Sie sind auf Seite 1von 10

Economic Geology

Vol. 98, 2003, pp. 147156

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 liquid. 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 worlds largest Cu-Au resource. Stable isotope compositions of anhydrous and hydrous alteration minerals from both deposits indicate
a magmatic source for both the prograde and retrograde stages: 18O averages 5.0 per mil for garnet (range,
3.47.2), 6.5 per mil for pyroxene (4.38.2), and 7.1 per mil for amphibole (4.38.7). The D 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 mixing with meteoric water. We conclude that the differences in fluid compositionprograde versus retrograde
stagesresulted from a magmatic fluid that intersected its solvus during the early stage, creating vapor and hypersaline 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
THE SOURCE of ore fluids is one of the most important topics
of applied geologic research (Hedenquist and Lowenstern,
1994; Pettke and Diamond, 1997; Harris and Golding, 2002).
Many studies have demonstrated multiple sources of fluids responsible for forming ore deposits. A variety of theories of ore
genesis invoke interactions among such fluids, particularly
those of magmatic and meteoric origin (Giggenbach, 1997;
and Taylor, 1997, respectively). Since most skarn ores have an
intimate spatial relationship with magmatic intrusions, their
minerals should provide a clear record of the fluid(s) that were
present in the intrusive environment during ore formation.
Skarn ore deposits are typically characterized by two distinct
alteration styles: an early prograde stage with anhydrous minerals, such as garnet and pyroxene, which forms from relatively high-temperature, hypersaline liquid (Kwak, 1986); and
a later retrograde stage which consists of hydrous minerals,
such as epidote, amphibole, and chlorite, and forms from lower
temperature, lower salinity fluids (Kwak, 1986). These two alteration stages commonly are thought to reflect a dominance
Corresponding

author: e-mail, meinert@wsu.edu


*Address after August 1, 2003: Department of Geology, Smith College,
Northampton, MA 01063; e-mail, Lmeinert@smith.edu.
0361-0128/01/3320/147-10 $6.00

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


1981). A review of stable isotope systematics in skarn systems
by Bowman (1998) indicates a dominance of magmatic waters
during the early prograde stage, in which anhydrous minerals
dominate, and a range from largely magmatic to largely meteoric waters during the later retrograde stage in which hydrous
minerals dominate.
We present data from two copper skarn deposits in Irian
Jaya and Canada of different ages, host rocks, and geologic
terrane. In both deposits, there was a transition from hightemperature, hypersaline liquid associated with prograde garnet-pyroxene skarn to lower temperature, lower salinity liquid
associated with retrograde skarn alteration. Based upon data
presented here, we interpret this transition to have been
caused by a common parent magmatic fluid that experienced
a change in pressure-temperature path during ascent.
Big Gossan Deposit, Ertsberg District, Irian Jaya
The Ertsberg district contains multiple Cu-Au skarn and
porphyry deposits, including the original Ertsberg discovery,
the GBT/IOZ/DOZ orebodies, Dom, Big Gossan, Kucing
Liar, and the largest of all, Grasberg, plus several other mineral occurrences in various stages of exploration and development (Fig. 1). Individually, Grasberg is the largest gold mine

147

148

MEINERT ET AL.
Ertsberg District

50

Papua
New
Guinea

70

Ti

45
Irian Jaya

Lembah
Tembaga

Ti

25Cairns
Australia

*
S4

Tk

Grasberg

Tk
Pacific
Ocean

Kkel
Sydney

Tw

Yello
w Va
lley S
yncli
ne

Kucing
Liar

Fig. 2A
Big Gossan

GBT

*
* *
*
Dom
Ertsberg

70

45

Puncak
Jaya
4883m

70
Intrusion (Ti)
Thrust fault
Tertiary
carbonate rocks
Cretaceous
clastic rocks
0

km

Fig. 2B

40
75

60
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
of the Ertsberg district collectively constitute the worlds
largest Cu-Au resource. At the end of 2001, proven and probable reserves were estimated at 2,600 Mt of ore at an average
grade of 1.13 percent Cu, 1.05 g/t Au, and 3.72 g/t Ag, containing 23.8 Mt of recoverable copper, 64.5 Moz of recoverable gold, and 151.6 Moz of recoverable silver (Freeport McMoRan Annual Report, 2001).
The Big Gossan Cu-Au skarn deposit is the highest grade
copper deposit in the Ertsberg district, with current reserves
of 37.4 Mt grading 2.69 percent Cu, 1.02 g/t Au, and 16 g/t Ag
(Meinert et al., 1997). Ore is associated with a series of 3 to 4
m.y. granodioritic dikes (McMahon, 1994 a, b, c) that have intruded close to the near-vertical faulted contact between the
Shale Member of the Cretaceous Ekmai Formation and the
stratigraphically overlying carbonate sequences of the Paleocene Waripi and Eocene Faumai Formations. Most mineralized and altered sequences occur in the purer carbonate
rocks of the Waripi Formation, although biotite and calc-silicate hornfels alteration also occur in the clastic footwall rocks
adjacent to skarn ore. The scale of the Big Gossan skarn system is more than 1 km along strike, more than 500 m vertically (open at depth), and up to 200 m in width for an aggregate volume of more than 0.1 km3 (Fig. 2).
The calcic magnesian skarn assemblage in the Waripi Formation hosts the bulk of the Big Gossan orebody and is characterized by relatively coarse-grained garnet and pyroxene,
0361-0128/98/000/000-00 $6.00

with an average mineral ratio of about 1:2. The average composition of all analyzed garnets is andradite84.7grossularite13.5spessartine1.5pyrope0.3. Garnet higher in the deposit and
on the western and eastern margins shows a slight increase in
iron relative to that in deeper, more central locations. The
iron content of pyroxene is zoned in both space and time. The
pale, proximal, and early pyroxene is nearly pure diopside
(Di) and the dark green, distal, and late pyroxene ranges up
to 75 mole percent hedenbergite (Hd) with minor johannsenite (Jo). In agreement with general skarn zonation
patterns (Meinert, 1992, 1997), pyroxene becomes more ironand manganese-rich toward the margins of the known deposit, suggesting that the hydrothermal system was centered
in the middle of the deposit, coincident with the largest mass
of dike. A similar zonation occurs vertically; the average pyroxene composition for the highest third of the skarn system
is Di58Hd38Jo4, whereas the average for the deepest third is
Di86Hd13Jo1.
Amphibole in the Big Gossan skarn system ranges from
actinolite to cummingtonite, with the main substitution being
Fe-Mg-Mn for Ca. Most amphibole occurs as an alteration
product of pyroxene, typically with quartz, carbonate, anhydrite, and sulfides. Subcalcic amphiboles occur on the margins of skarn where pyroxene has been totally replaced. In
these occurrences, amphibole is intimately intergrown with
quartz and carbonate, the latter ranging in composition from
calcite to manganiferous siderite, suggesting that the location

148

149

FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS

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

BGU 10-6
2400

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

of the pyroxene-amphibole transition was affected by higher


XCO2 near the marble front (Meinert et al., 1997). Thus, at a
given temperature near the skarn-marble contact, pyroxene
was altered to amphibole plus carbonate according to a reaction (based upon measured compositions) such as: pyroxene
+ H2O + CO2 = subcalcic amphibole + quartz + carbonate
7Ca(Fe0.5Mg0.4Mn0.1)Si2O6 + H2O + 7CO2 = Ca(Fe2.8 Mg2.6
Mn0.6)Si8O22 (OH)2 + 6SiO2 + 7(Ca0.86Fe0.1 Mg0.03Mn0.01)CO3.
Marble occurs for tens to hundreds of meters beyond the
Big Gossan skarn in the Waripi Formation hanging wall.
Within this aureole, grain size decreases systematically away
from skarn and dike contacts. Numerous planar to wavy veinlets, usually less than 1 mm thick and oriented perpendicular
to the skarn front, appear to represent paleofluid conduits.
The dark centerline of these veinlets is marked by a concentration of carbon, pyrite, sphalerite, galena, chlorite, serpentine, and/or clay.
The fluid associated with prograde skarn at Big Gossan is a
high-temperature NaCl-KCl hypersaline liquid with a low
CO2 content, <0.05 mole percent (Meinert et al., 1997). Homogenization temperatures for fluid inclusions in pyroxene
range from 320 to 485C, average 410C, and formed at a
pressure of 50 MPa, equivalent to a depth of 2 km, under
lithostatic conditions. Most fluid inclusions in pyroxene
contain multiple daughter minerals including halite, sylvite,
chalcopyrite, hematite, and anhydrite, indicating that a complex brine was present (Fig. 3A, B). Total salinity ranges from
38 to 65 wt percent NaCl + KCl; mean salinities are 22 wt
0361-0128/98/000/000-00 $6.00

percent KCl and 35 wt percent NaCl. By contrast, fluid inclusions in quartz and anhydrite associated with retrograde alteration average 7.1 wt percent NaCl equiv and spatially associated fluid inclusions homogenize to liquid and vapor at
370 to 380C (Fig. 3C, D ). This corresponds to boiling at a
pressure of 20 MPa, equivalent to a depth of 2 km below the
paleowater table under hydrostatic conditions. Since these
samples came from 600 m beneath the present surface, the
data indicate that at least 1.4 km of erosion has occurred in
the Ertsberg district over the past 4 m.y.
Mines Gasp, Quebec
Like the Ertsberg district, Mines Gasp in Quebec has both
porphyry and skarn ore (Fig. 4); however, the Devonian host
rocks and porphyritic quartz monzonite stocks have U-Pb zircon ages of 385 2.6 Ma (Stephenson et al., 1998), which is
much older than the ages at Ertsberg. The Mines Gasp district was discovered in 1921 and has produced more than 150
Mt of ore averaging 0.9 percent Cu (with production grades
up to 3.4% Cu). It has new drilled resources at Porphyry
Mountain of 200 Mt grading 0.73 percent Cu and 0.08 percent Mo (Hussey and Bernard, 1998). The scale of the Porphyry Mountain skarn system is more than 1 km in diameter
and more than 800 m vertically (open at depth) for an aggregate volume of more than 0.8 km3 (Fig. 5).
The Devonian host rocks at Mines Gasp have been gently
folded and consist of carbonaceous silty limestone, calcareous

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
Gaspe

York River Fm.


Porphyry Brook
e
nclin
au Sy
p
m
a
Ch

Maine

Boston

1 km

Atlantic
Ocean

Porphyry
Mtn

920A
923
890
905 886
793

904

Indian Cove Fm.

20

900
909

Copper
Mtn A'

N65
Murdochville

Shiphead Fm.

15

metamorphic aureole
(limit of bleaching)

E6530

Forillon Fm.

FIG. 4. Location and geology of the Mines Gasp 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 Forillon, Shiphead, and Indian Cove Formations (Gower and
Walker, 1993). These units were intruded by the Copper
Mountain stock and the Porphyry Mountain stock, the latter
encountered only in drill core. The margins of both stocks are
irregular, with numerous small dikes and sills intruded along
0361-0128/98/000/000-00 $6.00

bedding planes. Both stocks have been intensely altered by


the addition of secondary K feldspar and biotite, with relatively minor and late quartz-sericite-pyrite.
Close to the Porphyry Mountain intrusion, argillaceous
limestone was altered to coarse-grained, dark reddish-brown
skarn, consisting of >90 vol percent garnet, whereas more

150

151

FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS

-500 m

Cu
Mo

-1000 m

Indian Cove Fm.

Fault subparallel to
Porphyry Brook
Fault Zone

Gar >> pyx skarn


Pyx > gar skarn

Forillon Fm. Shiphead Fm.

A'
Sea
Level

Hornfels

Dark green pyx


Light green pyx
Calc-silicate hornfels

Marble
Biotite hornfels
Quartz monzonite
porphyry
Limit of intense
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 (AA) 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 hornfels. With increasing distance from the intrusion, skarn in
limestone becomes more pyroxene rich and the calc-silicate
hornfels in more clastic-rich units becomes lighter in color.
All skarn was overprinted by veinlets and vein envelopes of
hydrous minerals such as amphibole, chlorite, and epidote,
usually associated with sulfide minerals. Surrounding the two
main mineralized stocks is a composite aureole of bleaching
and recrystallization marked by a relatively sharp transition
from recrystallized white marble to fine-grained, dark-gray
limestone about 10 to 200 m beyond the limit of skarn (Fig.
4). The distal thermal aureole can be detected for several
kilometers beyond visible alteration by a very subtle change in
illite crystallinity (Williams-Jones, 1986).
Previous fluid inclusion and stable isotope investigations of
the Copper Mountain porphyry copper deposit at Mines
Gasp did not distinguish prograde and retrograde alteration
stages but did find evidence for a predominance of magmatic
fluid, except for a relatively late, postmineralization influx of
meteoric water. The proximal porphyry ore at Copper Mountain formed from a fluid largely of magmatic origin (DH2O =
42 to 61, based on extraction of inclusion fluids; Shelton,
1983). Fluid inclusion results for skarn (Shelton, 1983)
indicate high temperatures (TH = 334506C) and high
salinities (1556 wt % NaCl equiv), similar overall to the values reported from the prograde stage of the Big Gossan copper skarn.
Stable Isotope Results
Recent advances in stable isotope analytical methods allow
analysis of considerably smaller samples (a few mg for laser
fluorination lines) than is required for conventional analyses
(tens to hundreds of mg). This is particularly important for
studies of skarn systems due to the finely intergrown nature of most skarn minerals and the complex reactions that
0361-0128/98/000/000-00 $6.00

commonly result in incomplete replacement of one phase by


another (e.g., Shimazaki and Kusakabe, 1990).
The determination of 18 O (VSMOW) compositions by
laser fluorination of 2- to 3-mg samples from Big Gossan
(Table 1) provided average values of 6.1 per mil for garnet
(range, 5.57.2), 7.4 per mil for pyroxene (7.08.2), and
8.6 per mil for amphibole (8.68.7), excluding one sample
with calcite inclusions that caused slightly higher 18 O values
(10.911.1). No systematic spatial variation in isotopic
compositions was determined for any of the mineral species.
There is an average pyroxene-garnet difference of 1.3 per mil.
If this difference reflects a bulk equilibrium signature of the
system, a calculated fractionation temperature (Zheng, 1993)
would be 370 70C (Table 1). This is in general agreement
with measured fluid inclusion temperatures, with all but one
value ranging from 400 to 455C (Table 1). Conventional
analysis of D via Zn reduction of 100- to 200-mg samples of
amphibole resulted in average values of 77.5 per mil (range,
77 to 78); no lower values were detected.
For Mines Gasp samples (Table 1), the 18O values average 4.2 per mil for garnet (3.45.6), 5.3 per mil for pyroxene (4.36.0), and 4.9 per mil for amphibole (4.35.2).
Again, if the average pyroxene-garnet difference of 1.1 per
mil reflects a bulk equilibrium signature of the system, the
calculated fractionation temperature is 450 100C, also in
general agreement with independent temperature estimates
of 450 to 500C and slightly higher than the temperature estimates for Big Gossan. The D value of primary igneous amphibole at Mines Gasp is 87 per mil, whereas skarn amphibole record values from 78 per mil through 115 to 143 per
mil.
Discussion
Unlike contact metamorphism that is the result of the intrusion of relatively dry magmas (cf. Valley, 1986), world-class

151

152

MEINERT ET AL.
TABLE 1. Stable Isotope and Temperature Data for the Big Gossan and Mines Gasp Skarn Deposits
Big Gossan

Sample no.

Mineral

1-6-600
1-6-600
1-6-600
8-1-213
8-1-213

Actinolite
Actinolite, duplicate
Actinolite, duplicate
Actinolite#
Actinolite #, duplicate

9-5-340
10-3-78
23-2-122
23-4-420
14-6-400

Pyroxene
Pyroxene
Pyroxene
Pyroxene
Pyroxene

26-4-250
14-6-491
23-2-165
9-5-465
8-5-187

Garnet
Garnet
Garnet
Garnet
Garnet

Measured
18O(VSMOW)
()

Calculated
18Ofluid
()

Measured
D(SMOW)
()

Calculated
Dfluid
()

Measured
H2O
(wt %)

Flinc
T
(C)

8.6
8.6
8.7
11.1
10.9

9.6
9.6
9.7
12.1
11.9

78

24

2.6

350

77

23

1.9

350
350
350

7.0
7.0
7.3
7.5
8.2

8.8
8.7
9.0
9.2
10.1

455
445
445
445
490

5.5
5.6
5.7
6.2
7.2
pyx-gar = 1.3

8.1
8.2
8.3
8.8
9.8

400
400
400
400
400
370 70

Gasp
Porphyry Mtn
923-4859
900-5660-1
900-5660-2
900-5660-2
793-2990

Hornblende
Hornblende
Actinolite
Actinolite
Actinolite*
Actinolite

5.0
4.3
5.2
5.2
4.9
5.0

7.4
5.3
6.2
6.2
5.9
6.0

87
78
115
143

70
24
61
89

2.5
1.8
1.6
1.6

130

76

1.8

700
350
350
350
350
350

923-5008
904-4011
904-4011
890-4085

Pyroxene
Pyroxene
Pyroxene, duplicate
Pyroxene

4.3
5.4
5.6
6.0

6.2
7.3
7.5
7.9

500
500
500
500

909-4245B
905-3210
904-3995
886-4351
920-5012
909-4245A

Garnet
Garnet
Garnet
Garnet
Garnet
Garnet

3.4
3.6
3.6
4.3
4.5
5.6
pyx-gar = 1.1

6.2
6.4
6.4
7.1
7.3
8.4

450
450
450
450
450
450
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 Gasp 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 >1,000C, reproducible to 0.1 wt % H2O;
this does notinclude absorbed water, which was removed by outgassing under vacuum pumping for 12 h at ~150C; Flinc = fluid inclusion, # = calcite
inclusions, * = HCl leached

skarn ore deposits form as the result of wall-rock interaction


with huge amounts of hydrothermal fluid. Such fluid has an
origin not only from the immediate dike or stock intrusion but
also comes from a parent magma chamber at a depth below
the ore deposit. This generalization of a magmatic fluid
source is accepted, at least during the prograde stage (Einaudi et al., 1981). As a result of the fluid domination of the
system, the original host rock in the immediate vicinity of the
intrusive focus of hydrothermal fluid flow will soon have little
0361-0128/98/000/000-00 $6.00

effect on the isotopic mass balance of the skarn minerals.


Thus, these skarn minerals should indicate the composition of
the hydrothermal fluid, at least those in a proximal position to
the intrusion.
The fluid associated with prograde skarn formation at both
Big Gossan and Mines Gasp was a high- temperature
(350500C), hypersaline liquid (up to 5060 wt % NaCl +
KCl). The calculated 18O value of the fluid in equilibrium
with garnet and pyroxene from both deposits is within the

152

FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS

range of values for magmatic fluids (Taylor and Sheppard,


1986), and even minerals with minor calcite inclusions deviate only slightly toward marine carbonate values of 21 to 25
per mil (Fig. 6). The D values of amphibole are also typical
of the range for igneous minerals (Taylor and Sheppard,
1986), although they deviate to relatively light compositions.
The fluid associated with the retrograde alteration stage is
lower temperature (~350400C) and lower salinity (avg 7
wt % NaCl equiv) than that of the prograde skarn fluid. Calculated 18O and D fluid compositions from Big Gossan amphibole (Table 1) are in the range typical of magmatic fluid
dissolved in a felsic melt (Taylor, 1986), with no clear trend
toward measured values for local meteoric water (Fig. 7). The
amphibole samples from Mines Gasp record a wider range
of D fluid compositions, from 24 to 89 per mil. This wide
range may record the degassing of the parent magma chamber, as this process results in fractionation such that the residual water in the melt has lighter D values (Suzuki and Epstein, 1976). As a result, the D values of fluid exsolved at a
later stage will be progressively lower in D whereas the 18O
values remain relatively constant (Taylor, 1986, 1988; Bowman, 1998; Hedenquist et al., 1998).
Interaction of the fluid with wall rock cannot account for
this large variation in D values where the 18O values remain
constant, except possibly at unreasonably low water-to-rock
ratios that may characterize only the very earliest stage of
-5

10

15

20

25

Big Gossan
Garnet

Amphibole
Amphibole
with calcite
inclusions

Gasp
Garnet
Pyroxene
Amphibole

Marine Carbonate

Pyroxene

Igneous
amphibole
Magmatic
fluid
-5

10

15

20

25

18 O (0/ 00, VSMOW)


FIG. 6. Calculated 18O (VSMOW) composition of fluids in equilibrium
with garnet, pyroxene, and amphibole from the Big Gossan and Mines Gasp
skarn deposits. See Table 1 for data and sources of fractionation factors. The
average composition of magmatic fluid and the marine carbonate values are
from Taylor (1986) and references therein.
0361-0128/98/000/000-00 $6.00

153

alteration, i.e., ratios <0.01 (Bowman, 1998). A small, <10


percent, meteoric water component could account for slight
variations in the range of 18O composition, <2.5 per mil at
each deposit; however, such a small diluent could not be responsible for the greater than six-fold decrease in salinity
from the prograde hypersaline fluid to 7 wt percent NaCl
fluid during the retrograde stage.
In some cases a large variation in salinity of the exsolved
fluid can occur according to the degree of crystallization, but
this is critically dependent on the pressure (Cline, 1995). We
argue that the porphyry situation modeled by Shinohara and
Hedenquist (1997), in which there is a relatively small variation in the salinity of the bulk fluid, may apply here during
much of the crystallization and exsolution. By contrast, the
trend in the D composition of residual water during crystallization will always be toward lower values (Suzuki and
Epstein, 1976). Indeed, the residual water fixed by primary
hydrous minerals, with intrusions typically having concentrations of 0.5 to 1 wt percent water, are observed to have D
values <100 per mil as the result of late open-system degassing (Taylor, 1986, 1988; Hedenquist and Richards, 1998).
This observation suggests that the formation of retrograde
alteration minerals in general, and amphibole- epidote in particular, can result from evolution of a single magmatic hydrothermal system, since there is no evidence in the Big Gossan 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 percent NaCl, can be explained if the later fluid followed a cooling path where it never intersected its solvus (Fig. 8). In this
situation, a hypersaline liquid would not form, and the salinity of the fluid that reaches the ore deposit is that of the bulk
salinity of the exsolved magmatic fluid (Meinert et al., 1997;
Shinohara and Hedenquist, 1997; Hedenquist et al., 1998).
Such a scenario would result in a zonation in both space and
time not only of the skarn mineralogy, as described earlier,
but more fundamentally of the fluid from which those minerals precipitated (Fig. 8).
Model of Evolution
Upon initial intrusion of the parent magma to shallow
depths the temperature of both magma and immediately
adjacent wall rocks was >400C, the normal limit of brittle
behavior. Thus, the rocks behaved in a ductile fashion, sealing
the system from significant interaction with connate and meteoric waters (Fournier, 1992). As the pluton cooled and crystallized at depths of, say, 4 to 5 km (Fig. 8), the melt eventually saturated with respect to an aqueous fluid, which was a
supercritical fluid with a moderate salinity of 6 to 8 wt percent
(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 600C (Burnham and
Ohmoto, 1980; Fournier, 1987). Subsequent fluid ascent to
shallower depths probably was concurrent with intrusion of
porphyry dikes to approximately 2 to 3 km beneath the paleosurface, with the dikes providing the permeability to focus the
advective ascent of exsolved magmatic fluid.
The combination of high fluid flux and thermal mass of the
dikes kept the early fluid relatively hot during ascent, with
only minor conductive cooling (Shinohara and Hedenquist,

153

154

MEINERT ET AL.

25

-25

Gaspe igneous
amphibole

-50

Degassing

D (0/ 00, VSMOW)

Initial water
dissolved in melt

Big Gossan skarn


amphibole
Gaspe skarn
amphibole

-75
-100
-125

Present-day
Ertsberg meteoric
water, snow, & ice

-150
-175
-25

-20

-15

-10

-5

10

15

18 O (0/ 00, VSMOW)


FIG. 7. Calculated 18O versus D compositions of fluids in equilibrium with amphibole (Table 1) from the Big Gossan
and Mines Gasp skarn deposits. The field for water initially dissolved in silicate melt is from Taylor (1992). As water exsolves
from magma, the D of residual water remaining in the melt becomes lighter (Suzuoki and Epstein, 1976). Therefore, the
D 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

TC

600

800

Na:K = 1:1

Vapor
+
Liquid

Liquid
2

Skarn
Retrograde
alteration with
qtz-anhydrite,
amphibole,
and sulfides

0.5

10

%
wt

60

7
80

0.2

1.0
60
wt%

Pressure (MPa)

40

1 1
0.

80

100

t%
40 w

20
NaCl in liquid

Depth (lithostatic)

Vapor + Salt
Depth (hydrostatic)

20

6-8 wt%
NaCl

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

154

FIG. 8. Composition of coexisting liquid and vapor as a function


of depth and temperature (after Fournier, 1987). Isopleths of
NaCl in liquid (dark lines) and vapor (light lines) are shown for
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 500C
and a pressure of 50 MPa (lithostatic pressure at a depth of about
2 km). Continued depressurization results in the separation of a
small amount of hypersaline liquid with a low-salinity vapor dominating the mass of the two-phase fluid. The later liquid that ascends along the retrograde trajectory intersects its saturated
vapor-pressure curve at about 370C and a hydrostatic pressure of
20 MPa, also equivalent to a depth of about 2 km. At this point the
fluid begins to boil, with a small amount of vapor separating to
leave the residual liquid slightly higher in salinity.

155

FORMATION OF Cu-Au SKARN BY MAGMATIC FLUIDS

1997). The highest temperature and lithostatic pressure deduced for the formation of these deposits during their prograde stage is 535C and 50 MPa, respectively (Meinert et al.,
1997). This is consistent with the steep isotherms around a
near-solidus magma that is exsolving an aqueous fluid (Shinohara and Hedenquist, 1997). A magmatic fluid with bulk
salinity of 7 wt percent NaCl will intersect its solvus at a 2-km
depth and separate into a 50 wt percent hypersaline liquid in
equilibrium with a vapor containing ~1 wt percent NaCl (Fig.
8). Based on the fluid inclusion evidence for temperature and
salinity, most prograde skarn formed from this hypersaline
liquid.
Upon aqueous phase separation, most of the SO2 and HCl
from the parent fluid will partition into the vapor phase (Candela and Piccoli, 1995; Scaillet et al., 1998; Keppler, 1999),
which being less dense forms a vapor plume overlying the
reservoir of hypersaline liquid that is generated (Henley and
McNabb, 1978). This creation of a hypersaline liquid with
high base metal solubilities and relatively low sulfur content
may have inhibited sulfide deposition during formation of the
prograde skarn, similar to Bodnars (1995) argument for porphyry deposits. Condensation of much of the vapor plume at
shallow depths would have formed a highly acidic, low-salinity liquid which most likely would have caused advanced
argillic alteration (Hedenquist, 1995) and dissolution of the
host carbonate rocks with attendant brecciation, again analogous to formation of volcanic-hosted lithocaps over porphyry
deposits (Hedenquist et al., 1998). This can be thought of as
the ground-preparation stage of a growing skarn system in
which the vapor plume produces cavernous porosity in carbonate rock, to be overrun subsequently by the high- salinity,
skarn-forming fluid as the system expands. Thus, the many
breccias that occur at both Big Gossan and Mines Gasp, as
well as at other skarn deposits in the porphyry environment
(Einaudi, 1982), are part of the normal early-stage hydrothermal evolution of such systems.
As the underlying vapor-saturated magma chamber continued to cool and crystallize, fluid continually exsolved, and this
later fluid may have been relatively similar in bulk Na-K-Cl
composition to that of the earlier stage fluid (Cline, 1995; Shinohara and Hedenquist, 1997). However, because of the lack
of concurrent dike intrusion and/or a sharply lower fluid flux
as the parent magma approached the end stages of stagnant
crystallization (Shinohara and Hedenquist, 1997), the ascending fluid followed a cooling path different from that of the
early stage. The lower rate of advection (Shinohara and
Hedenquist, 1997) meant that this late fluid cooled sufficiently during ascent so that it never intersected its solvus
(Fig. 8).
Cooling to a temperature below approximately 400C results
in a transition from ductile to brittle rock behavior, causing a
change from lithostatic to hydrostatic conditions (Fournier,
1991). Thus, even though the later stage fluid reached the
previously formed skarn at the same paleodepth of 2 km, the
pressure was 20 MPa under the hydrostatic conditions. A 7 wt
percent NaCl liquid at 20 MPa will intersect the saturated
vapor-pressure curve at 370C, resulting in boiling of the ascending liquid. Continued boiling and vapor loss will result in
progressive cooling accompanied by only a slight increase in
salinity due to vapor loss, e.g., from 7 to 10 wt percent NaCl
0361-0128/98/000/000-00 $6.00

on boiling from 370 to 320C (Henley et al., 1984). At this


shallow depth and relatively low temperature, dissolved gases
will be lost to the vapor, possibly explaining sulfide mineralization from the low-salinity fluid that occurs during this late
stage.
This general scenario (Fig. 8) illustrates only two snapshots of a dynamic system. In reality, multiple intrusions
into the parent chamber and as shallow dikescombined
with fluctuations in the water table due to structural, paleosurface, and/or climatic changes (e.g., Simmons, 1991)
would have caused a more complex overprinting of various
skarn and retrograde alteration stages. These changes are
seen in the rock record as a series of crosscutting features
and zonation patterns that are typical of skarn and intrusioncentered ore deposits worldwide (Einaudi et al., 1981; Bendez and Fontbot, 2002; Rusk and Reed, 2002). Nevertheless, despite this complexity we conclude that the fluid
responsible for prograde and retrograde alteration as well as
ore formation was dominantly magmatic in origin. Similar
conclusions have been reached for porphyry copper deposits which formed largely within igneous rocks (e.g.,
Kusakabe et al., 1990; Shinohara and Hedenquist, 1997;
Hedenquist et al., 1998; Watanabe and Hedenquist, 2001;
Harris and Golding, 2002). Here we extend such conclusions to skarn deposits in which hydrothermal fluids have escaped the igneous carapace and interacted extensively with
surrounding wall rocks.
Acknowledgments
This work was partially supported by National Science
Foundation grant EAR-9725437. The senior authors visit to
the Geological Survey of Japan was facilitated by a Japan Society for Promotion of Science short-term invitation fellowship, administered by the National Science Foundation. We
thank Steve Rowins, John Bowman, and John Dilles for helpful comments on earlier versions of this paper.
April 5, October 17, 2002
REFERENCES
Allcock, J.B., 1982, Skarn and porphyry copper mineralization at Mines
Gasp, Murdochville, Qubec: ECONOMIC GEOLOGY, v. 77, p. 971999.
Bendez, R., and Fontbot, L., 2002, Late timing for high sulfidation
Cordilleran base metal lode and replacement deposits in porphyry-related
districts: The case of Colquijirca, central Peru: Society for Geology Applied
to Mineral Deposits News, no. 13, p. 2, 913.
Bigeleisen, J., Perlman, M.L., and Prosser, H.C., 1952, Conversion of hydrogen materials to hydrogen for isotopic analysis: Analytical Chemistry, v. 24,
p. 13561357.
Bodnar, R.J., 1995, Fluid-inclusion evidence for a magmatic source for metals in porphyry copper deposits: Mineralogical Association of Canada Short
Course Series, v. 23, p. 139152.
Bottinga, Y., and Javoy, M., 1975, Oxygen isotope partitioning among the
minerals in igneous and metamorphic rocks: Reviews of Geophysics and
Space Physics, v. 13, p. 401418.
Bowman, J.R., 1998, Stable-isotope systematics of skarns: Mineralogical Association of Cananda Short Course, v. 26, p. 99145.
Burnham, C.W., 1979, Magmas and hydrothermal fluids, in Barnes H.L., ed.,
Geochemistry of hydrothermal ore deposits, 2nd edition: New York, John
Wiley and Sons, p. 71136
Burnham, C.W., and Ohmoto, H., 1980, Late-stage processes of felsic magmatism; Society of Mining Geologists of Japan Special Issue 8, p. 111
Candela, P.A., and Piccoli, P.M., 1995, Model of ore-metal partitioning from
melts into vapor and vapor/brine mixtures: Mineralogical Association of
Canada Short Course, v. 23, p. 101128.

155

156

MEINERT ET AL.

Cline, J.S., 1995, Genesis of porphyry copper deposts: The behavior of water,
chloride, and copper in crystallizing melts: Arizona Geological Society Digest, v. 20, p. 6982.
Einaudi, M.T., 1982, Descriptions of skarn associated with porphyry copper
plutons, southwestern North America, in Titley, S.R., ed., Advances in geology of the porphyry copper deposits, southwestern North America: Tucson, University of Arizona Press, p. 139184.
Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits:
ECONOMIC GEOLOGY 75TH ANNIVERSARY VOLUME, p. 317391.
Fournier, R.O., 1987, Conceptual models of brine evolution in magmatic-hydrothermal systems: U.S. Geological Survey Professional Paper 1350, p.
14871506.
1991, The transition from hydrostatic to greater than hydrostatic fluid
pressure in presently active continental hydrothermal systems in crystalline
rock: Geophysical Research Letters, v. 18, p. 955958.
1992, The influences of depth of burial and the brittle-ductile transition
on the evolution of magmatic fluids: Geological Survey of Japan Report
277, p. 5759.
Freeport McMoRan, 2001, Annual Report: New Orleans, Louisiana, 70 p.
Giggenbach, W.F., 1997, The origin and evolution of fluids in magmatic-hydrothermal systems, in Barnes, H.L., ed., Geochemistry of hydrothermal
ore deposits, 3rd edition: New York, John Wiley and Sons, p. 737796.
Gower, S.J., and Walker, J.A., 1993, Skarn-type, base-metal deposits in northern N.B. Geology and skarn occurrences: Geological Society of the CIM,
Annual Field Conference, 3rd, 1993, Bathurst, Trip 1 Guidebook, p. 521.
Harris, A.C., and Golding, S.D., 2002, New evidence of magmatic-fluid-related phyllic alteration: Implications for the genesis of porphyry Cu deposits: Geology, v. 30, p. 335338.
Harrison, J., Kyle, J.R., Pennington, J., and Kavalieris, I., 1999, Hydrothermal alteration and fluid evolution of the Grasberg Cu-Au deposit, Irian
Jaya, Indonesia [abs.]: Geological Society of America, Abstracts with Programs, v. 31, p. A403.
Hedenquist, J.W., 1995, The ascent of magmatic fluid: Discharge versus mineralization: Mineralogical Association of Canada Short Course, v. 23, p.
263290.
Hedenquist, J.W., and Lowenstern, J. B., 1994, The role of magmas in the
formation of hydrothermal ore deposits: Nature, v. 370, p. 519527.
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., 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.
Henley, R.W., and McNabb, A., 1978, Magmatic vapor plumes and groundwater interaction in porphyry copper emplacement: ECONOMIC GEOLOGY,
v. 73, p. 120.
Henley, R.W., Truesdell, A.H., Barton, P.B., Jr., and Whitney, J.A., 1984,
Fluid-mineral equilibria in hydrothermal systems: Reviews in Economic
Geology, v. 1, 267 p.
Hussey, J., and Bernard, P., 1998, Exploration of the Porphyry Mountain CuMo deposit: Mining Engineering, v. 50, p. 3644.
Keppler, H., 1999, Experimental evidence for the source of excess sulfur in
explosive volcanic eruptions: Science, v. 284, p. 16521654.
Kusakabe, M., Hori, M., and Matsuhisa, Y., 1990, Primary mineralization of
the El Teniente and Rio Blanco porphyry copper deposits, Chile: Stable
isotopes, fluid inclusions, and Mg2+/Fe2+/Fe3+ ratios of hydrothermal biotite: University of Western Australia, Geology Department Publication 23,
p. 244259.
Kwak, T.A.P., 1986, Fluid inclusions in skarns (carbonate replacement deposits): Journal of Metamorphic Geology, v. 4, p. 363384.
McMahon, T.P., 1994a, Pliocene intrusions in the Ertsberg (Gunung Bijih)
mining district, Irian Jaya, Indonesia: Petrography, geochemistry, tectonic
setting: Unpublished Ph.D. dissertation, University of Texas, Austin, 299 p.
1994b, Pliocene Cu-Au-bearing igneous intrusions of the Gunung Bijih
(Ertsberg) district, Irian Jaya, Indonesia: Petrography and mineral chemistry: International Geology Review, v. 36, p. 820849.
1994c, Pliocene Cu-Au-bearing igneous intrusions of the Gunung Bijih
(Ertsberg) district, Irian Jaya, Indonesia: Major and trace element chemistry: International Geology Review, v. 36, p. 925946.

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

Meinert, L.D., 1992, Skarns and skarn deposits: Geoscience Canada, v. 19, p.
145162.
1997, Application of skarn deposit zonation models to mineral exploration: Exploration and Mining Geology, v. 6, p. 185208.
Meinert, L.D., Hefton, K.K., Mayes, D., and Tasiran, I., 1997, Geology, zonation, and fluid evolution of the Big Gossan Cu-Au skarn deposit, Ertsberg
district, Irian Jaya: ECONOMIC GEOLOGY, v. 92, p. 509526.
Pettke, T., and Diamond, L.W., 1997, Oligocene gold quartz veins at Brusson, NW Alps: Sr isotopes trace the source of ore-bearing fluid to over a 10km depth: ECONOMIC GEOLOGY, v. 92, p. 389406.
Rusk, B., and Reed, M., 2002, Scanning electron microscope-cathodoluminescence analysis of quartz reveals complex growth histories in veins from
the Butte porphyry copper deposit, Montana: Geology, v. 30, p. 727730.
Scaillet, B., Clemente, B., Evans, B.W., and Pichavant, M., 1998, Redox control of sulfur degassing in silicic magmas: Journal of Geophysical Research,
v. 103, p. 23,93723,949.
Shelton, K.L., 1983, Composition and origin of ore-forming fluids in a carbonate-hosted porphyry copper and skarn deposit: A fluid inclusion and
stable isotope study of Mines Gasp, Qubec: ECONOMIC GEOLOGY, v. 78,
p. 387 421.
Shimazaki, H., and Kusakabe, M., 1990, Oxygen isotope study of the
Kamioka Zn-Pb skarn deposits, central Japan: Mineralium Deposita, v. 25,
p. 221229.
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.
Simmons, S.F., 1991, Hydrologic implications of alteration and fluid inclusion studies in the Fresnillo district, Mexico: Evidence for a brine reservoir
and a descending water table during the formation of hydrothermal Ag-PbZn orebodies: ECONOMIC GEOLOGY, v. 86, p. 15791601.
Stephenson, E.M., Meinert, L.D., Mortensen, J.K., and Hussey, J., 1998, Age
and skarn alteration of the Porphyry Mountain Cu-Mo deposit, Mines
Gasp, Qubec [abs.]: Geological Society of America Abstracts with Programs, v. 30, p. A-372.
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.
1988, Degassing of rhyolitic magmas: Hydrogen isotope evidence and
implications for magmatic- hydrothermal ore deposits: Canadian Institute
of Mining and Mineralogy Special Volume 39, p. 3349.
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 279, p. 190194.
Taylor, H.P., Jr., 1997, Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits, in Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd edition: New York, John Wiley and Sons, p.
229302.
Taylor, H.P., Jr., and Sheppard, S.M.F., 1986, Igneous rocks: I. Processes of
isotopic fractionation and isotope systematics: Reviews in Mineralogy, v. 16,
p. 227271.
Valley, J.W., 1986, Stable isotope geochemistry of metamorphic rocks: Reviews in Mineralogy, v. 16, p. 445489.
Watanabe, Y., and Hedenquist, J.W., 2001, Mineralogic and stable isotope
zonation at the surface over the El Salvador porphyry copper deposit,
Chile: ECONOMIC GEOLOGY, v. 96, p. 17751797.
Williams-Jones, A.E., 1986, Low-temperature metamorphism of the rocks
surrounding les Mines Gasp, Quebec: Implications for mineral exploration: ECONOMIC GEOLOGY, v. 81, p. 466470.
Yang, K., and Bodnar, R.J., 1994, Magmatic-hydrothermal evolution in the
Bottoms of porphyry copper systems: Evidence from silicate melt and
aqueous fluid inclusions in granitoid intrusions in the Gyeongsang basin,
South Korea: International Geology Review, v. 36, p. 608628.
Zheng, Y.F., 1993, Calculation of oxygen isotope fractionation in anhydrous
silicate minerals: Geochimica et Cosmochimica Acta, v. 57, p. 10791091.

156

Das könnte Ihnen auch gefallen