured with inclusions remaining in the same assemblage after LA-ICP-MS or
on similar assemblages from other parts of the same section.
Average homogenization temperatures of these were then

1748

0361-0128/98/000/000-00 $6.00

1749

51-61.1/6
51-61.1/11
47-55.3/10
51-61.1/5
45-61.3/13
47-55.3/7
47-55.3/9
51-52.2 /9
51-52.2 /11
51-52.2/13
51-52.2/14
47-53/48
49-50.1/4
49-50.1/8
51-61.1/8

50.3-59.4/8
47-53/72
49-52/3
49-52/19
47-53/25
47-53/30
47-53/62
49-52/10
49-52/21
49-50.1/1
49-61.4/16
49-61.4/19
50.3-59.4/6
49.2-46.3/4
49.2-46.3/5
49.2-46.3/8
49-52/9
49-52/14
49-51/1
49-51/3
49-51/5

BLA 97

BLA 18

AK
AM
J
AJ
A
H
I
AF
AG
AH
AI
D
M
N
AL

AE
F
R
V
B
C
E
T
W
L
Y
Z
AD
AA
AB
AC
S
U
O
P
Q

AO

AN

Surface

Surface

146
413
226
334
118
128
311
260
337
354
277
283
89
278
279
292
242
315
182
184
192

524
547
147
523
428
110
146
204
211
212
224
226
357
371
537

Depth
(m below
surface)

0.74
0.51
0.05
0.10
0.06
0.10
0.30
0.08
0.05
0.05
0.58
1.40
0.72
0.80
0.80
0.77
0.06
0.14
0.94
0.85
0.91

0.02
0.01
1.40
0.02
1.40
0.50
1.40
0.01
0.04
0.04
0.11
0.56
0.09
0.05
0.01
1.28
0.75
0.14
0.06
0.10
0.17
0.23
0.06
0.05
0.02
0.80
1.94
2.12
0.53
0.53
0.71
0.05
0.16
1.55
1.50
1.48

0.03
0.02
1.95
0.03
2.43
0.78
1.95
0.02
0.09
0.09
0.11
0.40
0.04
0.05
0.07

Ore grades
Cu
Au
(wt %) (ppm)

Feldspar destructively overprinted,


Quartz-Eye Porphyry
Quartz phenos in feldspar destructively overprinted,
Quartz-Eye Porphyry

Quartz-magnetite, Early P3
Quartz-magnetite, Early P3
Quartz-magnetite, Early P3
Potassic with quartz-magnetite, Early P3
Potassic, Early P3
Potassic, Early P3
Potassic, Early P3
Potassic, Early P3
Potassic, Early P3
Potassic, Early P3
Feldspar destructively overprinted, Early P3
Feldspar destructively overprinted, Early P3
Feldspar destructively overprinted, Early P3
Feldspar destructively overprinted, Early P3
Feldspar destructively overprinted, Early P3
Feldspar destructively overprinted, Early P3
Slightly feldspar destructively overprinted, Early P3
Slightly feldspar destructively overprinted, Early P3
Slightly feldspar destructively overprinted, Early P3
Slightly feldspar destructively overprinted, Early P3
Slightly feldspar destructively overprinted, Early P3

Quartz-magnetite, P2 Porphyry
Quartz-magnetite, P2 Porphyry
Quartz-magnetite, P2 Porphyry
Quartz-magnetite with calcite filling, P2 Porphyry
Potassic with quartz-magnetite, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry
Potassic, P2 Porphyry

Porphyry type and alteration

B
B

10

50

5
10
15
5
40
55
10
10
15
15
30
5
10
20
20
15
30
5
60
10
35

10
10
40
25
35
50
25
60
50
50
45
10
55
40
25

Polyphase
(%)

15

15
5

25
10

5
20

30
30
10
20
5
40
50
10
25
5

15
25
15

10
5
5

25
5
25
10

Opaque
bearing
(%)

5
5

10
5
50

10
5
5

15
5
5
10
15

15

10
60
25
20

10
45

15
15
20
30

10
5

35
60

Aqueous
(%)

30
5
45
35

15
15

10

5
5
15
20
10
5

25

Brine
(%)

50

20

5
5
15
5
5
5
10
10
25
25
20
20
10
5
10
30
20

25
10
10

5
15
20

10
10
40
15
10
5
5

10

25

10

60
50
20
10
10
15
5
15
25
15
55
60
50
25
60
55
20
15
30
40
40

20
5
40
30
10
35
50
25
10
15
10
25
15
30
20

Vapor inclusions
HighLowdensity
density
(%)
(%)

Selected samples that were used to delineate the distribution of the different inclusion types in the deposit; the abundance of each inclusion type was estimated optically from double-polished sections; sample ID refers to the frequency diagrams of Figure 3; B = sample with boiling assemblage

Drill hole

Sample
ID
(Fig. 3)

Table 2. Distribution of Fluid Inclusions in Samples Used in This Study

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

1749

1750

ULRICH ET AL.
TABLE 3. Nomenclature of Fluid Inclusion Measurements

Th(L)
Tm(halite) > Th
Th(V)
Tm(ice)
Tm(CO2)
Tm(halite)

Final homogenization temperature to liquid


Final homogenization by halite dissolution
Final homogenization to vapor
Last ice melting temperature
Melting temperature of CO2
Melting temperature of halite

assumed for the analyzed inclusions. Measuring several assemblages of identical petrographic appearance within one
section generally showed consistent homogenization temperatures, indicating that this assumption is valid.
Results
The highest temperature was measured in an assemblage of
opaque-bearing brine inclusions that homogenize to the liquid at 771 16C (assemblage 92, Table 4). In general, highest homogenization temperatures of polyphase brine and

opaque-bearing brine inclusions are around 710 to 730C


(Figs. 4A, 5F, G). Among the simple brine inclusions, only the
assemblage of sample BLA97 homogenized above 600C, at
671 6C (assemblage 104). Two high-temperature boiling
assemblages (assemblages 114, 115) showed consistent homogenization of the polyphase brine inclusions at 720C, and
this value was adopted for assemblages 113 and 117, which
did not homogenize below the upper temperature limit of the
heating stage (600C). Homogenization of the polyphase,
opaque-bearing, and brine inclusion assemblages occurs either by halite melting or to liquid (Tm(halite) > Th(L) or Th(L)).
Polyphase brine inclusions: The data for all polyphase brine
inclusions in Figure 5A and F show a wide range in homogenization temperature and salinity values with a main mode at
400 to 500C and a consistent high-temperature mode above
700C, which mainly represents the assemblages of sample
BLA97. The assemblage with the lowest Th(L) homogenizes at
432 6C and has a salinity of 47 1 wt percent NaCl equiv,
whereas the assemblage (62) with the highest salinity (71 1
wt % NaCl equiv) homogenizes by halite dissolution (Tm(halite)
> Th(L) = 580 3C). Within some assemblages, inclusions

TABLE 4. Microthermometry Data and Pressure Estimates


Assemblage
21
4
20
25
110
100
101
106
108
111
109
73
112
54
116
58
59
107
57
62
47
104
113
117
115
114
61
67
66
68
93
91
92

Sample
49-52 19A
46-59QZE
49-52 14
51-61.1 II,5/1
49-52 3/3
49-52, 19B/3
49-52, 19b/3 IV
45-61.3, 13/1
49-50,1, 1/8
49-52 3/3
49-50,1, 4/6
47-53, 62/1
49-52,21/3
49-51, 5/5
BLA1/97A I,2
51-61.1, 5/ 3
51-61.1, 5 IV
47-53/25B
51-61.1 5 I
BLA1/97B,2
47-53 25B
BLA1/97B I,1
51-52,2, 13
BLA1/97B,2
BLA1/97A,1
BLA1/97A,1
BLA1/97B2
BLA IV 97, 3/1
Avg BLA IV 97, 3
BLA IV 97, 2/6
BLA2/97 B2
BLA 1/97B II,1
BLA2/97B2

Th(L) and Th(halite)


(C)

Salinity (wt %
NaCl equiv)
3 0.3
7.1 4.2
1.4 0.3
17.7 0.1
35 0.5
41.4 3.9
42.1 3.7
43.9 2.8
41 4.7
48.4 4.0
46 3.8
44 0.3
48.4 6.0
50.9 4.3
52.7 3.0
49.9 1.8
50 0.6
61.7 2.5
45.1 3.3
71.1 0.2
61.4 2.9
36.9 2.8
52 0.4
65.5 2.1
63.8 1.9
58 1.1
66.1 3.6
64 1.6
62 1.2
56 4.0
53 1.5
36.9 2.8
46.5 3.0

240 0.3
242 2.6
271 20.7
292 10.0
305 5.0
368 39.6
387 43.2
397 6.5
400 14.1
419 26.2
421 1.0
423 22.5
448 3.5
463 19.0
470 14.1
471 18.4
499 12.6
530 0.3
540 69.1
580 3.0
6652
671 6.0
7203
7203
7202
7202
7222
7223
7223
7223
7302
7303
771 16.3

P (bars)

Basis for P
estimate

50
50
80
80
80
2810/1501
150
180
200
250
220
220
270
320
320
340
380
370
530
1135/3801
620
1100
1050
680
750
900
680
750
800
950
1080
1280
1320

LV curve
LV curve
LV curve
LV curve
Boiling curve
Halite disappearance
LV curve
Boiling curve
Boiling curve
Boiling curve
Boiling curve
LV curve
Boiling curve
LV curve
Boiling curve
LV curve
LV curve
Boiling curve
LV curve
Halite disappearance
LV curve
LV curve
Boiling curve
Boiling curve
Boiling curve
Boiling curve
LV curve
LV curve
LV curve
LV curve
LV curve
LV curve
LV curve

Microthermometric data and pressure estimate of boiling assemblages and single-phase assemblages that are shown in Figure 6 and which were subsequently analyzed by LA-ICP-MS; sorting of the table is based on increasing homogenization temperatures (roughly corresponding to a sequences of late to
early assemblages); pressures for nonboiling assemblages are minimum pressures
1 Pressure calculated according to Bodnar (1994) for halite homogenizing inclusions and the assumed pressure, respectively (see text for explanation)
2 Single analysis due to decrepitation of inclusions from different assemblages
3 Extrapolated values from measurements of similar but different assemblages (see text for explanation)
0361-0128/98/000/000-00 $6.00

1750

1751

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA


Deposit-scale distribution of fluid inclusion types

WSW

ENE

3
6.
-4
-.2
49

AO

Gypsum
Fault

50
.3
-5
9.
4
51
-6
45 1.
1
-6
49 1.4
-6
1.
4
47
-6
2

AN

1.

3
5.
-5
47
5
-5
47

3
-5
47 .2
2
-5
51 -52
49 2

-5 1
47 9-5
4
1
0.
-5
49

m a.s.l.
2600

AD

2500

AF

AH

AE
AG
AI

2400

C
S

AA

P
Y

AB

Z
W

2300

quartz-magnetite
alteration
strong K-feldsparbiotite alteration
moderate and
weak K-feldsparbiotite alteration
secondary
biotite alteration
weak secondary
biotite alteration in
late porphyries
chlorite-epidote
alteration
late feldspardestructive
alteration
Cu-grade contour
0.5 wt%
approximate
boundaries of
main porphyries
against surrounding
andesites

AC
M

ve

Ste

N
A

s
Fa

AJ

ult

2200

AK

AL

AM

average abundance
high-density
vapor incl.

polyphase
brine incl.

low-density
vapor incl.
aqueous incl.

coexisting polyphase
brine and
vapor inclusions

opaque-bearing
brine incl.

low-grade qtz-mag
alteration
core

brine incl.

potassic

high-grade feldsparore zone destructive

22 samples 9 samples 16 samples 16 samples 13 samples

FIG. 3. Schematic illustration of section 47


through the deposit, with alteration and ore-grade
zones from Ulrich and Heinrich (2001) and pie diagrams showing the relative abundance of the inclusion types in single samples. Samples with boiling assemblages are indicated by white centers,
and ticks show the coexistence of low- or high-density vapor inclusions with polyphase brine inclusions. Key shows average overall abundance of the
different inclusion types in the low-grade core (22
samples), quartz-magnetite alteration (9), potassic
alteration (16), the high-grade ore zone (16), and
the feldspar destructively altered parts of the deposit (13). Letters in the pie diagrams correlate
with sample ID of Table 2.

80

Salinity (wt% NaCl equiv.)

Polyphase brine inclusions

64

70

Opaque-bearingbrine inclusions

60

Brine inclusions
Aqueous inclusions

50

High density
vapor inclusions
103
100

40

60

77
85

98
75
110 99

30

79

61

115/117

62
56
49

65-68

70

107
86
50
5155
48
81-83/87116
80
47
94 59
101 52112
96
78
95/54
106
58
74
84
109
57
90
89
88 53
97
102 76
105
108/111

69

63

72
71
114

93

113
92

91

104

5
22
5a

20

anhydrite
14
anhydrite

10

1
calcite

100

34
7

39

30

25
11/18
28/29
27
10/15/19
2624
16
36
12 40
21 4
13
6

17

8 23 20

31/32

37
38
43-46

41/42
116

33

300
200
400
Th L or by halite
melting (C)

500

600

700

800

62

0361-0128/98/000/000-00 $6.00

47

79
83

500

82

60
51

81

Tm halite (C)

FIG. 4. A. Salinity vs. final homogenization temperature (to liquid


or by halite melting) for all inclusion types of Alumbrera. Enclosing
lines represent single assemblages from different samples. Not all
temperatures shown above 600C have been explicitly measured,
some are adopted from similar assemblages (see Table 4) and slightly
shifted for clearer presentation. Stippled lines connect coexisting
vapor and liquid inclusions of boiling assemblages. The gap in salinity around 25 to 35 wt percent NaCl equiv may result from non-nucleation of halite (see Roedder, 1984) or may be the artifact of deviations of the real fluids from the NaCl-H2O model system used for
the salinity estimate (Vityk and Bodnar, 1998). B. Halite dissolution
temperatures as a function of vapor-liquid homogenization temperature to liquid for the halite-saturated inclusion types. Inclusions
plotting above the line undergo final homogenization by halite dissolution, whereas inclusions below the line homogenize to liquid.
Numbers next to the labels refer to the measured assemblages.

900

400
99

300
100

87 74
103
98

75
110 88

200
100
100

1751

86

56

65 61/64
63

55 107
48
50
95 80 96
59
94
112 58 116/54
84

101
73
109/52
106
77 85
53 90
102 76
108/111
97

68

49

70

66/67

69
71
72
115/117
114
93

78

113

92

57
89
91
105

104

Boiling
assemblages

200

300

B
400

500
Th L (C)

600

700

800

900

1752

ULRICH ET AL.
TABLE 5. Summary of the Microthermometric Data and Estimated Pressures of the Different Inclusion Types

Description of assemblages

Microthermometry data

Pressure estimate

Deposit-scale distribution

Polyphase brine assemblages:


Randomly distributed inclusions
of primary character; secondary
trails in phenocrystals and veins
are cut by boiling assemblages;
assemblages homogenizing by
halite melting are later than
those homogenized to liquid

Highest avg Th (L) ~722C


Lowest Th(L) = 430C
Salinity range 40 to 71 wt
% NaCl equiv

Highest P(min) = 950 bars


Lowest P(min) = 350 bars

Abundant in low-grade core,


quartz-magnetite and potassic
alteration

Boiling assemblages: Commonly


secondary trails; coexistence of
polyphase and vapor-rich
inclusions

Highest avg Th(L) ~720C


Lowest Th(L) = 300C
Salinity range 35 to 65 wt
% NaCl equiv

Highest P(min) = 1,050 bars


Lowest P(min) = 80 bars

Low-grade core and high-grade ore


zone, from present surface to at least
430 m deeper

Opaque-bearing brine
inclusions: Randomly distributed
inclusions of primary character in
phenocrystals; assemblages
homogenizing by halite
melting are later

Highest avg Th(L) ~771C


(one assemblage 92)
Lowest Th(L) = 360C
Salinity range 37 to 63 wt
% NaCl equiv

Highest P(min) = 1,320 bars


Lowest P(min) = 220 bars

Dominant in low-grade zone


high-grade ore zone

Brine assemblages:
Only small assemblages,
trails or randomly distributed

Highest avg Th(L) ~671C


(one assemblage 104)
Lowest Th (Tm (halite)) = 336C
Salinity range 36 to 46 wt
% NaCl equiv

Highest P(min) = 1,100 bars


Lowest P(min) = 150 bars

Subordinate in low-grade as well as


in high-grade zone

Aqueous assemblages:
Randomly and on planes,
irregular shaped to rounded
in samples with a feldspardestructive overprint

Highest avg Th(L) = 328C


Lowest Th(L) = 204C
Salinity range 0.2 to 25 wt
% NaCl equiv

P(min) <100 bars

Feldspar-destructive overprint,
peripheral areas of the deposit

High-density vapor assemblages:


Secondary trails cutting polyphase
brine trails; randomly distributed
inclusion of primary character

Highest avg Th(V) >600C


Lowest Th(V) ~370C
Salinity range 1 to 19 wt
% NaCl equiv

Highest P(min) = 760 bars


Lowest P(min) = 280 bars

High-grade ore zone, feldspardestructive overprint, minor in


low-grade zone

Low-density vapor assemblages:


Primary on late growth zones
related to feldspar-destructive
alteration; secondary trails cut
polyphase brine trails

Feldspar-destructive overprint,
peripheral areas of the deposit

homogenize both to liquid and by halite disappearance over a


small temperature interval (Fig. 4B). The exceptionally high
homogenization temperatures could indicate some H2O loss
due to postentrapment reequilibration (Sterner et al., 1995;
Audtat and Gnther, 1999), but no textural evidence for such
a process was observed in this sample. Even these extreme
values are probably real fluid temperatures, comparable to
those reported by Eastoe (1978) for the Panguna porphyry
deposit, Bougainville.
Melting temperatures of the daughter crystals other than
halite were observed at 100 to 150C for the first unknown
phase, which is rounded, weakly birefringent, and has a relatively low relief. After initial melting it reprecipitates on
cooling to 90C as an irregular aggregate with strong birefringence. The second daughter phase is orthorhombic,
weakly birefringent with a moderate relief, and it melts between 200 to 250C. The third phase is small, isometric,
round, and birefringent (erythrosiderite?; Wilson et al., 1980)
and it melts at 320 to 370C. A small cubic daughter crystal
0361-0128/98/000/000-00 $6.00

(sylvite?) melts at 460 to 490C always before halite (e.g., assemblage 67). Such high melting temperatures are rarely reported for sylvite but were previously found in the Endeavour
26 North porphyry Cu-Au deposit at Parkes, Australia (Heithersay and Walshe, 1995), implying very high K/Na ratios up
to 1.1 by weight (Roedder, 1971; Cloke and Kesler, 1979).
In polyphase brine and opaque-bearing brine inclusions,
the opaque phases never dissolve even up to 800C. The common presence of hematite daughter crystals in hematite-free
rocks indicates H2 diffusion out of the inclusions (Mavrogenes and Bodnar, 1994). Rare, birefringent rectangular
daughter crystals (anhydrite?) also did not dissolve to at least
600C, which also may be related to postentrapment oxidation of sulfur.
Opaque-bearing brine inclusion: These assemblages are
similar in salinity and homogenization temperatures to
polyphase brine inclusions (Figs. 4A and 5B, G). Apart from
common assemblages homogenizing to liquid, there are also
assemblages in which inclusions homogenize by halite disso-

1752

0361-0128/98/000/000-00 $6.00

Polyphase
brine inclusions

Boiling
assemblage

Boiling
assemblage

Polyphase
brine inclusions

Boiling
assemblage

Polyphase
brine inclusions

115

117

62

116

47

Boiling
assemblage

114

61

Polyphase
brine inclusions

68

Polyphase
brine inclusions

Brine
inclusions

104

67

Opaque-bearing
brine inclusions

93

Polyphase
brine inclusions

Opaque-bearing
brine inclusions

91

66

Opaque-bearing
brine inclusions

Type of
inclusion

92

Assemblage

1753

Early-P3 Porphyry, barren qtz


vein with vug, weak potassic
alteration, secondary trail,
coexisting vapor?

Strong fsp-destr altered QuartzEye Porphyry with qtz-py vein,


secondary trail
Quartz-Eye Porphyry, fsp-destr.,
secondary trail, low-density
vapor inclusions

Quartz-Eye Porphyry, strong


fsp-destr., negative crystal
shaped inclusions,
pseudosecondary trail
Fsp-destr. altered Quartz-Eye
Porphyry, qtz-py vein,
randomly distributed inclusions
Fsp-destr. altered Quartz-Eye
Porphyry, qtz-py vein, regular
inclusions, secondary trail
Quartz-Eye Porphyry, clustered,
round inclusions, close to an
assemblage of opaque-bearing
inclusions
Strong fsp-destr. altered QuartzEye Porphyry with qtz-py vein,
randomly distributed inclusions
Strong fsp-destr. Quartz-EyePorphyry, qtz-py vein, negative
crystal shaped, secondary trail,
one vapor 4.5wt%, decrepitation
at 712C, three inclusions
homogenized at 721C
Strong fsp-destr. altered QuartzEye Porphyry with qtz-py vein,
primary inclusions ?
Strong fsp-destr. altered QuartzEye Porphyry with qtz-py vein,
primary inclusions ?
Quartz-Eye Porphyry, strong
fsp-destr., partly open vein with
single crystals, secondary trail
Strong fsp-destr., Quartz-EyePorphyry, qtz-py vein, secondary
trail, low-density vapor, one
high-density vapor inclusion
with 10.5wt% NaCl eq.
Quartz-Eye Porphyry, fspdestr., secondary trail, lowdensity vapor inclusions

Petrographic description

511.5
21.8

445.2
24.7

580.3
3.0

541.3
14.6

529.2
13.5

553.9
19.5

529.3
11.1

515.5
8.3

489.7
8.4

490.0
30.0

281.2
36.4

450.1
11.7

292.3
36.4

390.5
28.1

Tm(halite)
or Tm(ice)

665.0

470.3
14.1

488.5
3.1

>600

>600
(720)

>600

>600

>600

>600
(720)

>600

671.0
6.0

>600

>600

771.5
16.3

Th(L)
(C)

61.4
2.9

52.7
3.0

71.1
0.1

65.5
2.1

63.8
1.9

66.1
3.6

63.8
1.6

61.9
1.2

58.3
1.1

58.5
4.0

36.9
2.8

53.3
1.5

37.9
2.8

46.5
3.0

0.02
0.01

40
50

240 60
20 20

n.a.

10
10

20
202

30
20

<30 <20

n.a.

<1

4
20 n.a.
2 30

90 110 n.a.
40 30

120 10
5
20 10 1

3
1

<1

2
1

1
1

1
1

510
280
20
10
130
20

<2

<300

<25

<30

40
10
<5
5
2
<30 40
2

<10

<5

0.3
0.14

0.78
0.26

70
25
2800
120
200
50
3600
470

0.56
0.1

<1.97

60 4500
10 1300

60 4400
10

20 2000
4 110

20 1,800
2 510

20 1700
8 650

20
50 3300
5 10 100

<30

<20

60

8
3

<10

30 1600
5 270
280 <1 <10 <2 140
470
10
n.a.
7
60 n.a. 2600
3 90
480

<2

500 <10 10
<2
2
600
10
3
n.a. 50 100 20
3
10 20 20 1
<40
3
10
<1
1
1 10
170 120 20 <300 70 170
6
8
20 40 5
7 30 2 2

20 <10
2
50
1
120 n.a. 30
10
30
10 <10
2

50

b <30 12.30 7.00 1.05 10.93 0.22 0.46 <30 360 90 <110 10 <380 20
1.20 2.20 0.14 0.33 0.12 0.11
60 30
1
20
v <30 1.10 0.50 0.07 0.59 0.12 0.04 <30 30
10 <10 <2 <200 2
0.50 0.30 0.04 0.31 0.04 0.02
10 10
1
150 8.10 15.40 4.20 16.90 0.32 0.95 50 550 120 40 n.a. n.a. 40
130 1.70 2.20 0.53 2.60 0.150.17
110 20 8
10

10
20
550
50
30
10
740
50

40 12.40 11.60 1.32 15.51 0.48 0.74 <30 690 130 70 <10 340
20 1.00 0.60 0.11 1.48 0.39 0.08
80 20 20
30

30 0.30 0.20 0.03 0.62 0.11 0.02 <3


40 0.001 0.04 0.01 0.71 0.17 0.01
b <70 10.00 13.00 2.30 22.12 0.05 0.89 n.a.
0.90 0.10 0.36 2.81 0.07 0.08
v <20 0.70 0.80 0.15 1.41 0.09 0.07 n.a.
0.05 0.10 0.04 0.74 0.04 0.02
90 10.70 13.80 3.90 22.10 0.12 1.20 n.a.
0.64 0.75 0.09 1.20 0.17 0.13

0.23
0.19

0.37
0.44

10

n.a.

20
50
4 20

20 490 <1
4 170

4 16.20 12.40 1.50 8.50 0.18 1.40 n.a. 750 80


70
40 n.a. 60
90 n.a.
20 2.00 2.70 0.29 1.90 0.14 0.26
170 30 60 70
20 20

25.20
0.50

24.30
0.60

40

50 1.10 0.40 0.06 0.31 0.05 <0.02 n.a.


40 0.10 0.20 0.02 0.09 0.02

n.a.

70 10.50 12.60 1.77 9.77 1.00 1.83 n.a. 960 120 220 <10
20 1.20 0.50 0.45 1.39 0.870.84
320 30 210

0.17
0.19

21.90
2.10

<50

n.a.

380 60 140 40 580 20


90 150
2
140 50 180 80 260 20 100 280 3

130 10.50 4.70 0.80 6.50 0.01 0.60 50 240 110 30


10
70 1.00 0.60 0.20 1.50 0.01 0.10 20 50 50 20 3

9.20
2.60

80 5.00 7.10 1.30 13.20 0.03 0.60 n.a.


4 0.70 0.50 0.20 2.40 0.03 0.10

120 10.6 5.30 1.10 8.60 0.01 0.60 50


130 1.40 1.90 0.40 1.80 0.01 0.20

Salinity wt %
Li
Na
K Mn Fe Cu Zn As Rb
Sr Mo Ag Sn Cs Ba La Ce
W
Tl
Pb Au
NaCl equiv (ppm) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm)

TABLE 6. Element Concentrations of Selected Elements in Fluid Inclusions (analyzed by LA-ICP-MS)

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

1753

Boiling
assemblage

Opaque-bearing
brine inclusions

Boiling
assemblage

Brine
inclusions

112

73

106

101

0361-0128/98/000/000-00 $6.00

1754

Aqueous
inclusions

Aqueous
inclusions

Aqueous
inclusions

Aqueous
inclusions
High-density
vapor inclusion

High-density
vapor inclusion

25

21

20

32

31

Brine
inclusions

100

54

58

59

Polyphase
brine inclusions
Polyphase
brine inclusions
Polyphase
brine inclusions
Polyphase
brine inclusions

Type of
inclusion

57

Assemblage

P2 Porphyry, clear quartz in vug


with late calcite, secondary trail
P2 Porphyry, clear quartz in vug
with late calcite, secondary trail
P2 Porphyry, clear quartz in vug
with late calcite, secondary trail
Early-P3 Porphyry, potassic
alteration, clear quartz crystals in
vug, secondary trails
Early P3 Porphyry, potassic,
coarse-grained barren qtz vein,
small secondary trail, mostly
low-density vapor
Early P3 Porphyry, moderate
potassic, randomly distributed
round inclusions
Early P3 or P2 Porphyry, py-cpy
cutting qtz vein, secondary trail,
vapor inclusions commonly
low-density
Early P3 Porphyry, qtz/mag,
potassic, slightly fsp-destr.,
randomly distributed inclusions,
close to opaque-bearing inclusions
Early P3 Porphyry, qtz vein,
potassic, not clearly defined
plane, negative crystal-shaped
to rounded inclusions
P2 Porphyry, clear qtz crystal
in vug with late calcite,
secondary trail
Phenocryst from Quartz-Eye
Porphryry, fsp-destr., randomly
distributed round to negative
crystal shaped inclusions
Early P3 Porphyry, qtz/mag,
potassic, slightly fsp-destr.,
cluster of rounded inclusions
Early P3 Porphyry, fsp-destr.
altered, secondary trail
Early P3 or P2 Porphyry?
With py-cpy vein cutting qtz
vein, potassic, secondary trail,
some CO2, magnetite solid
inclusions
Early P3 or P2 Porphyry?
With py-cpy vein cutting qtz
vein, potassic, secondary trail,
some CO2, magnetite solid
inclusions

Petrographic description

-2.0
0.4

-0.8
0.2
-2.2
0.2

-1.8
0.2

-4.6
2.9

-13.8
0.2

351.0
36.1

345.0
40.5

364.3
29.3

366.3
3.6

406.5
54.4

376.5
31.4
439.0
5.2
421.2
16.4
429.3
37.5

Tm(halite)
or Tm(ice)

430.0
8.3

270.8
20.7
399.0
23.0

240.0

242.0
2.6

292.2
10.0

219.5
27.6

386.8
43.2

397.3
6.5

423.1
22.5

447.5
3.5

540.8
69.1
499.0
12.6
471.1
18.4
462.8
19.0

Th(L)
(C)

3.4
0.7

1.4
0.3
3.6
0.4

3.0
0.3

7.1
4.2

17.6
0.1

41.4
3.9

42.1
3.7

43.9
2.8

44.0
0.3

48.3
6.0

45.1
3.3
51.9
0.6
49.9
1.8
50.9
4.3
500
300
520
110

n.a.

10
60 <10
10 70

n.a.

<3

5
5

<2

<2

3
4

490 <1
170

40
30 n.a.
30 40

<5

<10

160

210 0.98 0.94 0.31 1.30 0.25 0.10 n.a.


0.03 0.08 0.09 0.22 0.11 0.02

30
5

10
1

n.a.

<10 <100

3
1

<0.43

200
70

<1

90
60

3 1600
1 2340

10 1,500
6 590

20
20 1800
1 5 340
<1
10 <50 <20 100
5
25
410
4
10
40
20 1200
100 6 10
6 50

90
20
<10

110 630 n.a.


60 200

10
4

n.a. 1500
100

<10 <10 1700


1200
<10 <5
30
10
5
110 10 2000
5 20 6 860

30
1
<1

1400 80
20 250 <1
610 20 2 340

<1 <100

<5

n.a.

4
2

10 2,100 3
4 3,300

<5

250
70

280 0.47 0.070.05 <0.34 <0.01 0.03 n.a. 20


30 <120 <20 <900 5
<30
3
<2 <30 <10 110
180 0.13
0.01
10
370 1.40 0.84 0.14 0.89 0.60 0.08 <40 56
24 n.a. <10 840 10
30 100 20 100 10 170
1.10 0.69 0.171.00 0.73 0.08
40 20
10
150 20 50
180

<10 0.95 0.37 0.05 0.16 0.01 0.02 30


0.12 0.22 0.05
0.02

80 2.00 0.72 0.22 0.33 0.03 0.13 n.a. 60


80
<3
30 1.50 0.54 0.35 0.22 0.01 0.15
40 130

260 3.30 0.90 0.20 <0.10 <0.0040.10 90 130 180 <30


100 0.09 0.24 0.15
0.03 10 30 30

<5

470 70
50
20 <250 40
50 20
20
10
30 <10 <100 <70 <850 5
1
330 250 <30
6 <400 70
60 30
10

320 10.00 5.30 2.60 1.90 0.05 1.20 n.a. 500 740 <20
220 0.90 1.10 0.80 1.10 0.04 0.60
110 570

b <830 10.00 8.00 2.00 10.82 0.52 0.57 <30


0.90 1.30 0.51 4.06 0.31 0.13
v <740 0.60 0.50 0.06 0.60 0.30 0.03 <300
0.04 0.10 0.01 0.33 0.23 0.002
900 9.70 4.10 2.10 4.20 0.07 0.50 150
2.20 0.20 0.20 0.50 0.01 0.10 5

5
4

100 n.a. 210 <80 n.a. 2400


100
150
320
130 430 20
20
20 1800
40 870 10 10 5 520

20
60 n.a.
5 10

100 90 n.a. n.a. 40


60 50
20
150 140
8 <100 40
10 60 3
7

390 50
80 n.a.
60 9 30

50 11.00 8.10 1.90 9.16 0.26 0.54 <20 490 100 130
5 <290 40
50
10
70 1.20 0.90 0.14 1.72 0.10 0.07
40
20 2
v n.a. 0.30 0.70 0.01 <0.07 0.03 0.01 <30 10
10 <30 <3 <30 <1 <15 <1
0.20 0.50
0.02
10 10
220 10.40 6.80 2.00 6.30 0.20 0.90 n.a. 400 110 60
20 n.a. 90 320 n.a.
170 0.70 0.70 0.30 1.10 0.10 0.30
100 30 15
10 40

110 6.50 10.20 3.20 10.50 0.32 0.60 10


20 0.56 0.69 0.45 0.96 0.16 0.10 3
19.80
0.10
140
1.10
0.09
70
170 8.00 9.90 2.30 13.00 0.55 0.99 50
100 5.00 6.00 0.32 1.10 0.10 0.16 50
80 13.50 6.80 1.50 7.70 0.44 0.47 6
110 1.60 1.80 0.26 2.20 0.23 0.09

Salinity wt %
Li
Na
K Mn Fe Cu Zn As Rb
Sr Mo Ag Sn Cs Ba La Ce
W
Tl
Pb Au
NaCl equiv (ppm) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm)

TABLE 6. (Cont.)

1754
ULRICH ET AL.

0361-0128/98/000/000-00 $6.00

1755

High-density
vapor inclusion

High-density
vapor inclusion

High-density
vapor inclusion

High-density
vapor inclusion

High-density
vapor inclusion

High-density
vapor inclusion

High-density
vapor inclusion

40

41

42

43

44

45

46

Early P3 Porphyry, moderate


potassic, qtz-py vein, randomly
distributed rounded inclusions
Quartz-Eye Porphyry, fsp-destr.,
negative crystal shaped to
rounded, some with opaque
daughter crystals, secondary trail
Fsp-destr. Quartz-Eye Porphyry,
qtz-py vein, secondary trail,
some with opaque crystal
Quartz-Eye Porphyry, feldspardestructive, regular inclusions,
some with opaque daughter
crystal, secondary trail
Fsp-destr. Quartz-Eye Porphyry,
qtz-py vein, secondary trail,
some with opaque crystal
Strong fsp-destr. Quartz-Eye
Porphyry, qtz-py vein, clustered,
few high-density with coexisting?
polyphase inclusions
Strong fsp-destr. Quartz-Eye
Porphyry, qtz-py vein, clustered,
few high-density with coexisting?
polyphase inclusions
Strong fsp-destr. Quartz-Eye
Porphyry, qtz-py vein, clustered,
few high-density with coexisting?
polyphase inclusions

Petrographic description

-0.5

-0.5

-0.7

-1.0
0.1

-3.2
0.3

-3.5
1.2

-2.1
0.4

-6.3
0.8

Tm(halite)
or Tm(ice)

>600

>600

>600

>600

422.5
17.8

432.6
39.1

386.5
27.7

395.0
21.2

Th(L)
(C)

1.3

1.7
0.2

5.3
0.4

5.7
1.8

3.5
0.7

9.5
1.0

0.39

0.39

0.39
0.20

0.79

0.17
0.27

0.05
0.04

0.18
0.25

0.52
0.39

<10

<10

30
20

n.a.

<150 1.70 0.72 0.14 1.30 3.30 0.12 n.a.


0.14 0.24 0.06 0.15 1.20 0.03

280 2.10 0.90 0.16 1.7 2.60 0.13 n.a.


0.48 0.30 0.06
2.10 0.07

180 1.30 0.33 0.01 0.34 2.90 0.10 <60


130 0.11 0.37 0.01 0.14 1.10 0.10

720 3.00 1.60 0.15 2.9 3.00 0.68 n.a.


490 0.17 0.44 0.24
3.00 0.45

30

30
3

9
5

70

10
7

<10

6
5

<20

n.a.

n.a.

n.a.

n.a.

700

n.a.

1
1

<10

<20

570

<2

260
2
400 1

n.a.

<30 <10

<10

20

20
30 3900 <5
20 20 1500

<20 <600 <3

<2

<70

40

22

<20

620
420

<5

230
70

180
75

8
40
6 40

<50 <10

<0.19

<0.11

<0.21

<0.46

Salinity wt %
Li
Na
K Mn Fe Cu Zn As Rb
Sr Mo Ag Sn Cs Ba La Ce
W
Tl
Pb Au
NaCl equiv (ppm) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm) (ppm)

LA-ICP-MS results of fluid inclusion assemblages shown in Figure 8; each line represents averages and 1s standard deviations from an assemblage with several inclusions of identical phase proportions. With the listed elements
some additional elements were analyzed such as Ca, Mg, Al, Ti, B, Bi, Cl, and Sb, but these numbers were commonly below detection limit or masked by mass interferences, b and v correspond to the coexisting brine and vapor
phase of boiling assemblages.

High-density
vapor inclusion

Type of
inclusion

35

Assemblage

TABLE 6. (Cont.)

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

1755

1756
20

ULRICH ET AL.
40

Polyphase brine inclusions

15

30

10

20

10
0
30
B 25
20
15
10
5
0
12
C 10
8
6
4
2
0
60
D 50
40
30
20
10
0
25

0
25
20

Opaque-bearing
brine inclusions

15
10
5
0
25

Brine inclusions

20
15
10
5
0
30
25
20
15
10
5
0
20
15

Aqueous inclusions

High-density vapor
inclusions

20

15

10

10

5
0
0 10 20

30 40 50 60 70

Salinity (wt% NaCl equiv.)

5
0

100 200 300 400 500 600 700 800


Th L and Tm halite>Th L (C)

FIG. 5. Frequency diagrams of salinity and homogenization temperature.


The modes at high homogenization temperatures of the polyphase brine and
opaque-bearing brine inclusions are real and mainly from high-temperature
fluids trapped in the vein from the Quartz-Eye Porphyry (sample BLA97).

lution or by near-simultaneous vapor and halite disappearance (Fig. 4B). Groups of inclusions (e.g., assemblage 91-93)
homogenizing above 600C were found exclusively in sample
BLA97. The assemblage with the lowest Th(L) (363 42C)
has a salinity of 38 1 wt percent NaCl equiv. The assemblage with the highest salinity (63 4 wt % NaCl equiv) homogenizes by halite dissolution at 524 25C.
Brine inclusion: Most assemblages homogenize to liquid,
and their salinity is near the lower end of the salinity range of
opaque-bearing brine inclusions (Figs. 4A and 5C). Homogenization temperature is distinctly lower compared to the
polyphase and opaque-bearing brine inclusions, mostly
around 340 to 400C. Assemblage 100 shows halite melting
at temperatures much higher than vapor to liquid homogenization temperatures, which may indicate trapping at high
pressure (Bodnar, 1994; Cline and Bodnar, 1994) or some
otherwise unrecognizable postentrapment modification. The
assemblage (104) with the lowest salinity (37 3 wt % NaCl
equiv) has the highest homogenization temperature of 671
6C, whereas the assemblage (101) with the highest salinity
(42 4 wt % NaCl equiv) homogenizes at 387 43C.
Aqueous inclusions: Within a few groups of simple twophase inclusions, hydrohalite was observed as the last melting
phase in individual inclusions (assemblages 5, 5a, 22, Fig. 4A).
0361-0128/98/000/000-00 $6.00

This phase showed metastable behavior as described by


Roedder (1984), melting above 0C (in one case up to 9.3C).
The brown color of the ice in some of the aqueous inclusions
is typical for a CaCl2-bearing fluid (Shepherd et al., 1985, p
109). In assemblages 4, 5, 11, and 18 of feldspar destructively
altered Quartz-Eye Porphyry as well as potassically altered P2
Porphyry and Early P3 Porphyry, unusually large salinity variations (e.g., up to 8 wt % NaCl equiv) can occur within one
assemblage of aqueous inclusions with constant filling ratios
and homogenization temperatures, suggesting mixing between fluids of different salinity.
Limited microthermometry was carried out on aqueous inclusions in calcite and anhydrite (Fig 4A, assemblages 1-3).
The homogenization temperature of assemblages in anhydrite
is 231 22C. The salinity of the aqueous inclusions is very
variable and ranges from 7 5 wt percent NaCl equiv in one
assemblage to 15 9 wt percent NaCl equiv in another assemblage. Aqueous inclusions in calcite are very rare and only
a few were suitable for heating and freezing experiments.
They have very low salinity with a melting point depression of
0.2 to 0.3C, which corresponds to a salinity of 0.4 to 0.5
wt percent NaCl equiv. Inclusions in calcite have the lowest
homogenization temperatures measured at Alumbrera (Th(L)
= 204 17C). These data are in agreement with the occurrence of calcite in the latest stage of the vein paragenesis. The
larger variations in homogenization temperature and salinity
of aqueous inclusion assemblages in anhydrite than assemblages in quartz is possibly due to the lower stability of anhydrite and/or stretching of the inclusion during heating and
freezing runs.
High-density vapor inclusions: Due to the small amount of
liquid and the difficulty of seeing any phase transitions in
high-density vapor inclusions, the determination of the homogenization temperature may vary several tens or even hundreds (150300C) of degrees (Bodnar et al., 1985). Therefore, homogenization temperatures shown in Figures 4A and
5J have a larger range and may be severely underestimated.
Similarly, ice melting as a measure of apparent salinity is difficult, and results vary considerably, probably because of
minor coentrapment of coexisting brine. The lowest salinity
within one assemblage was measured at 1.6 0.7 wt percent
NaCl equiv and the highest apparent salinity was 19 2 wt
percent NaCl equiv. For most boiling assemblages, the salinity of low- and high-density vapor inclusions had to be inferred from the coexisting polyphase brine inclusions, using
the NaCl-H2O phase diagram (Sourirajan and Kennedy,
1962; Urusova, 1975; Bodnar et al., 1985). High-density vapor
assemblages, containing inclusions with and without opaque
daughter crystals, show the same salinity and homogenization
temperature as assemblages where only opaque-bearing
vapor inclusions occur.
In a few assemblages (31-34), clathrates were observed
during freezing runs, which indicate the presence of some
CO2. Melting of CO2 occurred at Tm(CO2) = 56.1C, and
clathrate melting could only be measured in one assemblage
at +2.4C. The total CO2 content is probably well below 3 mol
percent, the lower limit for visible liquid CO2 (Nash, 1976).
Low-density vapor inclusions: These inclusions provided
no information during microthermometric experiments. No
phase changes were observed during freezing.

1756

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

Pressure Estimation and P-T-XNaCl Evolution

The homogenization temperature and salinity of the liquid


member of the boiling assemblages are plotted as filled symbols in the phase diagram of the system NaCl-H2O (Fig. 6;
dots for sample BLA97, squares for all other samples). The
assemblages (113115, 117) with the highest homogenization
temperatures (~720C) indicate pressures between 1,050 and
680 bars. Lower pressures (300400 bars) are represented by
two boiling assemblages (107 and 116), and several boiling
assemblages (106, 108112) indicate very low pressures of
100 to 250 bars at temperatures of 300 to 450C.
Minimum pressures for liquid assemblages homogenizing
by vapor disappearance to the liquid were estimated in the
single-phase liquid field of the system NaCl-H2O and are
shown in Figure 6, using open symbols with a bar extending
to higher pressure. The highest minimum pressures of >1.08
to >1.32 kbars are indicated by a high-temperature brine and
three opaque-bearing brine assemblages of sample BLA97

Given appropriate experimental data, a unique estimate of


temperature (T) and pressure (P) conditions can be determined from a boiling assemblage (e.g., Roedder and Bodnar,
1980). All other inclusion data provide information about
minimum pressures and temperatures only. The occurrence
of boiling assemblages as well as brine assemblages homogenizing by halite dissolution requires significant variations in P
and T along and outside the two-phase region of the multicomponent fluid system. Approximate constraints on P-TXNaCl conditions can be derived by reference to experiments
in the NaCl-H2O model system (Sourirajan and Kennedy,
1962; Urusova, 1975; Bodnar et al., 1985; Fig. 6). Table 4 and
Figure 6 give the pressure estimates of all boiling assemblages and the inclusion populations that were later analyzed
with LA-ICP-MS.

1400
sample BLA 97
other samples

1300

117: boiling
assemblage

1200

boiling
assemblages

92

91: fluid inclusion


assemblage

800C

91

minimum P estimate

104

93

1100
700C

"vapor"

"liquid"

pressure (bars)
pressure (bars)

1000

113
68

900

114

600C

66

800
700

44
45/46

550C

43

115

67

47

117

61

500C

600
500

critical curve

57

450C

400

32
33

300

400C

111

35

100
20

0
1

21

25

73

106

V+NaCl

Cl

101
100

Na

108

62

V+

31

107
116
112
109

L+

40

59
58
54

41/42

200

110

1 0

50

100

wt%
wt% NaCl
NaClequiv
equiv
FIG. 6. Phase diagram of the system NaCl-H2O (Sourirajan and Kennedy, 1962; Urusova, 1975; Bodnar et al., 1985),
showing all boiling assemblages and single-phase assemblages analyzed by LA-ICP-MS. Circles describe the P-T-X variation
of assemblages from sample BLA97, and squares represent other samples from the main mineralized porphyries. Filled labels indicate boiling assemblages (numbers underlined) and open symbols show lower limits for pressure estimates based on
brine-only assemblages. The salinity and homogenization temperature data of the vapor-rich inclusion assemblages have
larger errors due to the underestimation of homogenization temperatures and possible overestimation of the salinity due to
coentrapment of minor liquid. For example, assemblage 35 must have been trapped at pressures greater than 480 bars in the
vapor field.
0361-0128/98/000/000-00 $6.00

1757

1757

1758

ULRICH ET AL.

(9193, 104). Slightly lower minimum pressures are estimated for high-temperature polyphase brine inclusions that
fall close to the 700C isotherm on which the high-temperature boiling assemblages plot. In contrast, low-temperature
assemblages ( 500C) of the different inclusion types give
minimum pressures over the range of 400 to below 100 bars.
Several assemblages consistently show homogenization by
halite melting, and in some assemblages, individual inclusions
homogenize by liquid or by halite melting. Homogenization
by halite dissolution has been interpreted to indicate entrapment of a halite-saturated hydrothermal brine (Eastoe, 1978;
Cloke and Kesler, 1979; Wilson et al., 1980), homogeneous
trapping at high pressures (Bodnar, 1994; Cline and Bodnar,
1994), or postentrapment H2O loss or volume shrinkage
(Sterner et al., 1988). Halite-homogenizing inclusions at
Alumbrera never occur on boiling trails, which would be clear
evidence for postentrapment modification (Audtat and Gnther, 1999). The analyzed fluid inclusion assemblages showed
no petrographic evidence of postentrapment disturbance
(with the possible exception of assemblage 100). We therefore suggest that the halite-homogenizing inclusion assemblages record the entrapment of a single-phase brine, probably not halite saturated, at pressures well above the
liquid-vapor phase boundary (Bodnar, 1994; Cline and Bodnar, 1994). Minimum pressures can be derived for this type of
assemblage if the salinity is around 40 wt percent NaCl equiv,
where experimental isochore data are available (Bodnar,
1994). Two assemblages that homogenize by halite melting
(62, 100) give unreasonably high pressures using the calculation approach of Bodnar (1994). Assemblage 100, although
within the salinity range of the experimental data, gives pressures of around 2.8 kbars due to the large difference between
vapor homogenization and halite dissolution, but the microthermometric data vary over a large range, probably indicating postentrapment disturbance in this case. In the following plots, this assemblage is grouped next to assemblage (101)
of otherwise similar inclusions in the same sample. Assemblage 62 indicates a minimum pressure of 1,135 bars, but the
extrapolation of the experimental data to such high salinities
(70 wt % NaCl equiv) and the known deviation from the
NaCl-H2O model system may lead to errors in the pressure
estimate. When the final homogenization temperature and
the salinity are plotted, this assemblage falls inside the halitesaturated field at ~380 bars (Fig. 6).
Table 5 summarizes the main observations of the depositscale distribution and the microthermometric characteristics
of the fluid inclusion assemblages. Even though it is not possible to document the entire fluid history in time at any point
in the deposit, a generalized fluid evolution can be deduced
on the assumption that each intrusion and its fluid pulse (including the Quartz-Eye Porphyry) followed a similar range of
P-T-XNaCl paths. Significant salinity and density variations of
fluids present during quartz-magnetite and potassic alteration
and probably the initial stages of copper deposition can be explained by a P-T path of general cooling and decompression
of one magmatic fluid. The highest pressure fluids trapped
above 600C are interpreted as the earliest inclusion assemblages. Among these, there is a trend from nonboiling brines
with ~50 wt percent NaCl equiv toward even higher salinity,
essentially following a high-temperature isotherm of the
0361-0128/98/000/000-00 $6.00

two-phase boundary (Fig. 6). Several of these assemblages


are clear boiling trails, and the salinity variation at high pressure and temperature is therefore interpreted to reflect
minor water loss of an initial input brine to a high-density
vapor phase. These boiling assemblages exclusively contain
opaque-rich polyphase brine inclusions, giving a first indication of high ore metal concentrations. At the lower pressure
end of this evolution, halite saturation may have occurred at
least locally, as indicated by a reversal of decreasing salinity
with decreasing pressures below 400 bars. Throughout
quartz-magnetite alteration, potassic alteration, and the associated copper ore deposition, salinities seem to be controlled
dominantly by variations in P and T, in a fluid system dominated by magmatic liquid + vapor halite. Below 400C,
many of the liquid-rich inclusions contain only one or no
opaque daughter crystals, and microanalytical data presented
in the following chapter show that this later part of near-isobaric cooling is associated with the loss of metallic components by ore metal precipitation. Large variations in salinity
only occur in late-stage, low-temperature aqueous assemblages, at temperatures between 200 and 300C. This part of
the fluid evolution is associated with feldspar-destructive alteration and may involve an influx of a low-salinity fluid of external origin mixing with diminishing amounts of magmatic
vapor.
LA-ICP-MS Analysis of Fluid Compositions
Methods
After petrographic characterization of the fluid inclusion
types and the determination of salinity and some of the homogenization temperatures, the composition of selected inclusions was analyzed by LA-ICP-MS. Representative sampling of the bulk composition of the now heterogeneous
inclusions is achieved by stepwise increase in the diameter of
the ablation crater, using an Excimer laser ablation system
with imaging optics (Gnther et al., 1997, 1998). The entire
content of the inclusion, daughter crystals as well as liquid, is
transported as an aerosol into the plasma, using He as a carrier gas. Absolute quantification of single inclusion LA-ICPMS signals was carried out by integration of all elements including Na, correction for host mineral contributions (most
commonly Al, Ti, Li, Ca), comparison of intensity ratios with
an external NIST glass standard, and referencing the ratios to
the absolute concentration of Na estimated from microthermometry (Gnther et al., 1997, 1998; Audtat et al., 1998;
Gnther and Heinrich, 1999).
Some elements are commonly below detection or were not
measured for every inclusion assemblage (Mg, Ca, Al, Cl, As,
Ag, Mo, Sn, Sb, La, Tl, Bi, U, Au). Where available, upper
concentration limits for these elements are indicated in Table
6 and Figures 7 and 8. Analysis of gold concentrations in single fluid inclusions was a major achievement of this new
method and was conducted under optimized conditions of
the ICP-MS, for maximum response on 197Au and a reduced
element menu (Si, Na, Au, Cu, As; Ulrich et al., 1999). The
transient signals consistently show overlapping short peaks for
Cu and Au, demonstrating without any doubt that the gold is
contained within the inclusions and that it is adsorbed onto
(or intergrown with) a copper sulfide daughter crystal (Ulrich

1758

1759

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

100

F=27

Th L= 720C
P= 680 bars

10

Fe

Sn
Ti

Zn

100

Pb

10

Li

Sr

Tl
Cs
Bi

Ce

Ag

Ce

Assemblage 117

La

0.1
0.1
100

La

10

100

0.1

0.1

10

Rb

115

100

F=7

Th L= 470C
P= 320 bars

10

100

10

Ba
Ag
Sb Cs
Mo
Bi
Tl

Mn
Zn

Pb

Sr

Rb

Fe

Na

Cu

Mn

Ba

F=8

Na

Cu

0.1

Th L= 720C
P= 750 bars

0.1

100

10

Th L= 720C
P= 900 bars

F=1

Na

Na
K

Cu

Ti

0.1

Fe
Cu

Mn
La

100

Pb
Rb

Rb

Ba

Cs
Ce
TI

0.1
0.1

10

Zn

Sr
Cs

Bi

Ag Bi

Pb

Ba

Li

Sr

Mn

Zn

10

Fe

Tl

116

100

0.1

100

10

114

0.1

10

100

0.1

100

10

100

Th L= 397C
P= 180 bars

10

F=10

Th L= 448C
P= 270 bars

F=2
Fe

K Na
Cu

0.1

Na
K
Fe
Mn

100

Ag

Mo

Ce

Cs

Mo

Sr Ba

Rb

Bi

Ag

1
106

0.1
0.1

10

Cu

Zn
Pb

Tl

10

Ti

100

0.1

10

100

0.1

Sr
Tl Ba
Li
Ce
La
Cs

10

100

Mn

Zn
Pb
Rb

112

0.1

10

concentration in liquid (ppm/wt%)


FIG. 7. Element concentrations in liquid (polyphase brine inclusions) vs. concentrations in coexisting vapor inclusions.
The black line represents equal concentration in the two phases. The dashed lines connect elements with similar tendencies
to be fractionated in the liquid. F is the minimum degree of Cu fractionation based on the average value of elements clearly
enriched in the liquid (eq. 1 in text). Error bars indicate 1 of the variability within one assemblage.
0361-0128/98/000/000-00 $6.00

1759

100

1760

ULRICH ET AL.

80

70
60

salinity

1600
1400
1200

50

1000

40

800
Th

30
20

600
400

P
BLA 97: opaque-bearing
and brine inclusions

10

BLA 97: polyphase


brine inclusions

200

Other samples

100
10
1
0.1

NaCl eq.
Na
Mn
Zn
Pb
Rb

100
10

Cs
Tl

1
0.1
100
10

C
Fe
K

1
Ba
Cu

0.1
100
10 Sr
Mo

Au

0.1

Fluid inclusion assemblages


FIG. 8. Interpreted P-T-salinity evolution (A) and LA-ICP-MS data (B, C) of the chemical composition of fluid inclusions.
Evolution sequence (from left to right) is interpreted on the basis of petrographic timing evidence, decreasing Th, and decreasing pressures for assemblages with overlapping homogenization temperatures. The high-temperature assemblages of
sample BLA97 are further separated on the left side, based on petrographic timing evidence in a sample that probably represents a separate magma and fluid pulse. Inclusion types are schematically indicated at the top. A. P-T-XNaCl data separated
according to relative timing relationships (BLA97) and decreasing homogenization temperature (all other samples). B. and
C. The concentration variation for nonreactive and reactive elements, respectively. Stippled lines interpolate elements omitted from analytical runs with a reduced element menu optimized for Au. Error bars extending from the bottom of the figure indicate upper limits for analyses where element concentrations are below detection. Error bars show 1 variability
within an assemblage of similar inclusions.
0361-0128/98/000/000-00 $6.00

1760

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

et al., 1999). Limits of detection as low as 0.1 ppm Au in 25m inclusions were determined, corresponding to a total
mass of about 1015g Au per inclusion (Ulrich et al., 1999).
Detection limits (calculated from the standard deviation of
background intensity; Gnther et al., 1998, 1999) vary depending on inclusion volume and signal length, and many Au
analyses are close to the limit of detection. Typical uncertainties of 20 percent are estimated from analyses of several microthermometrically identical inclusions in one assemblage,
but they increase for small inclusions (<10 m) and element
concentrations near the detection limit. For the low-density
vapor inclusions in boiling assemblages, only poorly determined minimum concentrations can be given, because the
LA-ICP-MS signals had to be referenced to the theoretical
NaCl concentration of the pure vapor phase estimated from
the NaClH2O phase diagram, ignoring the likelihood of coentrapment of some saline liquid.
Element distribution between brine and vapor
The fluid composition was analyzed on the same assemblages that were used to define the P-T-XNaCl conditions in
the system. These assemblages cover a wide range of Th,
salinity, and density. In boiling assemblages both inclusion
types (liquid- and vapor-rich) were analyzed individually to
explore the fractionation of the different elements between
coexisting fluid phases. Elements including K, Na, Fe, Mn,
Pb, Zn, Rb, Sr, Ba, Cs, Bi, and Ce, and Tl are concentrated
in the liquid phase where they lie within errors on a parallel
array to the line of equal concentrations in both inclusion
types (Fig. 7). Cu, on the other hand, is always shifted toward
the line of equal concentrations or lies above (assemblage
117), i.e., fractionates into the vapor phase relative to the
other metals or even in terms of absolute concentration. The
number indicated in each plot of Figure 7 reports the minimum degree of Cu fractionation between vapor and liquid for
each boiling assemblage using the following ratio (Heinrich et
al., 1999):
CuV/MV
F =
,
CuL/ML

(1)

where Cu and M represent the concentrations of Cu and average concentrations of liquid-partitioning metals, M (Na, K,
Fe, Mn, Zn, Rb, Cs, Pb) in the superscripted fluid phases.
These values are minimum estimates for the actual exchange
constant for Cu between the two end-member fluids. Minor
coentrapment of liquid in the vapor inclusions is most likely
and probably contributes to the near-linear distribution of all
other elements in Figure 7. However, by normalizing the distribution factor with M, the value of F is independent of any
systematic errors due to the internal standardization of the
analyses using Na. Preferential fractionation of Cu into the
vapor phase was suggested and shown for other deposits
(Heinrich et al., 1992, 1999) and is supported by the high Cu
concentrations generally found in high- and low-density vapor
inclusions at Alumbrera, although they do not always demonstrably coexist with a liquid phase. The highest Cu concentration is 3.3 1.2 wt percent Cu, measured in a high-density
vapor assemblage of sample BLA97, which is one order of
magnitude higher than the average Cu concentrations of the
0361-0128/98/000/000-00 $6.00

1761

polyphase brine inclusions (Table 6). Gold was below the detection limit of around 0.5 ppm even in the best signals from
low- and high-density vapor inclusions.
Chemical composition of liquid-rich inclusions
The compositions of the brine and aqueous inclusions, averaged for all analyzed inclusions in each assemblage, are
shown in Figure 8 and Table 6. The assemblages from sample
BLA97 are separated on the left side of the diagram, and all
assemblages are sorted according to a general sequence of decreasing temperature, pressure, and interpreted petrographic
timing. Based on the mineralization and alteration geochemistry of the deposit (Ulrich and Heinrich, 2001) the analyzed
elements are grouped into nonreactive elements (Fig. 8B)
and reactive elements (Fig. 8C). The nonreactive elements
include Cl (not plotted), Na, Mn, Pb, Zn, Rb, Tl, and Cs.
Their concentration trends in the halite-saturated assemblages are parallel to each other and to the total salinity of the
fluid across all samples. These highly soluble elements are effectively conserved and maintain their initial magmatic ratios
in the fluid, as they are transported through the deposit without significant mass transfer by fluid-rock reactions or mineral precipitation. The concentrations of all these elements
remain high in all brine types irrespective of significant variations in pressure and temperature, indicating a common
parentage of the fluid salinity from a magmatic source. Near
the low-temperature end of the trend, the concentration of all
nonreactive elements decreases rapidly, in parallel with the
total fluid salinity, probably indicating late dilution of the
magmatic brine with a low-salinity aqueous fluid.
The reactive elements include K, Fe, Cu, Mo, Au, Sr, and
Ba, whose concentrations vary relative to those of the nonreactive elements and the total salinity as a result of fluid-rock
reaction and mineral precipitation. With the exception of the
very earliest and hottest assemblages in sample BLA97, all
polyphase and opaque-bearing brine inclusions have approximately constant and high concentrations of Cu (avg 0.33 wt
%, range 0.051.0 wt %) as well as Fe and K. However, as the
temperature reaches 400C or less, there is a dramatic drop
in the concentration of Cu (and to a lesser extent Fe and K),
while the salinity and the concentration of all nonreactive elements remain constant and high. This drop in Cu is petrographically reflected by the disappearance of chalcopyrite
daughter crystals (brine inclusion assemblages 100, 101).
Concurrent with the Cu and Fe decrease in the fluid, K is reduced, whereas Ba and Sr are increasing in the fluid. The
possibly earliest high-temperature brine assemblages in sample BLA97 do not match this general trend and were singled
out on the left side of Figure 8 (assemblages 104, 93, 91, 92).
They show much lower Cu concentrations despite very high
temperature and pressure of entrapment and otherwise similar composition to other opaque-bearing brine inclusions.
Molybdenum concentrations in the high-temperature input
fluid are about 100 ppm and drop below the relatively poor
detection limit at the stage where the Cu concentration decreases sharply.
Limited data for the Au concentrations in high-temperature polyphase brine inclusions (assemblages 6668) indicate
an average Au content of the magmatic brine of 0.55 ppm
(range 0.21.4 ppm; Table 6, Fig. 8). Figure 9 shows all Au

1761

1762

ULRICH ET AL.

secondary minerals including biotite, albite, sericite + minor


clay minerals, quartz, chlorite, and magnetite, were analyzed
to deduce the isotopic composition of the fluids associated
with various stages of alteration. Hydrogen analyses of fluid
inclusion extracts of the quartz-magnetite, potassic, and
feldspar-destructive alteration were analyzed for comparison
to the isotopic fluid signatures calculated from fluid-mineral
equilibrium.

10

Au/Cu=1.2* 10-4

inferred
input fluid
ore samples

bulk ore

0.1
Cu loss
polyphase
brine inclusions

0.01
0.01

0.1

10

Au (ppm)

FIG. 9. Au vs. Cu concentration for single liquid inclusion analyses (open


circles) representing some of the earliest and highest temperature inclusion
assemblages in the orebody (assemblages 6668). Small dots are whole-rock
assay from section 47 through the ore deposit and the white square represents the bulk ore-grade ratio of the deposit. Encircled liquid-rich inclusion
analyses match the bulk ore, whereas analyses with lower Cu concentrations
may have lost some Cu to a vapor phase or early sulfide mineral inclusions
(C. A. Heinrich, pers. commun., 2001, modified from Ulrich et al., 1999).

analyses of individual inclusions (open circles) in comparison


to the bulk ore ratio of the deposit (square) and ore sample
assays from section 47 (small dots, see also Ulrich and Heinrich, 2001, fig. 10). The encircled fluid analyses have an
Au/Cu ratio (0.8 ppm Au/0.7 wt % Cu = 1.2 104 by wt)
identical to the bulk ore deposit (0.65 ppm Au/0.51 wt % Cu
= 1.3 104; Mndez, 1997; Wall, 1997, updated 0.64 ppm
Au/0.54 wt % Cu = 1.2 104; David Keough, writ. commun.,
2001) and are interpreted as the closest approximation to the
input ore fluid (Ulrich et al., 1999). Many of the polyphase
brine inclusions contain similar Au contents but significantly
lower Cu concentrations. Fractionation of some Cu into the
vapor phase (possibly on a local scale) is the most likely explanation, consistent with the high Cu concentrations in many
analyzed vapor inclusions and with their poorly determined
Au/Cu ratio of less than 5 104 by weight (calculated from
average limits of detection). On the other hand, some Cu
fractionation may have occurred into magmatic sulfide phases
in some of these samples (C. A. Heinrich, pers. commun.,
2001).
Oxygen and Hydrogen Isotopes
To further constrain the fluid sources that were involved in
the formation of the Bajo de la Alumbrera deposit, a selection
of samples was analyzed for their oxygen and hydrogen isotope compositions. The D and 18O values of igneous biotite
and quartz of an unaltered Northwest Porphyry intrusion, and
0361-0128/98/000/000-00 $6.00

Procedure and experimental fractionation factors


The oxygen and hydrogen isotope data are expressed relative to standard mean ocean water (SMOW) in per mil (Table
7). The absolute errors of the measured data are about 0.2
per mil for the oxygen isotopes and about 5 per mil for the
hydrogen. The mineral-fluid equilibrium temperatures for
the calculation of the isotopic fluid composition were estimated from the fluid inclusion homogenization temperatures
(Table 7). The fractionation factors used for the oxygen isotopes of quartz and albite were according to Bottinga and
Javoy (1973) and Matsuhisa et al. (1979) for the respective
temperatures, and for biotite the fractionation factor was estimated by subtracting quartz-water from quartz-biotite (Javoy,
1977). The hydrogen fractionation factors for biotite and
sericite (muscovite) are from Suzuoki and Epstein (1976),
and the chlorite fractionation factor is from Graham et al.
(1987). However, the temperature estimates of around 300
and 350C (from fluid inclusion data) for some of the alteration stages are below the temperature range of experimentally based fractionation data (e.g., 400800C, Suzuoki and
Epstein, 1976). At low temperature (below 400C) the hydrogen fractionation curves are not well established except
for kaolinite (e.g., Gilg and Sheppard, 1996). For that reason,
the fractionation factors (1000 ln mineral-water) at low temperatures for sericite-kaolinite, chlorite, and biotite were extrapolated from Taylor (1997, see diagram on p. 234), and are as
follows: 23 per mil (muscovite), 15 per mil (kaolinite), 35
per mil (chlorite), 35 per mil (biotite). The analyzed samples
of the feldspar-destructive alteration consist of a mixture of
fine-grained sericite and clay minerals, therefore, an average
fractionation factor between the value for kaolinite and
sericite of 19 per mil was used. The fractionation factor for
biotite depends on Fe content (Suzuoki and Epstein, 1976).
An Mg-rich composition (XFe = 0.4) was assumed for the secondary biotite of the potassic alteration, and a more Fe-rich
(XFe = 0.7) composition was used for the igneous biotite in the
unaltered sample of the Northwest Porphyry, based on the
general compositional trends shown by Ague and Brimhall
(1988).
Results and isotopic composition of hydrothermal fluids
The measured isotopic ratios of the minerals and the calculated isotopic signature of the fluid are given in Table 7. The
variation of the measured D values among the different
alteration minerals of the porphyry samples is small, excepting the D values for the secondary biotite of the andesitic
wall rock and particularly the igneous biotite of the Northwest Porphyry that are depleted (Fig. 10A). This low D
value of the igneous biotite from the Northwest Porphyry is
typical for a melt reservoir that has already degassed much of
its fluid (Taylor, 1986), which is in agreement with the late

1762

1763

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA


TABLE 7. Oxygen and Hydrogen Isotope Composition of Minerals, Whole-Rock Samples, and Fluid Inclusions and the
Calculated D and 18O Values of the Fluid in Exchange Equilibrium with These Samples
D
( SMOW)

Measured

18O
( SMOW)

T(C)

Calculated fluid1
D
18O
( SMOW)
( SMOW)

Igneous minerals
Northwest Porphyry
LP

Bio

-109.8 (0.8)

Qtz

9.2 (0.1)

700

-65.9 (2.4)

8.5 (0.4)

-65.4 (9)

9.9 (0.9)
8.3 (0.3)

Hydrothermal minerals and fluid inclusions


Porphyry quartz-magnetite zone
49-60/93m
Fluid incl.

-65.4 (9)

49-61.4/357m

Qtz
Mag

10.6 (0.1)
3.1

700

Qtz
Mag

10.5 (0.04)
2.8

700

9.9 (0.9)
8.0 (0.3)

Porphyry potassic zone


51-52.2/308m

WR (bio)

-68.3 (2.7)

Qtz

8.5 (0.2)

500

-29.8 (4.7)

5.3 (0.9)

49.2-46.3/452m

WR (bio)

-80.1 (3.8)

Qtz

9.7 (0.05)

500

-41.7 (4.7)

6.5 (0.9)

Qtz

11.3 (0.2)

500

-55 (22.8)

Qtz

9.2 (0.2)

500

-55

6.1 (0.9)

-71.1

Qtz

10.3 (0.1)

300

-52.1 (10)

3.4 (1.8)

BLA 97
43-47.1/195m

Fluid incl.

Porphyry feldspar-destructive zone


49.2-46.3/265m
WR (ser)

8.1 (1.2)

57.5-60/51m

WR (ser)

-73.3 (0.2)

Qtz

10.7 (0.2)

300

-54.3 (10)

3.9 (1.8)

55.5-60/192m

Fluid incl.

-52

Qtz

8.3 (0.08)

300

-52

1.4 (1.8)

38-49/97m

Fluid incl.

-79 (19.5)

Qtz

10.2 (0.01)

300

-79

3.3 (1.8)

-76.6 (0.8)

Ab

10.6 (0.1)

300

-41.6 (5)

4.7 (1.7)

Porphyry chlorite-epidote zone


BLA55
Chl
BLA64
Andesite potassic
49-61.4/94m

Chl

-78.3

Ab

9.9 (0.1)

300

-43.3 (5)

4.0 (1.7)

Bio

-79.4 (6.7)

Bio

7.4 (0.04)

350

-44.4 (5)

9.4 (0.1)

37-49/275m

Bio

-90.1 (1.6)

Bio

6.5 (0.04)

350

-55.7 (5)

8.5 (0.1)

37-49/275m

WR (bio)

-83.1 (1.8)

WR (bio)

6.5 (0.04)

350

-48.1 (5)

8.5 (0.1)

Andesite feldspar-destructive zone


46-42.5/46m
WR (ser)

-69.3 (15.3)

Qtz

10.7 (0.2)

300

-50.3

3.9 (1.8)

-30

-5

-70 (3)

0.1 (3.2)

Present-day meteoric water2


Estimate of meteoric
water of Miocenc age
from the Faralln
Negro mine

Fluid incl.

-70 (3)

Qtz

11.70

200

Abbreviations: bio = biotite, chl = chlorite, mag = magnetite, qtz = quartz, ser = sericite, WR = whole rock
Measured 18O and D of alteration minerals and the calculated isotopic signature of the fluids in equilibrium with the rock; the numbers in brackets of
the measured isotopic signature correspond to 1 and the numbers in brackets for the calculated signature indicates the variation of the isotopic signature for
temperatures of 50C above and below the cited temperature
1Fractionation factors used for the calculation of the isotopic fluid signature in the respective temperature interval are from Bottinga and Javoy (1973),
Matsuhisa et al. (1979), and Javoy (1977) for the oxygen isotopes and from Suzuoki and Epstein (1976) and Graham et al. (1987) for the hydrogen isotopes;
since at low temperatures the experimental data are rare, the fractionation factors were taken from the extrapolated curves from the diagram of Taylor (1997)
2Average of present-day measurements of meteoric water at Salta, Argentina; see text for explanation

0361-0128/98/000/000-00 $6.00

1763

1764

ULRICH ET AL.

Measured Dand 18O values


-20
-40
potassic (porphyry)

feldspardestructive

quartz-magnetite

potassic (wall rock)

-60

quartz/sericite

quartz/biotite

-80

albite/
chlorite

biotite/biotite

chlorite-epidote

-100
Northwest Porphyry
-120

quartz/biotite

A
10

18O

12

Calculated isotopic composition of the fluid


-10

-20

potassic
(porphyry)
500C

"present-day
meteoric water"

-30

potassic
(wall rock)
350C

chlorite-epidote
300C

-40

feldspardestructive
300C

-50

-60

-70

Faralln Negro mine


200C

quartz-magnetite
700C

Northwest
Porphyry
700C

primary
magmatic water

-80

B
-90
-10

-5

10

15

Fig 10. Isotopic signatures of the fluids associated with the magmatic and the alteration stages. A. Analyses of mineral
pairs that were analyzed for hydrogen and oxygen, respectively. B. Calculated isotopic signature of fluids at the indicated
temperatures for the different alteration zones and the Northwest Porphyry. Closed symbols are calculated values, open symbols indicate fluid inclusion extracts for hydrogen. The white triangle is from a nearby epithermal quartz vein and contained
fluid inclusions. The inclined bars represent the variation of fluid composition as a result of a temperature uncertainty of
50C. See text for discussion.

intrusive stage of the Northwest Porphyry. The oxygen values


for quartz and albite range from 8 to 11 per mil for the porphyries and are lower for the secondary biotite of the andesites (6.57.5%). This may be due to the different closing
temperatures of biotite and quartz and the higher resistance
of quartz to later isotopic exchange during the feldspar-destructive overprint of the system. The D value for the
feldspar-destructive overprint of the andesite has a larger
variation, including poor reproducibility of the duplicate
analysis. Despite the different diffusion rates and closing
0361-0128/98/000/000-00 $6.00

temperatures for O and H of a given mineral (Giletti, 1985,


1986), we used temperature data of the fluid inclusion study,
because the homogenization temperatures of the fluid inclusions are close to or below the estimated closing temperatures
of the analyzed minerals (Zaluski et al., 1994), except for the
700C of the magmatic stage. Various processes influence the
isotopic composition of minerals and the fluid in a magmatichydrothermal system, such as fluid exsolution (Dobson et al.,
1989; Taylor, 1992), phase separation (e.g., Shmulovich et al.,
1999), open- and closed-system fractionation, fluid mixing,

1764

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

and temperature and salinity changes (Horita et al., 1995).


These will be discussed based on first estimates of the isotopic compositions of fluids associated with the stages of
alteration.
Local meteoric water: The isotopic composition of the
Miocene meteoric water was estimated from analyses on a
low-sulfidation epithermal vein sample from the Faralln
Negro gold mine about 15 km west of Alumbrera. This
quartz-pyrite-rhodochrosite vein was formed from aqueous
fluid represented by inclusions with a homogenization temperature of 191 to 220C and a low salinity of 0.4 wt percent
NaCl equiv and probably represents our closest approximation to the hydrogen isotope composition of the local meteoric water at the time of magmatic-hydrothermal activity. The
18O value of the fluid calculated from the quartz analyses at
200 50C is around 0 3 per mil. This cannot represent
pure meteoric water with a D value of 70 3 per mil, determined from fluid inclusion extracts. Present-day meteoric
water at the closest observation station at Salta (500 km north
of Alumbrera) is heavier with D = 30 per mil and 18O = 5
per mil (Fig. 10; International Atomic Energy Agency, 1999).
Magmatic fluids (igneous biotite and matrix quartz): Simple fractionation between water and minerals at magmatic
temperatures (700C) leads to an isotopic composition for the
magmatic fluid in equilibrium with the Northwest Porphyry
of D = 66 per mil and 18O = 8.5 per mil. Exsolution of
magmatic fluid is accompanied by an enrichment of the fluid
in deuterium of up to 40 per mil, as a function of the amount
of water in the melt, the hydrogen speciation (H2O, OH), and
the mode of degassing (Taylor, 1992). In the following calculations an enrichment of 20 per mil in D and closed-system
fractionation is assumed. Salinity decreases the fractionation
factors by about D = 10 per mil (Horita et al., 1995). Thus,
a net enrichment of D = 10 per mil of the calculated fluid in
equilibrium with magmatic biotite results in an isotopic composition of the late magmatic fluid of D = 56 per mil and
18O = 7 to 8 per mil (Fig. 10B).
Quartz-magnetite assemblage: The 18O analyses of coexisting quartz and magnetite of the quartz-magnetite alteration
were used to determine equilibration temperatures of 612 or
750C, depending on the fractionation factors used (Downs
et al., 1981; Matthews et al., 1983), broadly consistent with
the highest fluid inclusion homogenization temperatures of
this zone (Figs. 5, 6). Calculated 18O of the fluid (9.9) and
measured D of quartz-hosted fluid inclusions (65 9)
lie in the magmatic water field, confirming the magmatic
source of these high-temperature fluids.
Potassic alteration (quartz and secondary biotite): The calculated D and 18O of the fluid from the potassically altered
Early P3 Porphyry, considering pure H2O and 500C are at
D = 42 and 30 per mil and 18O = 5.3 and 6.5 per mil, respectively, plotting above the magmatic water box (Fig. 10B).
This fluid has undergone phase separation, and the vapor
phase will be enriched in deuterium by up to 28 per mil and
slightly depleted in 18O relative to the coexisting brine
(Shmulovich et al., 1999). Phase separation of an initial magmatic fluid with D ~ 56 per mil and 18O ~ 7 per mil (as
calculated above) can result in a vapor with D ~ 42 per mil
and 18O ~ 6 per mil and a brine with D ~ 70 per mil and
18O ~ 8 per mil. These D values are consistent with the
0361-0128/98/000/000-00 $6.00

1765

large range of the values directly obtained from bulk fluid inclusion extracts (D = 39 to 71), which represent unknown proportions of vapor and brine inclusions in the samples. The calculated fluid in equilibrium with biotite (D =
42 and 30) corresponds more closely to equilibrium with
the vapor phase. With the limited amount of analyses we cannot decide whether this is due to reequilibration of the biotite
with the vapor phase, due inadequately corrected effects of
biotite composition, or due to variations in the initial magmatic fluid composition.
The calculated fluid in equilibrium with the secondary biotite in the andesitic wall rock, assuming a formation temperature of 350C, lies within and above the magmatic water box
but is shifted to the right of the samples of the potassic alteration in the porphyries (Fig. 10B). This difference might be
due to incipient feldspar-destructive overprint of the porphyry samples or due to the uncertainty in the fractionation
factor of biotite at 350C.
Propylitic (chlorite-epidote) alteration: The fluid for the
chlorite-epidote alteration is calculated with an estimated D
fractionation factor of 35 per mil for chlorite at 300C. The
D values of 41 to 44 per mil are in the same range as the
potassic alteration fluid, but the oxygen isotope signature
(based on coexisting albite) is shifted to lighter values (45)
and plots left of the magmatic water box. No salt correction
was applied for the calculation, assuming that a low-salinity
fluid (aqueous inclusions) was related to this alteration stage.
Feldspar-destructive alteration (sericite and/or clay and
late vein quartz): Isotopic fluid compositions were based on a
temperature of 300C and a poorly constrained fractionation
factor of D = 19 per mil, considering the mixture of muscovite and clay minerals in this alteration. The fluid signatures
obtained from quartz and sheet silicate extracts are around
D = 53 per mil and 18O = 3.7 per mil. Three hydrogen
analyses of the low-salinity aqueous and vapor-rich inclusions
extracted from two veins with strong feldspar-destructive
overprint have poorly reproducible D values of 52 to 79
per mil, and yield 18O values of ~1.4 to 3.3 per mil for the
fluid coexisting with quartz. The consistent shift toward lower
oxygen values is more prominent than in the fluid associated
with the chlorite-epidote alteration and indicates either a
minor meteoric water component or intense exchange of
magmatic water with rocks at lower temperatures (Fig. 10B).
Discussion
The combination of P-T-XNaCl data (Fig. 6), the fluid compositional variations determined by LA-ICP-MS (Fig. 8), and
the constraints from stable isotopes (Fig. 10) allow a tentative
but consistent reconstruction of the hydrothermal processes
that contributed to the formation of the porphyry Cu-Au deposit at Bajo de la Alumbrera.
Input fluid and magmatic source
The most likely parental ore fluid is recorded by numerous
assemblages of polyphase and opaque-bearing brine inclusions
associated with quartz-magnetite and potassic alteration in all
mineralized porphyries. This fluid was hot (700C), locally
trapped at high pressure, and of dominantly magmatic origin.
It may or may not have boiled during its ascent from an underlying magma chamber, but it intersected the two-phase

1765

1766

ULRICH ET AL.

curve shortly after arriving at the deposit area. Its salinity was
52 to 58 wt percent NaCl equiv or a little less if the highest
pressure boiling assemblage (113) was already modified by
some vapor loss. This fluid contained similar concentrations
of Na (~16 wt %), K (~12 wt %), and Fe (~8 wt %) and high
copper concentrations (up to 1 wt %). The fluid was also gold
rich (up to 1.4 ppm Au), with an Au/Cu ratio (encircled analyses in Fig. 9) that closely matches the bulk Au/Cu ratio of the
entire economic orebody. Despite systematic search, no inclusions of intermediate salinity and near-critical density were
observed at Alumbrerain contrast, for example, to the porphyry stage at Butte, Montana (Roedder, 1971), and suggestions that such supercritical fluids might be the most likely
parental ore fluids in other porphyries (Bodnar and Cline,
1991; Bodnar, 1998).
Thermodynamic modeling (Cline and Bodnar, 1991; Cline,
1995) based on experimental data (e.g., Kilinc and Burnham,
1972; Candela and Piccoli, 1995) indicates that initial fluid
saturation from a Cl-OH-bearing calc-alkaline magma at
pressures above 1 kbar leads to preferential partitioning of Cl
into the first exsolving fluid phase. This permits efficient extraction of Cl-complexed metals (Cu and probably Au) into an
early-exsolving fraction of high-salinity fluid. The high salinity, the high metal content, and the highest minimum pressures of 1.3 kbars recorded by the fluid inclusions at Alumbrera are in excellent agreement with such a source process
operating in a magma chamber located a few kilometers
below the present exposure level. Ulrich et al. (1999) estimated the minimum volume of this inferred magma reservoir
as ~100 km3, based on a comparison of the gold concentration
in normal magmas with the known mass of gold in the deposit. The extreme salinity and very high metal concentration
of this early fluid fraction implies that a normal calc-alkaline
magma with a typical water content of 4 wt percent, as required for amphibole crystallization (Naney, 1983), would
have to lose a significant fraction of lower salinity fluid after
the main mineralization stage. This low-salinity fluid might be
recorded in the extensive feldspar-destructive alteration that
postdates ore formation at Alumbrera and could be evolved
from a collapsing magmatic plume that interacts with wall
rock at low temperatures and minor meteoric water (Hedenquist and Shinohara, 1997). However, the alternative possibility exists that a special magma with an unusually high Cl/OH
ratio, or with an exceptionally high Au/Cu ratio and Au content (>2 ppb, Connors et al., 1993) acted as the fluid source
for the Alumbrera deposit.
Ore fluid evolution
The fluid evolution recorded by sample BLA97 and samples from the mineralized P2 and Early P3 Porphyries (Fig. 6)
indicates an initial trend of decreasing pressure at high temperatures (>650C), with brine salinities increasing from 50
to over 70 wt percent NaCl equiv. This is a result of vapor phase
separation, as indicated by numerous boiling assemblages of simultaneously trapped brine and vapor inclusions in a single
microfracture. Variants in the exact P-T evolution of individual fluid packets are recorded by different homogenization
behavior of inclusion assemblages, including homogenization
by halite dissolution at pressures well above the two-phase
envelope (Figs. 4B, 6). These variants in P-T evolution and
0361-0128/98/000/000-00 $6.00

homogenization behavior reflect the dynamics of the early


high-temperature regime at the brittle-ductile transition, with
repeated hydrofracturing and sealing by mineral precipitation
during each of the intrusive and hydrothermal phases. As a
result, fluid pressure fluctuated between a near-lithostatic
value of ~1 kbar and a near-hydrostatic value of ~300 bars. At
pressures below ~400 bars, the general trend of fluid properties straddles the halite saturation curve, with a reversal toward decreasing salinity with decreasing P and T (Fig. 6).
This probably indicates local halite saturation in the hydrothermal system, although no petrographic evidence of
halite was encountered. Stable isotope data indicate that
these fluids of the early and central alteration zones (quartzmagnetite and potassic) are predominantly or exclusively
magmatic.
Chemical mass transfer to the rocks during this stage was
probably intense and mainly involved deposition of quartz
and magnetite at temperatures near 700C, as confirmed by
oxygen isotope fractionation between these two minerals.
Quartz-magnetite alteration involved major molar volume increase, oxidation of the wall rocks, and removal of alkalis and
alkaline earth elements (Ulrich and Heinrich, 2001). Fe, K,
and Na were present in the initial fluid in high concentrations, and no significant variations among major cation ratios
are detected by LA-ICP-MS analysis (Fig. 8). This implies
that a large flux of magmatic fluid passed through the barren
quartz-magnetite core, imposing fluid-dominated conditions
during alteration. We have no data on the concentrations of
sulfur or its oxidation state in the fluid, but mass-balance data
(Ulrich and Heinrich, 2001) indicate that SO2 acted as the
main oxidant for magnetite formation from aqueous FeCl2 according to a reaction such as 9FeCl2 + SO2 + 10H2O >
3Fe3O4 + H2S + 18HCl.
The resulting H2S activity increase was insufficient to cause
sulfide saturation at these high temperatures. This is shown
by the Cu concentration in the fluids, which stays high despite major variations in P, T, and XNaCl due to boiling, consistent with the general absence of copper sulfides in the barren
quartz-magnetite core of the hydrothermal system. A significant mass of Cu may have passed through the core and its
quartz veins without ever precipitating, although an unknown
fraction of the Cu-rich brine may have evolved toward chalcopyrite deposition in the upper and peripheral parts of the
stock.
The role of the vapor phase
Microanalytical data from boiling assemblages and pure
vapor assemblages show significant partitioning of Cu to the
high- and low-density vapor phase, relative to other metals
that are preferentially enriched in the brine. In contrast to reduced tin-tungsten systems (Heinrich et al., 1992, 1999), the
absolute concentration of Cu in the vapor phase of boiling assemblages at Alumbrera is similar to that in the high-temperature brine, on average about 0.3 percent in both phases
(Fig. 7). Deviations from this average indicate substantial
redistribution of Cu between the two fluid phases, at all scales
from individual inclusion trails to maybe large ore blocks. It is
also possible that the vapor phase contributed to local redistribution of metals in the ore volume, as suggested by Gammons and William-Jones (1997). However, selective vapor

1766

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

transport was probably not significant as a mineralizing


process at the scale of the entire deposit. Considering an average concentration of 0.3 percent Cu for both phases, the extent of boiling required to increase the salinity of the brine
from 50 to 70 wt percent will cause only a small fraction of the
total Cu to be transferred to the vapor, and some of this Cu
may be returned to the brine by recondensation as the vapor
phase cools (Eastoe, 1982). The same conclusion follows from
the observation that the polyphase brine inclusions in boiling
assemblages do not show systematically depleted Cu concentrations compared to polyphase brine-only assemblages,
across a large range of trapping conditions from >600 to
~400C and 1 kbar to 300 bars (Fig. 8).
The ore precipitation process
The subsequent fluid evolution below 400 bars and 400C
involves several concurrent processes, including further fluid
cooling, decompression to one-third of the initial lithostatic
pressure or even less, potassic alteration, chalcopyrite precipitation, and probably halite saturation. Mixing with external
fluids, however, still did not play any detectable role throughout this stage.
The variation in chemical composition and the petrographic type of the brine inclusions as a function of temperature and inferred trapping history (Fig. 8) show a dramatic
change in the temperature range from 400 down to about
320C. The Cu concentration drops by about an order of
magnitude, K and Fe decrease by a lesser fraction, while Ba
and Sr concentrations increase markedly, as the fluid inclusions change from opaque-bearing types to common brine inclusions. The increase in Ba and Sr, which are representatives
for the major alkaline earth elements, is interpreted as the
fluid compositional response to the destruction of Ca and Mg
silicates (plagioclase, amphibole, etc.) during potassic alteration. Other element concentrations and total salinity do not
decrease substantially and metal ratios among all nonreactive
elements remain essentially constant. Considering that the
early high-temperature brine inclusions contain Au/Cu in the
same ratio as the bulk orebody, and that fluid concentrations
including Cu do not change through the preore stage, the
drop in Cu concentration between 400 and 320C reflects
the efficient (~85%) copper ore deposition by chalcopyrite
precipitation from magmatic brines.
Since pressure fluctuations, vapor separation, and fluid
cooling at higher temperatures occurred throughout the preore fluid record, and yet did not lead to Cu precipitation, we
conclude that cooling of the magmatic brine below a saturation temperature near 400C was the key requirement for
chalcopyrite precipitation. Potassic alteration accompanied
chalcopyrite deposition and was also driven by temperature
decrease (Burnham and Ohmoto, 1980). Mass-balance observations indicate that potassic alteration is not an effective sink
for acidity (Ulrich and Heinrich, 2001), and potassic alteration was therefore not an active chemical driving force for
chalcopyrite precipitation (like phyllic alteration driving, for
example, cassiterite deposition in greisen deposits, Heinrich,
1990). In contrast to vein-type Sn-W deposits studied by Audtat et al., (1998; 2000a) mixing of magmatic brine with lowsalinity external fluids played no significant role in copper
precipitation at Alumbrera, as shown by stable isotopes and
0361-0128/98/000/000-00 $6.00

1767

the constantly high salinity with fixed proportions of the nonreactive elements despite the Cu decrease in the fluids. The
fluid chemical evidence for cooling as the prime cause for
copper sulfide deposition is in perfect agreement with the
thermodynamic analysis of the Sungun porphyry by
Hezarkhani et al. (1999), who used recent experimental data
on Cu complex stability to confirm the earlier conclusion by
Barnes (1979) that fluid cooling to relatively low temperatures is the key to chalcopyrite precipitation in porphyry copper deposits. The ore grades of gold and copper at Bajo de la
Alumbrera are controlled by the magmatic fluid source and
the relatively low temperature coprecipitation of chalcopyrite
and native gold, rather than by the mineral chemistry of auriferous bornite precipitated at high temperature as suggested in general by Simon et al. (2000).
The concentration variations of reactive elements in the
fluids permit a direct estimate of the mass transfer associated
with the ore deposition process. By comparing the chemical
concentration changes in the fluid with the gains and losses
associated with the alteration of the rocks, an estimate of the
integrated fluid/rock ratio associated with potassic alteration
and Cu mineralization can be calculated. For each of the elements, Cu, Fe, K, Ba and Sr, the concentration decrease or
increase from the early ore-bearing polyphase brine inclusions to the later simple brine inclusions (the spent ore fluid)
was divided by the respective gains or losses of the Early P3
Porphyry (Ulrich and Heinrich, 2001, eq. 3). The resulting ratios vary from 1.5 (Cu) to 0.1 (Fe), but gains by the rock always correspond to metal losses by the fluid and vice versa for
all reactive elements. A variation within an order of magnitude is well within the uncertainty of such estimates, which
assume that a single sample set is representative for a bulkrock mass exceeding 107 t. The best estimate is the one based
on Cu, where approximately 3.3 g Cu/kg fluid are precipitated
to obtain an orebody with an average grade of 0.5 wt percent
or 5 g Cu/kg ore. This implies a bulk fluid/rock ratio of 1.5,
which illustrates that a comparatively modest amount of
highly metal charged magmatic fluid is sufficient to cause the
major mass transfer associated with potassic alteration and
porphyry copper mineralization. Nevertheless, a high degree
of fluid focusing from a large magmatic source volume to the
porphyry stock is required, as shown by comparing the
fluid/rock ratio in the deposit volume (1.5 or 150%) with the
proportion of magmatic brine that can be exsolved from the
crystallizing source pluton: at most a few percent, but probably much less, considering the unusually high salinity and
high Au concentration of the brine (Ulrich et al., 1999).
While the precipitation mechanism for Au and Cu is highly
efficient and nearly complete, ore metal deposition is also
highly selective. The average Mo/Cu ratio in the input fluid
(~0.03) is significantly higher than the average Mo/Cu concentration ratio in the orebody (0.005). This is consistent with
petrographic evidence and grade distribution data (MIM,
unpub. data), indicating that a significant fraction of the Mo
is transported through the main ore shell (economically
defined by the Cu and Au grades) and dispersed into the
subeconomic outer fringes of the magmatic-hydrothermal
system. Pb and Zn occur in equally high concentrations in the
magmatic fluid as Cu, but they behave as nonreactive elements and are not precipitated within the orebody volume.

1767

1768

ULRICH ET AL.

They may eventually have been dispersed in the local ground


water without ever precipitating, but there is petrographic evidence that Pb and Zn were at least partly deposited as an
outer halo in the propylitic zone. This behavior is consistent
with the very high solubility of Pb and Zn as chloride complexes at elevated temperatures (Hemley et al., 1992) and
demonstrates that a large tonnage of Pb and Zn, possibly even
exceeding the total copper resource of a large porphyry Cu
deposit, is simply flushed through the system without forming
any conspicuous orebodies unless a specific depositional
process (such as carbonate replacement in manto-style deposits) traps these ore metals. Such highly selective precipitation of metals was also demonstrated in an Sn-W-rich granite
system (Mole Granite, Audtat et al., 2000a).
Postore phyllic alteration
Feldspar-destructive alteration to sericite and variable proportions of clay minerals clearly crosscuts the concentrically
zoned pattern of quartz-magnetite, potassic, and propylitic alteration. It encroached downward and overprinted the ore
shell, commonly along postore faults and fracture zones filled
with pyrite anhydrite (now mostly present as gypsum). It involved sulfidation and silicification in places, but hardly any
quartz veining, and leached rather than introduced any copper sulfides.
Fluids associated with feldspar-destructive alteration occur
mostly as secondary inclusions in earlier quartz veins. Lowdensity vapor and an aqueous liquid of highly variable salinity were trapped more or less together but never coexist on
single inclusion trails showing equilibrium coexistence.
Aqueous fluids reveal a small but distinct salinity gap to the
magmatic brine inclusions (Fig. 4A), but this might be an artifact of deviations of the real fluids from the NaCl-H2O
model system used for the salinity estimate. Metal concentrations vary with salinity but maintain constant proportions,
indistinguishable from the cation ratios in the higher temperature magmatic fluids. The temperature was 200 to
320C and is not correlated with salinity. Pressure was low
but is not well constrained. Limited O-H isotope data indicate a trend from high D magmatic fluid toward meteoric
water. These observations indicate that feldspar-destructive
alteration involved minor near-isothermal mixing of heated
meteoric water with magmatic fluids. An unknown proportion of low-density magmatic vapor probably condensed into
this mixture, contributing acid volatiles (HCl, H2S) to the
conversion of feldspars to sheet silicates. The proportion of
meteoric water in the fluid mixture could have been substantial, if extensive oxygen isotope reequilibration with
rocks occurred at elevated temperatures, but isotope data
also permit that the fluid was still dominated by magmatic
water. In this case, the salinity variation to very low values
could be caused by a variation at source, magmatic vapor
condensation, or the influx of a late magmatic fluid of lowsalinity and low Cu content (cf., Arribas, 1995; Hedenquist,
1995; Hedenquist et al., 1998).
Practical Implications and Future Research
Our integrated study of the geology and hydrothermal geochemistry of a large porphyry Cu-Au deposit has demonstrated the practicality of LA-ICP-MS for routine analysis of
0361-0128/98/000/000-00 $6.00

fluid inclusion compositions in complex magmatic-hydrothermal systems. Extensive data for ore metal concentrations in the
evolving fluids has provided new insights about the source of
the metals, their transport, and the mechanisms of their precipitation. Our study also illustrates the potential of future applications of this microanalytical technique in mineral exploration.
The geologic study of Bajo de la Alumbrera showed a complex evolution of intrusive and veining events but a comparatively simple zonation of wall-rock alteration and ore-grade
distribution. The fluid inclusion record is dominated by brine
inclusions of relatively constant composition throughout all
stages of quartz veining and high-temperature alteration. Ore
metal concentrations remained high, despite large pressure
and temperature fluctuations, resulting in variable fluid densities of boiling and nonboiling brines. Cu and Au were precipitated upon cooling of the brines below a saturation temperature near 400C, while other metals like Pb and Zn were
flushed through the Cu-Au-mineralized vein stockwork without ever reaching saturation. Inclusions of Pb-Zn-rich but
Cu-depleted (spent) ore fluid could provide a useful indicator
for the proximity of an undiscovered Cu ( Au Mo) deposit
during exploration of poorly exposed grounds.
Fluid cooling as the prime factor controlling Cu-Au deposition implies that the bulk hypogene ore grade of a porphyrytype orebody is mainly controlled by the thermal structure of
the fluid-flow system. A large and rich deposit will form
where a high flux of metal-rich brine is focused from a major
magma reservoir through a restricted volume of highly permeable rock (i.e., the fractured porphyry neck) and is at the
same time efficiently cooled through the critical temperature
interval of ore mineral saturation. Efficient cooling probably
requires stable convection of external fluids around the magmatic fluid plume (ideally without direct mixing and dilution
of the magmatic brine). Vapor phase separation may have an
additional effect on the heat budget, but a site of productive
porphyry copper mineralization is probably characterized by
only minor net loss of vapor and the relatively volatile ore
metals (As, Sb, Cu, Au) to an overlying epithermal environment. Fluid-flow patterns, heat budget, and degree of brinevapor separation will be primarily controlled by the pressure
evolution and, thus, by the permeability structure between
the magma chamber and the surface.
The near-constant composition of preore stage brines provides information about the igneous fluid source, an inferred
subjacent magma chamber of at least 100-km3 size in the case
of Bajo de la Alumbrera and other deposits of similar size.
Orebody composition is determined by the Cu/Au ratio of the
mineralizing ore fluid, and this fluid composition is in turn
controlled by the magmatic fluid source. A future challenge
therefore lies in the development of practical techniques and
fundamental understanding to read the record of ordinary igneous rocks, notably the composition of silicate melts inclusions in unmineralized intrusions and associated volcanic
rocks. This may become a valuable source of information to
identify prospective igneous complexes and to determine the
productive fluid-generating stages within them.
Acknowledgments
We are grateful to MIM Exploration and Minera Alumbrera for generous financial, logistic, and scientific support of

1768

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

this project. In particular, we would like to thank Vic Wall for


initiating this collaboration, and Juan Angera, Steve Brown,
Peter Forrestal, David Keough, and Rick Valenta for their
help and for interesting discussions at various stages of the
project. We also acknowledge comments and reviews by Alan
Clark, Jean Cline, Viktor Koeppel, John Proffett, John Mavrogenes, and Alexandra Skewes. We are grateful for oxygen isotope analyses made by Anita Andrew at CSIRO (Sydney) and
for help from Gretchen Bernasconi with hydrogen isotope
analyses. This work is part of the senior authors Ph.D. and
was supported by ETH Project grant 0-20-041-95 and the
Swiss National Science Foundation (grants 21-45548.95 and
20-52265.97).
July 28, 2000; June 18, 2001
REFERENCES
Ague, J.J., and Brimhall, G.H., 1988, Regional variations in bulk chemistry,
mineralogy and the compositions of mafic and accessory in the batholiths
of California: Geological Society of America Bulletin, v. 100, p. 891911.
Arribas, A., Jr., 1995, Characteristics of high-sulfidation epithermal deposits,
and their relation to magmatic fluid: Mineralogical Association of Canada
Short Course Series 23 p. 419454.
Audtat, A., and Gnther, D., 1999, Mobility and H2O loss from fluid inclusions in natural quartz crystals: Contributions to Mineralogy and Petrology,
v. 137, p. 114.
Audtat, A., Gnther, D., and Heinrich, C.A., 1998, Formation of a magmatic-hydrothermal ore deposit: Insights with LA-ICP-MS analysis of fluid
inclusions: Science, v. 279, p. 20912094.
2000a, Causes for large-scale metal zonation around mineralized plutons: Fluid inclusion LA-ICP-MS evidence from the Mole Granite, Australia: ECONOMIC GEOLOGY, v. 95, p. 15631581.
2000b, Magmatic-hydrothermal evolution in a fractionating granite: A
microchemical study of the Sn-W-F-mineralized Mole Granite (Australia):
Geochimica et Cosmochimica Acta, v. 64, p. 33733393.
Barnes, H.L., 1979, Solubilities of ore minerals, in Barnes, H.L., ed.,
Geochemistry of hydrothermal ore deposits, 2nd ed.: New York, Wiley,
p. 404460.
Batchelder, J.N., 1977, Light stable isotope and fluid inclusion study of the
porphyry copper deposit at Copper Canyon, Nevada: ECONOMIC GEOLOGY,
v. 72, p. 6070.
Beane, R.E., and Bodnar, R.J., 1995, Hydrothermal fluids and hydrothermal
alteration in porphyry copper deposits, in Wahl, P.W., and Bolm, J.G., eds.,
Porphyry copper deposits of the American Cordillera: Tucson, AZ, Arizona
Geological Society, p. 8393.
Bodnar, R.J., 1994, Synthetic fluid inclusions: XII: The system H2O-NaCl.
Experimental determination of the halite liquidus and isochores for a 40 wt
% NaCl solution: Geochimica et Cosmochimica Acta, v. 58, p. 10531063.
1995, Fluid-inclusion evidence for a magmatic source for metals in porphyry copper deposits: Mineralogical Association of Canada Short Course
Series 23, p. 139152.
1998, Fluid evolution in porphyry copper deposits [abs.]: Goldschmidt
Conference, Toulouse, 1998, Abstracts, p. 180181.
Bodnar, R.J., and Cline, J.S., 1991, Fluid inclusion petrology of porphyry copper deposits revisited: Re- interpretation of observed characteristics based
on recent experimental and theoretical data: Plinius, v. 5, p. 2425.
Bodnar, R.J., and Vityk, M.O., 1994, Interpretation of microthermometric
data for H2O-NaCl fluid inclusions, in DeVivo, B., and Frezzotti, M.L.,
eds., Fluid inclusions in minerals: Methods and applications: Blacksburg,
VA, Virginia Polytechnic Institute and State University, p. 117130.
Bodnar, R.J., Burnham, C.W., and Sterner, S.M., 1985, Synthetic fluid inclusions in natural quartz. III. Determination of phase equilibrium properties
in the system H2O-NaCl to 1000C and 1500 bars: Geochimica et Cosmochimica Acta, v. 49, p. 18611873.
Boiron, M.C., Dubessy, J., Andr, N., Briand, A., Lacour, J.L., Mauchien, P.,
and Mermet, J.M., 1991, Analysis of monoatomic ions in individual fluid inclusions by laser-produced plasma emission spectroscopy: Geochimica et
Cosmochimica Acta, v. 55, p. 917923.
Bottinga, Y., and Javoy, M., 1973, Comments on oxygen isotope geothermometry: Earth and Planetary Science Letters, v. 20, p. 250265.

0361-0128/98/000/000-00 $6.00

1769

Bowman, J.R., Parry, W.T., Kropp, W.P., and Kruer, S.A., 1987, Chemical and
isotopic evolution of hydrothermal solutions at Bingham, Utah: ECONOMIC
GEOLOGY, v. 82, p. 395428.
Burnham, C.W., and Ohmoto, H., 1980, Late-stage processes of felsic magmatism: Mining Geology Special Issue 8, p. 113.
Candela, P.A., and Holland, H.D., 1986, A mass transfer model for copper
and molybdenum in magmatic hydrothermal systems: The origin of porphyry-type ore deposits: ECONOMIC GEOLOGY, v. 81, p. 119.
Candela, P.A., and Piccoli, P.M., 1995, Model ore-metal partitioning from
melts into vapor and vapor/brine mixtures: Mineralogical Association of
Canada Short Course Series 23, p. 101127.
Chivas, A.R., and Wilkins, R.W.T., 1977, Fluid inclusion studies in relation to
hydrothermal alteration and mineralization at the Koloula porphyry copper
prospect, Guadalcanal: ECONOMIC GEOLOGY, v. 72, p. 153169.
Cline, J.S., 1995, Genesis of porphyry copper deposits: The behavior of
water, chloride, and copper in crystallizing melts: Arizona Geological Society Digest, v. 20, p. 6982.
Cline, J.S., and Bodnar, R.J., 1991, Can economic porphyry copper mineralization be generated by a typical calc-alkaline melt?: Journal of Geophysical Research, v. 96, p. 81138126.
1994, Direct evolution of brine from a crystallizing silicic melt at the
Questa, New Mexico, molybdenum deposit: ECONOMIC GEOLOGY, v. 89, p.
17801802.
Cloke, P.L., and Kesler, S.E., 1979, The halite trend in hydrothermal solutions: ECONOMIC GEOLOGY, v. 74, p. 18231831.
Connors, K.A., Noble, D.C., Bussey, S.D., and Weiss, S.I., 1993, Initial gold
contents of silicic volcanic rocks: Bearing on the behavior of gold in magmatic systems: Geology, v. 21, p. 937940.
Damman, A.R., Kras, M.S., Touret, L.R.J., Rieffe, C.E., Kramer, A.L.M., Vis,
D.R., and Pintea, I., 1996, PIXE and SEM analyses of fluid inclusions in
quartz crystals from the K-alteration zone of the Rosia Poieni porphyry-Cu
deposit, Apuseni Mountains, Rumania: European Journal of Mineralogy, v.
8, p. 10811096.
Diamond, l.W., Marshall, D.D., Jackman, J.A., and Skippen, G.B., 1990, Elemental analysis of individual fluid inclusions in minerals by secondary ion
mass spectrometry (SIMS): Application to cation ratios of fluid inclusions
in an Archaean mesothermal gold-quartz vein: Geochimica et Cosmochimica Acta, v. 54, p. 545552.
Dilles, J.H., Solomon, G.C., Taylor, H.P., Jr., and Einaudi, M.T., 1992, Oxygen and hydrogen isotope characteristics of hydrothermal alteration at the
Ann-Mason porphyry copper deposit, Yerington, Nevada: ECONOMIC GEOLOGY, v. 87, p. 4463.
Dobson, P.F., Epstein, S., and Stolper, E.M., 1989, Hydrogen isotope fractionation between coexisting vapor and silicate glasses and melts at low
pressure: Geochimica et Cosmochimica Acta, v. 53, p. 27232730.
Downs, W.F., Touysinhthiphoneeaxy, Y., and Deines, P., 1981, A direct determination of the oxygen isotope fractionation of quartz and magnetite at
600 and 800C and 5 kbar: Geochimica et Cosmochimica Acta, v. 45, p.
20652072.
Eastoe, C.J., 1978, A fluid inclusion study of the Panguna porphyry copper
deposit Bougainville, Papua New Guinea: ECONOMIC GEOLOGY, v. 73, p.
721748.
1982, Physics and chemistry of the hydrothermal system at the Panguna
porphyry copper deposit, Bougainville, Papua New Guinea: ECONOMIC
GEOLOGY, v. 77, p. 127153.
Etminan, H., 1977, Le porphyre cuprifere de Sar Cheshmeh (Iran); role des
phases fluides dans les mecanismes d`alteration et de mineralisation: Unpublished Ph.D. thesis, Nancy, Science de la Terre,Universite de Nancy, 249 p.
Gammons, C.H., and Williams-Jones, A.E., 1997, Chemical mobility of gold
in the porphyry-epithermal environment: ECONOMIC GEOLOGY, v. 92, p.
4559.
Giletti, B.J., 1985, The nature of oxygen transport within minerals in the
presence of hydrothermal water and the role of diffusion: Chemical Geology, v. 53, p. 197206.
1986, Diffusion effects on oxygen isotope temperatures of slowly cooled
igneous and metamorphic rocks: Earth and Planetary Science Letters, v.
77, p. 218228.
Gilg, H.A., and Sheppard, S.M.F., 1996, Hydrogen isotope fractionation between kaolinite and water revisited: Geochimica et Cosmochimica Acta, v.
60, p. 529533.
Goldstein, R.H., and Reynolds, T.J., 1994, Systematics of fluid inclusions in
diagenetic minerals: Society for Sedimentary Geology Short Course 31,
199 p.

1769

1770

ULRICH ET AL.

Graham, C.M., Viglino, J.A., and Harmon, R.S., 1987, Experimental study of
hydrogen-isotope exchange between aluminous chlorite and water and of
hydrogen diffusion in chlorite: American Mineralogist, v. 72, p. 566579.
Gnther, D., and Heinrich, C.A., 1999, Enhanced sensitivity in laser ablation
ICP-mass-spectrometry using helium-argon mixtures as aerosol carrier:
Journal of Analytical Atomic Spectroscopy, v. 14, p. 13631368.
Gnther, D., Frischknecht, R., Heinrich, C.A., and Kahlert, H.-J., 1997, Capabilities of an argon fluoride 193 nm Excimer laser for laser ablation inductively coupled plasma mass spectrometry microanalysis of geological
materials: Journal of Analytical Atomic Spectroscopy, v. 12, p. 939944.
Gnther, D., Audtat, A., Frischknecht, R., and Heinrich, C.A., 1998, Quantitative analysis of major, minor and trace elements in fluid inclusions using
laser ablation-inductively coupled plasma-mass spectrometry (LA-ICPMS): Journal of Analytical Atomic Spectroscopy, v. 13, p. 263270.
Gnther, D., Jackson, S.E., and Longerich, H.P., 1999, Laser ablation and
arc/spark solid sample introduction into inductively coupled plasma mass
spectrometers: Spectrochimica Acta, Part B, v. 54, p. 381409.
Hedenquist, J.W., 1995, The ascent of magmatic fluid: Discharge versus mineralization: Mineralogical Association of Canada Short Course Series 23, p.
263289.
Hedenquist, J.W., and Richards, J.P., 1998, The influence of geochemical
techniques on the development of genetic models for porphyry copper deposits: Reviews in Economic Geology, v. 10, p. 235256.
Hedenquist, J.W., and Shinohara, H., 1997, K-silicate- to sericite-stage transition in porphyry Cu deposits; collapse of magmatic plume, or overprint by
meteoric water? [abs.]: Geological Society of America Abstracts with Programs v. 29, p. 359.
Hedenquist, J.W., Arribas, A., Jr., and Reynolds, T.J., 1998, Evolution of an
intrusion-centered hydrothermal system: Far Southeast-Lepanto porphyry
and epithermal Cu-Au deposits, Philippines: ECONOMIC GEOLOGY, v. 93, p.
373404.
Heinrich, C., 1990, The chemistry of hydrothermal tin (-tungsten) ore deposition: ECONOMIC GEOLOGY, v. 85, p. 457481.
Heinrich, C.A., Ryan, C.G., Mernagh, T.P., and Eadington, P.J., 1992,
Segregation of ore metals between magmatic brine and vapor: A fluid
inclusion study using PIXE microanalysis: ECONOMIC GEOLOGY, v. 87, p.
15661583.
Heinrich, C.A., Gnther, D., Audtat, A., Ulrich, T., and Frischknecht, R.,
1999, Metal fractionation between magmatic brine and vapor, determined
by micro-analysis of fluid inclusions: Geology, v. 27, p. 755758.
Heithersay, P.S., and Walshe, J.L., 1995, Endeavour 26 North: A porphyry
copper-gold deposit in the Late Ordovician, Shoshonitic Goonumbla Volcanic Complex, New South Wales, Australia: ECONOMIC GEOLOGY, v. 90, p.
15061532.
Hemley, J.J., Cygan, G.L., Fein, J.B., Robinson, G.R., Jr., and dAngelo,
W.M., 1992, Hydrothermal ore-forming processes in the light of studies in
rock-buffered systems: I. Iron-copper-zinc-lead sulfide solubility relations:
ECONOMIC GEOLOGY, v. 87, p. 122.
Henley, R.W., and McNabb, A., 1978, Magmatic vapor plumes and groundwater interaction in porphyry copper emplacement: ECONOMIC GEOLOGY,
v. 73, p. 120.
Hezarkhani, A., and Williams-Jones, A.E., 1998, Controls of alteration and
mineralization in the Sungun porphyry copper deposit, Iran: Evidence from
fluid inclusions and stable isotopes: ECONOMIC GEOLOGY, v. 93, p. 651670.
Hezarkhani, A., Williams-Jones, A.E., and Gammons, C.H., 1999, Factors
controlling copper solubility and chalcopyrite deposition in the Sungun
porphyry copper deposit, Iran: Mineralium Deposita, v. 34, p. 770783.
Horita, J., Cole, D.R., and Weslowski, D.J., 1995, The activity-composition
relationship of oxygen and hydrogen isotopes in aqueous salt solutions: III.
Vapor-liquid water equilibration of NaCl solutions to 350C: Geochimica et
Cosmochimica Acta, v. 59, p. 11391151.
International Atomic Energy Agency (IAEA), 1999, Global network for isotopes in precipitation database. http://www.iaea.org/programs/ri/gnip/gnipmain.htm
Javoy, M., 1977, Stable isotopes and geothermometry: Journal of the Geological Society of London, v. 133, p. 609636.
John, D.A., 1989, Evolution of hydrothermal fluids in the Park Premier stock,
Central Wasatch Mountains, Utah: ECONOMIC GEOLOGY, v. 84, p. 879902.
Kilinc, I.A., and Burnham, C.W., 1972, Partitioning of chloride between a silicate melt and coexisting aqueous phase from 2 to 8 kilobars: ECONOMIC
GEOLOGY, v. 67, p. 231235.
Lindgren, W., 1905, The copper deposits of the Clifton-Morenci district, Arizona: U.S. Geological Survey Professional Paper, v. 43, 375 p.

0361-0128/98/000/000-00 $6.00

Malvicini, L., and Llambias, E.J., 1963, Mineralogia y origin de los minerales
manganeso y sus asociados en Farallon Negro, Alto de la Blenda y los Viscos, Hualfin, Catamarca: Revista Asociacin Geolgica [Argentina], v. 18, p.
177199.
Matsuhisa, Y., Goldsmith, J.R., and Clayton, R.N., 1979, Oxygen isotopic
fractionation in the system quartz-albite-anorthite-water: Geochimica et
Cosmochimica Acta, v. 43, p. 11311140.
Matthews, A., Goldsmith, J.R., and Clayton, R.N., 1983, Oxygen isotope fractionations involving pyroxenes: The calibration of mineral-pair geothermometers: Geochimica et Cosmochimica Acta, v. 47, p. 631-644.
Mavrogenes, J.A., and Bodnar, R.J., 1994, Hydrogen movement into and out
of fluid inclusions in quartz: Experimental evidence and geological implications: Geochimica et Cosmochimica Acta, v. 58, p. 141148.
Mavrogenes, J.A., Bodnar, R.J., Anderson, A.J., Bajt, S., Sutton, S.R., and
Rivers, M.L., 1995, Assessment of the uncertainties and limitations of
quantitative elemental analysis of individual fluid inclusions using synchrotron X-ray fluorescence (SXRF): Geochimica et Cosmochimica Acta, v. 59,
p. 39873995.
Mndez, V., 1997, Yacimiento Bajo La Alumbrera, Provincia de Catamarca,
Repblica Argentina: Revista de la Asociacin Argentina de Gelogos
Economistas, v. 11, p. 1530.
Naney, M.T., 1983, Phase equilibria of rock-forming ferromagnesian silicates
in granitic systems: American Journal of Science, v. 283, p. 9931033.
Nash, T.J., 1976, Fluid-inclusion petrologydata from porphyry copper deposits and applications to exploration: U.S. Geological Survey Professional
Paper, v. 907- D, p. D1D16.
Oreskes, N., and Einaudi, M.T., 1992, Origin of hydrothermal fluids at
Olympic Dam: Preliminary results from fluid inclusions and stable isotopes: ECONOMIC GEOLOGY, v. 87, p. 6490.
Preece, R.K., III, and Beane, R.E., 1982, Contrasting evolutions of hydrothermal alteration in quartz monzonite and quartz diorite wall rocks at
the Sierrita porphyry copper deposit, Arizona: ECONOMIC GEOLOGY, v. 77,
p. 16211641.
Quan, R.A., Cloke, P.L., and Kesler, S.E., 1987, Chemical analyses of halite
trend inclusions from the Granisle porphyry copper deposit, British Columbia: ECONOMIC GEOLOGY, v. 82, p. 19121930.
Reynolds, T.J., and Beane, R.E., 1985, Evolution of hydrothermal fluid characteristics at the Santa Rita, New Mexico, porphyry copper deposit: ECONOMIC GEOLOGY, v. 80, p. 13281347.
Roedder, E., 1971, Fluid inclusion studies on the porphyry-type ore deposits
at Bingham, Utah, Butte, Montana, and Climax, Colorado: ECONOMIC GEOLOGY, v. 66, p. 98120.
1984, Fluid inclusions: Reviews in Mineralogy, v. 12, 646 p.
Roedder, E., and Bodnar, R.J., 1980, Geologic pressure determinations from
fluid inclusion studies: Annual Review of Earth and Planetary Science, v. 8,
p. 263301.
Rye, R.O., 1993, The evolution of magmatic fluids in the epithermal environment: The stable isotope perspective: ECONOMIC GEOLOGY, v. 88, p. 733752.
Sasso, A.M., and Clark, A.H., 1998, The Faralln Negro group, northwestern
Argentina: Magmatic, hydrothermal and tectonic evolution and implications for Cu-Au metallogeny in the Andean back-arc: Society of Economic
Geologists Newsletter, v. 34, p. 1, 818.
Sheets, R.W., Nesbitt, B.E., and Muehlenbachs, K., 1996, Meteoric water
component in magmatic fluids from porphyry copper mineralization,
Babine Lake area, British Columbia: Geology, v. 24, p. 10911094.
Shepherd, T.J., and Chenery, S.R., 1995, Laser ablation ICP-MS elemental
analysis of individual fluid inclusions: An evaluation study: Geochimica et
Cosmochimica Acta, v. 59, p. 39974007.
Shepherd, T.J., Rankin, A.H., and Alderton, D.H.M., 1985, A practical guide
to fluid inclusion studies: London, Blackie and Son, Ltd., 239 p.
Shinohara, H., and Hedenquist, J.W., 1997, Constraints on magma degassing
beneath the Far Southeast porphyry Cu-Au deposit, Philippines: Journal of
Petrology, v. 38, p. 17411752.
Shmulovich, K.I., Landwehr, D., Simon, K., and Heinrich, W., 1999, Stable
isotope fractionation between liquid and vapour in water-salt systems up to
600C: Chemical Geology, v. 157, p. 343354.
Simon, G., Kesler, S.E., Essense, E.J., and Chryssoulis, S.L., 2000, Gold in
porphyry copper deposits: Experimental determination of the distribution
of gold in the Cu-Fe-S system at 400 to 700C: ECONOMIC GEOLOGY, v. 95,
p. 259270.
Solomon, G.C., 1983, 18O/16O and D/H characteristics of the Ann-Mason
porphyry copper deposit, Yerington, Nevada [abs.]: Geological Society of
America Abstracts with Programs, v. 15, p. 277.

1770

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA


Sourirajan, S., and Kennedy, G.C., 1962, The system H2O-NaCl at elevated
temperatures and pressures: American Journal of Science, v. 260, p.
115141.
Sterner, S.M., Hall, D.L., and Bodnar, R.J., 1988, Synthetic fluid inclusions.
V: Solubility relations in the system NaCl-KCl-H2O under vapor-saturated
conditions: Geochimica et Cosmochimica Acta, v. 52, p. 9891005.
Sterner, S.M., Hall, D.L., and Keppler, H., 1995, Compositional re-equilibration of fluid inclusions in quartz: Contributions to Mineralogy and
Petrology, v. 119, p. 115.
Suzuoki, T., and Epstein, S., 1976, Hydrogen isotope fractionation between
OH-bearing minerals and water: Geochimica et Cosmochimica Acta, v. 40,
p. 12291240.
Taylor, B.E., 1986, Magmatic volatiles: Isotopic variation of C, H, and S: Reviews in Mineralogy, v. 16, p. 185226.
1992, Degassing of H2O from rhyolite magma during eruption and shallow intrusion, and the isotopic composition of magmatic water in hydrothermal systems: Geological Survey of Japan Report, v. 279, p. 190194.
Taylor, H.P. Jr., 1974, The application of oxygen and hydrogen isotope studies to problems of hydrothermal alteration and ore deposition: ECONOMIC
GEOLOGY, v. 69, p. 843883.
1997, Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits, in Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed.: New York, Wiley, p. 229302.
Ulrich, T., and Heinrich, C.A., 2001, Geology and alteration geochemistry of
the porphyry Cu-Au deposit at Bajo de la Alumbrera Argentina: ECONOMIC
GEOLOGY, v. 96, p. 17191742.

0361-0128/98/000/000-00 $6.00

1771

Ulrich, T., Gnther, D., and Heinrich, C.A., 1999, Gold concentrations of
magmatic brines and the metal budget of porphyry copper deposits: Nature, v. 399, p. 676679.
Urusova, M.A., 1975, Phase equilibria and thermodynamic characteristics of
solutions in the systems NaCl-H2O and NaOH-H2O at 350-550C: Geochemistry International, v. 1974, p. 944950.
Vityk, M.O., and Bodnar, R.J., 1998, Statistical microthermometry of synthetic fluid inclusions in quartz during decompression reequilibration:
Contributions to Mineralogy and Petrology, v. 132, p. 149162.
Wall, V.J., 1997, Bajo de la Alumbrera (Argentina). A world class copper gold
deposit: Revista de la Asociacin Argentina de Gelogos Economistas, v.
11, p. 9293.
Weisbrod, W., 1981, Fluid inclusions in shallow intrusives: Mineralogical Association of Canada Short Course 6, p. 241271.
Wilson, J.W. Jr., Kesler, S.E., Cloke, P.L., and Kelly, W.C., 1980, Fluid inclusion geochemistry of the Granisle and Bell porphyry copper deposits,
British Columbia: ECONOMIC GEOLOGY, v. 75, p. 4561.
Zaluski, G., Nesbitt, B., and Muehlenbachs, K., 1994, Hydrothermal alteration and stable isotope systematics of the Babine porphyry Cu deposits,
British Columbia: Implications for fluid evolution of porphyry systems:
ECONOMIC GEOLOGY, v. 89, p. 15181541.

1771

1772

ULRICH ET AL.

APPENDIX 1
Samples Used for the Fluid Inclusion Analyses
The samples are listed according to the different types of
inclusions, except for sample BLA97 where different inclusion assemblages were analyzed from the same sample. The
fluid inclusion assemblages listed in Figures 6 and 8 and Tables 4 and 6 appear in parentheses.
47-53/25 (assemblage 47)

depth: 118 m

Early P3 Porphyry with low Cu and Au grades. The quartz


vein is barren and the wall rock potassically altered (secondary biotite, minor magnetite, and K feldspar). The vein is
about 4 cm long, straight, and about 5 mm wide and has a little vug. The vein quartz is coarse grained and clear. The
polyphase inclusions occur on a secondary plane.
49-51/5 (54)

depth: 192 m

Early P3 Porphyry, potassically altered, with some later


feldspar-destructive overprint (biotite partly chloritized),
high-grade. The vein is barren and about 5 to 7 mm wide,
partly small open space with clear crystals of 2- to 3-mm size.
The polyphase brine inclusions occur along a trail.
51-61.1/5 (57, 58, 59, 25)

depth: 523 m

P2 Porphyry with strong quartz-magnetite alteration, low


grade. The sample is an open-space vein with clear quartz
crystals of up to 1 cm and quartz-magnetite altered wall rock.
No mineralization is observed in the hand specimen. On top
of the quartz crystals some calcite was later crystallized in the
vug of the vein. Polyphase brine inclusion assemblages as well
as aqueous assemblages were analyzed. Polyphase assemblages are randomly distributed and on secondary trails,
whereas the aqueous inclusions lie on planes and are related
to the late stage of calcite infill.
47-53/62 (73)

depth: 311 m

Early P3 Porphyry, moderate potasscially altered, low to


medium grade. Quartz vein with some pyrite in small vugs.
Pyrite seems to be later deposited. The vein is about 6 mm
wide. The inclusion assemblage occurs as randomly distributed, rounded inclusions and is probably of primary origin.
49-52/19 (100, 101, 21)

depth: 334 m

Early P3 Porphyry, quartz-magnetite alteration (K feldspar,


magnetite). Crosscutting quartz veins. The veins are irregular
to straight and between 5 and 9 mm wide. The brine inclusion
assemblages occur as rounded to negative crystal-shaped inclusions in a randomly distributed array. No marked differences in terms of fluid inclusion population were found in the
different veins. The aqueous inclusions are rounded and
occur as a group of randomly distributed inclusions not on a
defined plane.
49-52/14 (20)

depth: 315 m

Early P3 Porphyry, with different quartz veins. Commonly


small 3 to 7 mm wide and crosscutting veinlets. The wall rock
is feldspar destructively altered and mafic minerals and plagioclase are partly replaced by chlorite and sericite. The small
0361-0128/98/000/000-00 $6.00

veinlets may be related to the late phyllic overprint. They


contain small- to medium-grained quartz with dominantly
vapor rich and aqueous inclusions. The analyzed assemblage
of aqueous inclusions occurs along a secondary trail.
46-59 QZE (4)

Surface sample

Quartz-Eye Porphyry, next to drill hole 46-59 in the south


of the deposit. Quartz phenocryst in a strongly feldspar destructively altered matrix. The aqueous inclusions occur on
small planes of few inclusions and small groups of randomly
distributed inclusions. They have rarely negative crystal shape
and are commonly rounded. No clear timing relationships
could be observed due to the large amount of inclusion in
these phenocrysts.
BLA 97 (61, 62, 66, 67, 68, 91, 92, 93, 104, 114-117)
Surface block of Quartz-Eye Porphyry
The vein is up to 8 cm thick and partly open, with freestanding quartz crystals up to 1 cm. Later pyrite and some
gypsum are on top of some of the quartz crystals and interstitial between quartz grains. The wall rock is strongly feldspar
destructively altered. No Cu mineralization is observed in this
sample. The fluid inclusion assemblages occur along secondary trails as well as clusters of possibly primary origin. The
inclusion assemblages referred as clearly boiling assemblages
occur along secondary trails.
45-61.4/13 (106, 31)

depth: 428 m

P2 or Early P3 Porphyry, quartz-magnetite and/or potassic


alteration. Some remnants of secondary biotite and chlorite,
some magnetite. Late pyrite-chalcopyrite vein cuts quartz
vein. The quartz vein is about 8 mm wide and coarse grained.
The boiling assemblage is along a secondary trail, but in the
sample pure high-density vapor trails occur (CO2 bearing)
and polyphase brine trails. The boiling assemblage contains
mainly low-density vapor inclusions.
47-53/25 (107)

depth: 118 m

Early P3 or Late P3 Porphyry, weakly potassically altered


(secondary biotite). Low-grade barren quartz vein with small
vug. The vein is straight and about 4 to 6 mm wide. The boiling assemblage contains mostly low-density vapor inclusions
and occurs along a trail.
49-50.1/1 (108)

depth: 353 m

Barren quartz vein in potassically altered Early P3 Porphyry. Some primary biotite in the wall rock is unaltered,
whereas the hornblende is altered to secondary biotite, minor
K feldspar. The quartz vein is coarse grained and about 7 mm
wide. Only one vapor inclusion in the boiling trail could be
measured, all others were of the low-density type.
49-50.1/4 (109)

depth: 357 m

P2 or Early P3 Porphyry with quartz-magnetite K


feldspar alteration, low-grade. Multiple quartz veins crosscutting and a pyrite-quartz vein is cutting all the quartz veins.

1772

EVOLUTION OF PORPHYRY Cu-Au DEPOSIT, BAJO DE LA ALUMBRERA, ARGENTINA

1773

The veins are between 3 to 6 mm wide. The boiling assemblage consists of polyphase brine coexisting with some highdensity vapor as well as low-density vapor inclusions.

grained quartz vein of about 7-mm width. The boiling assemblage contains mostly low-density vapor inclusions next to the
polyphase brine inclusions.

49-52/3 (110, 111)

51-52.2/13 (113)

depth: 242 m

Early P3 Porphyry, low-grade, potassically altered, some K


feldspar, and secondary biotite. Slightly feldspar destructively
overprinted (some biotite replaced by chlorite, sericite). The
6-mm-wide barren quartz vein is cut by a pyrite veinlet.
49-52/21 (112)

depth: 212 m

P2 or Early P3 Porphyry with coarse-grained quartz-magnetite veins. The magnetite occurs in the center of the veins.
The veins are 4 to 7 mm wide. The boiling assemblage consists
of polyphase brine and mostly low-density vapor inclusions.

depth: 337 m

Early P3 Porphyry, potassic alteration (secondary biotite,


few secondary K feldspar in the matrix). Barren, coarse-

APPENDIX 2
Description of Samples Used for the Stable Isotope Analyses
LP

Surface sample of Northwest Porphyry dike


in the center of the deposit

49.2-46.3/1-2

depth: 264-265 m

Unaltered porphyry sample with fresh plagioclase, hornblende, and biotite crystals up to 3 mm. Biotite was separated
by handpicking.

Intrusion with a strong feldspar-destructive overprint. Vein


quartz was separated for oxygen analyses and whole-rock
samples were used for hydrogen isotope analyses. The major
hydrous mineral is sericite with some clay minerals.

49-60/12

57.5-60/7

depth: 93 m

depth: 51 m

Intense quartz-magnetite alteration of P2 Porphyry. The


magmatic texture is completely replaced by quartz and magnetite. Quartz and magnetite were separated for oxygen
analyses and geothermometry. Fluid inclusions (mainly polyphase brine inclusions) were analyses for hydrogen isotopes.

Quartz-Eye Porphyry (?) with a strong feldspar-destructive


overprint. The porphyric texture is partly obliterated. The
matrix is quartz rich with sericite and clay minerals. Matrix
quartz was separated for the oxygen analyses and whole-rock
samples were used for the hydrogen analyses.

49-61.4/3-5

BLA 55

depth: 357 m

Three samples of intensely quartz-magnetite alteration of


P2 Porphyry. Quartz and magnetite were separated for oxygen analyses and geothermometry.
51-52.2/16

depth: 308 m

Potassically altered Early P3 Porphyry with secondary K


feldspar and biotite. Quartz separates from a veinlet were analyzed for oxygen isotopes and whole-rock samples were used
for the hydrogen analysis (secondary biotite minor chlorite).
49.2-46.3/12

BLA 97

Propylitically altered porphyry with plagioclase phenocrysts


partly altered to calcite epidote. Mafic minerals are altered
to chlorite. Albite was handpicked to avoid altered plagioclase
crystals for the oxygen analyses and chlorite separates were
analyzed for the hydrogen isotopes.
BLA 64

Surface sample of Quartz-Eye Porphyry

Strongly altered (feldspar-destructive) Quartz-Eye Porphyry with barren quartz veins and later pyrite infill. The vein
quartz was analyzed for its oxygen isotope signature and hydrogen isotopes were analyzed in fluid inclusions in the
quartz.
0361-0128/98/000/000-00 $6.00

Surface sample of a porphyritic dike in


the volcanics in the north part of the deposit.

Albite and chlorite were used for the isotope analyses.

depth: 452 m

Early P3 Porphyry, potassically altered with quartz in the


matrix and secondary biotite. The quartz was separated from
the matrix and whole-rock analyses were made for the hydrogen isotope with secondary biotite as the predominant hydrous mineral.

Surface sample of NE Porphyry dike,


north of the Bajo

49-61.4/7

depth: 94 m

Completely replaced andesite by secondary biotite (potassic alteration). Biotite separates were used for the oxygen and
hydrogen analyses because no quartz or feldspar could be
separated for the oxygen isotope analyses.
37-49/12

depth: 275 m

Completely replaced andesite. Secondary biotite was separated for one run of analyses and whole-rock samples were
used for another run, without any significant difference for
the hydrogen and oxygen results.

1773

1774
46-42.5/1

ULRICH ET AL.

depth: 46 m

Feldspar destructively altered andesite with a late quartzpyrite veinlet. The quartz of this vein was used for the oxygen
analyses and the sericite and clay minerals in the associated
alteration halo were used for the hydrogen analyses.
43-47.1/8

depth: 97 m

Faralln Negro mine sample

Unlocated sample of the


epithermal vein system at
Faralln Negro
(NW of Alumbrera)

Banded vein of quartz, rhodochrosite, and pyrite feldspar


in a phyllic to argillic altered wall rock. Quartz separates were
analyzed for the oxygen and aqueous fluid inclusions for the
hydrogen isotopes.

Andesite or Los Amarillos Porphyry with a strong feldspardestructive overprint and a small (34 mm wide), barren
quartz vein. Main types of inclusions are vapor-rich and aqueous types.

0361-0128/98/000/000-00 $6.00

depth: 192 m

Early P3 Porphyry with barren, straight quartz vein with a


narrow feldspar-destructive alteration halo. Main types of inclusions are vapor-rich and aqueous types.

depth: 195 m

Potassically altered Early P3 Porphyry with a quartz-magnetite vein cut by a late pyrite vein. Polyphase brine and vapor
inclusions are the most abundant type of inclusions that were
analyzed for hydrogen isotopes.
38-49/2

55.5-60/4

1774

Vous aimerez peut-être aussi