Sie sind auf Seite 1von 26

2009 Society of Economic Geologists, Inc.

Economic Geology, v. 104, pp. 223248

Stable Isotope Constraints on Ore Formation at the San Rafael


Tin-Copper Deposit, Southeast Peru
THOMAS WAGNER,
Institute of Isotope Geochemistry and Mineral Resources, ETH Zurich, NW F 82.4, Clausiusstrasse 25, CH-8092 Zrich, Switzerland

MICHAEL S. J. MLYNARCZYK,
Rathdowney Resources Ltd., Killaderry, Daingean, County Offaly, Ireland

ANTHONY E. WILLIAMS-JONES,
Department of Earth and Planetary Sciences, McGill University, 3450 University Street, Montral QC H3A 2A7, Canada

AND ADRIAN J. BOYCE


Scottish Universities Environmental Research Centre (SUERC), East Kilbride, Glasgow, G75 0QF, Scotland, United Kingdom

Abstract
The San Rafael tin-copper deposit in the Eastern Cordillera of the Peruvian Central Andes is the worlds
largest hydrothermal tin lode, with a total resource of about 1 million metric tons metal, at an average grade of
4.7 wt percent Sn. The mineralization is of the cassiterite-sulfide type and occurs in a vertically extensive vein-
breccia system centered on a shallow-level, late Oligocene granitoid stock. The tin ores form cassiterite-quartz-
chloritebearing veins and breccias hosted by several large fault-jogs at depth in the lode. By contrast, the cop-
per ores, which contain disseminated acicular cassiterite, are localized in the upper part of the system. Both
ore types are associated with a very distinctive, strong chloritic alteration, which was preceded by intense seric-
itization, tourmalinization, and tourmaline veining. The 34S values of the sulfides range between 2 and 6 per
mil, and vary very little with location in the deposit. This indicates that the hydrothermal system was large, with
a relatively homogeneous source of sulfur, likely of magmatic origin. This is confirmed by stability relationships
of ore minerals, which indicate that the ore fluids were initially reducing. Microthermometric studies of fluid
inclusions in cassiterite, quartz, tourmaline, and fluorite show that the fluids responsible for the early, barren
stage were hot, hypersaline brines (380540C, 3462% NaCl equiv), whereas the ore-stage fluids had mod-
erate to low salinity (021 wt % NaCl equiv), and were of moderate temperature (290380C). In addition to
the marked dilution of the ore fluids with evolution of the hydrothermal system, they became progressively
more oxidizing, as inferred by the local association of minor hematite with cassiterite and the ubiquitous re-
placement of pyrrhotite by pyrite and marcasite. The 18O values of the fluid decreased systematically with
time, as indicated by the 18O values of different generations of tourmaline, cassiterite, and quartz. This evo-
lution was paralleled by an increase in the D values of the fluid, inferred from the D values of tourmaline
and chlorite. This trend is consistent with mixing of the ore fluids with a cooler fluid that had substantially lower
18O, and cannot be explained by fluid boiling. Based on structural evidence for an opening of the vein system
and a transition from lithostatic to hydrostatic conditions at the onset of mineralization, we infer that ore de-
position was caused by an influx of hot groundwater of meteoric origin which mixed repeatedly with tin-bear-
ing magmatic brines. The oxidation, dilution, cooling, and acid neutralization of the ore fluids destabilized chlo-
ride complexes of tin and triggered the large-scale precipitation of cassiterite.

Introduction aqueous phase, the leaching of crystallized magmatic miner-


HYDROTHERMAL TIN LODES represent a major class of granite- als by later fluids may be another significant metal source
related tin deposits characterized by very high grade (about (Taylor, 1979; Lehmann, 1990). In addition, some metal could
15 wt % Sn) and high tonnage, which makes them an ideal have been derived from the sequences of metasedimentary
target in the exploration for economically viable primary tin rocks, which host the granitic plutons (Williamson et al.,
resources (Taylor, 1979; Menzie et al., 1988). However, de- 2000; LeBoutillier et al., 2002). A variety of ore fluids have
spite their having been mined since antiquity and being well been invoked, with emphasis on primary magmatic fluids, but
described in the literature, there is still considerable uncer- also including meteoric waters, basinal brines (formation wa-
tainty about key issues related to the genesis of these deposits, ters), metamorphic fluids, and their mixtures. With regard to
such as the source of the tin, the origin of the ore fluids, and the meteoric fluids, the timing of their circulation and inter-
the mechanism(s) of ore deposition. Although it is generally action with magmatic fluids may be relatively early or occur
believed that the bulk of the metal is derived directly from an only in the waning stages of the hydrothermal system
evolved granitic melt by being partitioned into a magmatic (Collins, 1981; Primmer, 1985; Sun and Eadington, 1987;
Alderton and Harmon, 1991; Sheppard, 1994; Linnen and
Corresponding author: e-mail: thomas.wagner@erdw.ethz.ch Williams-Jones, 1995; Wilkinson et al., 1995; Smith et al.,

0361-0128/09/3809/223-26 223
224 WAGNER ET AL.

1996; Walshe et al., 1996; Jackson et al., 2000). Finally, the 2002) and, therefore, was not affected by any later geologic
following mechanisms have been proposed for the precipita- processes. Following a study of the deposit geology (Mlynar-
tion of cassiterite: cooling, pressure decrease, wall-rock alter- czyk et al., 2003) and its alteration (Mlynarczyk and Williams-
ation, boiling, and fluid mixing (Kelly and Turneaure, 1970; Jones, 2006; Mlynarczyk et al., 2009, submitted), a compre-
Eadington, 1985; Heinrich and Eadington, 1986; Heinrich, hensive stable isotope (S, O, H, C) investigation of ore and
1990, 1995; Linnen and Williams-Jones, 1995). Among these, gangue minerals and a preliminary fluid inclusion study were
mixing between magmatic and meteoric waters is thought to performed to evaluate the source of the fluids and metals, as
be the most effective means of producing the unusually high well as the timing and the mechanism of ore deposition. Our
ore grades of some deposits (Heinrich, 1990). However, this data indicate that ore deposition was caused by an influx of
hypothesis has not been properly tested because the contri- hot groundwaters of meteoric origin and their mixing with the
bution of meteoric water to hydrothermal systems depositing tin-bearing magmatic brines.
lode tin mineralization has been poorly documented.
San Rafael, the worlds richest primary Sn-Cu deposit Geologic Setting
(Kontak and Clark, 2002; Mlynarczyk et al., 2003), which is
located in the high Andes of the Eastern Cordillera of south- Regional geology
eastern Peru and is currently exploited by MINSUR S.A., is The San Rafael deposit lies in the northwesternmost exten-
an ideal natural laboratory in which to investigate the various sion of the Central Andean Sn-W(-Ag) metallogenic province
issues of lode tin genesis raised above. With a total resource (e.g., Turneaure, 1960a, b; Kelly and Turneaure, 1970;
of about 1 million metric tons (Mt) Sn, and an average grade McBride et al., 1983; Mlynarczyk and Williams-Jones, 2005)
of 4.7 wt percent Sn, the San Rafael lode, which has been ex- and crops out on the glaciated slopes of the Cordillera
posed by mining activity over a vertical extent of more than Carabaya, which forms part of the Eastern Cordillera of
1,300 m, allows an unprecedented opportunity to advance our southern Peru (Fig. 1). The geology of the region is domi-
understanding of the formation of this important deposit nated by a thick sequence of Lower Paleozoic, marine clastic
type. In addition, the deposit is very young (about 2224 Ma; metasedimentary rocks, which experienced extensive crustal
Clark et al., 1983; Kontak et al., 1987; Kontak and Clark, deformation in the Cenozoic. The unexposed Precambrian

N
PERU
PACIFIC OCEAN

Lima
VIC

SAN RAFAEL
MA
TO

RIA
R IA

JO
NO
MARIA

RG
E

VI GU
M CE IL
AR LE
ELEN

IA NT
E RM
NE SA O
LA
N
A

RA
FA
DES S E
ANUANO C
P E R AT L
A
LI
N
A
PE
DR
O

500 m

Meta-sediments (Upper Ordovician): Granitoids (Late Oligocene): Structures:


Phyllites (Sanda Fm.) Coarse-grained granitoids Mineralized veins
Quartzites (Sanda Fm.) Fine-grained granitoids Major faults
Slates and hornfelses (Sanda Fm.) Dyke of granitoid porphyry
Tourmalinized granite Access to the mine

FIG. 1. Geologic map of the western part of the San Rafael district (modified after Arenas, 1980).

0361-0128/98/000/000-00 $6.00 224


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 225

gneissic basement is covered by a 10- to 15-km-thick se- age) to 24.7 0.2 Ma (U-Pb monazite age). Ar/Ar data show
quence of Ordovician to Siluro-Devonian metapelites and that the upper part of the stock had cooled down to the clo-
metapsammites (the San Jos, Sanda, and Ananea Forma- sure temperature of biotite (about 300350C: Harrison et
tions), followed by about 3 to 4 km of late Paleozoic psam- al., 1985) at about 23.7 0.2 Ma. The main Sn-Cu mineral-
mites and carbonates (Mississippian Ambo Group, Pennsyl- ization appears to be slightly younger than this cooling age,
vanian Tarma Group, and Permian Copacabana Group) and 3 based on Ar/Ar ages ranging between 22.7 0.7 and 21.9
km of mid-Permian to Triassic red beds and intercalated vol- 0.2 Ma, obtained from hydrothermal adularia and muscovite
canics (Mitu Group). These rocks are, in turn, overlain by (Clark et al., 1983; Kontak et al., 1987; Kontak and Clark,
roughly 1 km of Cretaceous psammites and carbonates (Co- 2002).
tacucho Group) and 800 m of Miocene-Pliocene felsic ign- The San Rafael lode crops out on the eastern flanks of the
imbrites and red beds (Crucero Supergroup). A detailed re- glacier-capped Quenamari mountain, at an altitude between
view of the geology of the area is provided by Laubacher 4,500 and 5,100 m, and has a length of over 3 km. It has been
(1978), Clark et al. (1990), and Sandeman et al. (1996). exposed by mining over a vertical extent of about 1,300 m
Granitic and granodioritic plutons (some of which host signif- (from 5,100 to 3,800 m above sea level). Current mining ac-
icant Sn-W-base metal mineralization) were emplaced be- tivity is restricted to the lower 500 m of the deposit. The
tween the late Devonian to early Carboniferous (Kontak and upper part of the lode is hosted by metasedimentary rocks,
Clark, 1988) and the late Tertiary (Kontak et al., 1987; Kon- attains a width of up to 2 m, and contains appreciable chal-
tak and Clark, 2002). Examples of these are the large Permo- copyrite and subordinate needle-tin cassiterite. By contrast,
Triassic granitic batholiths, which host the Sarita polymetallic the lower part of the lode is hosted by granitic rocks, opens
deposit, and the middle- to late-Tertiary granitic stocks, into a series of subvertical shoots up to 50 m wide, and con-
which host the San Rafael tin-copper and Palca XI tungsten tains a large volume of high-grade cassiterite ore (Fig. 2).
deposits (Clark et al., 1983; Kontak and Clark, 1988; Kontak The ore shoots are composed of sets of veins that have thick-
et al., 1990). The peraluminous Tertiary granitic intrusions nesses in the range of 5 to 30 cm. Quartz and chlorite are the
are thought to originate from deep-crustal partial melting of principal gangue minerals, and are common in both the
metasediments related to magmatic underplating (Sandeman upper and lower parts of the deposit. The lode has an aver-
et al., 1995; Kontak and Clark, 2002). In addition to grani- age strike of 330, dips 40 to 75NE, and is texturally quite
toids, there are also subordinate peralkaline syenites and gab- complex, having undergone multiple episodes of vein re-
bro-diorites. However, most rocks of mafic composition are opening and brecciation. Mlynarczyk et al. (2003), who pro-
volcanic and form localized occurrences of shoshonite, vided a detailed description of the deposit geology, proposed
minette, absarokite, and high K calc-alkaline basalt (Kontak that the large ore shoots at depth in the lode represent fault
et al., 1986; Sandeman et al., 1995). jogs formed during sinistral-normal strike-slip movement
along the plane of the lode, which was synchronous with cir-
Deposit geology culation of the ore fluids.
The bulk of the tin in the San Rafael-Quenamari mining
district is hosted by a single, large vein-breccia system called Alteration and Vein Paragenesis
the San Rafael lode. This lode is part of an array of a few A detailed campaign of core-logging and underground
dozen subparallel veins that host Sn-(W)-Cu-Zn-Pb-Ag min- mapping in the San Rafael lode outlined as many as 15 dis-
eralization. Typically, the veins are planar, 500 to 3,500 m tinct vein types, which were arranged chronologically, based
long, have a northwestern strike and dip moderately to on crosscutting relationships (Mlynarczyk et al., 2009, sub-
steeply to the northeast. They are centered on a small (about mitted). The mineralogy of these veins and the associated
15 km2), high-level, late Oligocene granitoid pluton and strad- wall-rock alteration are summarized briefly below, within the
dle the contact with the surrounding Lower Paleozoic meta- framework of four principal paragenetic stages, adopted from
sedimentary rocks, which consist dominantly of Ordovician a subdivision originally proposed by Palma (1981).
phyllites and quartzites of the Sanda Formation (Fig. 1). The
pluton is a polyphase intrusion composed of coarse- to (I) Early, barren tourmaline stage
medium-grained K-feldspar megacrystic, cordierite-biotite- The early, barren stage was initiated by pervasive sericitic
(garnet-sillimanite) granite, leucogranite and granodiorite, alteration, followed by strong tourmalinization, which pro-
with local minor enclaves of tonalite and quartz diorite. The duced wide alteration halos around early tourmaline-quartz
magma was strongly peraluminous, S-type in character, and veins and tourmaline stringers. These early veins, which cut
clearly formed due to the partial melting of metasediments through large parts of the granitic pluton, are typically sealed
(Mlynarczyk et al., 2003). However, it was not geochemically and lack cassiterite. Locally, however, they host abundant ar-
evolved, and the different igneous phases do not represent a senopyrite. The tourmaline is mainly dravite and Fe-rich
fractional crystallization series (Dolejs et al., 2009, submit- dravite, but other varieties (foitite, Mg foitite) were also ob-
ted). The pluton is elongated to the northeast and corre- served. A detailed description of the mineralogy, chemistry,
sponds to a prominent topographic high (Nevado Quenamari, and temporal relationships of the different varieties of vein
5,300 m), but is only exposed locally, although its horizontal and alteration tourmaline can be found in Mlynarczyk and
extent is clearly evident from the distribution of the overlying Williams-Jones (2006).
contact metamorphic rocks (hornfelses and slates). The avail- In addition to the tourmaline-rich veins, laterally extensive
able geochronological data (Kontak and Clark, 2002) demon- (up to several meters wide) veins of hydrothermal micro-
strate that the pluton was emplaced at 24.6 0.2 (U-Pb zircon breccia are very common. These are composed of strongly

0361-0128/98/000/000-00 $6.00 225


226 WAGNER ET AL.

SE Nevado Quenamari NW

5000 m

R165
R709 R708

R725 R726 R729

R428
R429 4500 m
R137
R65 R28 R27 R113
R404 R511 R55 R57 R119
R573 R123
R399
Copper stopes D3-27
Tin orebodies D3-39 A13
D3-69 R17
500m Ore Shoot 4000 m
Contact Orebody a.s.l.

FIG. 2. Longitudinal section through the San Rafael lode, showing the location of hand samples collected during this
study. The location of drill core samples (elevations below about 4,300 m) is indicated in Tables 1 to 3.

tourmalinized rock flour and subordinate quartz, generally consists of cassiterite, chlorite, and quartz (Fig. 3a). Cassi-
occur in series, and contain the bulk of the tourmaline in the terite occurs mainly as massive, dark-brown, locally botryoidal
deposit. Commonly, they are crosscut by a dense swarm of layers (up to 7 cm wide) or as beige-colored, collomorphic
younger quartz veins and are associated with subordinate sili- wood tin. Chlorite is dark green and moderately to strongly
cification which, locally, could have been coeval with tourma- enriched in iron, corresponding compositionally to ripidolite
line precipitation. Both the tourmaline-rich veins and micro- and daphnite (Mlynarczyk et al., 2009, submitted). Locally,
breccia veins occur throughout the vertical extent of the lode, the cassiterite-chlorite-quartz veins also contain minor pro-
although they are more abundant in its central and upper portions of Fe-rich wolframite or arsenopyrite (Fig. 3b),
parts. Petrographic observations indicate that reopening of the which is the essential sulfide mineral associated with this
early tourmaline veins was quite common, and that there were stage. The larger, high-grade veins are complexly banded,
many episodes of brecciation. It is also apparent that some of with multiple generations of alternating quartz, cassiterite,
the major tourmaline-bearing veins, which clearly follow the and chlorite, that commonly record a consistent temporal se-
lode along strike and were responsible for establishing its ini- quence of mineralization that can be distinguished on the
tial structural framework, subsequently reopened during the basis of crystal morphology and color (Fig. 4). Some of these
ore stage (Mlynarczyk et al., 2003). The principal chemical veins have quartz-rich centers with a high proportion of sul-
changes associated with the strong sericite-tourmaline alter- fides (e.g., chalopyrite, arsenopyrite, pyrite) that were formed
ation were a nearly complete removal of alkalis and alkali- during the later sulfide stage, as deduced from crosscutting
earth elements (Na, K, Ca, Ba, Rb, Sr, Cs, Li, and Cl) from the vein relationships. Finally, traces of hematite are associated
wall rock and a marked addition of B, Mg, and Fe, and possi- with chlorite and cassiterite in the brecciated ores at about an
bly Sn, W, and In, though the latter could represent a late elevation of 4,400 m. No tourmaline occurs in the cassiterite-
overprint (Mlynarczyk et al., 2009, submitted). chlorite-quartz veins, but hairline veinlets of a late, Fe-rich
variety of dark-blue or dark-green tourmaline crosscut the
(II) Main cassiterite stage early tourmaline veins and microbreccias of the barren stage.
The initiation of the ore stage was related to a transition The ubiquitous cassiterite-chlorite-quartz veins are sur-
from a relatively closed to an open vein system, as all ore and rounded by wide halos of strong chloritic alteration and dis-
later vein types are characterized by textures consistent with play evidence of frequent reopening (crack-and-seal textures)
filling of open fractures. The earliest phase of the ore stage is and cockscomb textures, producing a characteristic recurrent
marked by a rare type of tourmaline-cassiterite-chlorite-ar- alternation of cassiterite, chlorite, and quartz layers.
senopyrite vein, in which an obvious but continuous transition The bulk of the tin ore is restricted to the lower, granite-
in mineral chemistry is observed. An early variety of orange- hosted part of the San Rafael lode, especially the wide, sub-
colored dravite was progressively replaced by green-colored vertical fault-jogs, which form elongated sinuous zones up to
schorl (a strongly Fe-rich tourmaline), and was accompanied 50 m wide. The highest ore grades (up to 45 wt % Sn) are as-
by simultaneous precipitation of cassiterite and chlorite, fol- sociated with several-meter-wide breccia zones, which either
lowed by massive arsenopyrite (Mlynarczyk and Williams- contain fragments of cassiterite veins cemented by chlorite
Jones, 2006). and quartz, or fragments of chloritized wall rock cemented
The majority of the veins and breccias of the main tin ore by cassiterite and quartz. Very high grade cassiterite miner-
stage, however, are characterized by a simple mineralogy that alization also occurs in the major (12 m wide) footwall and

0361-0128/98/000/000-00 $6.00 226


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 227

chl qz
apy
cp sp qz

tou cas
cas

cas
5.5cm chl 2.5cm
a b

sp cp

cp qz
qz

3.5cm 3.5cm
c d
FIG. 3. Photographs of representative ore hand specimens. a. Example of typical tin ore from a fault-jog in the lower part
of the deposit. Layers of massive, dark-brown cassiterite (cas) overgrown on fragments of early quartz (qz)-tourmaline (tou)
breccia and chloritized (chl) wall rock (sample SAR-R464, contact orebody, level 4350 m). b. Complex, open space-filling;
quartz-cassiterite-sulfide vein (the arrow indicates the direction of growth). Near the strongly chloritized margin of the vein
there are alternating layers of cassiterite, chlorite, and quartz. The center of the vein consists of large quartz crystals capped
by arsenopyrite (apy), with minor grains of sphalerite (sp) and chalcopyrite (cp) (sample SAR-R511, ore shoot, level 4310 m).
c. Typical copper ore from the upper part of the deposit, composed dominantly of chalcopyrite, with subordinate quartz and
pyrite (sample SAR-R710, San Rafael lode, level 4820 m). d. Broken fragments of sphalerite and chalcopyrite, cemented by
quartz, in the central part of a quartz-chalcopyrite vein (sample SAR-R722, San Rafael lode, level 4730 m).

hanging-wall veins of the lode structure, as well as in loca- (III) Sulfide stage
tions where veins are deflected, branch, or intersect. It The bulk of the sulfide ore is restricted to the middle and
should be emphasized that the bulk of the tin is contained in upper parts of the San Rafael lode, and is most strongly con-
the veins, veinlets, and breccias, as opposed to being dissem-
centrated where the lode straddles the intrusion-slate contact
inated in the chloritized wall rock (Mlynarczyk et al., 2003,
and within the slates. The sulfides clearly postdated cassi-
2009, submitted).
terite deposition, and are characterized by a great mineralog-
The entire structure of the San Rafael lode is enveloped by
a 2- to 10-m-wide zone of very strong, texturally destructive, ical diversity. The dominant sulfide mineral is chalcopyrite
chloritic alteration, which affected the host rocks during this (Fig. 3c), which is associated with subordinate amounts of
and the following stage. Chloritization is the dominant type of needle-tin cassiterite and stannite, yielding a low grade (0.51
alteration in the deposit, affecting its entire vertical extent wt %) tin ore. Other sulfides, which locally are very abundant,
and strongly overprinting all earlier alteration types. Locally, are pyrrhotite, arsenopyrite, pyrite, sphalerite, and galena. In
chloritization is associated with silicification, but the latter addition, hematite, marcasite, bismuthinite, and native bis-
may have been developed late. Studies of the whole-rock muth have also been observed. The main gangue minerals are
chemistry indicate that chloritization leached most of the con- quartz, chlorite, and fluorite. The minerals that formed dur-
tent of alkali and alkali-earth elements from the granitic wall ing the sulfide stage occur either in distinct quartz-sulfide
rocks (Na, K, Ca, Ba, Rb, Sr, Cs, and F, Cl, S) and introduced veins (commonly with a marked predominance of either
large amounts of Fe, H2O, Mn, Sn, W, In, and Mg (Mlynar- quartz or sulfides) or as a late filling (or reopening) of stage II
czyk et al., 2009). cassiterite-quartz-chlorite veins. The structural orientation of

0361-0128/98/000/000-00 $6.00 227


228 WAGNER ET AL.

inside large vugs in massive pyrrhotite ore, and siderite vein-


3 cm lets cutting across pyrrhotite, that most of the pyrite, marca-
site, and siderite originated by hypogene replacement of
qz
qz8 pyrrhotite.
9.8 Mineralogical and textural observations suggest that the
sulfide ores of San Rafael formed at comparatively high tem-
peratures, as indicated by the widespread presence of well-
developed sphalerite stars in chalcopyrite (Wiggins and Craig,
py 1980; Kojima and Sugaki, 1984; Sugaki et al., 1990) and com-
cp mon sphalerite blebs enclosed in other sulfides. Examples of
10.8 some of these ore textures are shown in Figure 5. However,
although specific paragenetic sequences can be determined
cas5 for individual samples, these parageneses generally cannot be
2.7 extended to the entire deposit, as most sulfides (aside from
9.7 pyrite and marcasite) occur in multiple generations and are
qz7 broadly coeval.
qz
(IV) Late, barren quartz-carbonate stage
cas4
10.1 2.2
4.6 The last stage of veining is represented by numerous bar-
qz6 3.0 ren quartz carbonate (calcite, siderite) veins with trace
qz5 amounts of chalcopyrite and sphalerite but no cassiterite.
3.3 They have narrow, strongly chloritized margins. These veins
crosscut all other types of veins and breccias, locally produc-
cas ing dense stockworks, and are commonly quite thick (0.21
m), implying massive precipitation of quartz in the waning
cas3 stages of the hydrothermal system. On this basis, the bulk of
the silicification observed in the deposit is tentatively assigned
3.3 to this stage.
cas In addition to the four hypogene stages, a fifth, supergene
stage is well developed in the uppermost, copper-rich part of
the lode. This resulted in the alteration of the primary ores to
Fe oxides and hydroxides (limonite) and secondary Cu min-
qz4 10.6 erals such as covellite and chalcocite. No secondary enrich-
qz 2.9 ment of tin minerals was observed.
Fluid inclusions
chl We carried out a reconnaissance fluid inclusion study of the
qz3 San Rafael deposit in order to provide important constraints
qz2
10.6 on the temperature and composition of the ore-forming flu-
cas2 ids. The minerals studied were quartz, cassiterite, tourmaline,
3.4 and fluorite, for which over two dozen fluid inclusion sections
qz1 R 119 were made. Microthermometric measurements were per-
formed at McGill University on a Linkam THMSG-600 fluid
FIG. 4. Photograph of a large, banded cassiterite-quartz-chlorite-chal- inclusion stage, using synthetic fluid inclusions (SYNFLINC)
copyrite vein (sample R 119, contact orebody, level 4310 m). Also shown are
the locations of crystals analyzed for their oxygen isotope composition. The as temperature standards (calibration temperatures: 56.6,
arrow indicates the direction of vein growth; different generations of quartz 21.2, 10.7, 0.0, and 374.1C). Heating rates were
and cassiterite are identified by labels at the left side of the figure. Note that 0.1C/min when phase transitions were approached. Errors
the oldest cassiterite generation (cas1) is not shown in the photograph, but is were 0.2 and 5C for final ice melting and homogeniza-
present in the sample. The minerals analyzed are quartz (white label), cassi-
terite (yellow), chlorite (green) and wolframite (pink).
tion temperatures, respectively. The total salinity of the inclu-
sion fluids (expressed as wt % NaCl equiv) was calculated
using the equation of Bodnar (1993) for low- and moderate-
the veins formed during the sulfide stage is the same as that salinity inclusions, and the equation of Sterner et al. (1988)
of stage II veins, suggesting a common mechanism of forma- for hypersaline inclusions. Fluid inclusion densities and iso-
tion. Moreover, the sulfide stage veins are associated with the chores were calculated using the computer package FLUIDS
same, pervasive chloritic alteration as the stage II veins (Mly- (Bakker, 2003).
narczyk et al., 2003). In addition, the lode hosts breccias,
which either contain sulfide fragments (Fig. 3d) or are ce- Fluid inclusion petrography
mented by sulfides. It also hosts large (1030 cm) vugs filled The studied samples contain abundant fluid inclusions,
by pyrite, marcasite, and rare hematite. We infer from the oc- ranging in diameter between 2 and 200 m, many of which
currence of cubes of pyrite (several centimeters in diameter) clearly represent different generations. Representative fluid

0361-0128/98/000/000-00 $6.00 228


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 229

cas
chl

qz cas cp
chl
qz

a 500 m b 580 m

py cp
sph po cp
py apy sp

cas stn py
chl
cas po
c chl 300 m d 460 m

mc
py
po
qz
py
sp
sp
cp
cp

e apy gn 580 m f 200 m

py cp
sp

stn py
sp
qz
qz
cp
g 80 m h 300 m

FIG. 5. Photomicrographs in reflected light showing representative textures of ore assemblages. a. Alternating layers of
chlorite (chl), cassiterite (cas) and quartz (qz), typical of the rich tin ores hosted by the fault-jogs in the deeper parts of the
deposit (sample R119, vein selvage, contact orebody, level 4310 m). b. Aggregates of needle-tin cassiterite crystals (seen in
cross section), intergrown with chlorite, quartz, and chalcopyrite (cp) in ore from the upper part of the deposit (sample R725,
San Rafael lode, level 4730 m). c. Needle-tin cassiterite, rimmed and crosscut by stannite (stn), associated with chlorite, chal-
copyrite, and pyrite (py), which contains inclusions of sphalerite (sp) and pyrrhotite (po) (sample R119, center of the vein,
contact orebody, level 4310 m). d. Euhedral arsenopyrite (apy) overgrown by chlorite, chalcopyrite, and cassiterite. The host
chalcopyrite contains abundant blebs of pyrrhotite (locally replaced by pyrite) and veinlets of sphalerite (sample R55, ore
shoot, level 4330 m). e. Arsenopyrite replaced by sphalerite and galena (gn), and overgrown by chalcopyrite and pyrrhotite.
Pyrrhotite is partly replaced by pyrite and marcasite (mc) (sample A1-1, orebody 150, level 4100 m). f. Sphalerite with abun-
dant chalcopyrite inclusions (chalcopyrite disease), veined by coarse-grained chalcopyrite, quartz, and pyrite. The latter sul-
fides contain fine sphalerite inclusions (sample R28, ore shoot, level 4330 m). g. Sphalerite rimmed by stannite, enclosed in
quartz, and neighboring chalcopyrite and pyrite (sample R119, center of the vein, contact orebody, level 4310 m). h. Euhe-
dral crystals of late pyrite in small vugs hosted by chalcopyrite, and crosscut by late sphalerite veinlets (copper ore, charac-
teristic of the upper levels of the deposit. sample R710, San Rafael lode, level 4820 m).

0361-0128/98/000/000-00 $6.00 229


230 WAGNER ET AL.

inclusion assemblages are shown in Fig. 6. The vast majority are in fact secondary in origin. As a result, clusters of fluid in-
of the inclusions are liquid-vapor (LV) and liquid-vapor-solid clusions displaying highly variable salinities, likely unrelated to
(LVS) types, with the latter type commonly containing more primary entrapment, are quite common. We have addressed
than one solid. On the basis of EDS analyses of opened LVS these issues by a particularly careful selection of primary and
inclusions, these solids are (1) a ubiquitous, highly birefrin- pseudosecondary fluid inclusions, where we adhered strictly to
gent mineral, which in some inclusions forms bundles and the criteria for these fluid inclusion types established by Roed-
could be a hydrous phyllosilicate (chlorite or phengitic der (1984) and Goldstein and Reynolds (1994). Moreover, in
sericite-illite), (2) aggregates of radiating birefringent crystals, establishing the final data set reported in this paper, we re-
(3) transparent, platy solids, some of which are birefringent stricted ourselves to fluid inclusion assemblages, i.e., groups of
(probable Fe-Al-Na-Ca-bearing chlorides and carbonates), texturally well-constrained fluid inclusions that yielded highly
(4) single tiny specks of an equant opaque mineral (likely consistent microthermometric results.
chalcopyrite), (5) cubes of halite (in hypersaline inclusions),
(6) relatively uncommon, finely prismatic crystals of cassi- Microthermometric results
terite (needle-tin variety), (7) anhedral grains of cassiterite Several fluid inclusion assemblages (in the sense of Gold-
(occurring also as solid inclusions), and (8) bladed crystals of stein and Reynolds, 1994) in each sample were analyzed mi-
calcite (in late quartz-calcite veins). With the exception of crothermometrically using the cycling technique method for
halite, these minerals do not redissolve on heating and likely determination of Tmice and Th (Haynes, 1985). The prelimi-
represent accidently trapped crystals. In addition to the solid- nary results of fluid inclusion microthermometry are plotted
bearing inclusions, vapor-rich and vapor-only fluid inclusions on a homogenization temperature (Th) versus salinity diagram
are locally present, but these are restricted to the early, bar- (Fig. 7). Although NaCl is the dominant solute, cassiterite-
ren tourmaline stage (I). No aqueous-carbonic inclusions hosted primary and pseudosecondary inclusions contain ap-
were observed in the minerals studied and preliminary gas preciable CaCl2 and lesser amounts of KCl. This was inferred
chromatographic analyses of inclusion fluids indicate very low from the microthermometric measurements (temperature of
proportions of CO2 and CH4. initial ice melting between 60 and 50C, temperature of
Because multiple generations of secondary inclusions are hydrohalite dissolution close to 30C) and, more directly, by
very abundant in most samples and their trails are locally very EDS analysis of fluid inclusion decrepitate mounds. In addi-
dense, it appears possible that some of the original primary tion, FeCl2 was also present in the ore fluids, as demonstrated
fluid inclusions were subsequently refilled. In addition, some by comparatively low but consistent amounts of FeCl2 de-
of the secondary fractures in the host minerals (especially tected in the analyzed decrepitates.
quartz) have healed very well, producing a scattered distribu- Stage I quartz- and tourmaline-hosted inclusions have a
tion of those inclusions, which texturally appear isolated but salinity range of 34 to 62 wt percent NaCl equiv (with most

a 30 m 30 m
b

c 25 m
d 50 m

FIG. 6. Photomicrographs of representative fluid inclusion assemblages from a complex cassiterite-quartz-sulfide vein
(sample R-119, contact orebody, level 4310 m). a. Cluster of primary fluid inclusions in stage II quartz. b. Primary fluid in-
clusions in stage II cassiterite developed along growth zones. c. Large primary fluid inclusion in stage II quartz. d. Densely
distributed very low salinity secondary fluid inclusions in stage IV quartz.

0361-0128/98/000/000-00 $6.00 230


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 231

600
LVH inclusions (homogenize by halite dissolution)
LVH inclusions (homogenize by vapor disappearance)
Homogenization temperature (C) Primary / pseudosecondary LV inclusions Stage I
500 quartz / tourmaline
Secondary LV inclusions
g
oilin
i cb
isod
Stage II cassiterite ep
400 Stage III
quartz / fluorite

e
rv
cu
n
tio
ra
tu
300

sa
e
lit
ha
Stage II
Stage IV quartz
200 quartz
necking
down

100
0 10 20 30 40 50 60 70

Total salinity (wt.% NaCl eq.)


FIG. 7. Homogenization temperature versus salinity of fluid inclusions from San Rafael ores. The host minerals investi-
gated were quartz, fluorite, cassiterite, and tourmaline. Abbreviations: L = liquid, V = vapor, H = halite.

data clustering around 4156 wt %) and homogenize by dis- bands relative to those in cassiterite layers indicate marked,
appearance of either halite or vapor in the range 240 to periodic fluctuations in the composition of the fluids circulat-
535C, but mainly between 340 and 475C. Based on textural ing in the vein system.
evidence, it is inferred that the inclusions, which have a very Primary inclusions in stage III quartz and fluorite are char-
high salinity and homogenization temperature, formed due to acterized by a salinity in the range of 0.513 wt % NaCl equiv
local boiling of the early fluids, although necking down after (mostly 36 wt %) and a Th in the range of 265340C. Sig-
a phase change could also explain the highly variable phase nificantly, primary and pseudosecondary inclusions in quartz
ratios observed within fluid inclusion assemblages. The data associated with chalcopyrite in the upper zone of the deposit
obtained are in broad agreement with those of Kontak and (sample R709, elev 4,820 m) have, on average, a Th lower by
Clark (2002). about 40C than inclusions hosted by a very similar variety of
Stage II, cassiterite-hosted, primary and pseudosecondary quartz associated with the chalcopyrite-rich centers of
fluid inclusions were measured in several generations of cas- deeper-seated stage II-III veins (sample R119, elev 4,310 m).
siterite (cas1 to cas3, Fig. 4), and have a very narrow range in Another important observation is the lower Th limit of about
salinity (1721 wt % NaCl equiv) and Th (354361C). 265C consistently displayed by all (primary-secondary) stage
Quartz-hosted inclusions from the same stage exhibit consis- II and stage III inclusions. Fluid inclusions hosted by stage IV
tently lower temperature and salinity, which vary depending quartz have the lowest salinity (02.5 wt % NaCl equiv) and
on the generation of the host quartz (1116 wt % NaCl equiv homogenization temperature (235265C).
and 265305C for qz1 and qz2, and 2.55.5 wt % NaCl When the microthermometric data for all four paragenetic
equiv and 265295C for the later qz3 and qz4, Fig. 4). The stages are considered jointly (Fig. 7), it is evident that the
salinity of the fluids which precipitated stage II quartz is likely sets of fluid temperature and salinity recorded for each stage
to have spanned the entire range between 2.5 and 16 wt per- form a broad linear trend of salinity with homogenization
cent NaCl equiv, as suggested by the behavior of fluids temperature. Quartz-hosted primary-pseudosecondary stage
trapped in secondary inclusions. However, the most remark- II inclusions form two populations at lower and salinity than
able feature of the quartz-hosted inclusions from this stage is stage II cassiterite- (the main cassiterite stage) hosted inclu-
a population of inclusions of very low salinity (Tmice as high as sions, whereas stage III quartz-fluorite-hosted inclusions
1.3C), which are quite similar in composition and tempera- have higher Th values than stage II quartz-hosted inclusions.
ture to the fluids that precipitated the barren quartz from Overall, the fluid inclusion data point to repeated fluid mix-
stage IV (Fig. 7). In addition, the character of the ore veins, ing during the evolution of the San Rafael deposit, with the
in which bands of cassiterite, chlorite, and quartz repetitively hotter and more saline fluids of the main cassiterite stage
alternate, and the consistently lower homogenization temper- more closely reflecting the composition of the primary ore
ature and salinity of the fluid inclusions trapped in quartz fluid.

0361-0128/98/000/000-00 $6.00 231


232 WAGNER ET AL.

Mineral Geothermometry for sulfur, oxygen, hydrogen, and carbon isotope analysis. The
Mlynarczyk et al. (2009) have tested two empirical chlorite samples cover a wide range of textures of hydrothermal sul-
thermometers (Cathelineau, 1988; Jowett, 1991) on different fides, oxides, and silicates, and permit characterization of the
samples of alteration and vein chlorite from stage II of the isotopic evolution of the system along the paragenetic se-
paragenesis. The chlorite temperatures obtained are in the quence. In addition to samples that cover the different min-
range of 335415C. The thermometer of Jowett (1991) eralization stages, we also performed selected small-scale
yields values about 515C higher than those from that of studies of individual veins employing a high sample density to
Cathelineau (1988), and the values for alteration chlorite are, resolve isotopic effects at the micro- and mesoscales.
on average, roughly 50C higher than those for vein chlorite.
The theoretical geothermometer of Walshe (1986) yields Analytical techniques
temperatures in the range 295 to 320C for the alteration Separates of hydrothermal carbonate (siderite, calcite),
chlorite (Mlynarczyk et al., 2009). These temperatures are oxide, and silicate minerals (quartz, cassiterite, wolframite,
consistently below the Th values of most stage II fluid inclu- chlorite, tourmaline) were prepared by careful handpicking
sions. We have not further considered these results, because under a binocular microscope, followed by cleaning in doubly
the thermodynamic model of Walshe (1986) has been shown distilled water. Sulfide minerals (pyrite, pyrrhotite, arsenopy-
to be inconsistent with more recent experimental and theo- rite, chalcopyrite, sphalerite, galena) were analyzed by in situ
retical studies (Holland et al., 1998; Vidal et al., 2001; Aja, laser combustion from standard polished blocks. Whole-rock
2002; Aja and Dyar, 2002;). Overall, the results suggest that oxygen isotope compositions were analyzed from rock pow-
chlorite (and, by inference, the associated cassiterite) formed ders prepared by standard crushing and milling techniques.
in the temperature range 350420C. All mineral separates and whole-rock samples were dried at
The stage III sulfide ores commonly contain microscopic 80C for at least 6 hours prior to isotopic analysis. The sulfur,
overgrowths of stannite on sphalerite (Fig. 5g), as well as less oxygen, and carbon isotope analysis was conducted in the sta-
common intergrowths between these two minerals. It has ble isotope lab at SUERC (East Kilbride, UK), whereas hy-
been shown that the equilibrium compositions of sphalerite drogen isotope analysis was carried out at the isotope labora-
and stannite can be used successfully as a geothermometer tory of the Department of Geological Sciences at Queens
(Nekrasov et al., 1976, 1979; Shimitzu and Shikazono, 1985; University (Kingston, Canada).
Bortnikov et al., 1990). Carefully selected sphalerite-stannite Sulfide minerals were combusted in an oxygen atmosphere
intergrowths in representative ore samples were analyzed by using a SPECTRON LASERS 902Q CW Nd:YAG laser (1 W
electron microprobe at McGill University. Representative re- power); typical spot diameters in this study were 200 to 400
sults and average data are given in Table 1. The equation of m. Laser extraction was followed by cryogenic purification
Nakamura and Shima (1982) was used to calculate the for- of the SO2 gas and subsequent on-line mass-spectrometric
mation temperature of the intergrowths, which is interpreted analysis. Details of the laser extraction technique, calibration,
to represent the lower temperature limit of stage III ore for- and correction procedures are provided by Kelley and Fallick
mation. The resulting temperatures are generally in the range (1990), Kelley et al. (1992) and Wagner et al. (2002). Laser
270 to 290C, i.e., they are slightly lower than fluid inclusion calibration data for arsenopyrite were taken from the study of
homogenization temperatures from this stage. Wagner et al. (2004). Reproducibility of the analytical results,
and mass spectrometer calibration, were monitored through
Stable Isotope Studies replicate measurements of international standards NBS-123
In order to test our hypothesis that fluid mixing was the (34SV-CDT: 17.1) and IAEA-S-3 (34SV-CDT: 31.0), as
dominant ore-forming process, we performed a comprehen- well as the internal lab standard CP-1 (34SV-CDT: 4.6).
sive stable isotope study on the main ore and gangue miner- The analytical precision (1) was about 0.2 per mil. All sul-
als. Representative samples from the principal mineralization fur isotope compositions are reported in standard delta nota-
stages (stages IIV) of the San Rafael deposit were selected tion relative to V-CDT.

TABLE 1. Composition of Sphalerite-Stannite Intergrowths and Equilibrium Temperatures Calculated


Using the Geothermometer of Nakamura and Shima (1982)

Sample no. Location Mineral Cu (wt%) Fe (wt%) Zn (wt%) Sn (wt%) S (wt%) Total XFe/XZn T (C)

R27 Ore shoot 4330 m Sphalerite 0.13 11.14 54.55 0.04 33.19 99.05 0.20 282
Stannite 28.34 14.94 2.11 25.56 30.13 101.08 7.08
R119 Contact orebody 4310 m Sphalerite 0.20 11.79 53.69 0.01 33.45 99.14 0.22 277
Stannite 27.93 15.24 1.80 24.99 30.02 99.98 8.47
R447 Contact orebody 4270 m Sphalerite 0.51 11.47 52.80 0.02 33.05 97.85 0.22 269
Stannite 28.25 14.40 1.43 26.51 29.62 100.21 10.07
R404 S contact orebody 4310 m Sphalerite 0.10 12.70 52.83 0.01 34.20 99.84 0.24 305
Stannite 28.53 13.75 2.59 26.85 30.41 102.13 5.31
A1-1 No. 150 orebody 4100 m Sphalerite 0.82 13.69 50.6 0.00 33.91 99.02 0.27 288
Stannite 28.27 14.23 1.69 25.4 29.52 99.11 8.42
R65 Ore shoot 4390 m Sphalerite 0.42 11.32 53.55 0.07 33.61 98.97 0.21 269
Stannite 28.52 13.95 1.42 26.32 30.10 100.31 9.82

0361-0128/98/000/000-00 $6.00 232


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 233

Oxygen was extracted from mineral separates and whole- analytical results was monitored through replicate measure-
rock powders by reacting 1 to 5 mg of sample with purified ments of the internal Mab2b standard (13CV-PDB: 2.48 ;
chlorine trifluoride in a laser fluorination system, based on 18OV-PDB: 2.40) before and after each batch of samples.
techniques of Sharp (1990) and Mattey and Macpherson Accuracy was controlled by replicate measurements of inter-
(1993). The oxygen was converted to CO2 by reaction with a national standards IAEA-CO-1 (13CV-PDB: 2.48; 18OV-PDB:
heated graphite rod; the isotopic composition of the cryo- 2.44) and IAEA-NBS-19 (13CV-PDB: 1.95; 18OV-PDB:
genically purified CO2 gas was measured on-line with a VG 2.20 ). External precision (1) was found to be better than
PRISM III mass spectrometer. Analytical precision was con- 0.2 for both carbon and oxygen isotope compositions.
trolled through replicate measurements of the internal labo- Carbon and oxygen isotope data are reported relative to V-
ratory standard SES-2 (18OV-SMOW: 10.2) during the PDB and V-SMOW, respectively.
course of the study. The latter was calibrated against interna- Hydrogen isotope analysis of chlorite and tourmaline was
tional standards IAEA-NBS-28 (18OV-SMOW: 9.6) and carried out by elemental analysis continuous-flow isotope
IAEA-NBS-30 (18OV-SMOW: 5.1). Precision (1) was found ratio mass spectrometry (CF-IRMS), using a high-tempera-
to be better than 0.2 per mil for the whole dataset. Some ad- ture reduction method (Sharp et al., 2001). This method in-
ditional tourmaline samples were analyzed by conventional volves passing the sample in an He stream through a carbon-
techniques following the procedures of Clayton and Mayeda packed furnace heated to 1450C, and releasing the water,
(1963). All O isotope data are reported relative to V-SMOW. which is then reduced to H2. Hydrogen gas is purified by
Hydrothermal carbonates were analyzed for their C and passing it through a 5A molecular sieve gas chromatographic
O isotope composition using an Analytical Precision column, and subsequently analyzed with a Finnigan Mat
AP2003 continuous-flow mass spectrometer, equipped Delta XL Plus gas source mass spectrometer. Precision (1)
with an automated carbonate preparation system. About 1 was found to be better than 3 per mil. The hydrogen isotope
mg of sample powder was placed in a 6 ml vacutainer, then data are reported relative to V-SMOW.
sealed and loaded onto the autosampling unit. Each sam-
ple was flushed with helium, then a predetermined Sulfur isotope data
amount of 93 percent phosphoric acid was injected into A total of 64 in situ laser sulfur isotope analyses were per-
each tube, following the procedure of McCrea (1950). The formed on polished blocks from 20 samples, spanning both
acid reaction was conducted at a temperature of 70 the paragenesis and the vertical extent of the deposit. The
0.1C; reaction times were 24 hours for pure calcite sam- majority of the samples are from stage III (sulfide stage),
ples and 120 hours for all other carbonates. After comple- where sulfides are abundant and generally several sulfide
tion of the reaction, the samples were transferred to the minerals are present, in contrast to stages I and II for which
processing system and analyzed with an AP2003 mass sulfides are scarce and are represented only by arsenopyrite.
spectrometer. Oxygen isotope data for siderite were cor- The sulfides analyzed were arsenopyrite, chalcopyrite,
rected using an acid fractionation factor calculated from pyrrhotite, pyrite, sphalerite, and galena. Typically, several
Rosenbaum and Sheppard (1986). Reproducibility of the coexisting sulfides were analyzed in each sample (Fig. 8). The

tou
qz cp
3.7 5.0
gn
5.3 5.1
apy 5.9 py py
3.6 sp qz
cp
5.5
sp py
3.8
5.4 4.2
qz D3-27 R 404 R 119

gn
cp po 5.1
3.9
gn 6.3 2.4
apy apy sp
1.2
4.8 gn qz
3.0
qz sp 5.4 cp 1.4 sp
po py
qz 4.8 cp
apy 5.0 3.7
4.4
po
4.0
R 511a J 410 R 428
FIG. 8. Images showing millimeter-scale textural relationships and sulfur isotope data for representative polished sections.
The average diameter of the polished blocks is 2.5 cm. (apy = arsenopyrite, cp = chalcopyrite, gn = galena, po = pyrrhotite,
py = pyrite, qz = quartz, sp = sphalerite, tou = tourmaline).

0361-0128/98/000/000-00 $6.00 233


234 WAGNER ET AL.

TABLE 2. Sulfur Isotope Data for the San Rafael Sn-Cu Deposit (n = 64)

Sample no. Mineral Textural relationships 34SV-CDT ()

Early, barren stage


D3-39-1 Arsenopyrite Subhedral, in early tourmaline-quartz vein (DDH no. 3 to the Mariano vein, 39 m) 2.0
D3-39-2 Arsenopyrite Subhedral, in early tourmaline-quartz vein (DDH no. 3 to the Mariano vein, 39 m) 3.0
D3-27-1 Arsenopyrite Subhedral, in very early quartz veinlet (DDH no. 3 to the Mariano vein, 27 m) 3.6
D3-27-2 Sphalerite Anhedral, with arsenopyrite, in early quartz veinlet (DDH no. 3 to the Mariano vein, 27 m) 5.4
D3-27-3 Arsenopyrite Subhedral, with tourmaline, in early quartz veinlet (DDH no. 3 to the Mariano vein, 27 m) 3.7
D8-50-1 Arsenopyrite Early veinlets in slate, associated with tourmaline and quartz (DDH #8, 50.5 m) 3.8
D8-50-2 Arsenopyrite Early veinlets in slate, associated with tourmaline and quartz (DDH #8, 50.5 m) 3.7
D3-182b-1 Arsenopyrite Early zone, massive, in tourmaline-cassiterite vein (DDH no. 3 to the Mariano vein, 182.5 m) 2.4
D3-182b-2 Arsenopyrite Late zone, massive, in tourmaline-cassiterite vein (DDH no. 3 to the Mariano vein, 182.5 m) 6.2
D3-182b-3 Arsenopyrite Late zone, massive, in tourmaline-cassiterite vein (DDH no. 3 to the Mariano vein, 182.5 m) 3.6
Sulfide ore stage
511b-1 Arsenopyrite Large subhedral crystals (early zone) center of quartz-chlorite-cassiterite-arsenopyrite vein 3.9
511b-2 Sphalerite Anhedral grain, with arsenopyrite, center of quartz-chlorite-cassiterite-arsenopyrite vein 4.2
511a-1 Arsenopyrite Large subhedral crystals (latest zone) center of quartz-chlorite-cassiterite-arsenopyrite vein 3.9
511a-2 Arsenopyrite Large subhedral crystals (late zone) center of quartz-chlorite-cassiterite-arsenopyrite vein 4.4
A13-1 Galena Inclusions in massive pyrite ore 5.4
A13-2 Pyrite Coarse-grained massive ore 4.0
R729-1 Pyrite Narrow veinlet in chalcopyrite ore 3.8
R729-2 Chalcopyrite Coarse-grained, massive ore 2.3
R17-1 Chalcopyrite Vug fillings in quartz vein 4.2
R17-2 Pyrite Anhedral grains in quartz 4.7
R573-1 Chalcopyrite Rim around pyrrhotite 3.7
R573-2 Pyrrhotite Massive vein, with quartz 3.0
R573-3 Pyrrhotite Massive vein, with quartz 3.1
R404-1 Pyrite Fine-grained, replacing pyrrhotite, in quartz vein 3.8
R404-2 Pyrite Euhedral, replacing fine pyrite in quartz vein 4.2
R404-3 Sphalerite Subhedral, in quartz vein 5.9
R404-4 Galena Subhedral, in quartz vein 5.3
R165-1 Chalcopyrite Coarse-grained massive ore 4.8
R165-2 Pyrite Euhedral crystals in chalcopyrite ore 4.9
R165-3 Chalcopyrite Coarse-grained massive ore 3.9
G4-1 Arsenopyrite Euhedral crystals in massive chalcopyrite ore (San German vein, level 4650 m) 3.9
G4-2 Pyrrhotite Veinlets crosscutting chalcopyrite (San German vein, level 4650 m) 3.4
G4-3 Chalcopyrite Coarse-grained massive ore (San German vein, level 4650 m) 3.7
G4-4 Pyrrhotite Veinlets crosscutting chalcopyrite (San German vein, level 4650 m) 4.0
R428-1 Galena Subhedral inclusions in sphalerite, in quartz-chalcopyrite-sphalerite vein 1.4
R428-2 Chalcopyrite Coarse-grained, overgrowing quartz, in quartz-chalcopyrite-sphalerite vein 3.0
R428-3 Sphalerite Coarse-grained, overgrowing chalcopyrite, in quartz-chalcopyrite-sphalerite vein 6.3
R428-4 Galena Subhedral inclusions in sphalerite, in quartz-chalcopyrite-sphalerite vein 5.1
R428-5 Sphalerite Coarse-grained, overgrowing chalcopyrite, in quartz-chalcopyrite-sphalerite vein 5.4
R428-6 Sphalerite Coarse-grained, overgrowing chalcopyrite, in quartz-chalcopyrite-sphalerite vein 4.8
R428-7 Galena Subhedral inclusions in sphalerite, in quartz-chalcopyrite-sphalerite vein 1.2
R428-8 Galena Subhedral inclusions in sphalerite, in quartz-chalcopyrite-sphalerite vein 2.4
R726-1 Galena Anhderal inclusions in quartz, in quartz-chalcopyrite vein 3.1
R726-2 Sphalerite Subhedral inclusions in quartz, in quartz-chalcopyrite vein 5.6
R726-3 Chalcopyrite Massive, coarse-grained ore, with quartz 5.1
R119-1 Pyrite Veinlets crosscutting chalcopyrite 5.1
R119-2 Chalcopyrite Massive, in the central part of cassiterite-quartz-chlorite-chalcopyrite vein 5.5
R119-3 Chalcopyrite Massive, in the central part of cassiterite-quartz-chlorite-chalcopyrite vein 5.0
R65b-1 Pyrite Massive, replacing pyrrhotite, in chalcopyrite-arsenopyrite-pyrite vein 3.4
R65b-2 Pyrite Massive, replacing pyrrhotite, in chalcopyrite-arsenopyrite-pyrite vein 3.8
R65b-3 Pyrite Fine-grained, replacing pyrrhotite, in chalcopyrite-arsenopyrite-pyrite vein 4.5
R55-1 Arsenopyrite Euhedral crystal in chalcopyrite, in quartz-chalcopyrite-pyrrhotite-arsenopyrite vein 5.0
R55-2 Pyrrhotite Anhedral inclusions in chalcopyrite, in quartz-chalcopyrite-pyrrhotite-arsenopyrite vein 5.8
R55-3 Chalcopyrite Coarse-grained, massive ore, in quartz-chalcopyrite-pyrrhotite-arsenopyrite vein 4.9
R55-4 Pyrrhotite Anhedral inclusions in chalcopyrite, in quartz-chalcopyrite-pyrrhotite-arsenopyrite vein 5.6
R55-5 Arsenopyrite Euhedral crystal in chalcopyrite, in quartz-chalcopyrite-pyrrhotite-arsenopyrite vein 5.1
R28a-1 Galena Anhedral crystals in quartz-chalcopyrite-pyrite-sphalerite-galena vein 3.3
R28a-2 Sphalerite Anhedral crystals in quartz-chalcopyrite-pyrite-sphalerite-galena vein 3.9
R28a-3 Sphalerite Anhedral crystals in quartz-chalcopyrite-pyrite-sphalerite-galena vein 4.5
R28a-4 Galena Anhedral crystals in quartz-chalcopyrite-pyrite-sphalerite-galena vein 3.7
J410-1 Pyrrhotite Inclusions in large pyrite crystal (Jorge vein, level 4000 m) 3.7
J410-2 Pyrite Large, blocky crystal replacing pyrrhotite (Jorge vein, level 4000 m) 4.8
J410-3 Pyrrhotite Coarse-grained, massive ore (Jorge vein, level 4000 m) 4.0
J410-4 Pyrrhotite Coarse-grained, massive ore (Jorge vein, level 4000 m) 5.0

Notes: All samples were analyzed in situ by a laser combustion system; unless indicated otherwise, the samples are from the San Rafael vein and their
location is shown in Figure 2

0361-0128/98/000/000-00 $6.00 234


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 235

results of the analyses are listed and illustrated in Table 2 and Oxygen isotope data
Figure 9, respectively, and show that the sulfide ores have a Thirty-nine oxygen isotope analyses were performed on
relatively uniform 34S composition, ranging between 2 and 6 carefully handpicked mineral separates of oxides and silicates
per mil. Although the range of 34S values is narrow, as might
from 17 samples, including a complexly banded, 5-cm-thick,
be expected from a deposit associated with a S-type granitoid
quartz-chlorite-cassiterite-wolframite-chalcopyrite-pyrite
(Ohmoto and Goldhaber, 1997), minerals within individual
samples commonly display variable 34S values. For example, vein, which was sampled in detail (Fig. 4). In addition, 5
the 34S values of arsenopyrite in sample D3-182b vary within whole-rock samples of granitoids (ranging from fresh to
the same growth band from 3.6 to 6.2 per mil, whereas the strongly chloritized) and 12 vein carbonates (see section
34S of galena in sample R428 varies from 1.2 to 5.1 per mil below) were analyzed. The oxide and silicate minerals ana-
within only a few millimeters (Fig. 8; Table 2). There is, how- lyzed were tourmaline, quartz, chlorite, cassiterite, and wol-
ever, little variation in the 34S values of the sulfide minerals framite, and the results are summarized in Figure 10 and
with respect to their location in the deposit. Arsenopyrite (the listed in Table 3. The 18O values for the individual minerals
only sulfide that occurs throughout the entire paragenesis) are all positive and have relatively narrow ranges of 9.7 to
displays a distinct trend of increasing 34S values from stage I 11.9 per mil for tourmaline, 9.7 to 14.4 per mil for quartz, 3.9
(2.03.8) to stage III (3.85.1%). to 4.6 per mil for chlorite, 1.8 to 3.4 per mil for cassiterite,
and 1.7 to 3.0 per mil for wolframite. The transition from the
early, barren stage I to the tin-rich stage II is characterized by
a marked decrease in the 18O values of tourmaline. Within
8
the same vein the 18O value of early, orange-colored tour-
6 maline is 11.7 per mil, whereas the succeeding dark-green
4 tourmaline that is associated with abundant cassiterite and
Arsenopyrite chlorite (the first to form in the paragenesis) has a 18O value
2
a of 9.7 per mil (sample D3-182, Table 3). In the subsequent
0
0 1 2 3 4 5 6 7 8 9 10 stages (II and III), the 18O values of cassiterite and quartz
are variable, but nevertheless provide evidence for an overall
6 decrease of 18O with time, which for the ore vein investi-
4 gated in detail corresponds to a change from 3.4 to 2.7 per mil
2 for cassiterite, and from 10.6 to 9.8 per mil for quartz (sample
b Chalcopyrite R 119, Fig. 4; Table 3). The latest cassiterite in the paragene-
0
0 1 2 3 4 5 6 7 8 9 10 sis (stage III needle-tin cassiterite, intergrown with chalcopy-
rite from the upper zone of the deposit), has an even lower
6
18O value of 1.8 per mil (sample R725, Table 3). By contrast,
4 quartz from the paragenetically latest stage (IV) has markedly
higher 18O values, in the range of 11.3 to 14.4 per mil.
Frequency

2
c Pyrrhotite Whole-rock 18O values of fresh granitic rocks range from 9.6
0
0 1 2 3 4 5 6 7 8 9 10 (biotite-granodiorite) to 10.9 per mil (leucogranite). A se-
6
quence of four slabs of granodiorite with increasing chloritic
alteration was investigated in detail and shows a systematic
4 decrease in 18O with alteration, down to 8.0 per mil for the
2 strongly chloritized granodiorite and 4.9 per mil for the chlo-
d Pyrite
0 ritite, which occupies the center of the vein (Table 3).
0 1 2 3 4 5 6 7 8 9 10 Although a lack of complete isotopic equilibrium between
6 coexisting quartz and cassiterite has been reported for a num-
ber of hydrothermal tin lodes (Alderton, 1989), several iso-
4 topic geothermometers were tested on the silicate-oxide as-
2 semblages (quartz, cassiterite, wolframite, chlorite) from the
e Sphalerite
0
banded ore vein investigated in detail (sample R119, Fig. 4).
0 1 2 3 4 5 6 7 8 9 10 Because quartz and cassiterite could not always be separated
6
from the same growth band, we calculated oxygen isotope
temperatures based on data for pairs of minerals intimately
4 intergrown in a single band, as well as for pairs where in
2 which the minerals were in adjacent bands. Overall, the geo-
f Galena thermometers (quartz-cassiterite, quartz-wolframite, quartz-
0
0 1 2 3 4 5 6 7 8 9 10 chlorite), which are all based on carefully conducted experi-
ments (Matsuhisa et al., 1979; Zhang et al., 1994; Cole and
34SV-CDT () Ripley, 1999) give average equilibrium temperatures in the
FIG. 9. Histograms displaying the sulfur isotope composition of San range of 370 to 450C, about 40 to 60 C higher than the
Rafael sulfides. fluid inclusion homogenization temperatures. The error on

0361-0128/98/000/000-00 $6.00 235


236 WAGNER ET AL.

6
4
2
a Tourmaline
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
6
4
2
b Quartz
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Frequency

6
4
2
c Chlorite
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

6
4
2
d Cassiterite
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

6
4
2
e Wolframite
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

18OV-SMOW ()
FIG. 10. Histograms displaying the oxygen isotope composition of San Rafael silicate and oxide minerals.

the calculated equilibrium temperatures, considering the an- two successive generations of siderite were studied, but did
alytical uncertainty of the 18O values, is on the order of about not yield a consistent trend.
30C. When different stages of the vein evolution are com-
pared, the calculated oxygen isotope temperatures (e.g., the Discussion
quartz-cassiterite pair for which most data are available) show
a general decrease from the early stage II assemblage (450C) Conditions of ore deposition
to the later complex cassiterite-quartz-wolframite-chlorite as- Based on our geologic and textural observations, the hy-
semblage (400420C). drothermal history of the ore-forming system at San Rafael
comprised two contrasting episodes. The first of these (stage
Hydrogen isotope data I) was characterized by the formation of numerous tourma-
Four tourmaline and two chlorite samples were analyzed line-quartz ( arsenopyrite) veins, veinlets, and stringers,
for their hydrogen isotope composition and the results are with which sericitic and tourmaline alteration are associated.
listed in Table 4. The D values of tourmaline lie in a narrow It is important to note that these structures are typically quite
range between 75 and 73 per mil and are essentially iden- narrow, locally discontinuous, and invariably sealed, likely in-
tical within errors. The D values of the two chlorite samples dicating pressures approaching lithostatic. Stratigraphic re-
are 88 and 81 per mil. construction based on regional geologic data (e.g., Clark et
al., 1990) suggests that about 2.5 km of rock has been eroded
Carbon and oxygen isotope data of carbonates since the emplacement of the deposit (at about 25 Ma). This
The oxygen and carbon isotope data for late-stage vein car- translates into a lithostatic fluid pressure of 700 to 800 bars,
bonates (5 samples of calcite and 7 of siderite) are shown in which is consistent with pressure estimates given by Kontak
Figure 11 and presented in Table 5. The 18O and 13C val- and Clark (2002). Fluid inclusion data demonstrate that the
ues of calcite samples are in the range of 8.6 to 14.2 per mil fluids responsible for this early, barren stage were hot, hyper-
and of 8.5 to 3.2 per mil, respectively. The data for siderite saline brines (fluid trapping temperatures of 380540C
are more variable, and are in the range of 13.8 to 23.1 per mil based on isochoric projection of fluid inclusion Th values to
and 6.5 to 4.3 per mil, respectively (Fig. 11). In two samples, 700800 bars, salinity of 3462 wt % NaCl equiv) and the

0361-0128/98/000/000-00 $6.00 236


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 237

TABLE 3. Oxygen Isotope Data for Quartz, Cassiterite, Wolframite, Chlorite, Tourmaline, and Whole-Rock Samples
from the San Rafael Sn-Cu Deposit (n = 44)

Sample no. Mineral Textural relationships 18OV-SMOW ()

Early, barren stage


D5-70 Tourmaline Tourmaline vein, euhedral, prismatic tourmaline (DDH # 5 to the San Rafael vein, 70 m) 11.5
D5-94 tou Tourmaline Quartz-tourmaline vein (DDH #5 to the San Rafael vein, 94 m) 11.9
D5-94 qz Quartz Quartz-tourmaline vein (DDH #5 to the San Rafael vein, 94 m) 12.3
D1-280 Tourmaline Tourmaline-quartz vein (DDH #1 to the San Rafael vein, 280 m) 10.8
D3-69 Quartz Quartz-tourmaline veinlet: the quartz is very early and borders the tourmaline vein 12.0
(DDH no. 3 to the Mariano vein, 69 m)
D3-182 a Tourmaline Tourmaline-cassiterite-chlorite-arsenopyrite vein (early, orange tourmaline) 11.7
(DDH no. 3 to the Mariano vein, 182.5 m)

Main ore stage (tin and copper)


D3-182 b Tourmaline Tourmaline-cassiterite-chlorite-arsenopyrite vein (late, green tourmaline, associated with 9.7
cassiterite) (DDH no. 3 to the Mariano vein, 182.5 m)
D3-182 b Cassiterite Tourmaline-cassiterite-chlorite-arsenopyrite vein (earliest cassiterite in the deposit) 2.0
(DDH no. 3 to the Mariano vein, 182.5 m)
R119 Banded quartz-chlorite-cassiterite-wolframite-chalcopyrite-pyrite vein (Fig. 5)
R119 qz1 Quartz Euhedral quartz associated with chlorite, in between cassiterite bands (Qtz IB) 10.6
R119 qz2 Quartz Large, euhedral quartz crystals, preceding the deposition of massive cassiterite (Qtz IV) 10.6
R119 qz3 Quartz Small, euhedral quartz crystals, with chlorite and needle-tin cassiterite (Qtz VI) 10.1
R119 qz4 Quartz Large, euhedral quartz crystals, preceding needle-tin cassiterite deposition (Qtz VII) 9.7
R119 qz5 Quartz Large, subhedral quartz crystals, following needle-tin cassiterite deposition (Qtz VIII) 10.8
R119 qz6 Quartz Very late, barren, discordant quartz-chlorite veinlet (Qtz X) 9.8
R119 cas1 Cassiterite Earliest layer of massive, dark-brown cassiterite (Cas I) 3.4
R119 cas2 Cassiterite Earliest zone of a 7-cm-thick layer of dark-brown, massive cassiterite (Cas IIIA) 2.9
R119 cas3 Cassiterite Middle zone of a 7-cm-thick layer of dark-brown, massive cassiterite (Cas IIIB) 3.3
R119 cas4 Cassiterite Latest zone of a 7-cm-thick layer of dark-brown, massive cassiterite (Cas IIIC) 3.3
R119 cas5 Cassiterite Early needle-tin cassiterite (Cas IV) 2.2
R119 cas6 Cassiterite Late, needle-tin cassiterite, associated with chalcopyrite (Cas V) 2.7
R119 chl Chlorite Chlorite, associated with quartz and needle-tin cassiterite 4.6
R119 wol Wolframite Wolframite, associated with chlorite and quartz 3.0
R113 Wolframite Same occurrence as R119 wol 2.1
R137 Cassiterite, quartz and chlorite-cemented breccia of chloritized wall-rock fragments
R137 qz1 Quartz Euhedral quartz preceding cassiterite deposition 10.8
R137 cas Cassiterite Layer of massive, dark-brown cassiterite 2.7
R137 qz2 Quartz Euhedral quartz following cassiterite deposition 11.2
D4 cas1 Cassiterite Early layers of light-colored wood-tin cassiterite (San Rafael Lode, elev 4500 m) 2.3
D4 cas2 Cassiterite Late layers of dark wood-tin cassiterite (San Rafael Lode, elevation 4500 m) 3.4
R511 chl Chlorite Cassiterite-chlorite-quartz-arsenopyrite vein 3.9
R511 qz Quartz Large, euhedral quartz crystals from a cassiterite-chlorite-quartz-arsenopyrite vein 12.2
D1-219 wol Wolframite Quartz-chlorite-cassiterite-wolframite vein (DDH no. 1 to the San Rafael vein, 219 m) 1.7
D1-219 qz Quartz Quartz-chlorite-cassiterite-wolframite vein (DDH no. 1 to the San Rafael vein, 219 m) 10.1
R725 Cassiterite Needle tin cassiterite-chalcopyrite-quartz vein (tin ore from the upper zone) 1.8
R429 Quartz Anhedral quartz from a large quartz-sphalerite vein 9.8
R708 Quartz Subhedral quartz from a major quartz-chalcopyrite-fluorite-siderite vein 11.1

Late, barren stage


R399 qz1 Quartz Late-stage, very large crystal of euhedral vug quartz (core) 14.0
R399 qz2 Quartz Late-stage,very large crystal of euhedral vug quartz (outer zone) 13.1
R709 Quartz Late-stage, barren, very large quartz vein 14.4
D5M-155 Quartz Quartz-calcite fill of a reopened tourmaline vein (DDH no. 5 to the Mariano vein, 155 m) 11.3

Whole-rock samples
I20 Whole rock Fresh alkali-feldspar cordierite-biotite leucogranite (Ramp, elev 3930 m) 10.9
D1M-61d Whole rock Fresh biotite-granodiorite (DDH no. 1 to the Mariano vein, 61 m) 9.6
D1M-61c Whole rock Mildly altered biotite-granodiorite (DDH no. 1 to the Mariano vein, 61 m) 9.6
D1M-61b Whole rock Strongly chloritized biotite-granodiorite (DDH no. 1 to the Mariano vein, 61 m) 8.0
D1M-61a Whole rock Chloritite adjacent to the center of a chloritic vein (DDH no. 1 to the Mariano vein, 61 m) 4.9

Notes: Unless indicated otherwise, the samples are from the San Rafael vein and their location is shown in Figure 2

alteration minerals that formed indicate that they were acidic an open fracture-filling character, with ubiquitous vugs, likely
(Mlynarczyk et al., 2009, submitted). indicative of hydrostatic conditions. Evidence for multiple
By contrast, the subsequent stages (i.e., tin-mineralized vein-opening events suggests that the pressure must have
stage II, copper-mineralized stage III, and late, barren quartz fluctuated from slightly above lithostatic (700800 bars) to
stage IV) are all associated with a distinctive, strong chloritic near hydrostatic (200300 bars) during ore deposition (as-
alteration (and locally, silicification). The veins invariably have suming a roughly similar depth to that during stage I). The

0361-0128/98/000/000-00 $6.00 237


238 WAGNER ET AL.

TABLE 4. Hydrogen Isotope Data for Tourmaline and Chlorite from the San Rafael Sn-Cu Deposit (n = 6)

Sample no. Mineral Textural relationships DV-SMOW ()

Early, barren stage


D5-70 Tourmaline Tourmaline vein, euhedral, prismatic tourmaline (DDH no. 5 to the San Rafael vein, 70 m) 75
D1-280 Tourmaline Tourmaline-quartz vein (DDH #1 to the San Rafael vein, 280 m) 75
D3-182 a Tourmaline Tourmaline-cassiterite-chlorite-arsenopyrite vein (early, orange tourmaline) 74
(DDH no. 3 to the Mariano vein, 182.5 m)

Main ore stage


D3-182 b Tourmaline Tourmaline-cassiterite-chlorite-arsenopyrite vein (late, green tourmaline, associated with 73
cassiterite) - DDH no. 3 to the Mariano vein, 182.5 m
R119 chl Chlorite Associated with quartz and needle-tin cassiterite, from a complexly banded quartz-chlorite 88
-cassiterite-wolframite-chalcopyrite-pyrite vein (Fig. 5)
R511 chl Chlorite Cassiterite-chlorite-quartz-arsenopyrite vein 81

Notes: Unless indicated otherwise, the samples are from the San Rafael vein and their location is shown in Figure 2

fluids circulating in the system were much cooler (260 are considered less reliable than the pressure-corrected fluid
380C, calculated from isochoric projection of fluid inclusion inclusion temperatures.
Th values) and had a moderate to low salinity (021 wt % The transition between the early, barren stage I and the
NaCl equiv). During the main cassiterite stage, fluid temper- subsequent tin and copper stages is recorded by rare tourma-
atures were in the range of 370380C, whereas quartz was line-cassiterite-chlorite-arsenopyrite veins, in which the early
deposited at consistently lower temperatures of 290330C. and main stage mineral assemblages overlap, indicating that
Because the vein textures indicate multiple repetition of cas- stage II may have occurred very soon after stage I (Mlynar-
siterite and quartz deposition, the observed temperature dif- czyk and Williams-Jones, 2006). We emphasize that these
ference between cassiterite- and quartz-hosted fluid inclu- veins are of the open fracture-filling type and are surrounded
sions likely reflects a sequence of distinct pulses of hotter by envelopes of strong chloritic alteration, identical to the
magmatic and cooler meteoric fluids that mixed repeatedly in ore-stage veins. This observation is very important, as it im-
a hydrologically very dynamic system. plies that the onset of ore deposition (stage II) and the
The quartz-cassiterite oxygen isotope temperatures calcu- marked change in vein and alteration mineralogy were di-
lated for this stage (400420C) are in reasonable agree- rectly related to a change in structural style, i.e., a transition
ment with the temperature estimate obtained from pressure- from a closed vein system (lithostatic conditions) to an open
corrected fluid inclusion data. However, as it was commonly vein system (hydrostatic conditions).
necessary to base the isotopic temperatures on cassiterite Although the tin stage (stage II) and the copper stage (stage
and quartz from adjacent layers, which as noted above III) were part of the same, protracted hydrothermal event,
formed at different temperatures, the isotopic temperatures and were associated with the same chloritic alteration, the
physico-chemical conditions of the ore fluid were markedly
different, and, therefore, resulted in the formation of con-
6 trasting mineral assemblages. In order to constrain the depo-
Calcite sitional conditions for each stage, we have thermodynamically
4 Siderite modeled them in fO2-pH and fO2-aS space, using phase dia-
grams combining aqueous and mineral equilibria in the Cu-
2 Fe-Sn-As-S-O-H system (Fig. 12). In addition, we have mod-
()

0 eled the speciation of tin in the system Sn-Na-Cl-O-H, to


V-PDB

constrain the total solubility of Sn as a function of pH, and the


-2 oxidation state of the hydrothermal fluid (Fig. 13). The details
of the calculation methods and the source of thermodynamic
13C

-4 data used are listed in the Appendix.


-6 As discussed above, the temperature and pressure during
cassiterite precipitation (stage II) were approximately 370
-8 to 380C and 200 to 300 bars, respectively, and the fluid had
a salinity of about 20 wt percent NaCl equiv. The equilibrium
-10
mineral assemblage was very simple and consisted of cassi-
5 10 15 20 25
terite, Fe-rich chlorite (daphnite), and quartz. Subordinate
18O () arsenopyrite, which occurs locally, was the only sulfide pre-
V-SMOW
sent, and in one sample minor hematite was observed to
FIG. 11. A plot of 13C versus 18O for different carbonates from the San grow coevally with cassiterite and chlorite (Mlynarczyk et al.,
Rafael deposit. The arrows link successive generations of siderite from the 2009, submitted). Based on the observed mineral assem-
same sample. The two data points in the lower right corner of the diagram
correspond to samples from the uppermost part of the deposit (elev 4820 m); blages and the activity diagrams (Fig. 12), it is concluded that
the remaining samples originate from or below an elevation of 4330 m. the conditions during precipitation of cassiterite evolved

0361-0128/98/000/000-00 $6.00 238


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 239

TABLE 5. Carbon and Oxygen Isotope Data for Vein Carbonates from the San Rafael Sn-Cu Deposit (n = 12)

Sample no. Mineral Textural relationships 13CV-PDB () 18OV-SMOW ()

Late, barren stage


D1-118 Calcite Quartz-calcite-sphalerite-chalcopyrite infill of a reopened tourmaline vein 8.5 8.6
(DDH no. 1 to the San Rafael vein, 118.5 m)
D5-137 Calcite Quartz-calcite infill of a reopened tourmaline vein 4.0 9.4
(DDH no. 5 to the San Rafael vein, 137.9 m)
D5-165 Calcite Quartz-calcite infill of a reopened tourmaline vein 7.8 9.6
(DDH no. 5 to the San Rafael vein, 165.8 m)
D4-230 Calcite Quartz-calcite-arsenopyrite infill of a reopened tourmaline vein 3.2 14.2
(DDH no. 4 to the San Rafael vein, 230.5 m)
D5M-155 Calcite Quartz-calcite infill of a reopened tourmaline vein 3.5 9.1
(DDH no. 5 to the Mariano vein, 155 m)
R27 sid1 Siderite Early, anhedral, reddish, opaque siderite from a quartz- 0.3 18.8
sphalerite-pyrite-marcasite-chalcopyrite-galena-siderite vein
R27 sid2 Siderite Late, euhedral, greenish, translucent siderite from a quartz- 0.0 15.7
sphalerite-pyrite-marcasite-chalcopyrite-galena-siderite vein
R28 sid Siderite Late, euhedral, greenish, translucent siderite from a quartz- 0.3 13.8
sphalerite-pyrite-marcasite-chalcopyrite-galena-siderite vein
R57 sid Siderite Late, orange, translucent siderite from a quartz-cassiterite- 0.3 17.7
chlorite-siderite vein
R123 sid Siderite Greenish, translucent siderite from a quartz-fluorite-pyrite- 4.3 15.7
marcasite-siderite vein
R708 sid2 Siderite Early, anhedral, reddish, opaque siderite from a quartz- 6.5 23.2
chalcopyrite-fluorite-siderite vein
R708 sid1 Siderite Late, euhedral, greenish, translucent siderite from a quartz- 4.9 22.0
chalcopyrite-siderite veinlet in chloritized wall rock

Notes: Unless indicated otherwise, the samples are from the San Rafael vein and their location is shown in Figure 2

from moderately reduced (log fO2 about 28 to 26, upper pyrrhotite was clearly late, the common presence of pyrite in
part of the arsenopyrite stability field) to more oxidizing (log stage III ores indicates that there must have been large fluc-
fO2 above 25, magnetite-hematite boundary). The pH tuations in fO2, and that the hydrothermal environment be-
evolved from acidic to more alkaline, with the final pH at the came progressively more oxidizing (i.e., moved from the
end of massive cassiterite precipitation being slightly above pyrrhotite to the pyrite stability field in Fig. 12c) with time.
about 4. At this pH value, the cassiterite solubility contours Furthermore, the ubiquitous replacement of pyrrhotite by
flatten out considerably (Fig. 13), i.e., the solubility decreases pyrite might indicate that the sulfidation state of the system
only very slightly upon further increase in pH. The sulfur con- increased with time (Fig. 12d), a feature that has been ob-
centration in the fluid is difficult to constrain, but a minimum served for many magmatic-hydrothermal systems (Einaudi et
value of aS of about 0.005 to 0.01 can be estimated from the al., 2003). Additional information about the oxidation state
stability field of arsenopyrite (Fig. 12b). and the fluid evolution during stage III mineralization comes
During the sulfide stage (stage III), the temperature and from the sulfur isotope data. The calculated sulfur isotope
pressure were approximately 300C (based on fluid inclusion contours show that an increase in oxidation state above a log
trapping temperatures and stannite-sphalerite thermometry) fO2 of 31 to 29 would result in a considerable shift in the
and 200 to 300 bars (assuming hydrostatic conditions and a 34S of precipitated sulfides towards more negative values
roughly similar depth to that during the previous mineraliza- (Fig. 12). From the relatively narrow range in 34S values of
tion stages), and the fluid salinity was around 5 wt percent the San Rafael sulfides it can be concluded that although the
NaCl equiv. The mineral assemblage consists mainly of chal- conditions during stage III became more oxidizing with time,
copyrite, pyrrhotite, Fe-rich chlorite (daphnite), quartz, and they did not reach as high as the pyrite-hematite boundary
minor cassiterite, but locally other sulfides (e.g., arsenopyrite, (Fig. 12c).
sphalerite, galena, pyrite) are also present. Figure 12 shows
the most relevant mineral stability relations, as well as the sul- Source of sulfur
fur isotope contours that were calculated for chalcopyrite (see The 34S values of sulfide minerals from the San Rafael de-
below). It can be deduced from the stability of chlorite (i.e., posit range between 2 and 6 per mil, and show relatively lit-
the daphnite end member as calculated from the measured tle variation with respect to the location in the deposit (see
chlorite composition) and pyrrhotite that during most of stage above). These rather uniform values point to a large-scale hy-
III, the oxidation state was lower (log fO2 below 35, upper drothermal system with a homogeneous source of sulfur,
daphnite stability limit in Fig. 12c) than during stage II, while which was likely of magmatic origin. As argued by Hattori and
the pH was likely similar (around 4). It should be noted, how- Keith (2001) for porphyry systems, the very narrow range in
ever, that in many locations in the deposit, pyrite is an abun- 34S values displayed by giant, granite-related deposits is best
dant sulfide of stage III mineralization, and pyrrhotite dis- explained by a single magmatic source of sulfur, rather than
plays an incipient or advanced replacement by pyrite, marcasite, by homogenization of the sulfur from a variety of sources in
or rare hematite. Although in some case this replacement of local country rocks having diverse S isotope compositions. It

0361-0128/98/000/000-00 $6.00 239


240 WAGNER ET AL.

Stability field of arsenopyrite T=380 C a Stability field of arsenopyrite T=380 C b


-18 P=300 bar -18 P=300 bar
HSO4- Qtz saturation
Qtz saturation
aH2O=0.873 aH2O=0.873
-20 (20wt.% NaCl eq.) -20 (20wt.% NaCl eq.)
aS = 0.01 pH=4

-22 SO42- -22 hematite


hematite
HSO4-
-24 pyrite -24 H2 S
Log fO2

Log fO2
pyrite
-26 -26
magnetite
magnetite asas
-28 -28 p

as
asp
-30 as -30 as
pyrrhotite loe loe pyrrhotite
-32 -32
as
asp p
loe
H2S loe HS-
-34 -34
0 2 4 6 8 10 12 -4 -3 -2 -1 0
pH Log aS
T=300 C c T=300 C d
P=300 bar P=300 bar
Stability field of chalcopyrite Qtz saturation Stability field of chalcopyrite Qtz saturation
-24 aH2O=0.973 -24 aH2O=0.973
HSO4 - prl ms kfs (5 wt.% NaCl eq.) (5 wt.% NaCl eq.)
ams=0.517 aK+= 0.005 pH=4
aS = 0.01 34Stot=+3.0
-26 -26
34Stot=+3.0
hematite
-28 -28 -17.5
HSO4-
hematite -15.0
-10.0
bn+py H2S -5.0
-30 -30 0.0 bn
cp SO42- +
+2.8 cp py
Log fO2
Log fO2

-32 pyrite -32 pyrite


-17.5
-15.0
-10.0
-5.0 magnetite
-34 0.0 -34
+2.8
+2.9

-36 -36 pyrrhotite


+3.0

dap
adap=0.18 magnetite
pyrrhotite
-38
cp
-38 bn cp
+p
bn+po HS-
o
H2S
-40 -40
0 2 4 6 8 10 12 -5 -4 -3 -2 -1
pH Log aS

FIG. 12. Log fO2-pH and log fO2-aS diagrams showing stability relationships in the system Fe-O-S, for stage II (oxide as-
semblage, diagrams a and b) and stage III (sulfide assemblage, diagrams c and d) of the paragenesis. The temperature and
fluid salinity for which the diagrams were drawn were constrained from fluid inclusion microthermometry, whereas pressure
was estimated. Diagrams a and b also show stability relationships for the Fe-As-S system (asp = arsenopyrite, as = native ar-
senic, loe = loellingite). Diagrams c and d show the stability relationships in the Cu-Fe-S system (cp = chalcopyrite, bn =
bornite, py = pyrite) and of the principal alteration minerals (dap = daphnite, prl = pyrophyllite, ms = muscovite, kfs = K
feldspar). The activities of alteration minerals were calculated from microprobe data (see Appendix for details), and the
chemical reactions were balanced assuming conservation of Al and quartz saturation. Diagrams c and d (sulfide stage) also
show superimposed S isotope contours, calculated for chalcopyrite, assuming aS = 0.01 and 34Stotal= 3.0 , inferred from
the S isotope composition of the sulfides from this stage (see text for details).

0361-0128/98/000/000-00 $6.00 240


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 241

1.0 20
SMOW
0

10-1
log fO2 = - 32 -20
Total Sn solubility (mol/kg)

Tertiary Andean Groundwater

DV-SMOW ()
meteoric water ?
- 31 -40 Magmatic
water box
10-2

e
- 30

r lin
-60

ate
Present-day
- 29

cw
Andean
-80 meteoric

ori
10-3 water
Chlorite
Tourmaline

te
- 28

Me
- 27 -100
- 22 - 26
10-4 - 23 - 24 - 25 -120
-20 -10 0 10 20
18O ()
V-SMOW

10-5
3.0 3.5 4.0 4.5 5.0 5.5 6.0
qz wol cas tou
pH
FIG. 13. Total solubility of tin (mol/kg) in the system Sn-Na-Cl-O-H cal-
culated as a function of pH and fO2 at 380C and 300 bar (the inferred con-
-20 -10 0 10 20
ditions for stage II). For details of the calculations, see the Appendix.
chl
18O
should be noted though that the San Rafael sulfides have Ore fluid ()
slightly higher 34S values than some other tin deposits (Kon- FIG. 14. Plot of 18O versus D for tourmaline (open squares) and chlo-
tak, 1990), implying an initial relative enrichment in the 34S rite (open circles) from the San Rafael deposit and calculated water in iso-
content of the sulfur source. This is not unexpected for S-type topic equilibrium with these minerals (black squares and circles). The thick
granitic magmas, which acquire most of their sulfur through dashed trendline from tourmaline to chlorite represents the inferred evolu-
tion of fluid (water) during the formation of the deposit. The lower diagram
assimilation of country rocks. Thus, for such plutons, the 34S (18O axis) shows the distribution of the calculated 18O values for water in
values of ore sulfides are very similar to those of local igneous equilibrium with the different ore and gangue minerals (cas = cassiterite, chl
rocks and local country rocks, and reliable discrimination be- = chlorite, qz = quartz, tou = tourmaline, wol = wolframite). The calculations
tween magmatic and sedimentary sulfur is not possible applied isotopic fractionation equations that are mostly based on well-con-
(Ohmoto and Goldhaber, 1997). strained experiments in the appropriate temperature range (Matsuhisa et al.,
1979; Graham et al., 1984, 1987; Carothers et al., 1988; Kotzer et al., 1993;
Source of ore fluids Zheng, 1993; Zhang et al., 1994; Cole and Ripley, 1999).

We have calculated the oxygen and hydrogen isotope com-


position of the ore fluids (Fig. 14) during different mineral- oxygen isotope compositions of successive generations of
ization stages to constrain the most likely fluid sources for the quartz and cassiterite (see earlier section), and also by miner-
San Rafael deposit. The isotopic composition of the early flu- als which formed late in the paragenesis, such as calcite and
ids, which were in equilibrium with stage I tourmaline, is very siderite.
close to magmatic water values, which, considering the high The decrease in ore fluid 18O with time is paralleled by
temperature and salinity of these fluids, supports the idea that marked corresponding decreases in fluid salinity and temper-
they were magmatic in origin. The water in equilibrium with ature, as indicated by the fluid inclusion data (Fig. 7). These
late, Fe-rich tourmaline, which formed at the onset of cassi- trends, which appear broadly aligned, could represent mixing
terite precipitation (single analysis in the upper right corner of the early, hot, hypersaline (presumably magmatic) brine
of the magmatic water box, Fig. 14) however, was enriched in with a relatively warm fluid that had consistently lower 18O
deuterium and had isotopically lighter oxygen. Compared to and D values. Considering the shallow level of emplacement
the tourmaline data, the chlorite data demonstrate a fluid of the San Rafael pluton and fluid inclusion evidence for very
evolution characterized by a marked decrease in 18O and a low salinity fluids (this could not be produced by simple boil-
slight increase in D with time. The water in equilibrium with ing of a magmatic brine; Kontak and Clark, 2002) circulating
ore-stage chlorite had 18O lower by about 7 per mil and D in the system, a good candidate for this external fluid would
higher by 10 per mil than that of the fluid in equilibrium with be heated groundwater of meteoric origin that had partly ex-
early tourmaline (Fig. 14). It should be noted, however, that changed its oxygen isotope composition with the host rocks.
the D of water in equilibrium with chlorite cannot be pre- Unfortunately, it is currently not possible to properly con-
cisely constrained due to the lack of accurate fractionation strain the isotopic composition of the late Oligocene, Andean
factors (Graham et al., 1987). The trend of decreasing fluid meteoric waters, coeval with formation of the deposit, to fur-
18O with time is, nevertheless, clearly indicated by the ther substantiate this hypothesis. This is because of the very

0361-0128/98/000/000-00 $6.00 241


242 WAGNER ET AL.

large uncertainty concerning the rate of the uplift in this part of fluid salinity and temperature (above 360C and about 21
of the Andes during the last 25 m.y. It is generally assumed wt % NaCl equiv during formation of massive layers of cassi-
that the elevation of the Central Andes in the late Oligocene terite, but below 300C and 216 wt % NaCl equiv during
was about a third of their current elevation (Gregory-Wodz- formation of intervening quartz layers) is readily explained by
icki, 2000; Anders et al., 2002), implying that ancient mete- incursions of heated groundwaters of meteoric origin which
oric waters must have had considerably higher 18O and D mixed with magmatic fluids (Mlynarczyk et al., 2003). A
values than their present-day counterparts, but the actual val- model invoking the mixing of magmatic fluids with waters
ues can only be roughly estimated (Fig. 14). Applying both having lower 18O and D could also explain the observed
closed- and open-system scenarios (Taylor, 1977, 1997) for systematic decrease of cassiterite and quartz 18O values with
oxygen isotope exchange between water of meteoric origin time. Assuming that hot groundwaters of meteoric origin
and typical Andean granites (e.g., Longstaffe et al., 1983), the would not have had sufficient time to fully equilibrate with
calculated 18O values of hot groundwater at temperatures of the host rocks, they likely would be cooler and much more ox-
200 to 250C (assuming moderate fluid/rock ratios between idizing than the magmatic fluids. Numerical simulation of
0.1 and 1.0) would be in the range between 2 and 2 per mil. complex fluid-granite equilibria (Dolejs and Wagner, 2008)
has shown that rock-buffered low-salinity aqueous fluids will
Ore deposition processes become much more oxidizing as temperature decreases from
The rich tin ores of the San Rafael deposit testify to an un- 400 to 200C. Therefore, mixing of hot groundwaters with
usually effective mechanism of ore deposition, which caused tin-bearing magmatic fluids would have oxidized the fluid sys-
substantial supersaturation of tin and focused mineralization tem, increased the overall pH, and decreased temperature
into several large fault-jogs, at depth in the lode (Mlynarczyk and ligand activity, all of which would have destabilized tin
et al., 2003). Experimental studies and chemical modeling chloride complexes and triggered cassiterite precipitation.
have shown that in natural hydrothermal systems the solubil- A further evaluation of boiling and fluid mixing scenarios
ity of tin is highest in hot, reduced, saline, and acidic solutions and their relative importance is possible from the evolution of
(Fig. 13), and that the bulk of the tin is transported as stan- the oxygen isotope composition of the ore fluid. Boiling in-
nous (Sn2+) chloride complexes (Eugster and Wilson, 1985; creases the 18O of the remaining liquid because of preferen-
Pabalan, 1986; Taylor and Wall, 1993; Mller and Seward, tial partitioning of the light isotope into the vapor, and this is
2001). The precipitation of cassiterite (SnO2), in which tin is reflected in the paragenesis by a gradual increase in 18O with
in the tetravalent state, can, therefore, be induced by an in- time. By contrast, mixing of magmatic waters with groundwa-
crease in fO2 as well as by decreases in temperature and ligand ters of meteoric origin (isotopically light) should lead to pro-
(chloride) ion activity and an increase in pH. Possible mech- gressive decreases in the 18O values of the ore fluid. It
anisms for tin ore formation could include reaction of the ore should be noted, however, that the temperature decrease as-
fluids with the host rocks, boiling, redox-coupled precipita- sociated with mixing will increase the isotopic fractionation
tion, mixing of the ore fluids with fluids of a markedly differ- between the ore fluid and the precipitating minerals, which,
ent composition, or a combination of some of the above (Ead- in the case of the oxygen isotope composition of cassiterite
ington, 1985; Heinrich, 1990, 1995). and wolframite, will partly offset the effect of addition of low-
The fluid-rock interaction, which accompanied ore deposi- 18O groundwaters. Because of such complexities, quantita-
tion at San Rafael, produced a wide envelope of very strong tive modeling of boiling and fluid mixing mechanisms of tin
chloritization but is not considered to have influenced ore for- ore formation are required to discriminate between them.
mation as mass-balance calculations show that such alteration
increases the acidity of the fluid (Mlynarczyk et al., 2009, sub- Modeling of fluid evolution scenarios
mitted), leading to an increase in cassiterite solubility (Fig. To test the hypothesis of fluid mixing and compare it to one
13). Redox-coupled precipitation (e.g., coupled precipitation of boiling of the ore fluid, the oxygen isotope composition of
of arsenopyrite and cassiterite) is also not considered to have silicate and oxide minerals precipitating by each mechanism
played a significant role because it would have led to consid- was calculated as a function of temperature. Because the
erable heterogeneity in the compositions of the ore and alter- available experimental studies of isotopic liquid-vapor parti-
ation minerals, which was clearly not the case. tioning in the H2O-NaCl system are restricted to conditions
On the other hand, structural evidence for a transition from below the critical point of pure water (Horita et al., 1993,
a closed to an open vein system, and a concomitant change 1995), it was not possible to adequately model boiling for the
from lithostatic to hydrostatic conditions, implies that boiling temperature-composition space at San Rafael. Based on boil-
may have taken place. Boiling would have oxidized the ore ing models for low-salinity fluids (e.g., Truesdell and Nathen-
fluid, decreased its temperature and significantly increased its son, 1977; Wagner et al., 2005), it can be predicted that boil-
pH due to the loss of acidic volatiles. Thus, it would have pro- ing would produce a progressive increase in mineral 18O
moted the precipitation of cassiterite. However, vapor-rich values. This would be the consequence of preferential parti-
fluid inclusions are rare and are only found in stage 1 veins tioning of the lighter 16O isotope into the vapor phase (Horita
that formed prior to tin mineralization. Thus, if boiling oc- and Wesolowski, 1994), yielding an increasingly 18O-rich
curred during the main ore-forming events, it was probably of residual brine. Furthermore, the effect would be enhanced
limited extent. By contrast, the coincidence of the ore stage by the cooling of the residual brine as a consequence of en-
with the opening of the vein system and the widespread thalpy transfer from the liquid to the vapor that is required to
appearance of cooler, much more dilute fluids (down to al- maintain the enthalpy balance. Hence, the first-order effect
most zero salinity), as well as the ensuing periodic fluctuation of boiling on the 18O of the precipitated minerals is broadly

0361-0128/98/000/000-00 $6.00 242


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 243

similar to that of fluid cooling and can therefore be simulated with time are shown in Figure 15, together with the measured
by the latter. ranges of 18O values and fluid inclusion trapping tempera-
Based on the fluid inclusion data and the measured oxygen tures for ore stage quartz and cassiterite. From the calcula-
isotope data for stage I silicates, the parameters of the start- tions, it is evident that boiling and fluid mixing have a con-
ing magmatic fluid were assumed to be a temperature of trasting influence on the oxygen isotope composition of the
500C, a salinity of 45 wt percent NaCl equiv, and a 18O precipitating minerals. Whereas mixing with hot, low-18O
value of 11 per mil. One model considered that this fluid fluid produces a systematic decrease in mineral 18O values
cooled due to boiling down to a temperature of 200C, (followed by a minor increase below 250C), boiling results in
whereas the other considered its mixing with a cooler, dilute a systematic increase of the 18O values of the oxides and sil-
fluid (230C, zero salinity), assuming three possible 18O icates with time, a scenario clearly not supported by the par-
compositions of 2.0, 0.0, 2.0 percent representing the com- agenetic information and the isotopic data for San Rafael. In
position of hot groundwaters of meteoric origin that had ex- addition to the good agreement between the predicted and
changed their oxygen with their host rocks. A stepwise de- observed trend of decreasing mineral 18O values, the mixing
crease of ore fluid temperature took place in each model and model also reproduces closely the relatively narrow range of
in the case of mixing was accompanied by a decrease in salin- mineral 18O compositions observed in the deposit (Fig. 15).
ity and fluid 18O, proportional to the aliquot of diluting fluid It is, however, noteworthy that this is only the case when the
added. For simplicity, a linear relationship between the tem- diluting fluid is assumed to be relatively hot (230C), as mix-
perature and composition of the fluid mixture was assumed, ing of the magmatic brine with cold (25100C), dilute water
although it is actually the enthalpy, not the temperature, that would drive the mineral 18O compositions toward very high
varies linearly with composition during mixing (e.g., Reed and values.
Spycher, 1984; Spycher and Reed, 1989). The influence of The clear lack of a trend of increasing mineral 18O values
fluid salinity on the concentration of oxygen was taken into at San Rafael (apart from that observed in the lower-temper-
account (Schwinn et al., 2006) because the salinity of the ature stage IV, which is consistent with the mixing model),
magmatic end member was quite high but, in the absence of therefore, precludes a significant role for boiling in cassiterite
pertinent experimental data, the mineral-water fractionation deposition. By contrast, the mixing of a magmatic brine with
factors were assumed to be salinity independent. The latter heated, dilute groundwaters of meteoric origin reproduces
were calculated for every temperature step (using the equa- remarkably well the observed 18O composition of the ore
tions of Matsuhisa et al., 1979; Zhang et al., 1994; and Cole and gangue minerals. In our model, fluid mixing was simu-
and Ripley, 1999), enabling the 18O composition of quartz, lated as a continuous process to capture the first-order ef-
cassiterite, chlorite, and wolframite in equilibrium with the fects, whereas the complex banding of the veins with multiple
fluid to be determined. generations of the main ore and gangue minerals suggests
Selected modeling results are presented in Table 6 and the that fluid mixing occurred repeatedly in a hydrologically very
calculated changes in the 18O values of quartz and cassiterite dynamic system.

TABLE 6a. Calculated Oxygen Isotope Composition of Quartz, Cassiterite, Chlorite and Wolframite, in Equilibrium
with a Magmatic Brine (T = 500C, 18O = 11.0), which is Mixing with Meteoric Water (T = 230C, 18O = 2.0)

Tmix (C) 18O (fluid) 18O (mineral)


F = mass
Fraction brine Quartz Cassiterite Chlorite Wolframite

1.00 500 11.0 13.3 6.7 10.1 7.8


0.95 487 10.4 13.0 6.2 9.5 7.3
0.90 473 9.9 12.6 5.7 9.0 6.8
0.85 460 9.4 12.3 5.2 8.5 6.3
0.80 446 8.9 12.0 4.7 8.0 5.8
0.75 433 8.4 11.8 4.3 7.5 5.4
0.70 419 7.9 11.6 3.9 7.0 5.0
0.65 406 7.4 11.4 3.5 6.6 4.6
0.60 392 6.9 11.2 3.1 6.2 4.2
0.55 379 6.5 11.0 2.8 5.8 3.8
0.50 365 6.0 10.9 2.6 5.4 3.5
0.45 352 5.6 10.9 2.3 5.1 3.2
0.40 338 5.1 10.8 2.1 4.8 2.9
0.35 325 4.7 10.8 2.0 4.5 2.6
0.30 311 4.3 10.8 1.9 4.3 2.4
0.25 298 3.9 10.9 1.9 4.1 2.1
0.20 284 3.5 11.0 1.9 4.0 1.9
0.15 271 3.1 11.2 2.0 3.9 1.8
0.10 257 2.7 11.4 2.1 3.9 1.6
0.05 244 2.4 11.6 2.4 3.9 1.5
0.00 230 2.0 12.0 2.7 4.0 1.5

0361-0128/98/000/000-00 $6.00 243


244 WAGNER ET AL.

TABLE 6b. Calculated Oxygen Isotope Composition of Quartz, Cassiterite, Chlorite and Wolframite, in Equilibrium
with a Magmatic Brine (T = 500C, 18O = 11.0), Which is Progressively Cooling

18O (fluid) 18O (mineral)

T (C) Quartz Cassiterite Chlorite Wolframite

500 11.0 13.3 6.7 10.1 7.8


480 11.0 13.6 6.8 10.1 7.8
460 11.0 13.9 6.8 10.1 7.9
440 11.0 14.3 6.9 10.1 8.0
420 11.0 14.7 7.0 10.2 8.1
400 11.0 15.1 7.1 10.2 8.2
380 11.0 15.6 7.4 10.3 8.4
360 11.0 16.1 7.6 10.5 8.5
340 11.0 16.7 8.0 10.7 8.7
320 11.0 17.3 8.4 10.9 8.9
300 11.0 18.0 8.9 11.2 9.2
280 11.0 18.7 9.5 11.6 9.5
260 11.0 19.6 10.3 12.0 9.8
240 11.0 20.5 11.2 12.7 10.3
230 11.0 21.0 11.8 13.0 10.5

Concluding Remarks I). The modeling of tin solubility suggests that these fluids
The hydrothermal history of the San Rafael vein system could have transported high concentrations of tin, but they
began with hot, acidic, saline and reducing fluids, which had did not deposit any cassiterite.
an isotopic composition very close to that of typical magmatic The onset of cassiterite deposition came with a major
fluids. These fluids likely exsolved from a late granitic melt re- change in the plumbing of the hydrothermal system, inferred
lated to the San Rafael igneous center and produced exten- to correspond to a change from lithostatic to hydrostatic pres-
sive sericitic and tourmaline alteration of the wall rocks (stage sure conditions, upon a major reopening of the San Rafael
lode. The change in structural style, indicated by the ubiqui-
tous open fracture-filling character of stage II-IV veins, was
22 associated with a drop in temperature and major changes in
20 a
fluid chemistry. This is evident from the strong chloritic al-
18 teration and the massive deposition of cassiterite and, subse-
Boiling (cooling)
16 quently, sulfide minerals. The ore fluids were considerably
()

Stage IV quartz
14
Stage II-III quartz
less saline and somewhat cooler than the early brines, were
+2.0
more oxidizing, and had lower 18O and D values.
Quartz

12
0.0 The precipitation of cassiterite was most likely caused by a
10
18O

8 -2.0 marked increase in the oxidation state of the ore fluid (sup-
Mixing ported by evidence of rare hematite), an increase in pH, and
6 High-T fluid:
4
Low-T fluid:
O
500 C
O
to a lesser extent, decreases in temperature and chloride ac-
230 C 45 wt.% NaCl eq.
2 0 wt.% NaCl eq. 18O=+11 tivity. The two depositional mechanisms that could explain the
geologic evidence are boiling of the magmatic brine or its mix-
0
250 300 350 400 450 500 ing with cooler, oxidizing meteoric waters. Quantitative mod-
TOre fluid (C) eling of mineral-water oxygen fractionation predicts the ob-
served decrease of fluid 18O values with time for the case of
18 fluid mixing, whereas this modeling suggests that boiling will
16 b produce the opposite trend in 18O values. Also, the general
14
()

12
10
FIG. 15. Diagrams showing the effect of boiling (cooling) and mixing of
Cassiterite

Boiling (cooling)
8 magmatic and meteoric water derived fluids on the oxygen isotope composi-
6 tion of quartz (a) and cassiterite (b). Because no experimental data are avail-
4 Stage II-III cassiterite able for isotopic liquid-vapor fractionation in the H2O-NaCl system at condi-
18O

+2.0 tions above the critical point of water, the boiling model has only considered
2 0.0 the cooling effect (resulting from enthalpy transfer from the liquid to the
0 -2.0 Mixing vapor). Therefore, the boiling curves are labeled as boiling (cooling). The
-2 starting conditions for the magmatic brine were a temperature of 500C, a
salinity of 45 wt percent NaCl equiv, and a 18O value of 11 per mil. Mixing
-4 lines were calculated assuming that groundwater had a temperature of
250 300 350 400 450 500 230C, zero salinity and 18O values between 2.0 and 2.0 per mil (labels on
TOre fluid (C) the mixing lines).

0361-0128/98/000/000-00 $6.00 244


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 245

evolution of the hydrothermal system toward domination by on the metallogenic evolution of the Andes of southeastern Peru: ECO-
NOMIC GEOLOGY, v. 85, p. 15201583.
strongly dilute, cooler fluids is consistent with an incursion of
Clayton, R.N., and Mayeda, T.K., 1963, The use of bromine pentafluoride in
hot groundwaters of meteoric origin. As the bulk of the tin the extraction of oxygen from oxides and silicates for isotopic analysis:
ore at San Rafael is restricted to large fault-jogs at depth in Geochimica et Cosmochimica Acta, v. 27, p. 4352.
the lode, we conclude that these jogs provided sites favorable Cole, D.R., and Ripley, E.M., 1999, Oxygen isotope fractionation between
for repeated episodic mixing between magmatic and meteoric chlorite and water from 170 to 350C: A preliminary assessment based on
partial exchange and fluid/rock experiments: Geochimica et Cosmochimica
fluids, which resulted in unusually efficient, structurally fo- Acta, v. 63, p. 449457.
cused cassiterite deposition. Collins, P.L.F., 1981, The geology and genesis of the Cleveland tin deposit,
Western Tasmania: Fluid inclusion and stable isotope studies: ECONOMIC
Acknowledgments GEOLOGY, v. 76, p. 365-392.
The authors wish to thank Ing. Fausto Zavaleta Cruzado, Dolejs, D., and Wagner, T., 2008, Thermodynamic modeling of non-ideal
mineral-fluid equilibria in the system Si-Al-Fe-Mg-Ca-Na-K-H-O-Cl at el-
General Manager of MINSUR S.A., Ing. Luis Alva Florian evated temperatures and pressures: Implications for hydrothermal mass
and Ing. Otto Velarde Junes, successive managers of the San transfer in granitic rocks. Geochimica et Cosmochimica Acta, v. 72, p.
Rafael mine, Ing. Julver Alvarez Romero, Ing. Pastor Luque 526554.
Malag, Ing. Ladislao Guilln Cardenas, and Ing. Luis San- Dolejs, D., Mlynarczyk, M.S.J., and Williams-Jones, A.E., 2009, S-type pera-
talla Medrano, mine geologists, as well as other staff of the luminous granitic rocks associated with a world-class tin deposit: a petro-
genetic study of the San Rafael stock, southeastern Peru. J. Petrol. (sub-
San Rafael mine, for logistic support and helpful collabora- mitted)
tion during the field work. The assistance of Harold E. Waller Eadington, P.J., 1985, The solubility of cassiterite in hydrothermal solutions
and Nestor Roldan from EMINASA and Mario Arenas in relation to some lithological and mineral associations of tin ores, in Tay-
Figueroa, consulting geologist, were much appreciated. The lor, R.P., and Strong, D.F., eds., Recent advances in the geology of granite-
related mineral deposits: Canadian Institution of Mining and Metallurgy,
authors would also like to thank the reviewers, B. Lehmann Special Volume 39, p. 2532.
and A. Campbell, as well as associate editor J. Muntean and Einaudi, M.T., Hedenquist, J.W., and Inan, E.E., 2003, Sulfidation state of
Editor L. Meinert for their constructive comments, which fluids in active and extinct hydrothermal systems: Transitions from por-
helped to improve this paper. The research was funded by phyry to epithermal environments: SEG Special Publications, v. 10, p.
NSERC and FQRNT grants to AEW-J and NERC support of 285313.
Eugster, H.P., and Wilson, G.A., 1985, Transport and deposition of ore-form-
the Isotope Community Support Facility at SUERC. ing elements in hydrothermal systems associated with granites, in Halls, C.,
ed., High heat production (HHP) granites, hydrothermal circulation and
November 6, 2008; February 18, 2009 ore genesis: Institution of Mining and Metallurgy Conference, London, p.
8798.
REFERENCES Goldstein, R.H., and Reynolds, T.J., 1994, Systematics of fluid inclusions in
Aja, S.U., 2002, The stability of Fe-Mg chlorites in hydrothermal solutions. diagenetic minerals: Society of Sedimentary Geology Short Course, v. 31,
II. Thermodynamic properties: Clays and Clay Minerals, v. 50, p. 591600. 199 p.
Aja, S.U., and Dyar, D.M., 2002, The stability of Fe-Mg chlorites in hy- Graham, C.M., Atkinson, J., and Harmon, R.S., 1984, Hydrogen isotope frac-
drothermal solutions. I. Results of experimental investigations: Applied tionation in the system chlorite-water [abs.]: Natural Environment Re-
Geochemistry, v. 17, p. 12191239. search Council (NERC) Publication Series D, no. 25, p. 139.
Alderton, D.H.M., 1989, Oxygen isotope fractionation between cassiterite Graham, C.M., Viglino, J.A., and Harmon, R.S., 1987, Experimental study of
and water: Mineralogical Magazine, v. 53, p. 373376. hydrogen isotope exchange between aluminous chlorite and water and of
Alderton, D.H.M., and Harmon, R.S., 1991, Fluid inclusion and stable iso- hydrogen diffusion in chlorite: American Mineralogist, v. 72, p. 566579.
tope evidence for the origin of mineralizing fluids in southwest England: Gregory-Wodzicki, K.M., 2000, Andean paleoelevation estimates: A review
Mineralogical Magazine, v. 55, p. 605611. and critique: Geological Society of America Bulletin, v. 112, p. 10911105.
Anders, M.H., Gregory-Wodzicki, K.M., and Spiegelman, M., 2002, A criti- Grnvold, F., and Stlen, S., 1992, Thermodynamics of iron sulfides. II. Heat
cal evaluation of late Tertiary accelerated uplift rates for the Eastern capacity and thermodynamic properties of FeS and of Fe0.875S at tempera-
Cordillera, Central Andes of Bolivia: Journal of Geology, v. 110, p. 89100. tures from 298.15 K to 1000 K, of Fe0.98S from 298.15 K to 800 K, and of
Arenas, M.J., 1980, Mapa geologico superficial del Distrito Minero San Fe0.89S from 298.15 K to about 650 K. Thermodynamics of formation: Jour-
Rafael, Puno: Unpublished Report MINSUR Archives. nal of Chemical Thermodynamics, v. 24, p. 913936.
Bakker, R.J., 2003, Package FLUIDS. 1. Computer programs for analysis of Harrison, T.M., Duncan, I., McDougall, I., 1985, Diffusion of 40Ar in biotite:
fluid inclusion data and for modelling bulk fluid properties: Chemical Ge- Temperature, pressure and compositional effects: Geochimica et Cos-
ology, v. 194, p. 323. mochimica Acta, v. 49, p. 24612468.
Barton, P.B., 1969, Thermochemical study of the system Fe-As-S: Geochim- Haynes, F.M., 1985, Determination of fluid inclusion compositions by se-
ica et Cosmochimica Acta, v. 33, p. 841857. quential freezing: ECONOMIC GEOLOGY, v. 80, p. 14361439.
Bodnar, R.J., 1993, Revised equation and table for determining the freezing Hattori, K.H., and Keith, J.D., 2001, Contribution of mafic melt to porphyry
point depression of H2O-NaCl solutions: Geochimica et Cosmochimica copper mineralization: Evidence from Mount Pinatubo, Philippines, and
Acta, v. 57, p. 683684. Bingham Canyon, Utah, USA: Mineralium Deposita, v. 36, p. 799806.
Bortnikov, N.S., Zaozerina, O.N., Genkin, A.D., and Muravitskaya, G.N., Heinrich, C.A., 1990, The chemistry of hydrothermal tin (-tungsten) ore de-
1990, Stannite-sphalerite intergrowthspossible indicators of conditions position: ECONOMIC GEOLOGY, v. 85, p. 457481.
of ore deposition: International Geology Review, v. 32, p. 11321144. 1995, Geochemical evolution and hydrothermal mineral deposition in
Cathelineau, M., 1988, Cation site occupancy in chlorites and illites as a func- Sn (-W-base metal) and other granite-related ore systems: some conclu-
tion of temperature: Clay Minerals, v. 23, p. 471485. sions from Australian examples, in Thompson, J.F.H., ed., Magmas, fluids
Carothers, W.W., Adami, L.H., and Rosenbauer, R.J., 1988, Experimental oxy- and ore deposits: Mineralogical Association of Canada (MAC) Short
gen isotope fractionation between siderite-water and phosphoric acid liber- Course, v. 23, p. 203220.
ated CO2-siderite: Geochimica et Cosmochimica Acta, v. 52, p. 24452450. Heinrich, C.A., and Eadington, P.J., 1986, Thermodynamic predictions of the
Clark, A.H., Palma, V.V., Archibald, D.A., Farrar, E., Arenas, M.J., and hydrothermal chemistry of arsenic, and their significance for the parage-
Robertson, R.C.R., 1983, Occurrence and age of tin mineralization in the netic sequence of some cassiterite-arsenopyrite-base metal sulfide de-
Cordillera Oriental, Southern Peru: ECONOMIC GEOLOGY, v. 78, p. 514520. posits: ECONOMIC GEOLOGY, v. 81, p. 511529.
Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., Arenas Figueroa, M.J., Holland, T.J.B., and Powell, R., 1998, An internally consistent thermody-
France, L.J., McBride, S.L., Woodman, P.L., Wasteneys, H.A., Sandeman, namic data set for phases of petrological interest: Journal of Metamorphic
H.A., and Archibald, D.A., 1990, Geologic and geochronologic constraints Geology, v. 16, p. 309343.

0361-0128/98/000/000-00 $6.00 245


246 WAGNER ET AL.

Holland, T., Baker, J., and Powell, R., 1998, Mixing properties and activity- Longstaffe, F.J., Clark, A.H., McNutt, R.H., and Zentilli, M., 1983, Oxygen
composition relationships of chlorites in the system MgO-FeO-Al2O3-SiO2- isotopic compositions of Central Andean plutonic and volcanic rocks, lati-
H2O: European Journal of Mineralogy, v. 10, p. 395406. tudes 26-29 south: Earth and Planetary Science Letters, v. 64, p. 918.
Horita, J., and Wesolowski, D.J., 1994, Liquid-vapor fractionation of oxygen Matsuhisa, Y., Goldsmith, J.R., and Clayton, R.N., 1979, Oxygen isotopic
and hydrogen isotopes of water from the freezing point to the critical tem- fractionation in the system quartz-albite-anorthite-water: Geochimica et
perature: Geochimica et Cosmochimica Acta, v. 58, p. 34253447. Cosmochimica Acta, v. 43, p. 11311140.
Horita, J., Wesolowski, D.J., and Cole, D.R., 1993, The activity-composition Mattey, D.P., and Macpherson, C.G., 1993, High-precision oxygen isotope
relationship of oxygen and hydrogen isotopes in aqueous salt solutions. I. microanalysis of ferromagnesian minerals by laser fluorination: Chemical
Vapor-liquid water equilibration of single salt solutions from 50 to 100C: Geology, v. 105, p. 305318.
Geochimica et Cosmochimica Acta, v. 57, p. 27972817. McBride, S.L., Robertson, R.C.R., Clark, A.H., Farrar, E., 1983, Magmatic
Horita, J., Cole, D.R., and Wesolowski, D.J., 1995, The activity-composition and metallogenic episodes in the northern tin belt, Cordillera Real, Bolivia:
relationship of oxygen and hydrogen isotopes in aqueous salt solutions. III. International Journal of Earth Sciences, v. 72, p. 685713.
Vapor-liquid equilibration of NaCl solutions to 350 C: Geochimica et Cos- McCrea, J.M., 1950, On the isotopic chemistry of carbonates and a
mochimica Acta, v. 59, p. 11391151. palaeotemperature scale: Journal of Chemical Physics, v. 18, p. 849857.
Jackson, P., Changkakoti, A., Krouse, H.R., and Gray, J., 2000, The origin of Menzie, W.D., Reed, B.L., and Singer, D.A., 1988, Models of grades and ton-
greisen fluids of the Foleys zone, Cleveland tin deposit, Tasmania, Aus- nages of some lode tin deposits, in Hutchison, C.S., ed., Geology of tin de-
tralia: ECONOMIC GEOLOGY, v. 95, p. 227236. posits in Asia and the Pacific; mineral concentrations and hydrocarbon ac-
Johnson, J.W., Oelkers, E.H., and Helgeson, H.C., 1992, SUPCRT92: A soft- cumulations in the ESCAP region: Selected papers from the International
ware package for calculating the standard molal thermodynamic properties Symposium on the Geology of Tin Deposits, Nanning, China, October
of minerals, gases, aqueous species, and reactions from 1 to 5000 bar and 0 1984, p. 7388.
to 1000 C: Computers and Geosciences, v. 18, p. 899947. Migdisov, A.A., Williams-Jones, A.E., Lakshtanov, L.Z., and Alekhin, Y.V.,
Jowett, E.C., 1991, Fitting iron and magnesium into the hydrothermal chlo- 2002, Estimates of the second dissociation constant of H2S from the surface
rite geothermometer: GAC-MAC-SEG Joint Meeting, Program with Ab- sulfidation of crystalline sulfur: Geochimica et Cosmochimica Acta, v. 66, p.
stract, p. A62. 17131725.
Kelley, S.P., and Fallick, A.E., 1990, High precision spatially resolved analy- Mlynarczyk, M.S.J., and Williams-Jones, A.E., 2005, The role of collisional
sis of 34S in sulphides using as laser extraction technique: Geochimica et tectonics in the metallogeny of the Central Andean tin belt: Earth and
Cosmochimica Acta, v. 54, p. 883888. Planetary Science Letters, v. 240, p. 656667.
Kelley, S.P., Fallick, A.E., McConville, P., and Boyce, A.J., 1992, High preci- 2006, Zoned tourmaline associated with cassiterite: implications for
sion, high spatial resolution analysis of sulfur isotopes by laser combustion fluid evolution and tin mineralization in the San Rafael Sn-Cu deposit, SE
of natural sulfide minerals: Scanning Microscopy, v. 6, p. 129138. Peru: Canadian Mineralogist, v. 44, p. 347365.
Kelly, W.C., and Turneaure, F.S., 1970, Mineralogy, paragenesis and geot- Mlynarczyk, M.S.J., Sherlock, R.L., and Williams-Jones, A.E., 2003, San
hermometry of tin and tungsten deposits of the eastern Andes, Bolivia: Rafael, Peru: geology and structure of the worlds richest tin lode: Mineral-
ECONOMIC GEOLOGY, v. 65, p. 609680. ium Deposita, v. 38, p. 555567.
Kojima, S., and Sugaki, A., 1984, Phase relations in the Cu-Fe-Zn-S system Mlynarczyk, M.S.J., Wagner, T., Williams-Jones, A.E., and Dolejs, D., 2009,
between 500 and 300C under hydrothermal conditions: ECONOMIC GE- Geology and geochemistry of alteration at the San Rafael Sn-Cu deposit,
OLOGY, v. 80, p. 158171. SE Peru: ECONOMIC GEOLOGY (in review)
Kontak, D.J., 1990, A sulfur isotope study of main-stage tin and base metal Mller, B., and Seward, T.M., 2001, Spectrophotometric determination of
mineralization at East Kemptville tin deposit, Yarmouth County, Nova Sco- the stability of tin (II) chloride complexes in aqueous solution up to 300C:
tia, Canada: evidence for magmatic origin of metals and sulfur: Economic Geochimica et Cosmochimica Acta, v. 65, p. 41874199.
Geology, v. 85, p. 399-407. Nakamura, Y., and Shima, H., 1982, Fe and Zn partitioning between spha-
Kontak, D.J., and Clark, A.H., 1988, Exploration criteria for tin and tungsten lerite and stannite [abs.]: Joint Meeting of the Society of Mining Geology
mineralization in the Cordillera Oriental of southeastern Peru, in Taylor, of Japan, Association of Mineralogy, Petrology and Economic Geology, and
R.P., and Strong, D.F., eds., Recent advances in the geology of granite-re- Mineralogical Society of Japan, p. A8.
lated mineral deposits: Canadian Institution of Mining and Metallurgy, Nekrasov, I.Y., Sorokin, V.I., and Osadchii, E.G., 1976, Partition of iron and
Special Volume 39, p. 157169. zinc between sphalerite and stannite at T = 300 to 500C and P = 1kb: Dok-
2002, Genesis of the giant, bonanza San Rafael lode tin deposit, Peru: lady Earth Sciences, v. 226, p. 136138.
origin and significance of pervasive alteration: Economic Geology, v. 97, p. 1979, Fe and Zn partitioning between stannite and sphalerite and its ap-
1741-1777. plication in geothermometry: Physics and Chemistry of the Earth, v. 11, p.
Kontak, D.J., Clark, A.H., Farrar, E., Pearce, T.H., Strong, D.F., and Baads- 739742.
gaard, H., 1986, Petrogenesis of a Neogene shoshonite suite, Cerro Moro- Oelkers, E.H., and Helgeson, H.C., 1990, Triple-ion anions and polynuclear
moroni, Puno, southeastern Peru: Canadian Mineralogist, v. 24, p. compexing in supercritical electrolyte solutions: Geochimica et Cos-
117135. mochimica Acta, v. 54, p. 727738.
Kontak, D.J., Clark, A.H., Farrare, E., Archibald, D.A., and Baadsgaard, H., Ohmoto, H., 1972, Systematics of sulfur and carbon isotopes in hydrothermal
1987, Geochronological data for tertiary granites of the southeast Peru seg- ore deposits: ECONOMIC GEOLOGY, v. 67, p. 551578.
ment of the Central Andean tin belt: ECONOMIC GEOLOGY, v. 82, p. Ohmoto, H., and Goldhaber, M.B., 1997, Sulfur and carbon isotopes, in
16111618. Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed:
Kontak, D.J., Cumming, G.L., Krstic, D., Clark, A.H., and Farrar, E., 1990, New York, Wiley, p. 517611.
Isotopic composition of lead in ore deposits of the Cordillera Oriental, Pabalan, R.T., 1986, Solubility of cassiterite (SnO2) in NaCl solutions from
southeastern Peru: ECONOMIC GEOLOGY, v. 85, p. 15841603. 200C350C, with geologic applications: Unpublished Ph.D. thesis, Penn-
Kotzer, T.G., Kyser, T.K., King, R.W., and Kerrich, R., 1993, An empirical sylvania State University, 141 p.
oxygen- and hydrogen-isotope geothermometer for quartz-tourmaline and Palma, V.V., 1981, The San Rafael tin-copper deposit, Puno, SE Peru: Un-
tourmaline-water: Geochimica et Cosmochimica Acta, v. 57, p. 34213426. published M.Sc. thesis, Kingston, Canada, Queens University, 235 p.
Laubacher, G., 1978, Estudio geologico de la region norte del Lago Titicaca: Pashinkin, A.C., Muratova, V.A., Moiseyev, N.V., and Bazhenov, J.V., 1991,
Peru, Instituto Geologia y Mineria, v. 5, 120 p. Heat capacity and thermodynamic functions of iron diarsenide in the T
LeBoutillier, N.G., Camm, G.S., Shail, R.K., Bromley, A.V., Jewson, C., and range 5 K to 300 K: Journal of Chemical Thermodynamics, v. 23, p.
Hoppe, N., 2002, Tourmaline-quartz-cassiterite mineralization of the 827830.
Lands End granite at Nanjizal, west Cornwall: Proceedings of the Ussher Primmer, T.J., 1985, Discussion on the possible contribution of metamorphic
Society, v. 10, p. 312318. water to the mineralising fluid of south-west England: Preliminary stable
Lehmann, B., 1990, Metallogeny of tin: Lecture Notes in Earth Sciences, v. isotope evidence: Proceedings of the Ussher Society, v. 6, p. 224228.
32, Berlin,Springer-Verlag, 211 p. Reed, M.H., and Spycher, N.F., 1984, Calculation of pH and mineral equi-
Linnen, R.L., and Williams-Jones, A.E., 1995, Genesis of a magmatic meta- libria in hydrothermal waters with application to geothermometry and
morphic hydrothermal system: The Sn-W polymetallic deposits at Pilok, studies of boiling and dilution: Geochimica et Cosmochimica Acta, v. 48, v.
Thailand: ECONOMIC GEOLOGY, v. 90, p. 11481166. 14791492.

0361-0128/98/000/000-00 $6.00 246


STABLE ISOTOPE CONSTRAINTSSAN RAFAEL SN-CU DEPOSIT, SE PERU 247

Robie, R.A., and Hemingway, B.S., 1995, Thermodynamic properties of min- Sun, S., and Eadington, P.J., 1987, Oxygen isotope evidence for the mixing of
erals and related substances at 298.15 K and 1 Bar (105 Pascals) pressure magmatic and meteoric waters during tin mineralization in the Mole gran-
and at higher temperatures: U.S. Geological Survey Monograph, v. 2131, ite, New South Wales, Australia: ECONOMIC GEOLOGY, v. 82, p. 4352.
461 p. Sverjensky, D.A., Shock, E.L., and Helgeson, H.C., 1997, Prediction of the
Robie, R.A., Seal, R.R. II, and Hemingway, B.S., 1994, Heat capacity and en- thermodynamic properties of aqueous metal complexes to 1000C and 5
tropy of bornite (Cu5FeS4) between 6 and 760 K and the thermodynamic kb: Geochimica et Cosmochimica Acta, v. 61, p. 13591412.
properties of phases in the system Cu-Fe-S: Canadian Mineralogist, v. 32, Taylor, H.P., 1977, Water/rock interactions and the origin of H2O in granitic
p. 945956. batholiths: Journal of the Geological Society London, v. 133, p. 509558.
Roedder, E., 1984, Fluid inclusions: Reviews in Mineralogy, v. 12, 646 p. 1997, Oxygen and hydrogen isotope relationships in hydrothermal min-
Rosenbaum, J., and Sheppard, S.M.F., 1986, An isotopic study of siderites, eral deposits, in Barnes, H.L., ed., Geochemistry of hydrothermal ore de-
dolomites and ankerites at high temperatures: Geochimica et Cosmochim- posits, 3rd ed.: New York, Wiley, p. 229302.
ica Acta, v. 50, p. 11471150. Taylor, J.R., and Wall, V.J., 1993, Cassiterite solubility, tin speciation and
Ryzhenko, B.N., Shvarov, Y.V., and Kovalenko, N.I., 1997, The Sn-Cl-F-C-S- transport in a magmatic aqueous phase: ECONOMIC GEOLOGY, v. 88, p.
H-O-Na system: Thermodynamic properties of components within the 437460.
conditions of the Earths crust: Geochemistry International, v. 35, p. Taylor, R.G., 1979, Geology of tin deposits: Developments in Economic Ge-
10161020. ology, v. 11, 543 p.
Sandeman, H.A., Clark, A.H., and Farrar, E., 1995, An integrated tectono- Truesdell, A.H., and Nathenson, M., 1977, The effects of boiling and dilution
magmatic model for the evolution of the Southern Peruvian Andes (13- on the isotopic compositions of Yellowstone thermal waters: Journal of
20S) since 55 Ma: International Geology Review, v. 37, p. 10391073. Geophysical Research, v. 82, p. 36943704.
Sandeman, H.A., Clark, A.H., Farrar, E., and Arroyo-Pauca, G., 1996, A crit- Turneaure, F.S., 1960a, A comprarative study of major ore deposits in Bolivia:
ical appraisal of the Cayconi Formation, Crucero Basin, southeastern Peru: ECONOMIC GEOLOGY, v. 55, p. 217254.
Journal of South American Earth Sciences, v. 9, p. 381392. 1960b, A comparative study of major ore deposits in Bolivia. Part II:
Schwinn, G., Wagner, T., Markl, G., and Baatartsogt, B., 2006, Quantification ECONOMIC GEOLOGY, v. 55, p. 574606.
of mixing processes in ore-forming hydrothermal systems by combination Vidal, O., Parra, T., and Trottet, F., 2001, A thermodynamic model for Fe-Mg
of stable isotope and fluid inclusion analyses: Geochimica et Cosmochim- aluminous chlorite using data from phase equilibrium experiments and nat-
ica Acta, v. 70, p. 965982. ural pelitic assemblages in the 100-600C, 1-25 kb range: American Jour-
Sharp, Z.D., 1990, A laser-based microanalytical method for the in situ de- nal of Science, v. 301, p. 557592.
termination of oxygen isotope ratios of silicates and oxides: Geochimica et Wagner, T., Boyce, A.J., and Fallick, A.E., 2002, Laser combustion analysis of
Cosmochimica Acta, v. 54, p. 13531357. 34S of sulfosalt minerals: Determination of the fractionation systematics
Sharp, Z.D., Atudorei, V., and Durakiewicz, T., 2001, A rapid method for de- and some crystal-chemical considerations: Geochimica et Cosmochimica
termination of hydrogen and oxygen isotope ratios from water and hydrous Acta, v. 66, p. 28552863.
minerals: Chemical Geology, v. 178, p. 197210. Wagner, T., Boyce, A.J., Jonsson, E., and Fallick, A.E., 2004, Laser micro-
Sheppard, S.M.F., 1994, Stable isotope and fluid inclusion evidence for the probe sulphur isotope analysis of arsenopyrite: experimental calibration
origin and evolution of hercynian mineralizing fluids, in Seltmann, R., and application to the Boliden Au-Cu-As massive sulphide deposit: Ore
Kmpf, H., and Mller, P., eds., Metallogeny of collisional orogens: Czech Geology Reviews, v. 25, p. 311325.
Geological Survey, Prague, p. 4960. Wagner, T., Williams-Jones, A.E., and Boyce, A.J., 2005, Stable isotope-based
Shimitzu, M., and Shikazono, N., 1985, Iron and zinc partitioning between modeling of the origin and genesis of an unusual Au-Ag-Sn-W epithermal
coexisting stannite and sphalerite: A possible indicator of temperature and system at Cirotan, Indonesia: Chemical Geology, v. 219, p. 237260.
sulfur fugacity: Mineralium Deposita, v. 20, p. 314320. Walshe, J.L., 1986, A six-component chlorite solid-solution model and the
Shock, E.L., Oelkers, E.H., Johnson, J.W., Sverjensky, D.A., and Helgeson, conditions of chlorite formation in hydrothermal and geothermal systems:
H.C., 1992, Calculation of the thermodynamic properties of aqueous ECONOMIC GEOLOGY, v. 81, p. 681703.
species at high pressures and temperatures: Journal of the Chemical Soci- Walshe, J.L., Halley, S.W., Anderson, J.A., and Harrold, B.P., 1996, The in-
ety Faraday Transactions, v. 88, p. 803826. terplay of groundwater and magmatic fluids in the formation of the cassi-
Shock, E.L., Sassani, D.C., Willis, M., and Sverjensky, D.A., 1997, Inorganic terite-sulfide deposits of western Tasmania: Ore Geology Reviews, v. 10, p.
species in geological fluids: Correlations among standard molal thermody- 367387.
namic properties of aqueous ions and hydroxide complexes: Geochimica et Wiggins, L.B., and Craig, J.R., 1980, Reconnaissance of the Cu-Fe-Zn-S sys-
Cosmochimica Acta, v. 61, p. 907950. tem: sphalerite phase relationship: ECONOMIC GEOLOGY, v. 75, p. 742751.
Shvarov, Y.V., 1978, Minimization of the thermodynamic potential of an open Williamson, B.J., Spratt, J., Adams, J.T., Tindle, A.G., and Stanley, C.J., 2000,
chemical system: Geochemistry International, v. 15, p. 200203. Geochemical constraints from zoned hydrothermal tourmalines on fluid
1981, A general equilibrium criterion for an isobaric-isothermal model evolution and tin mineralization: an example from fault breccias at Roche,
of a chemical system: Geochemistry International, v. 18, p. 3845. SW England: Journal of Petrology, v. 41, p. 14391453.
Shvarov, Y.V., and Bastrakov, E., 1999, HCh: A software package for geo- Wilkinson, J.J., Jenkin, G.R.T., Fallick, A.E., and Foster, R.P., 1995, Oxygen
chemical equilibrium modeling. Users guide: Australian Geological Survey and hydrogen isotopic evolution of Variscan crustal fluids, south Cornwall,
Organization, 61 p. U.K: Chemical Geology, v. 123, p. 239254.
Smith, M., Banks, D.A., Yardley, B.W.D., and Boyce, A., 1996, Fluid inclu- Zhang, L., Liu, J., Chen, Z., and Zhou, H., 1994, Experimental investigations
sion and stable isotope constraints on the genesis of the Cligga Head Sn-W of oxygen isotope fractionation in cassiterite and wolframite: ECONOMIC
deposit, S.W. England: European Journal of Mineralogy, v. 8, p. 961974. GEOLOGY, v. 89, p. 150157.
Spycher, N., and Reed, M.H., 1989, Evolution of a Broadlands-type epither- Zhang, X., and Spry, P.G., 1994, FO2PH: a quickBASIC program to calculate
mal ore fluid along alternative P-T paths: Implications for the transport and mineral stabilities and sulphur isotope contours in log fO2-pH space: Min-
deposition of base, precious, and volatile metals: ECONOMIC GEOLOGY, v. eralogy and Petrology, v. 50, p. 287291.
84, p. 328359. Zheng, Y.F., 1993, Calculation of oxygen isotope fractionation in hydroxyl-
Sterner, S.M., Hall, D.L., and Bodnar, R.J., 1988, Synthetic fluid inclusions. bearing silicates: Earth and Planetary Science Letters, v. 120, p. 247263.
V. Solubility relations in the system NaCl-KCl-H2O under vapor-saturated
conditions: Geochimica et Cosmochimica Acta, v. 52, p. 9891005.
Sugaki, A., Kitakaze, A., and Kojima, S., 1990, Sphalerite stars in chalcopy-
rite; are they always the results of an unmixing process; discussion: Miner-
alium Deposita, v. 25, p. 8283.

0361-0128/98/000/000-00 $6.00 247


248 WAGNER ET AL.

APPENDIX

Thermodynamic Calculations from the original sources (pyrrhotite: Grnvold and Stlen,
We modeled the depositional conditions prevalent during 1992; bornite: Robie et al., 1994) were fitted with a four-term
the main cassiterite stage and the late sulfide stage at San polynominal of the form used by Holland and Powell (1998).
Rafael by constructing phase diagrams showing aqueous and It should be noted that the entropy values for bornite and
mineral equilibria in the Cu-Fe-Sn-As-S-O-H system and pyrrhotite given in Robie and Hemingway (1995) also origi-
contours of sulfur isotope fractionation. In addition, we car- nate from Grnvold and Stlen (1992) and Robie et al.
ried out a series of speciation calculations in the model system (1994), thereby ensuring internal consistency of these datasets.
Sn-Na-Cl-O-H in order to obtain the total solubility of Sn as Data for calculating the stability fields of arsenic phases (ar-
a function of pH and the oxidation state of the hydrothermal senopyrite, loellingite, native arsenic) were compiled from
fluid. All the calculations were carried out with the HCh soft- Robie and Hemingway (1995) and Barton (1969). The heat
ware package (Shvarov and Bastrakov, 1999), which models capacity function for loellingite was fitted to the original data
heterogeneous equilibria and reaction progress by minimiza- given in Pashinkin et al. (1991), which were the source of the
tion of the Gibbs free energy of the total system (Shvarov, entropy data tabulated in Robie and Hemingway (1995). All
1978, 1981). Thermodynamic data for most aqueous species calculations of individual activity coefficients of aqueous
were taken from the SUPCRT92 database and subsequent species applied an extended Debye-Hckel model using the
updates (Johnson et al., 1992; Shock et al., 1997; Sverjensky b-gamma equation for NaCl as the background electrolyte
et al., 1997). Data for a number of aqueous Sn species came (Oelkers and Helgeson, 1990; Shock et al., 1992). The set of
from Ryzhenko et al. (1997). Thermodynamic data for rock- equations of Zhang and Spry (1994) and the most recent set
forming silicate and oxide minerals were taken from the in- of isotopic fractionation factors given in Ohmoto and Gold-
ternally consistent dataset of Holland and Powell (1998). haber (1997) were used to calculate the sulfur isotope con-
The data for cassiterite, pyrite, chalcopyrite, bornite, and tours. The model of Zhang and Spry (1994), which excludes
pyrrhotite (not contained in this dataset) were compiled from the aqueous S2 species, is preferred over the original formal-
Robie and Hemingway (1995). As heat capacity functions for ism of Ohmoto (1972), because the second dissociation con-
pyrrhotite and bornite are not given in Robie and Hemingway stant of H2S is too small for S2 to be a significant species at
(1995), the experimentally determined heat capacity data geologically realistic values of pH (Migdisov et al., 2002).

0361-0128/98/000/000-00 $6.00 248

Das könnte Ihnen auch gefallen