You are on page 1of 10

Economic Geology

Vol. 98, 2003, pp. 147–156

Formation of Anhydrous and Hydrous Skarn in


Cu-Au Ore Deposits by Magmatic Fluids
L. D. MEINERT,†,*
Department of Geology, Washington State University, Pullman, Washington 99164-2812

J. W. HEDENQUIST,
Department of Geology and Geological Engineering, Colorado School of Mines, Golden, Colorado 80401-1887

H. SATOH,
Geological Survey of Japan, AIST Tsukuba Central 7, 1-1-1 Higashi, Tsukuba 305-8567, Japan

AND Y. MATSUHISA
Geological Survey of Japan, AIST Tsukuba Central 7, 1-1-1 Higashi, Tsukuba 305-8567, Japan

Abstract
Most skarn ore deposits are characterized by two distinctly different alteration styles. An early prograde stage
with anhydrous minerals, such as garnet and pyroxene, forms from relatively high-temperature, hypersaline liq-
uid. A later retrograde stage with hydrous minerals, such as epidote, amphibole, and chlorite plus sulfide ore
minerals, forms from lower temperature, lower salinity fluids. These two alteration stages commonly have been
thought to reflect a dominance of magmatic and meteoric water, respectively, with relevance to the source of
ore metals. We report data from two different skarn systems, one being part of the world’s largest Cu-Au re-
source. Stable isotope compositions of anhydrous and hydrous alteration minerals from both deposits indicate
a magmatic source for both the prograde and retrograde stages: δ18O averages 5.0 per mil for garnet (range,
3.4–7.2‰), 6.5 per mil for pyroxene (4.3–8.2‰), and 7.1 per mil for amphibole (4.3–8.7‰). The δD values of
late amphibole are more complex, with magmatic values (–77 to –78‰) for one deposit and both magmatic
and lighter values for another deposit that could be explained either by magmatic degassing or by limited mix-
ing with meteoric water. We conclude that the differences in fluid composition—prograde versus retrograde
stages—resulted from a magmatic fluid that intersected its solvus during the early stage, creating vapor and hy-
persaline liquid, whereas in the later stage this magmatic fluid did not intersect its solvus because it followed
a different cooling path. This late, low- salinity liquid only boiled once its vapor-pressure curve was reached,
causing sulfide ore to precipitate during the retrograde stage.

Introduction of magmatic and meteoric water, respectively (Einaudi et al.,


THE SOURCE of ore fluids is one of the most important topics 1981). A review of stable isotope systematics in skarn systems
of applied geologic research (Hedenquist and Lowenstern, by Bowman (1998) indicates a dominance of magmatic waters
1994; Pettke and Diamond, 1997; Harris and Golding, 2002). during the early prograde stage, in which anhydrous minerals
Many studies have demonstrated multiple sources of fluids re- dominate, and a range from largely magmatic to largely mete-
sponsible for forming ore deposits. A variety of theories of ore oric waters during the later retrograde stage in which hydrous
genesis invoke interactions among such fluids, particularly minerals dominate.
those of magmatic and meteoric origin (Giggenbach, 1997; We present data from two copper skarn deposits in Irian
and Taylor, 1997, respectively). Since most skarn ores have an Jaya and Canada of different ages, host rocks, and geologic
intimate spatial relationship with magmatic intrusions, their terrane. In both deposits, there was a transition from high-
minerals should provide a clear record of the fluid(s) that were temperature, hypersaline liquid associated with prograde gar-
present in the intrusive environment during ore formation. net-pyroxene skarn to lower temperature, lower salinity liquid
Skarn ore deposits are typically characterized by two distinct associated with retrograde skarn alteration. Based upon data
alteration styles: an early prograde stage with anhydrous min- presented here, we interpret this transition to have been
erals, such as garnet and pyroxene, which forms from rela- caused by a common parent magmatic fluid that experienced
tively high-temperature, hypersaline liquid (Kwak, 1986); and a change in pressure-temperature path during ascent.
a later retrograde stage which consists of hydrous minerals, Big Gossan Deposit, Ertsberg District, Irian Jaya
such as epidote, amphibole, and chlorite, and forms from lower
temperature, lower salinity fluids (Kwak, 1986). These two al- The Ertsberg district contains multiple Cu-Au skarn and
teration stages commonly are thought to reflect a dominance porphyry deposits, including the original Ertsberg discovery,
the GBT/IOZ/DOZ orebodies, Dom, Big Gossan, Kucing
†Corresponding
Liar, and the largest of all, Grasberg, plus several other min-
author: e-mail, meinert@wsu.edu
*Address after August 1, 2003: Department of Geology, Smith College, eral occurrences in various stages of exploration and develop-
Northampton, MA 01063; e-mail, Lmeinert@smith.edu. ment (Fig. 1). Individually, Grasberg is the largest gold mine

0361-0128/01/3320/147-10 $6.00 147


148 MEINERT ET AL.

Ertsberg District
50° Papua
New
70° Guinea
Ti
45°
Irian Jaya Tk
Pacific
Ocean
Ti 25°Cairns
Lembah Tk
Grasberg
Tembaga Australia

* Kkel
Sydney

* Tw

Yello 70°
Kucing w Va
lley S
* Liar
GBT
yncli
ne

S4° Fig. 2A *
Ertsberg 45° Puncak

Big Gossan * * Jaya


4883m

*
Dom
70°

Intrusion (Ti)
Thrust fault
Tertiary
Fig. 2B 40°
75°
carbonate rocks
Cretaceous
clastic rocks
60°
0 km 2 E137°

FIG. 1. Location and geology of the Ertsberg district, Irian Jaya, showing major deposits and location of cross sections of
Figure 2 (modified from Meinert et al., 1997).

and the third largest copper mine in the world. The deposits with an average mineral ratio of about 1:2. The average com-
of the Ertsberg district collectively constitute the world’s position of all analyzed garnets is andradite84.7grossular-
largest Cu-Au resource. At the end of 2001, proven and prob- ite13.5spessartine1.5pyrope0.3. Garnet higher in the deposit and
able reserves were estimated at 2,600 Mt of ore at an average on the western and eastern margins shows a slight increase in
grade of 1.13 percent Cu, 1.05 g/t Au, and 3.72 g/t Ag, con- iron relative to that in deeper, more central locations. The
taining 23.8 Mt of recoverable copper, 64.5 Moz of recover- iron content of pyroxene is zoned in both space and time. The
able gold, and 151.6 Moz of recoverable silver (Freeport Mc- pale, proximal, and early pyroxene is nearly pure diopside
MoRan Annual Report, 2001). (Di) and the dark green, distal, and late pyroxene ranges up
The Big Gossan Cu-Au skarn deposit is the highest grade to 75 mole percent hedenbergite (Hd) with minor jo-
copper deposit in the Ertsberg district, with current reserves hannsenite (Jo). In agreement with general skarn zonation
of 37.4 Mt grading 2.69 percent Cu, 1.02 g/t Au, and 16 g/t Ag patterns (Meinert, 1992, 1997), pyroxene becomes more iron-
(Meinert et al., 1997). Ore is associated with a series of 3 to 4 and manganese-rich toward the margins of the known de-
m.y. granodioritic dikes (McMahon, 1994 a, b, c) that have in- posit, suggesting that the hydrothermal system was centered
truded close to the near-vertical faulted contact between the in the middle of the deposit, coincident with the largest mass
Shale Member of the Cretaceous Ekmai Formation and the of dike. A similar zonation occurs vertically; the average py-
stratigraphically overlying carbonate sequences of the Pale- roxene composition for the highest third of the skarn system
ocene Waripi and Eocene Faumai Formations. Most miner- is Di58Hd38Jo4, whereas the average for the deepest third is
alized and altered sequences occur in the purer carbonate Di86Hd13Jo1.
rocks of the Waripi Formation, although biotite and calc-sili- Amphibole in the Big Gossan skarn system ranges from
cate hornfels alteration also occur in the clastic footwall rocks actinolite to cummingtonite, with the main substitution being
adjacent to skarn ore. The scale of the Big Gossan skarn sys- Fe-Mg-Mn for Ca. Most amphibole occurs as an alteration
tem is more than 1 km along strike, more than 500 m verti- product of pyroxene, typically with quartz, carbonate, anhy-
cally (open at depth), and up to 200 m in width for an aggre- drite, and sulfides. Subcalcic amphiboles occur on the mar-
gate volume of more than 0.1 km3 (Fig. 2). gins of skarn where pyroxene has been totally replaced. In
The calcic ± magnesian skarn assemblage in the Waripi For- these occurrences, amphibole is intimately intergrown with
mation hosts the bulk of the Big Gossan orebody and is char- quartz and carbonate, the latter ranging in composition from
acterized by relatively coarse-grained garnet and pyroxene, calcite to manganiferous siderite, suggesting that the location

0361-0128/98/000/000-00 $6.00 148


FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS 149

SW NE SW NE Elev. (m)
BGU 14-2
BGU 10-1
2900

Ertsberg Intrusion
BGU 14-3
BGU 10-3
Intrusive Breccia
2800

Massive sulfide cap


Pyx > Gar skarn
2700
Gar > Pyx skarn
Big Gossan dikes
BGU 14-6
2600
Marble (Tw)
Kkeh marker shale
BGU 10-5
Kembelangan Group
2500

BGU 14-7

A B BGU 10-6

2400

FIG. 2. Cross sections through the Big Gossan skarn system showing the distribution of garnet and pyroxene skarn. Lo-
cations shown in Figure 1.

of the pyroxene-amphibole transition was affected by higher percent KCl and 35 wt percent NaCl. By contrast, fluid in-
XCO2 near the marble front (Meinert et al., 1997). Thus, at a clusions in quartz and anhydrite associated with retrograde al-
given temperature near the skarn-marble contact, pyroxene teration average 7.1 wt percent NaCl equiv and spatially as-
was altered to amphibole plus carbonate according to a reac- sociated fluid inclusions homogenize to liquid and vapor at
tion (based upon measured compositions) such as: pyroxene 370° to 380°C (Fig. 3C, D ). This corresponds to boiling at a
+ H2O + CO2 = subcalcic amphibole + quartz + carbonate— pressure of 20 MPa, equivalent to a depth of 2 km below the
7Ca(Fe0.5Mg0.4Mn0.1)Si2O6 + H2O + 7CO2 = Ca(Fe2.8 Mg2.6 paleowater table under hydrostatic conditions. Since these
Mn0.6)Si8O22 (OH)2 + 6SiO2 + 7(Ca0.86Fe0.1 Mg0.03Mn0.01)CO3. samples came from 600 m beneath the present surface, the
Marble occurs for tens to hundreds of meters beyond the data indicate that at least 1.4 km of erosion has occurred in
Big Gossan skarn in the Waripi Formation hanging wall. the Ertsberg district over the past 4 m.y.
Within this aureole, grain size decreases systematically away
from skarn and dike contacts. Numerous planar to wavy vein-
lets, usually less than 1 mm thick and oriented perpendicular Mines Gaspé, Quebec
to the skarn front, appear to represent paleofluid conduits. Like the Ertsberg district, Mines Gaspé in Quebec has both
The dark centerline of these veinlets is marked by a concen- porphyry and skarn ore (Fig. 4); however, the Devonian host
tration of carbon, pyrite, sphalerite, galena, chlorite, serpen- rocks and porphyritic quartz monzonite stocks have U-Pb zir-
tine, and/or clay. con ages of 385 ± 2.6 Ma (Stephenson et al., 1998), which is
The fluid associated with prograde skarn at Big Gossan is a much older than the ages at Ertsberg. The Mines Gaspé dis-
high-temperature NaCl-KCl hypersaline liquid with a low trict was discovered in 1921 and has produced more than 150
CO2 content, <0.05 mole percent (Meinert et al., 1997). Ho- Mt of ore averaging 0.9 percent Cu (with production grades
mogenization temperatures for fluid inclusions in pyroxene up to 3.4% Cu). It has new drilled resources at Porphyry
range from 320° to 485°C, average 410°C, and formed at a Mountain of 200 Mt grading 0.73 percent Cu and 0.08 per-
pressure of 50 MPa, equivalent to a depth of 2 km, under cent Mo (Hussey and Bernard, 1998). The scale of the Por-
lithostatic conditions. Most fluid inclusions in pyroxene phyry Mountain skarn system is more than 1 km in diameter
contain multiple daughter minerals including halite, sylvite, and more than 800 m vertically (open at depth) for an aggre-
chalcopyrite, hematite, and anhydrite, indicating that a com- gate volume of more than 0.8 km3 (Fig. 5).
plex brine was present (Fig. 3A, B). Total salinity ranges from The Devonian host rocks at Mines Gaspé have been gently
38 to 65 wt percent NaCl + KCl; mean salinities are 22 wt folded and consist of carbonaceous silty limestone, calcareous

0361-0128/98/000/000-00 $6.00 149


150 MEINERT ET AL.

FIG. 3. Photomicrographs of inclusions in skarn minerals. A. Multiphase fluid inclusion in pyroxene containing vapor
bubble (V), halite (H), sylvite (S), hematite? (Hm), and unknown mineral (X). B. Multiphase fluid inclusion in pyroxene
containing vapor bubble (V), halite (H), sylvite (S), and chalcopyrite (Cpy). C. Vapor-rich fluid inclusion in quartz. D. Part
of a large pyroxene crystal which has been completely altered to amphibole (amph) and anhydrite. The anhydrite contains
hundreds of vapor-rich fluid inclusions, most of which homogenize to vapor.

Quebec
Mines York River Fm.
Gaspe Porphyry Brook
e
nclin
ampau Sy
Ch
Maine
Porphyry A
Atlantic Mtn
Boston Ocean
920A Indian Cove Fm. 20°

890 923 900


905 886
1 km 793 904 909

Copper
Mtn A' N65°

Shiphead Fm. Murdochville


15°

metamorphic aureole Forillon Fm.


(limit of bleaching) E65°30‘

FIG. 4. Location and geology of the Mines Gaspé district (modified from Allcock, 1982). Crosshatched areas are surface
projections of porphyry plutons. Sample locations are surface projections of drill holes.

siltstone and shale, and minor marker tuff beds of the Foril- bedding planes. Both stocks have been intensely altered by
lon, Shiphead, and Indian Cove Formations (Gower and the addition of secondary K feldspar and biotite, with rela-
Walker, 1993). These units were intruded by the Copper tively minor and late quartz-sericite-pyrite.
Mountain stock and the Porphyry Mountain stock, the latter Close to the Porphyry Mountain intrusion, argillaceous
encountered only in drill core. The margins of both stocks are limestone was altered to coarse-grained, dark reddish-brown
irregular, with numerous small dikes and sills intruded along skarn, consisting of >90 vol percent garnet, whereas more

0361-0128/98/000/000-00 $6.00 150


FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS 151

A' A Gar >> pyx skarn


Sea
Level Pyx > gar skarn

Indian Cove Fm.


Fault subparallel to Dark green pyx
Porphyry Brook
Fault Zone
Light green pyx

Calc-silicate hornfels

Hornfels

Forillon Fm. Shiphead Fm.


-500 m
Marble

Biotite hornfels
Cu
Quartz monzonite
Mo porphyry

Limit of intense
-1000 m K-silicate alteration
FIG. 5. Cross section with no vertical exaggeration through the Porphyry Mountain skarn system showing the distribution
of garnet and pyroxene skarn. Location (A’–A) shown in Figure 4. Top of cross section (sea level) is about 900 m below the
surface of Porphyry Mountain. Arrows indicate direction of increasing Cu and Mo grades relative to K silicate-altered core
of the intrusion. Gar = garnet, Pyx = pyroxene.

clastic-rich units were altered to dark- green pyroxene horn- commonly result in incomplete replacement of one phase by
fels. With increasing distance from the intrusion, skarn in another (e.g., Shimazaki and Kusakabe, 1990).
limestone becomes more pyroxene rich and the calc-silicate The determination of δ18 O (VSMOW) compositions by
hornfels in more clastic-rich units becomes lighter in color. laser fluorination of 2- to 3-mg samples from Big Gossan
All skarn was overprinted by veinlets and vein envelopes of (Table 1) provided average values of 6.1 per mil for garnet
hydrous minerals such as amphibole, chlorite, and epidote, (range, 5.5–7.2‰), 7.4 per mil for pyroxene (7.0–8.2‰), and
usually associated with sulfide minerals. Surrounding the two 8.6 per mil for amphibole (8.6–8.7‰), excluding one sample
main mineralized stocks is a composite aureole of bleaching with calcite inclusions that caused slightly higher δ18 O values
and recrystallization marked by a relatively sharp transition (10.9–11.1‰). No systematic spatial variation in isotopic
from recrystallized white marble to fine-grained, dark-gray compositions was determined for any of the mineral species.
limestone about 10 to 200 m beyond the limit of skarn (Fig. There is an average pyroxene-garnet difference of 1.3 per mil.
4). The distal thermal aureole can be detected for several If this difference reflects a bulk equilibrium signature of the
kilometers beyond visible alteration by a very subtle change in system, a calculated fractionation temperature (Zheng, 1993)
illite crystallinity (Williams-Jones, 1986). would be 370° ± 70°C (Table 1). This is in general agreement
Previous fluid inclusion and stable isotope investigations of with measured fluid inclusion temperatures, with all but one
the Copper Mountain porphyry copper deposit at Mines value ranging from 400° to 455°C (Table 1). Conventional
Gaspé did not distinguish prograde and retrograde alteration analysis of δD via Zn reduction of 100- to 200-mg samples of
stages but did find evidence for a predominance of magmatic amphibole resulted in average values of –77.5 per mil (range,
fluid, except for a relatively late, postmineralization influx of –77 to –78‰); no lower values were detected.
meteoric water. The proximal porphyry ore at Copper Moun- For Mines Gaspé samples (Table 1), the δ18O values aver-
tain formed from a fluid largely of magmatic origin (δDH2O = age 4.2 per mil for garnet (3.4–5.6‰), 5.3 per mil for pyrox-
–42 to –61‰, based on extraction of inclusion fluids; Shelton, ene (4.3–6.0‰), and 4.9 per mil for amphibole (4.3–5.2‰).
1983). Fluid inclusion results for skarn (Shelton, 1983) Again, if the average pyroxene-garnet difference of 1.1 per
indicate high temperatures (TH = 334°–506°C) and high mil reflects a bulk equilibrium signature of the system, the
salinities (15–56 wt % NaCl equiv), similar overall to the val- calculated fractionation temperature is 450° ± 100°C, also in
ues reported from the prograde stage of the Big Gossan cop- general agreement with independent temperature estimates
per skarn. of 450° to 500°C and slightly higher than the temperature es-
timates for Big Gossan. The δD value of primary igneous am-
Stable Isotope Results phibole at Mines Gaspé is –87 per mil, whereas skarn amphi-
Recent advances in stable isotope analytical methods allow bole record values from –78 per mil through –115 to –143 per
analysis of considerably smaller samples (a few mg for laser mil.
fluorination lines) than is required for conventional analyses
(tens to hundreds of mg). This is particularly important for Discussion
studies of skarn systems due to the finely intergrown na- Unlike contact metamorphism that is the result of the in-
ture of most skarn minerals and the complex reactions that trusion of relatively dry magmas (cf. Valley, 1986), world-class

0361-0128/98/000/000-00 $6.00 151


152 MEINERT ET AL.

TABLE 1. Stable Isotope and Temperature Data for the Big Gossan and Mines Gaspé Skarn Deposits

Big Gossan

Measured Calculated Measured Calculated Measured Flinc


Sample no. Mineral δ18O(VSMOW) δ18Ofluid δD(SMOW) δDfluid H2O T
(‰) (‰) (‰) (‰) (wt %) (°C)

1-6-600 Actinolite 8.6 9.6 –78 –24 2.6 350


1-6-600 Actinolite, duplicate 8.6 9.6
1-6-600 Actinolite, duplicate 8.7 9.7 350
8-1-213 Actinolite# 11.1 12.1 –77 –23 1.9 350
8-1-213 Actinolite #, duplicate 10.9 11.9 350

9-5-340 Pyroxene 7.0 8.8 455


10-3-78 Pyroxene 7.0 8.7 445
23-2-122 Pyroxene 7.3 9.0 445
23-4-420 Pyroxene 7.5 9.2 445
14-6-400 Pyroxene 8.2 10.1 490

26-4-250 Garnet 5.5 8.1 400


14-6-491 Garnet 5.6 8.2 400
23-2-165 Garnet 5.7 8.3 400
9-5-465 Garnet 6.2 8.8 400
8-5-187 Garnet 7.2 9.8 400
∆pyx-gar = 1.3 370 ± 70

Gaspé

Porphyry Mtn Hornblende 5.0 7.4 –87 –70 2.5 700


923-4859 Hornblende 4.3 5.3 –78 –24 1.8 350
900-5660-1 Actinolite 5.2 6.2 –115 –61 1.6 350
900-5660-2 Actinolite 5.2 6.2 –143 –89 1.6 350
900-5660-2 Actinolite* 4.9 5.9 350
793-2990 Actinolite 5.0 6.0 –130 –76 1.8 350

923-5008 Pyroxene 4.3 6.2 500


904-4011 Pyroxene 5.4 7.3 500
904-4011 Pyroxene, duplicate 5.6 7.5 500
890-4085 Pyroxene 6.0 7.9 500

909-4245B Garnet 3.4 6.2 450


905-3210 Garnet 3.6 6.4 450
904-3995 Garnet 3.6 6.4 450
886-4351 Garnet 4.3 7.1 450
920-5012 Garnet 4.5 7.3 450
909-4245A Garnet 5.6 8.4 450
∆pyx-gar = 1.1 450 ± 100

Mineral separates were purified by a combination of heavy liquids and magnetic methods followed by hand picking under a binocular microscope; purity
was checked optically and by X-ray diffraction; all oxygen and all deuterium, except for the Porphyry Mountain sample, were analyzed at the Geological
Survey of Japan using analytical methods described in Hedenquist et al. (1998); deuterium of the Porphyry Mountain sample was analyzed at Oregon State
University using methods described by Bigeleisen et al. (1952); temperatures for Big Gossan pyroxene and amphibole are based on pressure-corrected fluid
inclusion measurements (Meinert et al., 1997); temperatures of skarn minerals at Mines Gaspé are based on limited fluid inclusion and phase equilibria
data from Allcock (1982) and Shelton (1983); calculated fluid compositions for amphibole are based on fractionation factors of Bottinga and Javoy (1975)
and Suzuoki and Epstein (1976), and for garnet and pyroxene, the fractionation factors of Zheng (1993); the water content of amphibole was determined
by manometric measurement of H2 after reduction of the structural H2O released during stepwise heating to >1,000°C, reproducible to 0.1 wt % H2O;
this does notinclude absorbed water, which was removed by outgassing under vacuum pumping for 1–2 h at ~150°C; Flinc = fluid inclusion, # = calcite
inclusions, * = HCl leached

skarn ore deposits form as the result of wall-rock interaction effect on the isotopic mass balance of the skarn minerals.
with huge amounts of hydrothermal fluid. Such fluid has an Thus, these skarn minerals should indicate the composition of
origin not only from the immediate dike or stock intrusion but the hydrothermal fluid, at least those in a proximal position to
also comes from a parent magma chamber at a depth below the intrusion.
the ore deposit. This generalization of a magmatic fluid The fluid associated with prograde skarn formation at both
source is accepted, at least during the prograde stage (Ein- Big Gossan and Mines Gaspé was a high- temperature
audi et al., 1981). As a result of the fluid domination of the (350°–500°C), hypersaline liquid (up to 50–60 wt % NaCl +
system, the original host rock in the immediate vicinity of the KCl). The calculated δ18O value of the fluid in equilibrium
intrusive focus of hydrothermal fluid flow will soon have little with garnet and pyroxene from both deposits is within the

0361-0128/98/000/000-00 $6.00 152


FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS 153

range of values for magmatic fluids (Taylor and Sheppard, alteration, i.e., ratios <0.01 (Bowman, 1998). A small, <10
1986), and even minerals with minor calcite inclusions devi- percent, meteoric water component could account for slight
ate only slightly toward marine carbonate values of 21 to 25 variations in the range of δ18O composition, <2.5 per mil at
per mil (Fig. 6). The δD values of amphibole are also typical each deposit; however, such a small diluent could not be re-
of the range for igneous minerals (Taylor and Sheppard, sponsible for the greater than six-fold decrease in salinity
1986), although they deviate to relatively light compositions. from the prograde hypersaline fluid to 7 wt percent NaCl
The fluid associated with the retrograde alteration stage is fluid during the retrograde stage.
lower temperature (~350°–400°C) and lower salinity (avg 7 In some cases a large variation in salinity of the exsolved
wt % NaCl equiv) than that of the prograde skarn fluid. Cal- fluid can occur according to the degree of crystallization, but
culated δ18O and δD fluid compositions from Big Gossan am- this is critically dependent on the pressure (Cline, 1995). We
phibole (Table 1) are in the range typical of magmatic fluid argue that the porphyry situation modeled by Shinohara and
dissolved in a felsic melt (Taylor, 1986), with no clear trend Hedenquist (1997), in which there is a relatively small varia-
toward measured values for local meteoric water (Fig. 7). The tion in the salinity of the bulk fluid, may apply here during
amphibole samples from Mines Gaspé record a wider range much of the crystallization and exsolution. By contrast, the
of δD fluid compositions, from –24 to –89 per mil. This wide trend in the δD composition of residual water during crystal-
range may record the degassing of the parent magma cham- lization will always be toward lower values (Suzuki and
ber, as this process results in fractionation such that the resid- Epstein, 1976). Indeed, the residual water fixed by primary
ual water in the melt has lighter δD values (Suzuki and Ep- hydrous minerals, with intrusions typically having concentra-
stein, 1976). As a result, the δD values of fluid exsolved at a tions of 0.5 to 1 wt percent water, are observed to have δD
later stage will be progressively lower in δD whereas the δ18O values <–100 per mil as the result of late open-system de-
values remain relatively constant (Taylor, 1986, 1988; Bow- gassing (Taylor, 1986, 1988; Hedenquist and Richards, 1998).
man, 1998; Hedenquist et al., 1998). This observation suggests that the formation of retrograde
Interaction of the fluid with wall rock cannot account for alteration minerals in general, and amphibole- epidote in par-
this large variation in δD values where the δ18O values remain ticular, can result from evolution of a single magmatic hy-
constant, except possibly at unreasonably low water-to-rock drothermal system, since there is no evidence in the Big Gos-
ratios that may characterize only the very earliest stage of san skarn system for mixing with a significant component of
meteoric water. The sharp decrease in salinity, from prograde
hypersaline brine to a retrograde salinity of about 7 wt per-
-5 0 5 10 15 20 25 cent NaCl, can be explained if the later fluid followed a cool-
ing path where it never intersected its solvus (Fig. 8). In this
Big Gossan situation, a hypersaline liquid would not form, and the salin-
ity of the fluid that reaches the ore deposit is that of the bulk
Garnet salinity of the exsolved magmatic fluid (Meinert et al., 1997;
Shinohara and Hedenquist, 1997; Hedenquist et al., 1998).
Pyroxene Such a scenario would result in a zonation in both space and
time not only of the skarn mineralogy, as described earlier,
Marine Carbonate

Amphibole but more fundamentally of the fluid from which those miner-
als precipitated (Fig. 8).
Amphibole
with calcite Model of Evolution
Gaspé inclusions Upon initial intrusion of the parent magma to shallow
Garnet depths the temperature of both magma and immediately
adjacent wall rocks was >400°C, the normal limit of brittle
Pyroxene behavior. Thus, the rocks behaved in a ductile fashion, sealing
the system from significant interaction with connate and me-
Amphibole teoric waters (Fournier, 1992). As the pluton cooled and crys-
Igneous tallized at depths of, say, 4 to 5 km (Fig. 8), the melt eventu-
amphibole ally saturated with respect to an aqueous fluid, which was a
Magmatic supercritical fluid with a moderate salinity of 6 to 8 wt percent
fluid (Burnham, 1979; Yang and Bodnar, 1994; Bodnar, 1995). This
fluid ponded near the crystalline shell of the magma chamber
at a subsolidus temperature of about 600°C (Burnham and
-5 0 5 10 15 20 25 Ohmoto, 1980; Fournier, 1987). Subsequent fluid ascent to
shallower depths probably was concurrent with intrusion of
δ18 O (0/ 00, VSMOW) porphyry dikes to approximately 2 to 3 km beneath the paleo-
surface, with the dikes providing the permeability to focus the
FIG. 6. Calculated δ18O (VSMOW) composition of fluids in equilibrium advective ascent of exsolved magmatic fluid.
with garnet, pyroxene, and amphibole from the Big Gossan and Mines Gaspé
skarn deposits. See Table 1 for data and sources of fractionation factors. The The combination of high fluid flux and thermal mass of the
average composition of magmatic fluid and the marine carbonate values are dikes kept the early fluid relatively hot during ascent, with
from Taylor (1986) and references therein. only minor conductive cooling (Shinohara and Hedenquist,

0361-0128/98/000/000-00 $6.00 153


154 MEINERT ET AL.

25
Initial water
0 Big Gossan skarn dissolved in melt
amphibole
δD (0/ 00, VSMOW) -25 Gaspe skarn
amphibole
-50 Gaspe igneous

Degassing
amphibole
-75

-100

-125
Present-day
-150 Ertsberg meteoric
water, snow, & ice
-175
-25 -20 -15 -10 -5 0 5 10 15
δ18 O (0/ 00, VSMOW)
FIG. 7. Calculated δ18O versus δD compositions of fluids in equilibrium with amphibole (Table 1) from the Big Gossan
and Mines Gaspé skarn deposits. The field for water initially dissolved in silicate melt is from Taylor (1992). As water exsolves
from magma, the δD of residual water remaining in the melt becomes lighter (Suzuoki and Epstein, 1976). Therefore, the
δD of late exsolved water is lower (Taylor, 1986; Hedenquist and Richards, 1998). Values for present-day meteoric waters
from the Ertsberg district are from Harrison et al. (1999).

200 400
T°C 600 800

Na:K = 1:1
1 Vapor
Liquid +
Liquid
20 2 Vapor + Salt
1 1
0.
Depth (hydrostatic)

40 4 0.2
7
Pressure (MPa)

5 Skarn 2
80

0.5
wt

Retrograde 10
%

60 6 alteration with
qtz-anhydrite,
amphibole, 1.0
FIG. 8. Composition of coexisting liquid and vapor as a function
and sulfides
Depth (lithostatic)

of depth and temperature (after Fournier, 1987). Isopleths of


60

3 NaCl in liquid (dark lines) and vapor (light lines) are shown for
wt%

80 compositions of interest. Two different cooling paths are shown for


an initial liquid composition of 7 wt percent NaCl. The liquid that
ascends along the “skarn” trajectory hits its solvus at about 500°C
20 and a pressure of 50 MPa (lithostatic pressure at a depth of about
40 w

2 km). Continued depressurization results in the separation of a


NaCl in liquid small amount of hypersaline liquid with a low-salinity vapor domi-
t%

100 4 nating the mass of the two-phase fluid. The later liquid that as-
cends along the “retrograde” trajectory intersects its saturated
vapor-pressure curve at about 370°C and a hydrostatic pressure of
6-8 wt% 20 MPa, also equivalent to a depth of about 2 km. At this point the
NaCl fluid begins to boil, with a small amount of vapor separating to
leave the residual liquid slightly higher in salinity.

0361-0128/98/000/000-00 $6.00 154


FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS 155

1997). The highest temperature and lithostatic pressure de- on boiling from 370° to 320°C (Henley et al., 1984). At this
duced for the formation of these deposits during their pro- shallow depth and relatively low temperature, dissolved gases
grade stage is 535°C and 50 MPa, respectively (Meinert et al., will be lost to the vapor, possibly explaining sulfide mineral-
1997). This is consistent with the steep isotherms around a ization from the low-salinity fluid that occurs during this late
near-solidus magma that is exsolving an aqueous fluid (Shino- stage.
hara and Hedenquist, 1997). A magmatic fluid with bulk This general scenario (Fig. 8) illustrates only two snap-
salinity of 7 wt percent NaCl will intersect its solvus at a 2-km shots of a dynamic system. In reality, multiple intrusions—
depth and separate into a 50 wt percent hypersaline liquid in into the parent chamber and as shallow dikes—combined
equilibrium with a vapor containing ~1 wt percent NaCl (Fig. with fluctuations in the water table due to structural, paleo-
8). Based on the fluid inclusion evidence for temperature and surface, and/or climatic changes (e.g., Simmons, 1991)
salinity, most prograde skarn formed from this hypersaline would have caused a more complex overprinting of various
liquid. skarn and retrograde alteration stages. These changes are
Upon aqueous phase separation, most of the SO2 and HCl seen in the rock record as a series of crosscutting features
from the parent fluid will partition into the vapor phase (Can- and zonation patterns that are typical of skarn and intrusion-
dela and Piccoli, 1995; Scaillet et al., 1998; Keppler, 1999), centered ore deposits worldwide (Einaudi et al., 1981; Ben-
which being less dense forms a vapor plume overlying the dezú and Fontboté, 2002; Rusk and Reed, 2002). Neverthe-
reservoir of hypersaline liquid that is generated (Henley and less, despite this complexity we conclude that the fluid
McNabb, 1978). This creation of a hypersaline liquid with responsible for prograde and retrograde alteration as well as
high base metal solubilities and relatively low sulfur content ore formation was dominantly magmatic in origin. Similar
may have inhibited sulfide deposition during formation of the conclusions have been reached for porphyry copper de-
prograde skarn, similar to Bodnar’s (1995) argument for por- posits which formed largely within igneous rocks (e.g.,
phyry deposits. Condensation of much of the vapor plume at Kusakabe et al., 1990; Shinohara and Hedenquist, 1997;
shallow depths would have formed a highly acidic, low-salin- Hedenquist et al., 1998; Watanabe and Hedenquist, 2001;
ity liquid which most likely would have caused advanced Harris and Golding, 2002). Here we extend such conclu-
argillic alteration (Hedenquist, 1995) and dissolution of the sions to skarn deposits in which hydrothermal fluids have es-
host carbonate rocks with attendant brecciation, again analo- caped the igneous carapace and interacted extensively with
gous to formation of volcanic-hosted lithocaps over porphyry surrounding wall rocks.
deposits (Hedenquist et al., 1998). This can be thought of as
the ground-preparation stage of a growing skarn system in Acknowledgments
which the vapor plume produces cavernous porosity in car- This work was partially supported by National Science
bonate rock, to be overrun subsequently by the high- salinity, Foundation grant EAR-9725437. The senior author’s visit to
skarn-forming fluid as the system expands. Thus, the many the Geological Survey of Japan was facilitated by a Japan So-
breccias that occur at both Big Gossan and Mines Gaspé, as ciety for Promotion of Science short-term invitation fellow-
well as at other skarn deposits in the porphyry environment ship, administered by the National Science Foundation. We
(Einaudi, 1982), are part of the normal early-stage hydrother- thank Steve Rowins, John Bowman, and John Dilles for help-
mal evolution of such systems. ful comments on earlier versions of this paper.
As the underlying vapor-saturated magma chamber contin- April 5, October 17, 2002
ued to cool and crystallize, fluid continually exsolved, and this
later fluid may have been relatively similar in bulk Na-K-Cl REFERENCES
composition to that of the earlier stage fluid (Cline, 1995; Shi- Allcock, J.B., 1982, Skarn and porphyry copper mineralization at Mines
nohara and Hedenquist, 1997). However, because of the lack Gaspé, Murdochville, Québec: ECONOMIC GEOLOGY, v. 77, p. 971–999.
of concurrent dike intrusion and/or a sharply lower fluid flux Bendezú, R., and Fontboté, L., 2002, Late timing for high sulfidation
as the parent magma approached the end stages of stagnant Cordilleran base metal lode and replacement deposits in porphyry-related
districts: The case of Colquijirca, central Peru: Society for Geology Applied
crystallization (Shinohara and Hedenquist, 1997), the ascend- to Mineral Deposits News, no. 13, p. 2, 9–13.
ing fluid followed a cooling path different from that of the Bigeleisen, J., Perlman, M.L., and Prosser, H.C., 1952, Conversion of hydro-
early stage. The lower rate of advection (Shinohara and gen materials to hydrogen for isotopic analysis: Analytical Chemistry, v. 24,
Hedenquist, 1997) meant that this late fluid cooled suffi- p. 1356–1357.
ciently during ascent so that it never intersected its solvus Bodnar, R.J., 1995, Fluid-inclusion evidence for a magmatic source for met-
als in porphyry copper deposits: Mineralogical Association of Canada Short
(Fig. 8). Course Series, v. 23, p. 139–152.
Cooling to a temperature below approximately 400°C results Bottinga, Y., and Javoy, M., 1975, Oxygen isotope partitioning among the
in a transition from ductile to brittle rock behavior, causing a minerals in igneous and metamorphic rocks: Reviews of Geophysics and
change from lithostatic to hydrostatic conditions (Fournier, Space Physics, v. 13, p. 401–418.
Bowman, J.R., 1998, Stable-isotope systematics of skarns: Mineralogical As-
1991). Thus, even though the later stage fluid reached the sociation of Cananda Short Course, v. 26, p. 99–145.
previously formed skarn at the same paleodepth of 2 km, the Burnham, C.W., 1979, Magmas and hydrothermal fluids, in Barnes H.L., ed.,
pressure was 20 MPa under the hydrostatic conditions. A 7 wt Geochemistry of hydrothermal ore deposits, 2nd edition: New York, John
percent NaCl liquid at 20 MPa will intersect the saturated Wiley and Sons, p. 71–136
vapor-pressure curve at 370°C, resulting in boiling of the as- Burnham, C.W., and Ohmoto, H., 1980, Late-stage processes of felsic mag-
matism; Society of Mining Geologists of Japan Special Issue 8, p. 1–11
cending liquid. Continued boiling and vapor loss will result in Candela, P.A., and Piccoli, P.M., 1995, Model of ore-metal partitioning from
progressive cooling accompanied by only a slight increase in melts into vapor and vapor/brine mixtures: Mineralogical Association of
salinity due to vapor loss, e.g., from 7 to 10 wt percent NaCl Canada Short Course, v. 23, p. 101–128.

0361-0128/98/000/000-00 $6.00 155


156 MEINERT ET AL.

Cline, J.S., 1995, Genesis of porphyry copper deposts: The behavior of water, Meinert, L.D., 1992, Skarns and skarn deposits: Geoscience Canada, v. 19, p.
chloride, and copper in crystallizing melts: Arizona Geological Society Di- 145–162.
gest, v. 20, p. 69–82. ——1997, Application of skarn deposit zonation models to mineral explo-
Einaudi, M.T., 1982, Descriptions of skarn associated with porphyry copper ration: Exploration and Mining Geology, v. 6, p. 185–208.
plutons, southwestern North America, in Titley, S.R., ed., Advances in ge- Meinert, L.D., Hefton, K.K., Mayes, D., and Tasiran, I., 1997, Geology, zona-
ology of the porphyry copper deposits, southwestern North America: Tuc- tion, and fluid evolution of the Big Gossan Cu-Au skarn deposit, Ertsberg
son, University of Arizona Press, p. 139–184. district, Irian Jaya: ECONOMIC GEOLOGY, v. 92, p. 509–526.
Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits: Pettke, T., and Diamond, L.W., 1997, Oligocene gold quartz veins at Brus-
ECONOMIC GEOLOGY 75TH ANNIVERSARY VOLUME, p. 317–391. son, NW Alps: Sr isotopes trace the source of ore-bearing fluid to over a 10-
Fournier, R.O., 1987, Conceptual models of brine evolution in magmatic-hy- km depth: ECONOMIC GEOLOGY, v. 92, p. 389–406.
drothermal systems: U.S. Geological Survey Professional Paper 1350, p. Rusk, B., and Reed, M., 2002, Scanning electron microscope-cathodolumi-
1487–1506. nescence analysis of quartz reveals complex growth histories in veins from
——1991, The transition from hydrostatic to greater than hydrostatic fluid the Butte porphyry copper deposit, Montana: Geology, v. 30, p. 727–730.
pressure in presently active continental hydrothermal systems in crystalline Scaillet, B., Clemente, B., Evans, B.W., and Pichavant, M., 1998, Redox con-
rock: Geophysical Research Letters, v. 18, p. 955–958. trol of sulfur degassing in silicic magmas: Journal of Geophysical Research,
——1992, The influences of depth of burial and the brittle-ductile transition v. 103, p. 23,937–23,949.
on the evolution of magmatic fluids: Geological Survey of Japan Report Shelton, K.L., 1983, Composition and origin of ore-forming fluids in a car-
277, p. 57–59. bonate-hosted porphyry copper and skarn deposit: A fluid inclusion and
Freeport McMoRan, 2001, Annual Report: New Orleans, Louisiana, 70 p. stable isotope study of Mines Gaspé, Québec: ECONOMIC GEOLOGY, v. 78,
Giggenbach, W.F., 1997, The origin and evolution of fluids in magmatic-hy- p. 387– 421.
drothermal systems, in Barnes, H.L., ed., Geochemistry of hydrothermal Shimazaki, H., and Kusakabe, M., 1990, Oxygen isotope study of the
ore deposits, 3rd edition: New York, John Wiley and Sons, p. 737–796. Kamioka Zn-Pb skarn deposits, central Japan: Mineralium Deposita, v. 25,
Gower, S.J., and Walker, J.A., 1993, Skarn-type, base-metal deposits in north- p. 221–229.
ern N.B. Geology and skarn occurrences: Geological Society of the CIM, Shinohara, H., and Hedenquist, J.W., 1997, Constraints on magma degassing
Annual Field Conference, 3rd, 1993, Bathurst, Trip 1 Guidebook, p. 5–21. beneath the Far Southeast porphyry Cu-Au deposit, Philippines: Journal of
Harris, A.C., and Golding, S.D., 2002, New evidence of magmatic-fluid-re- Petrology, v. 38, p. 1741–1752.
lated phyllic alteration: Implications for the genesis of porphyry Cu de- Simmons, S.F., 1991, Hydrologic implications of alteration and fluid inclu-
posits: Geology, v. 30, p. 335–338. sion studies in the Fresnillo district, Mexico: Evidence for a brine reservoir
Harrison, J., Kyle, J.R., Pennington, J., and Kavalieris, I., 1999, Hydrother- and a descending water table during the formation of hydrothermal Ag-Pb-
mal alteration and fluid evolution of the Grasberg Cu-Au deposit, Irian Zn orebodies: ECONOMIC GEOLOGY, v. 86, p. 1579–1601.
Jaya, Indonesia [abs.]: Geological Society of America, Abstracts with Pro- Stephenson, E.M., Meinert, L.D., Mortensen, J.K., and Hussey, J., 1998, Age
grams, v. 31, p. A403. and skarn alteration of the Porphyry Mountain Cu-Mo deposit, Mines
Hedenquist, J.W., 1995, The ascent of magmatic fluid: Discharge versus min- Gaspé, Québec [abs.]: Geological Society of America Abstracts with Pro-
eralization: Mineralogical Association of Canada Short Course, v. 23, p. grams, v. 30, p. A-372.
263–290. Suzuoki, T., and Epstein, S., 1976, Hydrogen isotope fractionation between
Hedenquist, J.W., and Lowenstern, J. B., 1994, The role of magmas in the OH-bearing minerals and water: Geochimica et Cosmochimica Acta, v. 40,
formation of hydrothermal ore deposits: Nature, v. 370, p. 519–527. p. 1229–1240.
Hedenquist, J.W., and Richards, J.P., 1998, The influence of geochemical Taylor, B.E., 1986, Magmatic volatiles: Isotopic variation of C, H, and S: Re-
techniques on the development of genetic models for porphyry copper de- views in Mineralogy, v. 16, p. 185–226.
posits: Reviews in Economic Geology, v. 10, p. 235–256. —— 1988, Degassing of rhyolitic magmas: Hydrogen isotope evidence and
Hedenquist, J.W., Arribas, A., Jr., and Reynolds, T.J., 1998, Evolution of an implications for magmatic- hydrothermal ore deposits: Canadian Institute
intrusion-centered hydrothermal system: Far Southeast-Lepanto porphyry of Mining and Mineralogy Special Volume 39, p. 33–49.
and epithermal Cu-Au deposits, Philippines: ECONOMIC GEOLOGY, v. 93, p. ——1992, Degassing of H2O from rhyolite magma during eruption and shal-
373–404. low intrusion, and the isotopic composition of magmatic water in hy-
Henley, R.W., and McNabb, A., 1978, Magmatic vapor plumes and ground- drothermal systems: Geological Survey of Japan Report 279, p. 190–194.
water interaction in porphyry copper emplacement: ECONOMIC GEOLOGY, Taylor, H.P., Jr., 1997, Oxygen and hydrogen isotope relationships in hy-
v. 73, p. 1–20. drothermal mineral deposits, in Barnes, H.L., ed., Geochemistry of hy-
Henley, R.W., Truesdell, A.H., Barton, P.B., Jr., and Whitney, J.A., 1984, drothermal ore deposits, 3rd edition: New York, John Wiley and Sons, p.
Fluid-mineral equilibria in hydrothermal systems: Reviews in Economic 229–302.
Geology, v. 1, 267 p. Taylor, H.P., Jr., and Sheppard, S.M.F., 1986, Igneous rocks: I. Processes of
Hussey, J., and Bernard, P., 1998, Exploration of the Porphyry Mountain Cu- isotopic fractionation and isotope systematics: Reviews in Mineralogy, v. 16,
Mo deposit: Mining Engineering, v. 50, p. 36–44. p. 227–271.
Keppler, H., 1999, Experimental evidence for the source of excess sulfur in Valley, J.W., 1986, Stable isotope geochemistry of metamorphic rocks: Re-
explosive volcanic eruptions: Science, v. 284, p. 1652–1654. views in Mineralogy, v. 16, p. 445–489.
Kusakabe, M., Hori, M., and Matsuhisa, Y., 1990, Primary mineralization of Watanabe, Y., and Hedenquist, J.W., 2001, Mineralogic and stable isotope
the El Teniente and Rio Blanco porphyry copper deposits, Chile: Stable zonation at the surface over the El Salvador porphyry copper deposit,
isotopes, fluid inclusions, and Mg2+/Fe2+/Fe3+ ratios of hydrothermal bio- Chile: ECONOMIC GEOLOGY, v. 96, p. 1775–1797.
tite: University of Western Australia, Geology Department Publication 23, Williams-Jones, A.E., 1986, Low-temperature metamorphism of the rocks
p. 244–259. surrounding les Mines Gaspé, Quebec: Implications for mineral explo-
Kwak, T.A.P., 1986, Fluid inclusions in skarns (carbonate replacement de- ration: ECONOMIC GEOLOGY, v. 81, p. 466–470.
posits): Journal of Metamorphic Geology, v. 4, p. 363–384. Yang, K., and Bodnar, R.J., 1994, Magmatic-hydrothermal evolution in the
McMahon, T.P., 1994a, Pliocene intrusions in the Ertsberg (Gunung Bijih) “Bottoms” of porphyry copper systems: Evidence from silicate melt and
mining district, Irian Jaya, Indonesia: Petrography, geochemistry, tectonic aqueous fluid inclusions in granitoid intrusions in the Gyeongsang basin,
setting: Unpublished Ph.D. dissertation, University of Texas, Austin, 299 p. South Korea: International Geology Review, v. 36, p. 608–628.
——1994b, Pliocene Cu-Au-bearing igneous intrusions of the Gunung Bijih Zheng, Y.F., 1993, Calculation of oxygen isotope fractionation in anhydrous
(Ertsberg) district, Irian Jaya, Indonesia: Petrography and mineral chem- silicate minerals: Geochimica et Cosmochimica Acta, v. 57, p. 1079–1091.
istry: International Geology Review, v. 36, p. 820–849.
——1994c, Pliocene Cu-Au-bearing igneous intrusions of the Gunung Bijih
(Ertsberg) district, Irian Jaya, Indonesia: Major and trace element chem-
istry: International Geology Review, v. 36, p. 925–946.

0361-0128/98/000/000-00 $6.00 156

You might also like