Vous êtes sur la page 1sur 38

Economic Geology Vol. 97, 2002, pp.

17411777

Genesis of the Giant, Bonanza San Rafael Lode Tin Deposit, Peru: Origin and Significance of Pervasive Alteration
DANIEL J. KONTAK
Nova Scotia Department of Natural Resources, P.O. Box 698, Halifax, Nova Scotia B3J 2T9, Canada
AND

ALAN H. CLARK

Department of Geological Sciences and Geological Engineering, Queens University, Kingston, Ontario K7L 3N6,Canada

Abstract
The San Rafael Sn-Cu deposit, located at lat 1413'58" S, long 7019'18" W, on the upper slopes of the Cordillera de Carabaya, northern Puno Department, southeastern Peru, yielded 38,182 t of fine tin in 2001 and, with reserves of 14.460 Mt at 5.14 percent Sn, the main San Rafael Lode constitutes, both at present and historically, the largest-known bonanza-grade cassiterite repository. Sn and subordinate Cu mineralization were confined to laterally and vertically extensive brittle shear zones generated by regional tectonism that transect a small 24.65 0.20 Ma, epizonal stock of strongly peraluminous, Lachlan S-type cordierite-biotite monzogranite and granodiorite. Early, barren, stage I quartz-tourmaline veins and breccias were emplaced at 24.10 0.15 Ma by high-temperature (580C), boiling, saline fluids that plausibly directly exsolved from the granitic melt, whereas both cassiterite (stage II) and, at higher elevations, chalcopyrite (stage III) ores were precipitated from largely nonboiling, cooler (Th = 215420C), and less saline (020 wt % NaCl equiv) fluids at 21.9 to 22.7 circa 0.5 Ma. Only narrow (2 m) envelopes of chlorite and weaker sericite and silica are associated with stages II and III veins, but more than 80 percent of the upper 750 m of the San Rafael stock experienced quasipervasive hydrothermal alteration. Early K metasomatism is recorded by widespread replacement of magmatic plagioclase and perthitic alkali feldspar by Na-free orthoclase, while broadly coeval alteration to albite more erratically affected groundmass plagioclase and the rims of alkali feldspar phenocrysts. Despite the intensity of the alkali metasomatism, only minor redistribution of dispersed, magmatic Sn occurred, largely during alteration to orthoclase. Other petrographically similar granitic bodies in the region experienced more intense Na metasomatism, and they host argentian base metal rather than Sn-dominated veins. Superimposed hydrolytic alteration, most intense within 20 to 30 m of the San Rafael Lodes, converted biotite to Fe-rich chlorite and converted alkali feldspar and plagioclase to fine-grained muscovite, but there is no significant dispersed cassiterite or chalcopyrite. Secondary fluid inclusions trapped in quartz phenocrysts of the granitic rocks range in salinity from 0 to 65 wt percent NaCl equiv and in Th from 200 to 530C; fluid boiling is evident in the higher-temperature populations. First-melting temperatures and decrepitate analyses reveal, in addition to NaCl, widely variable concentrations of KCl, CaCl2 (attaining 40 wt % of the solute) and, apparently most abundant in proximity to the lodes, FeCl2. Entrapment pressures are estimated to have fluctuated in the range 150 to 615 bars for much of the hydrothermal history, but fluid overpressuring to at least 2 kbar occurred, probably cyclically, during the evolution of the hydrothermal system. The salinity vs. homogenization temperature range for the phenocrysthosted inclusions closely matches that of primary inclusions in quartz and cassiterite in the lodes and, remarkably, small clusters of the former record the entire thermal and compositional trajectory of the mineralizing fluids. Early magmatic brines are inferred to have permeated much of the stock whereas, circa 2 m.y. later, Snand Cu-rich, lower-temperature, lower-pH fluids were largely channeled along the evolving shear zones where they mixed with cool, nonsaline, tectonically driven ground water. In conjunction with fluid neutralization through hydrolytic alteration, this channeling resulted in catastrophic precipitation of botryoidal and coarsely crystalline cassiterite and, subsequently, chalcopyrite. The hiatus between initial retrograde boiling and ore deposition, a feature documented in several world-class, high-grade lithophile-element deposits, may record protracted storage of metal-rich brines at depth, perhaps in association with small volumes of highly fractionated melt, such as is represented by small bodies of tourmaline leucogranite exposed on the upper margin of the San Rafael stock. However, the parallel evolution in temperature and salinity exhibited by the early nonmineralizing and later fertile hydrothermal fluids may also indicate that Sn and Cu were largely introduced into the subjacent magma chamber at a late stage.

Introduction HYDROTHERMAL alteration assemblages have long been recognized as providing a record of the compositional evolution and migration paths of magmatic-hydrothermal aqueous fluids, as well as enlarged and, ideally, systematically zoned targets in

Corresponding author: e-mail, charbon@geol.queensu.ca

mineral exploration. Predictably, therefore, they have been extensively documented (e.g., Lentz, 1994). Less attention has been paid to the interrelationships between the nature and intensity of alteration and the scale of ore metal concentration although, in the lithophile-metal field, the subject of the present discussion, Eugster and Wilson (1985) and, in more detail, Halter et al. (1996) modeled the deposition of

0361-0128/01/3307/1741-37 $6.00

1741

1742

KONTAK AND CLARK

cassiterite as a direct result of proton metasomatism (greisening) of quartzofeldspathic host rocks and concomitant fluid neutralization. In such systems, a broad correlation between the intensity and/or extent of feldspar-destructive alteration and Sn ore-grade and/or tonnage would be predicted. Heinrich (1990), however, emphasized the limitations inherent in such a cassiterite precipitation mechanism, and the deposits (e.g., East Kemptville, Nova Scotia) on which the modeling was based are indeed either small or low grade (Kontak et al., 1995), and they would probably not now constitute realistic exploration targets. Heinrich also noted that such constraints could be overridden by the abrupt mixing of a Sn-bearing magmatic brine with cool meteoric water, under which circumstances there are essentially no chemical limitations on tin ore grade..... (op. cit., p. 457). Herein we document aspects of the hydrothermal alteration associated with a giant, bonanza lode tin deposit, i.e., San Rafael, southeastern Peru, by far the most important single hard-rock source of cassiterite during the past decade. Minsur S.A.s underground mine yielded 37,409 and 38,182 tonnes of fine tin-in-concentrate in 2000 and 2001, respectively, and it has reported in situ reserves (Dec. 31, 2001) of 14.460 Mt at 5.14 percent Sn and 0.04 percent Cu, at a cutoff grade of 2.63 percent Sn; the mineable ore grades are probably at least 5.6 percent Sn. This study represents part of an ongoing investigation (Palma, 1981; Kontak, 1985; Yamamura, 1990; Sandeman, 1995) of the geologic history, metallogeny, and ore genesis of the southeast Peru segment of the central Andean inner arc (Clark et al., 1983a, 1984). An overview of this region and the great variety of ore deposits therein is provided by Clark et al. (1990a). The petrogenesis of parental magmas, the structural and mineralogic evolution of the lode system, and the stable isotopic chemistry of the ores and host rocks are topics that await further investigation. The granitoid rocks that host the mineralized structures at San Rafael superficially appear unaltered but are extensively and intensely metasomatized. As emphasized by Kontak and Clark (1988), the field recognition of such cryptic alteration permits a preliminary evaluation of the economic potential of the numerous felsic intrusions of the region. The alteration both preceded and accompanied the deposition of cassiterite, and it developed quasipervasively through at least 1.5 km3 of the epizonal San Rafael granite stock. However, no dispersed, porphyry-style Sn mineralization has been documented (cf. Sillitoe et al., 1975), and the genetic relationships between widespread alteration and structurally focused ore deposition require definition, representing a key facet of the ore-genetic model for an exceptionally rich hydrothermal center. The geology of this exceptional deposit has been outlined by Arenas (1980), Palma and Clark (1982), Clark et al. (1990a) and Kontak et al. (1995). Arenas (1999) provides a succinct overview but does not document the alteration that is the main subject of the present contribution. Because the tectonic and petrogenetic setting and evolution of the San Rafael Lodes differ markedly from those of the widely documented greisen-associated cassiterite vein systems (e.g., Halter et al., 1996), we summarize the salient features of the deposit and its surroundings to provide a basis for discussion of the alteration relationships.
0361-0128/98/000/000-00 $6.00

The San Rafael Deposit The San Rafael deposit is centered at lat 1413' 58" S., long 7019'18" W, in northern Puno Department, southeast Peru (Fig. 1a). The vein system crops out at altitudes of up to 5,000 m a.s.l. on the southern slopes of the Cordillera de Carabaya, a segment of the central Andean Cordillera Oriental. The deposit lies close to the apparent northwestern termination of the Bolivian Sn-Ag belt (Fig. 1a), which flanks the eastern border of the Altiplano (Turneaure, 1971). All economic cassiterite deposition occurred within a series of broadly northwest-striking brittle shear zones (Fig. 1b) that transect the San Rafael stock (Arenas, 1980; Clark et al., 1983b). This small epizonal granitic body, which has an outcrop of less than 1 km2 on the glaciated southern slopes of Nevado San Bartolom de Quenamari, the southwest spur of Nevado San Francisco de Quenamari (5,284 m a.s.l.), is a member of the upper Oligocene-lower Miocene (2226 Ma) Picotani intrusive suite, the hypabyssal component of the volcano-sedimentary Picotani Group, which records the initiation of Cenozoic magmatism in the region (Clark et al., 1990a; Sandeman et al., 1997). It is hosted (Palma, 1981) by clastic rocks of the Ordovician Sandia Formation, which had experienced prehnitepumpellyite facies regional metamorphism and polyphase deformation in the Early Carboniferous, early Hercynian orogeny (Laubacher, 1978; Clark et al., 1990b). Southeast of the mine, the contact between the lower Paleozoic metaclastic rocks and the overlying diagenetic-grade (A.H. Clark, unpub. data) and openly folded Mississippian Ambo Group is interpreted, on the basis of minor structures in the former (A.H. Clark, unpub. data), as a thrust fault that dips southsoutheastsoutheast at circa 15 to 20(Fig.1b). Steeply dipping, broadly northwest-striking transcurrent faults delimit the eastern and western boundaries of the inlier (Fig. 1b), and the San Rafael stock and lode system were thus emplaced into a tectonic dome or arch. Intrusion of granitoid magma took place in the latest Oligocene, at 24.6 to 24.7 0.2 Ma (respectively, 206Pb/238U single zircon and monazite dates: Clark et al., 2000), while 40 Ar/39Ar age spectra show that the upper 500 m of the stock cooled through the Ar-retention temperature of biotite (ca. 310C : Harrison et al., 1985) at 23.7 0.2 Ma (all dates herein are quoted at 2 error). The intrusion is dominated by coarse, porphyritic monzogranitic and finer-grained, less porphyritic monzogranitic to granodioritic facies (Fig. 1b, c), both characterized by phenocrysts of alkali feldspar, quartz, biotite and, a striking feature in outcrop, pinitized cordierite. The strongly peraluminous granitoid rocks conform closely to Chappell and Whites (1974) original S-type clan, termed Lachlan S-type by Sylvester (1998) and peraluminous S-type by Patio-Douce (1999), in contradistinction to lower-temperature and more hydrous Himalayan-type muscovite-rich leucogranites. Ore-metal zonation The mineralized structures in the San Rafael mine (Fig. 1b), among which the Veta San Rafael (San Rafael Lode) is by far the most important, possess (Arenas, 1980; Palma, 1981; Clark et al., 1983b) a dramatic vertical ore-metal zonation (Fig. 2a), comparable to that of the lodes of the

1742

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1743

a)
12S 72W 68 200 km

b)

N
0 km 2

Study area
14

Bolivia Peru
16

Sn -A g
B

t el

18
Pacific Ocean

Altiplano

Chile

Veta San Rafael

QG
Mina Quenamari

moraines, scree, ice tourmaline leucogranite granitic dikes fine- to medium-grained monzogranite/granodiorite coarse-grained monzogranite Ambo Group (Mississippian) Sandia Formation (Ordovician)

SRG

Mina San Rafael

faults

veins

c)
Quenamari
213 206 217,218 244 195,194,196 203 202 197 198 190 239 187A

1 km
154 403,406 156 248 249A 150,151 165,166A 408 140,141 250,252 193

San Rafael

FIG. 1. (a) Location of the San Rafael Sn-Cu deposit at the apparent northwest extremity of the central Andean Sn-Ag-polymetallic belt. (b) Simplified geologic map of the San Rafael-Quenamari mining district (after Arenas, 1980, 1999; Palma, 1981; A.H. Clark, unpub. data). The San Rafael (SRG) and Quenamari (QG) stocks are considered to represent cupolas on a single steep-walled pluton, circumscribed by the granitic ring dikes. The coarsegrained facies is largely monzogranitic, while the finer straddles the monzogranite-granodiorite field boundary. The largely north-northweststriking lodes and veins developed during thrusting of the Ambo Group over the Sandia Formation. The northwest-verging thrust in the southeast sector is poorly exposed but its location and sense of displacement are in conformity with structures in the underlying Sandia Formation rocks. (c) Sketch maps showing the locations of surface granitoid samples specifically cited herein. N.B. San Rafael samples 140, 141, and 156 are from the small bodies of tourmaline leucogranite at the margin of the main stock. The majority of underground samples were taken from the 4533 m level. Surface samples in which fluid inclusions were studied in most detail are in italic type.

major Camborne-Redruth district of the Cornubian province, southwest England (Dines, 1956; Clark et al., 1995), in which a chalcopyrite-rich zone overlies one dominated by cassiterite. This clear relationship is not known in other major deposits in the central Andean tin belt and is indeed unusual globally. Sphalerite, galena, and Ag sulfides and sulfosalts, as well as ankerite, siderite, and calcite, are modestly enriched at the lateral, particularly the southern (Cuerpo Polimetalico: Fig. 2a), extremities of the vein system, but the vein becomes
0361-0128/98/000/000-00 $6.00

narrow and discontinuous in the hornfelsic rocks adjacent to the stock. The San Rafael Lode has been developed throughout a vertical interval of more than a kilometer, from immediately above 5,000 m to 3,950 m a.s.l., and along a horizontal distance of 2.4 km (Fig. 2a). It yielded chalcopyrite concentrates above circa 4730 m, but almost all recent mining has been restricted to the underlying Sn zone, where Sn/Cu ratios exceed 25 (Arenas, 1999). Palma (1981) and Palma and Clark (1982) demonstrated that the ore-metal zoning records the

1743

1744

KONTAK AND CLARK

a) SSE NNW

500 m
5

5000 m (a.s.l.)

San Rafael Stock 3 1


4.6% t@4.6% 1.6Mt@ 6 6.0% 1.9 1.9Mt@

Sandia Fm.

5 3
Cuerpo Polimetalico (Pb,Zn,Cu, Sn,Ag) Cuerpo Contacto Sur

4500 m

1
t@ 3.2% 0.8Mt@3.2%
% % t 6.2 t@ 6.2 0.6M

5.8Mt@5.5 %

150-S Brecha Rampa 410

150 310-S Ore Shoot

250-S Cuerpo Contacto Brecha

4000 m

Cu grade contours (%) (%)

b)
437
35 7 ,2 0

N
0m 431 420

Ore Shoot Brecha

0-E

8,4 28 ,20 0N

0m

0m
8,4 28 ,40 0N

100m

FIG. 2. (a) Longitudinal section of the San Rafael Lode showing Cu isograde lines and the main Sn orebodies, each of which is located within a dilational fault-jog controlled by combined sinistral-transcurrent and reverse displacement during paragenetic stage II (see Fig. 3; simplified after Minsur S.A., unpub. reports, 2000). (b) Level plans (42004370 m) showing the form of the Ore Shoot and contiguous Brecha orebodies on the San Rafael Lode. The Ore Shoot comprises, from northeast to southwest, a high-grade hanging wall cassiterite-quartz-chlorite vein; polyphase hydrothermal breccias; central stockwork; hydrothermal breccias; and a weak footwall vein (simplified after Minsur S.A., unpub. reports, 2000; R. Mason and D.J. Kerr, unpub. report to Minsur S.A., 2000).

juxtaposition of earlier cassiterite-dominated and later chalcopyrite-rich assemblages (Fig. 3), concentrated in, respectively, the deeper and shallower intervals of the lode. The San Rafael Lode system occupies the southwest half of a mining district that also incorporates the much smaller Quenamari, or Carabaya, mine, 3.5 km to the northeast (Fig. 1b). Chalcopyrite-cassiterite ore occurs in the latter area but is subordinate to Ag-rich chalcopyrite-sphalerite-galena(-carbonate) veins comparable to the distal zones of the San Rafael Lode (Hannington, 1983; Kontak, 1985; Clark et al., 1990a). Ore at Quenamari is associated with a series of granitoid stocks and dikes (Fig. 1b), petrographically similar to, but probably representing a shallower erosion level than, the San Rafael stock. The San Rafael and Quenamari granites almost certainly represent cupolas of a single pluton. Several other Sn-base metal vein systems, coeval with San Rafael, are exposed on the southern slopes of the Cordillera de Carabaya (Clark et al., 1983b, 1990a, Kontak et al., 1987) but, with the possible exception of Santo Domingo, 22 km west of San
0361-0128/98/000/000-00 $6.00

Rafael (Bateman, 1982), none has been shown to be of economic importance. The large and high-grade Palca 11 (Regina) ferberite-scheelite lode (Clark et al., 1990a; Yamamura, 1991), formerly mined in the eastern part of the broader Cordillera de Carabaya district, is coeval with the San Rafael deposit (Farrar et al., 1990), but only traces of scheelite and wolframite have been confirmed from the San Rafael Lode (Palma, 1981; A.H. Clark, unpub. data), although the darker zones of cassiterite grains are enriched in W as well as in Fe (M.S.J. Mlynarczyk, unpub. report for Minsur S.A., 1999). Alteration-mineralization relationships The paragenetic evolution of the San Rafael Lode system, first established by Palma (1981) and modified by A.H. Clark (unpub. data), is illustrated in Figure 3. The initial mineralization in the San Rafael stock, assigned to stage I (Palma, 1981; Fig. 3), generated swarms of tourmaline-quartz veins and hydrothermal breccias, with minor arsenopyrite, lllingite, and hexagonal pyrrhotite (Palma, 1981; A.H. Clark, unpub. data)

1744

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1745

Hypogene Paragenetic Relationships of San Rafael Lode and its Immediate Envelope
Tourmaline Quartz Arsenopyrite Lllingite Pyrrhotite Muscovite Chlorite Cassiterite (botryoidal) Cassiterite (coarse-grained) Scheelite Wolframite Chalcopyrite Pyrite Cassiterite (acicular) Sphalerite Galena Stannite Acanthite Matildite Bismuth Adularia Marcasite Siderite Ankerite Calcite Amethyst Fluorite Stibnite Gudmundite Mackinawite

24.10 0.15 Ma

STAGE I

22.4 - 22.7 0.7 Ma

STAGE II

21.9 - 22.0 0.2 Ma

STAGE III

STAGE IV

FIG. 3. Paragenetic and age relationships of the San Rafael and associated lodes (Palma, 1981; Kontak, 1985; A.H. Clark, unpub. data). N.B. Stages I (precursor tourmaline-quartz veins) and II (main cassiterite veins) are shown as a continuum but probably represent discrete events. 40Ar/39Ar ages determined by Y. Chen and E. Farrar (Clark et al., 2000).

but neither cassiterite nor chalcopyrite. These veins and breccias are most widely developed in the upper part of the stock but persist at depth at least to 3,850 m a.s.l.. Narrow (generally < 50 cm) alteration selvages exhibit a temporal evolution from pervasive tourmalinization to intense quartz > muscovite alteration, evidence for strong leaching of Ca, Na, Al, and K (cf. Zweng and Clark, 1995). The superimposed stage II assemblages (Fig. 3) incorporate more than 97 percent of the cassiterite in the San Rafael and associated lodes. In the upper circa 150 m of the Sn zone of the San Rafael Lode (4,3104,460 m a.s.l.), cassiterite occurs (Palma, 1981) predominantly as botryoidal, finely laminated estao madera (wood-tin), intimately intergrown with chlorite and quartz. The remarkable development of such textures, unparalleled elsewhere in extent, probably indicates that the early stage II fluids were markedly supersaturated in Sn. Below circa 4,310 m, coarsely crystalline cassiterite predominates, but crudely boytroidal forms are widespread, and reversion to the woodtin habit occurred episodically as stage II evolved. The
0361-0128/98/000/000-00 $6.00

youngest stage II assemblages, dominating the San Rafael Lode below circa 4,050 m, record a marked overall increase in quartz deposition and concomitant decreases in cassiterite and, less pervasively, chlorite as vein minerals. However, the succeeding Stage III sulfide-dominated assemblages again incorporate abundant chlorite (Fig. 3). This stage is characterized by chalcopyrite, largely hexagonal pyrrhotite, pyrite, and minor stannite above 4,730 m, but the iron sulfides predominate below that elevation. Acicular cassiterite (needle-tin) occurs in direct association with chalcopyrite in the lower part of the Cu-rich zone. Stage III therefore exhibits single-pass (monoascendant) ore-metal zoning. Numerous lode segments exposed on the intermediate levels of the mine contain both coarse cassiterite and sulfides, including chalcopyrite, but megascopic and microscopic relationships indicate that in such cases the sulfides occur as veinlets cutting brecciated cassiterite, or as axial or marginal reopenings of cassiteriterich veins. Locally, chalcopyrite-acicular cassiterite veins cut bodies rich in wood-tin. Stages II and III are not, therefore,

1745

1746

KONTAK AND CLARK

considered gradational at the scale of the overall lode structure. The final stage III event was the development of quartzchlorite veins lacking sulfides and was most intense below circa 4,300 m. A minor terminal veining event, Stage IV (Palma, 1981), emplaced (Fig. 3) veins dominated by calcite and quartz (in part amethystine), with erratic fluorite, stibnite, gudmundite, and exceptionally Ni- and Cr-rich mackinawite (A.H. Clark, unpub. data). The abundance of vein chlorite associated with stage II cassiterite and stage III chalcopyrite is paralleled by moderate to intense alteration to chlorite of the granitic host rocks, but this hydrolytic alteration only locally extends more than 2 m from the lodes. Mass exchange calculations have not been carried out for the alteration envelopes, but the abundance (to 40 modal %) and Fe-rich composition of the chlorite (see below) imply that, in addition to almost complete leaching of Ca, K, and Na, intense Fe-metasomatism occurred (Palma, 1981). A decrease in the deposition of cassiterite toward the termination of stage II is paralleled by a decrease in the degree and extent of chloritization, but there is no systematic relationship overall between the local abundance of cassiterite and such hydrolytic alteration. These proximal chloritic zones grade outward to weak to moderate sericitic alteration, largely affecting plagioclase in the host granites. Muscovite-quartz assemblages, however, locally replace plagioclase adjacent to stage II cassiterite-rich veinlets lacking evidence for either stage I or III activity. Silicified envelopes also occur locally on both stage II and stage III veins. Structural relationships In the upper 300 m of the San Rafael stock, stage I vein systems and breccia bodies widely possess either stockwork or sheeted structures (Palma, 1981); their crudely radial and concentric orientations suggest a control by cooling stresses in the pluton. Late in stage I, however, tourmaline-quartz deposition became largely restricted to steeply dipping, quasicontinuous, conjugate shears striking on average 300 (westsouthwest dipping) and 335 (east-northeast dipping). Several late stage I tourmaline-quartz vein segments record dextraltranscurrent displacement, but it is unclear whether vertical movement was normal or reverse. Stage II fractures and Sn deposits were preferentially, but not exclusively, developed along the stage I 335 structures (tourmaline leaders), although at depth both 300 and 335 orientations host economic concentrations of cassiterite on the San Rafael Lode, and the extremely high-grade northern segment of the lode is developed along an intermediate C shear that strikes 310 to 325. The lode dips east-northeast at 40 to 70 (Fig. 2b), averaging 60, but steepens to subvertical below 4,000 m a.s.l. That stage II cassiterite mineralization in the San Rafael Lode took place during left-lateral strike-slip movement was demonstrated by Palma (1981): the southern contact of the stock is offset by 5 to 30 m on various levels. Vertical displacement during stage II was reverse. Much of the cassiterite in the San Rafael Lode is hosted by major orebodies (Fig. 2a) plunging east-northeast at 60 to 70 and exhibiting classic dilational jog geometries (Fig. 2b; cf. Cox et al., 2001) resulting from largely sinistral-transcurrent displacement, a
0361-0128/98/000/000-00 $6.00

relationship first documented by J. Alvarez (unpub. report to Minsur S.A., 1978) and substantiated by Palma (1981) and D.J. Kerr (unpub. report to Minsur S.A., 2002). It has been estimated (D.J. Kerr, unpub. report to Minsur S.A., 2001) that 80 to 82 percent of the Sn ore reserves occurs in these orebodies. The form of the first to be discovered, and hence termed the Ore Shoot, is illustrated in Figure 2b. The orebodies record multistage hydrothermal brecciation and cassiterite deposition, and they incorporate zones, particularly in the hanging wall, attaining more than 35 percent Sn over widths of at least 3 m. Structural relationships, therefore, played a major role in the development of the large bonanza orebodies that are responsible for the preeminence of San Rafael among hard-rock Sn deposits. Kontak and Clark (1988) and Clark et al. (1990a) emphasize that the San Rafael Lode and associated brittle shear zones formed in response to regional, northwest-southeast contraction. This contraction was initiated in the closing phases of stage I, and structural relationships in the Quenamari mine area demonstrate that shear propagation and at least stage III mineralization were contemporaneous with the thrusting of the Carboniferous cover rocks over the Ordovician lower plate (A.H. Clark, unpub. data). On a smaller scale, R. Mason and D.J. Kerr (unpub. report to Minsur S.A., 2000) argue that the left (west)-stepping dilational jogs on the San Rafael Lode were controlled by the intersection of the developing shear and preexisting, early stage I, radial and concentric fractures concentrated close to the eastern margin of the stock. Ore formation conditions Fluid-inclusion microthermometric studies of vein quartz (Palma, 1981; Kontak, 1985; A.H. Clark, unpub. data) demonstrate that stage I fluids were highly saline and boiled at temperatures of up to 580C, becoming cooler and less saline with time. The most intense quartz > muscovite alteration is associated with the cooler fluids. Stages II (Sn) and III (Cu Sn) formed at lower temperatures, each possessing an overall homogenization temperature range of 215 to 420C. The predominant ore-depositing fluids were nonboiling brines with moderate to very low salinity (<220 wt % NaCl equiv). Quartz coprecipitated with both botryoidal wood-tin and coarsely crystalline cassiterite in stage II, and stage III acicular cassiterite was deposited from fluids with circa 10 to 20 wt percent NaCl equiv and Th = circa 300420C. However, microthermometric data for alternating bands of coarse cassiterite and quartz from the Ore Shoot (4310 m level) indicate that the cassiterite was deposited from significantly hotter (ca. 2030C) and more saline (ca. 510 wt % NaCl equiv) fluids than the associated quartz (A.H. Clark, unpub. data). Palma (1981) demonstrated that, unexpectedly, no major changes in fluid temperature, i.e., Th, or salinity attended the transition from Sn- to Cu-dominant mineralization in the upper part of the San Rafael Lode. However, the local occurrence of vapor-rich inclusions indicates that the fluid sporadically boiled during chalcopyrite deposition, a feature not evident in stage II cassiterite assemblages. Stage IV veins formed at lower temperatures (Th = 170300C) and from dilute fluids (5 wt % NaCl equiv).

1746

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1747

Age relationships of ore deposition The best of several 40Ar/39Ar age plateaus for hydrothermal muscovites directly associated with stage I tourmaline-quartz veins yields an age of 24.10 0.15 Ma, in permissive agreement with an origin through retrograde boiling of the granitic magma. However, incremental-heating 40Ar/39Ar age spectra of aggregates of muscovite (2M1) flakes that replace plagioclase in proximity to stage II-dominated segments of the San Rafael Lode yield markedly younger plateau ages in the range 22.4 to 22.7 0.7 Ma (Clark et al., 2000). These dates overlap with the troughs, 22.1 to 23.3 Ma, in the age spectra of magmatic K feldspar (orthoclase structure) in the stock (Clark et al., 2000), suggesting that cassiterite mineralization, at 420C, occurred when the upper part of the stock had cooled to circa 230C (Berger and York, 1981). The evidence that economic hydrothermal activity significantly postdated emplacement (24.624.7 Ma) and initial cooling of the epizonal intrusion is supported by 40Ar/39Ar plateau dates of 21.9 to 22.0 0.2 Ma determined for hydrothermal adularia (lowsanidine structure) from a stannite-bearing stage III assemblage from the Jorge lode, which is contiguous with the San Rafael Lode (Clark et al., 2000). These age data imply that the upper part of the San Rafael stock experienced retrograde boiling and high-T, barren hydrothermal activity shortly after its emplacement, but that economic mineralization was significantly delayed and was, moreover, thermally prograde with respect to the contiguous granitic host rocks. Stages II and III in the San Rafael deposit were contemporaneous with terminal Picotani Group intrusive and volcanic activity in the wider Carabaya district (Sandeman et al., 1997). Focus of the study In the present contribution, the hydrothermal evolution of this immensely fertile system is examined on two bases: the secondary fluid inclusions widely trapped in quartz phenocrysts of the granitic host rocks, and the concomitant subsolidus mineralogic and geochemical processes in the upper 300 m of the stock. Particular attention is paid to evidence for the involvement of plausibly magmatic brines, on the one hand, and meteoric waters on the other (Kontak and Clark, 1997a). The San Rafael and Quenamari Granites Almost all exposures of the San Rafael and Quenamari granitoid rocks exhibit phenocrysts of alkali feldspar (to 12 cm in length), quartz (average 0.8 cm), extensively pinitized and hence megascopically melanocratic cordierite (average 0.8 cm), and biotite (average 0.4 cm). Plagioclase, an abundant matrix constituent, locally occurs as phenocrysts attaining 0.6 cm in length. None of the major minerals exhibits significant vertical changes in modal percent within the 1 km exposure interval of the SG stock. The euhedral forms of the original cordierite grains (Fig. 4b, h), as well as the overall sparsity of metasedimentary xenoliths, strongly imply a phenocrystic origin (Palma, 1981). However, sillimanite, an erratic minor constituent, occurs as both magmatic microphenocrysts and as fibrolitic aggregates intergrown with spinel in extremely rare mineral clots, plausibly of restitic origin (Palma, 1981; Kontak, 1985). Muscovite does not occur as a magmatic phase.
0361-0128/98/000/000-00 $6.00

The San Rafael granite stock and similar Quenamari granite intrusions (Fig. 1b) are circumscribed by fine-grained, variably porphyritic ring dikes. These dikes are petrologically similar to the central granites, exhibiting megacrysts of K feldspar, cordierite and biotite (Fig. 4d, e, g). Outcrop (Arenas, 1980) and underground exposures reveal that the upper 750 m of the San Rafael granite is dominated by two granitoid facies (Fig. 1b): an abundantly phenocrystic monzogranite with a coarse-grained felsic matrix; and a more sparsely phenocrystic phase with a finer-grained, mesocratic matrix, which straddles the monzogranite-granodiorite boundary in the alkali feldspar-plagioclase-quartz modal diagram and dominates the higher-elevation, northeastern outcrops of the stock. (Fig. 1b). These facies, however, occur widely as mutual enclaves whose diameters range from 5 cm to hundreds of meters and whose contacts range from planar and sharp, through protrusive and gradational, to cuspate-sinuous (cf. Powell et al., 1999). Such relationships strongly suggest that the facies represent two distinct magmas, with varying rheology contrasts, that commingled in the upper part of the stock. In addition, both facies widely incorporate mesocratic to melanocratic ellipsoidal enclaves, averaging 10 to 20 cm in maximum diameter. These enclaves include (A.H. Clark, unpub. data) microgranular biotite granodioritic to monzodioritic, as well as less abundant dioritic, diabasic/gabbroic, lamprophyric (minette, in part vesicular), and moderately micaceous to surmicaceous (>65% dark mica) members. The enclave relationships demonstrate that the two major granitoid magmas making up much of the San Rafael stock additionally incorporated pillows of diverse mafic magmas (A.H. Clark, unpub. data). This intimate coexistence of mantle-derived and anatectic magmas is a salient characteristic of the Picotani Group and intrusive suite (Kontak et al., 1986; Sandeman and Clark, 1993; Sandeman et al., 1997; Kontak and Clark, 1997b). Sandeman et al. (1995) propose that this igneous suite, emplaced during the Aymar tectonic event prior to major mid-Miocene, Quechuan crustal contraction and thickening, records shallow crustal anatexis caused by the incursion of mafic melts generated largely in the lithospheric mantle, in turn resulting from perturbation of the underlying asthenosphere above a foundering detached slab. High heat flow in relatively thin crust was focused along major structures relict from the Incaic, upper Eocene-lower Oligocene, Zongo-San Gabn tectonic zone (Kontak et al., 1990; Sandeman et al., 1995). The porphyritic granitoid rocks at San Rafael are associated with small bodies of distinctive tourmaline leucogranite, exposed along the northwest roof and wall of the stock (Fig. 1b, c; samples COCA [herein abbreviated CC]-156), and on its southwest margin (too restricted to indicate in Fig. 1b, c; CC141), as well as in scattered locations in Sandia Formation country rocks above the main stock (also too small to be recorded in Fig. 1b). The absence of tourmaline veinlets and clots argues against a metasomatic origin for this rock (e.g., Fig. 4f), which has potential implications for the evolution of the granitic magmas and hence for ore genesis. The contact of the leucogranite and the main intrusive phases is obscured on surface, but its main exposure is probably plug like; however, a single 6 cm vein has been observed to cut monzogranite in drill core from circa 4,300 m a.s.l.

1747

1748

KONTAK AND CLARK

5 cm d e f

5 cm

5 cm g h

5 cm i

5 cm

5 cm

5 cm

5 cm

FIG. 4. Hand samples of San Rafael and Quenamari granites. (a) Medium-grained, K feldspar-megacrystic cordierite-biotite monzogranite. Coin for scale is 2 cm in diameter; (b) cordierite-rich monzogranite from main San Rafael granite intrusion with abundant pinitized cordierite (dark subequant grains). (c) Fine-grained nonphenocrystic phase of the San Rafael granite; (d) Porphyritic granite from ring dike at the southwest margin of the San Rafael granite (see Fig.1b). Three dark grains at upper left are subhedral cordierite. (e) Sample of fine-grained dike rock from Quenamari area showing sparse phenocrysts. (f) Fine-grained tourmaline-bearing leucogranite (CC-156) phase of the San Rafael granite, from the upper contact of the intrusion. Small dark grains are tourmaline. (g, h, i) Porphyritic Quenamari granite from central intrusion (see Fig. 1b) showing the variable replacement of K feldspar phenocrysts (pale gray) by secondary albite (white), resulting in pseudorapakivi texture, as discussed in the text.

The petrography and chemistry of the least-altered samples of the San Rafael granite provide a datum for the identification and delimitation of metasomatic processes. All accessible outcrops of the San Rafael granite and most exposures of the Quenamari granite were studied herein (Kontak, 1985). Underground exposures of the San Rafael granite are documented by Palma (1981) and A.H. Clark (unpub. data). Most samples specifically cited herein are from outcrop (Fig. 1c). Petrographic study reveals that approximately 15 percent of the upper 400 m of the intrusion preserves little-modified magmatic relationships. The texture of both monzogranite and granodiorite is porphyritic overall (Fig. 4) with subhedral to euhedral grains of quartz, plagioclase, alkali feldspar, biotite, and
0361-0128/98/000/000-00 $6.00

cordierite in a hypidiomorphic granular matrix with variable development of granophyre (Fig. 5a, b). Plagioclase widely preserves fine-scale oscillatory zones (Fig. 5c, d) as well as sievetextured zones (Fig. 5e, f), features characteristic of hypabyssal and volcanic rocks. Alkali feldspar is variably perthitic, and film perthite is more widely developed than flame and bleb types (Fig. 5g, h). Rare grains of apparently uniform, unexsolved alkali feldspar are observed. The igneous mica varies from dark red-brown to orangey brown and encloses accessory phases, including apatite, zircon, and monazite (Fig. 5i). Cordierite, locally an abundant phase, is invariably pinitized and hence dark green at shallow levels of the stock, but fresh, pale blue and mauve cores and sectors are preserved in phenocrysts at depth.

1748

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1749

Crd
1 mm 1 mm 1 mm

0.5 mm

0.6 mm

0.3 mm

1mm

0.2 mm

0.4 mm

FIG. 5. Photomicrographs of polished thin sections illustrating salient magmatic petrographic features of the San Rafael granite. (a) Graphic texture characteristic of much of the upper part of the San Rafael granite. Note pinitized cordierite (Crd) grain at left and strongly zoned, euhedral plagioclase grain at upper left. (b) Hypidiomorphic granular texture typical of fresh San Rafael granite. (c) Zoned plagioclase grains with incipient white mica alteration along a fracture in otherwise fresh granite. (d) Strongly zoned plagioclase with very calcic core (dark). In detail, this grain incorporates numerous reversals in zonation (calcic spikes). Note the xenomorphic texture of the surrounding matrix. (e, f) Glomerocryst of plagioclase with sieve texture in all grains. Close-up of the area is shown in f. (g) Orthoclase megacryst in dike rock, exhibiting film and flame perthite. Note the fracture localizing abundant and coarser perthite and the irregular nature of the grain boundary, that shows intergrowth between the megacryst and matrix. (h) Detail of previous field showing the phenocryst margin inundated with fluid inclusions. Inset shows well preserved film perthite adjacent to the area where coarsening to flame perthite occurs. (i) Typical dark red-brown biotite grain with radiation halos about accessory phases. 0361-0128/98/000/000-00 $6.00

1749

1750

KONTAK AND CLARK

Despite the epizonal context of the San Rafael granite, there are no miarolytic cavities, and there is therefore no evidence that the volume of magma making up the stock experienced retrograde boiling. Whole-rock geochemistry Examination of more than 200 polished thin sections from outcrop and the upper 600 m of the underground exposures in the San Rafael granite, and from the Quenamari granite, permits the distinction of rocks with negligible, minor, moderate, and strong hydrothermal alteration. It should, however, be emphasized that sanidine, the magmatic alkali feldspar species in entirely fresh epizonal members of the Picotani intrusive suite (Kontak and Clark, 1988; Sandeman, 1995), is absent in the San Rafael granite and Quenamari granite; in this respect, therefore, none of the San Rafael granite or

Quenamari granite samples represents a pristine igneous assemblage. Reference is also made below to the broadly coeval rocks of the Santo Domingo center and to the Antauta granitic dike, which crops out 12 km south-southeast of San Rafael (Kontak and Clark, 1988; Sandeman et al., 1997). Analysis of the negligibly altered granitoid rocks (Table 1; Figs. 6, 7) reveals (1) a range of SiO2 content from 64.4 to 70.7 wt percent in the larger stocks, but with some dike rocks exceeding 72 wt percent; (2) K2O/Na2O significantly exceeding unity (Fig. 6a); (3) an aluminum saturation index (moles Al2O3/moles CaO + Na2O + K2O) of 1.1 to 1.25 overall (Fig. 6b), the most Ca-Fe-Mgrich granodioritic samples having the lower values (1.11.15: Palma, 1981); (4) continuous trends in Harker-type variation diagrams, such as are typical of felsic suites (Kontak, 1985); (5) both whole-rock and biotite compositions indicating very reduced conditions, the latter

a)
Na2O wt. %
5
Field of fresh SRG

b)
K2O wt. % K2O wt. %
9 7 5 3 9 7 5 3
1.2 2 1.4 1.6 1.8

c)

1 1 3 5 7 9

20

40

60

80

Ba (ppm, X1000)

d)

K2O wt. %
CaO wt. %

e)

A/CNK
K2O wt. %
9 7 5 3

Sn (ppm)

f)

100

200

300

100

200

300

200

400

600

Sr (ppm)

g)
K2O wt. % K2O wt. %
9 7 5 3
1000 2000

h)
9 7 5 3

Sr (ppm)
40

Rb (ppm)

i)
Fractionation

Nb (ppm)
200

30 20 10

100

100

200

Ba (ppm)

Zr (ppm)

Zr (ppm)

Alteration Index

Negligible Minor

Moderate Strong

FIG. 6. Binary element plots for major and trace element data for fresh and variably altered San Rafael and Quenamari granites. The circled field in each plot is for fresh San Rafael granite. The arrow in the Nb-Zr plot shows the fractionation inferred to be responsible for two samples of tourmaline leucogranite (CC-141, 156), on the basis of the apparent immobility of these trace elements. A/CNK (b) is the aluminum saturation index, i.e., moles Al2O3/moles CaO + Na2O + K2O. The symbols refer to the different groupings of samples defined on the basis of the degree of alteration, determined petrographically as discussed in the text. 0361-0128/98/000/000-00 $6.00

1750

TABLE 1. Compositions of Fresh and Altered Granites, San Rafael-Quenamari Granite District, Southeast Peru Sample number and degree of alteration of granite COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA COCA 166A 403 203 151 195 249A 244 150 156 141 250 252 140 198 217 213 239 190 187A 218 206 W W W M M M M M M S S S S S S S S S S S S 68.10 0.40 15.31 3.07 NA 0.07 1.17 1.14 3.24 4.78 0.18 97.46 64.40 0.38 15.61 2.27 NA 0.05 1.30 0.95 4.08 3.23 0.18 92.45 68.50 0.37 15.80 2.57 NA 0.05 1.23 1.02 2.46 4.96 0.19 97.15 70.47 0.36 15.47 0.27 1.87 0.03 1.29 1.31 2.84 4.86 0.19 98.96 68.40 0.46 16.58 0.20 2.47 0.05 1.56 1.40 3.09 4.80 0.24 99.25 70.20 0.35 15.14 2.56 NA 0.12 1.03 0.54 3.47 2.68 0.20 96.29 69.05 68.40 73.67 74.11 69.61 69.80 74.39 0.37 0.38 0.12 0.10 0.33 0.40 0.11 15.74 15.90 14.85 14.41 15.66 15.14 15.23 2.28 2.60 0.27 ND 0.28 2.55 0.18 NA NA 0.48 2.11 2.90 NA 0.56 0.08 0.12 0.01 0.02 0.12 0.09 0.06 1.21 0.98 0.79 0.86 1.06 0.79 1.04 0.92 1.42 0.27 0.62 0.27 0.33 0.49 4.72 4.45 3.72 4.95 0.03 0.48 2.22 2.45 2.16 4.84 1.13 7.40 8.28 3.06 0.19 0.18 0.28 0.32 0.21 0.21 0.32 97.01 96.59 99.30 98.63 97.87 98.07 97.66 68.10 0.38 15.12 2.70 NA 0.03 1.26 0.40 2.57 5.73 0.20 96.49 66.80 0.48 15.60 2.95 NA 0.08 2.14 0.55 2.05 6.17 0.26 97.08 68.40 0.43 15.77 3.49 NA 0.12 1.46 1.25 3.91 2.68 0.19 97.70 68.50 0.39 15.36 2.86 NA 0.07 1.31 0.37 1.20 7.11 0.21 97.38 68.70 0.33 16.27 2.90 NA 0.10 1.20 0.34 1.12 5.84 0.18 96.98 64.90 0.50 15.67 3.40 NA 0.06 2.36 0.27 0.01 7.28 0.19 94.64 66.60 0.42 14.52 3.78 NA 0.45 1.77 0.48 0.46 7.45 0.27 96.20 65.90 0.36 15.19 2.99 NA 0.14 1.61 0.30 0.36 8.62 0.22 95.69

Compound or element

0361-0128/98/000/000-00 $6.00 69.11 0.40 16.29 0.34 2.17 0.05 1.63 1.54 2.91 4.83 0.21 99.48 68.40 0.43 15.32 2.74 NA 0.10 1.10 1.10 2.81 4.79 0.16 96.95 68.34 0.45 16.65 0.30 2.38 0.07 1.57 1.45 3.47 4.70 0.25 99.63 ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU 17 553 NA 329 227 122 10 16 59 130 9 NA 135.3 137.9 106.5 80.4 48.6 19.2 20.1 9.7 10.7 7.0 9.0 7.0 5.0 3.0 4.0 1.3 3.9 1.3 3.8 1.2 3.2 1.2 4.2 1.3 5.0 1.4 3.5 1.3 4.4 1.3 6.0 1.6 4.1 1.3 58.0 52.6 46.4 41.4 28.4 21.0 16.3 13.0 10.8 19.1 9.5 8.0 9.9 9.0 44.4 46.8 44.1 37.0 25.7 7.4 14.8 13.0 13.9 10.0 9.6 7.0 5.0 5.0 163.7 149.4 108.8 76.9 45.0 17.6 21.3 19.1 14.3 14.9 9.4 6.0 3.3 6.6 160.3 164.8 126.7 84.8 57.7 17.4 30.3 16.7 18.1 10.4 11.9 9.0 5.0 7.8 40.4 39.0 29.9 29.5 18.3 0.5 9.7 9.0 7.0 10.0 9.6 9.0 8.0 11.0 77.2 67.6 45.8 43.1 24.9 7.8 12.4 19.3 9.7 8.0 5.1 5.0 5.1 9.9 85.3 4.0 76.3 5.6 54.2 3.7 52.8 2.0 30.8 3.7 9.4 8.5 18.8 3.8 26.3 3.0 15.8 1.3 9.5 6.4 12.6 3.5 7.0 6.0 5.1 8.0 7.0 15.0 4.1 3.7 1.3 4.7 1.4 15.0 16.0 20.0 19.5 13.5 30.8 10.4 19.4 15.5 14.1 9.7 7.0 5.5 6.0 7.6 1.8 104.7 84.6 54.4 46.8 27.5 6.8 11.4 6.6 7.8 8.6 5.7 4.0 2.7 1.2 5.3 1.5 8.1 1.9 9 856 51 308 256 156 NA NA NA NA 13 47 13 9 870 1,573 NA 42 306 322 298 168 158 147 12 10 28 NA 93 NA 314 NA 8 23 NA 51 9 560 45 230 256 143 NA NA NA NA 7 61 7 530 40 308 168 134 NA NA NA NA 17 56 16 503 42 354 167 119 10 6 1 194 13 NA 14 762 54 321 285 156 12 28 79 203 4 NA 9 711 37 200 186 138 NA NA NA NA 4 52 8 438 47 183 255 153 NA NA NA NA 11 54 9 32 371 124 45 NA 183 630 264 32 174 28 NA NA NA 40 NA 49 NA 98 11 15 52 NA 39 11 118 1,502 NA 32 162 502 87 105 24 128 NA 10 14 15 670 675 925 56 21 36 NA NA 7 918 45 588 84 168 NA NA NA NA 4 NA 6 6 357 1,023 1,221 NA 51 71 446 380 376 13 226 191 23 135 167 10 NA NA 7 NA NA 69 NA NA 60 NA NA 61 22 14 NA 62 115 141.9 121.9 80.1 71.1 39.2 13.1 19.7 12.1 11.9 10.0 8.6 6.0 4.0 3.2 4.4 1.4 187.4 191.6 144.9 103.8 65.2 24.4 29.2 18.6 14.9 13.3 11.5 8.0 7.2 7.0 5.2 1.4 4.6 1.4 7 512 64 188 120 173 NA NA NA NA 85 43 8 9 8 9 9 946 1,392 1,839 1,777 1,317 46 37 82 58 64 460 478 462 465 509 78 100 156 201 197 154 149 193 146 166 NA NA NA NA NA NA NA NA NA NA NA NA NA NA NA NA NA NA NA NA 38 32 18 50 13 50 44 137 195 118 94.3 86.0 57.7 51.6 30.1 11.5 14.9 7.0 9.5 5.9 8.2 6.0 3.2 4.1 5.4 1.5 82.6 77.9 52.0 46.8 28.6 8.5 14.9 14.9 10.7 5.6 8.7 7.0 3.7 4.0 7.9 1.9 7.7 1.9 5.5 1.5 5.3 1.5

COCA COCA COCA COCA COCA COCA COCA COCA 154 165 248 406 408 194 202 197 F F F F F F F W

SiO2 TiO2 Al2O3 Fe2O3 FeO MnO MgO CaO Na2O K2O P2O5 Total

70.70 0.35 14.67 2.06 NA 0.07 0.91 0.89 2.86 4.66 0.16 97.33

69.50 0.34 15.50 2.74 NA 0.03 1.12 1.04 2.79 4.96 0.21 98.23

70.00 0.40 15.22 2.31 NA 0.05 1.02 1.07 3.00 4.82 0.18 98.07

66.40 0.48 16.10 2.62 NA 0.07 1.55 1.34 3.41 4.59 0.21 96.77

66.50 0.49 15.91 3.27 NA 0.10 1.44 0.78 2.71 5.64 0.21 97.05

1751

Trace elements (ppm) Nb 7 10 6 7 Ba 602 455 479 1,193 V 38 39 38 63 Rb 317 377 351 310 Sr 174 142 165 294 Zr 135 135 127 186 Ni NA NA NA NA Cu NA NA NA NA Pb NA NA NA NA Zn NA NA NA NA Sn 7 23 12 20 Cr 34 42 33 65

8 911 57 358 209 187 NA NA NA NA 4 48

Chondrite-normalized REE values La 219.6 Ce 203.2 Pr 158.0 Nd 111.0 Sm 64.3 Eu 25.0 Gd 24.4 Tb 22.3 Dy 14.5 Ho 16.8 Er 6.3 Tm 5.0 Yb 3.0 Lu 3.0

%C 3.7 A/CNK 1.3

4.2 1.3

3.6 1.3

3.6 1.2

4.4 1.3

1751

F = fresh, M = moderately altered, NA = not analyzed, ND = not detected, S = strongly altered, W = weakly altered

1752

KONTAK AND CLARK

a)

b)

Sn

Ab
c)

Or An
d)

Ab C

Or

Ab
e)

Or Ab Ba
f)

Or Zr

Sr
Alteration Index

Rb Sr
Negligible Minor Moderate Strong

Rb

FIG. 7. Ternary plots of normative (CIPW) proportions of minerals calculated for variably altered San Rafael and Quenamari granites.

corresponding to the Ni-NiO buffer in ternary Fe+2-Fe+3-Mg+2 space (Kontak, 1985), compatible with the occurrence of rare ilmenite as the sole Fe-Ti oxide phase, and its stoichiometric (Fe,Mg)TiO3 composition (A.H. Clark, unpub. data); and (6) a clustering in terms of both quartz-albite-orthoclase (Fig. 7a) and anorthite-albite-orthoclase (Fig. 7c) normative proportions close to minimum melting compositions. In its whole-rock chemistry, the San Rafael granite is more cafemic than other strongly peraluminous, S-type granites (White and Chappell, 1983; Tischendorf et al., 1987; Stimac et al., 1995). Rb contents fall in the narrow range of 350 50 ppm (Fig. 6f), and K/Rb ratios are confined to 120 to 140, values typical of granites (Shaw, 1968). Ba shows the greatest variation among the trace elements (479 to 1,193 ppm, and one very weakly altered rock at 1,573 ppm: Fig. 6g), whereas Sr exhibits a much narrower range (142294 ppm: Fig. 6d). There is limited dispersal of the data in ternary plots, and the fresh samples cluster tightly in Ba-Sr-Rb and Zr-Sr-Rb diagrams (Fig. 7e, f). Two moderately altered leucocratic granite samples (CC-141 and 156), however, are strongly enriched in both Zr and Nb (Fig. 6i), a relationship interpreted as evidence for strong fractionation of the melt. The rare earth element (REE) abundances (Fig. 8a) are similar to those of unaltered peraluminous granites in general
0361-0128/98/000/000-00 $6.00

(e.g., Muecke and Clarke, 1981), and the patterns are indistinguishable from those of other peraluminous granites of similar age in southeast Peru (Kontak and Clark, 1988). The granites exhibit chondrite-normalized light rare earth element (LREE) values of 80 to 120, strongly fractionated REE overall [(La/Yb)N = 2040], and negative Eu anomalies (EuN/Eu* = 0.50.9). Fluorine contents in the San Rafael granite and Quenamari granite are 0.08 to 0.18 wt percent (A.H. Clark, unpub. data), much lower than those in many granites associated with lithophile-metal mineralization (Tischendorf, 1977), but in conformity with the low and consistent F contents in biotite (Fig. 9). The Sn contents of the fresh rocks are anomalous, attaining 23 ppm (Table 2; Fig. 6c) and contrasting with the Clarke value of circa 4 ppm for granites sensu lato. It is, however, significant that neither of the analyzed samples of tourmaline leucogranite (CC 141 or 156) is Sn-enriched relative to the main monzogranites and granodiorites (Table 2). Mineral chemistry The compositions of biotite, plagioclase, and alkali feldspar in the freshest samples of the San Rafael granite and Quenamari granite provide a basis for the recognition of the effects of superimposed fluid-rock interaction. Phenocrystic biotite in the upper part of the stock (Table 2; Fig. 9) shows a limited

1752

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1753

a)

1000

b)
least altered granite (SRG)
Santo Domingo

1000

c)
minor alteration
CC-203 CC-403 CC-197 CC-166A

1000

moderate alteration

Rock/Chondrite

Rock/Chondrite

Rock/Chondrite

100

CC-406 CC-202

100

100

CC-150 CC-249A CC-151 CC-156

10

10

10

tourmaline leucogranite

field of fresh SRG


1 La Pr Sm Gd Dy Er Yb Ce Nd Eu Tb Ho Tm Lu

1 La Pr Sm Gd Dy Er Yb Ce Nd Eu Tb Ho Tm Lu

1 La Pr Sm Gd Dy Er Yb Ce Nd Eu Tb Ho Tm Lu

d)

1000

strong K alteration

Rock/Chondrite

100

CC-217 CC-198 CC-239 CC-190 CC-252 CC-250

10

1 La Pr Sm Gd Dy Er Yb Ce Nd Eu Tb Ho Tm Lu

FIG. 8. Chondrite-normalized rare-earth element plots for various populations of the San Rafael (SRG) and Quenamari granites, classified on the basis of the degree of alteration related to metasomatism. A fresh sample of the petrologically similar Santo Domingo granite is included in (a). Plots (c) and (d) illustrate the distinct pattern of a late-stage tourmaline leucogranite (CC-156), inferred to represent a more evolved melt fraction.

Si per formula unit

a)

b)
3.2
= Mineral separates

tourmaline leucogranite (COCA-156)

Si per formula unit

3.2 3.0 2.8 2.6 1.0 3.0 5.0

3.0 2.8 2.6 0.5 0.6 0.7 0.8

c)
wt. % F

Fe/(Fe+Mg)
4 3 2 1
0.5 0.6 0.7 0.8

d)
wt. % Al2O3

wt. % TiO2
30

S-type granites

20

S/I-type granites

10 4 6 8 10 12

Fe/(Fe+Mg)
CC-17 CC-156 CC-408 CC-154 CC-244

wt. % MgO
CC-151 CC-150

FIG. 9. Binary element plots for magmatic biotite in eight samples of San Rafael granite, as determined by electron microprobe and bulk analysis of mineral separates. Note that the data for the tourmaline leucogranite (no. 156) occupy distinct fields in all plots. The fields delineated for S- and I-type granites are based on data for well characterized granite suites elsewhere. 0361-0128/98/000/000-00 $6.00

1753

1754

0361-0128/98/000/000-00 $6.00 Granite, sample number, and point Antauta COCA-17 3 San Rafael COCA-150 12 San Rafael COCA-150 13 San Rafael COCA-150 14 San Rafael COCA-156 23 San Rafael COCA-156 24 San Rafael COCA-156 31 San Rafael COCA-156 32 San Rafael COCA-408 48 San Rafael COCA-408 49 37.88 2.89 19.87 0.10 16.36 0.27 8.98 0.10 0.99 8.32 0.61 96.37 0.26 96.11 2.795 1.205 0.523 0.160 0.006 1.010 0.017 0.988 0.008 0.142 0.783 2.687 1.313 0.338 0.190 0.007 1.217 0.017 1.058 0.011 0.053 0.849 2.680 1.320 0.355 0.190 0.009 1.172 0.014 1.078 0.007 0.056 0.858 2.680 1.320 0.332 0.197 0.009 1.253 0.018 0.999 0.008 0.053 0.891 3.227 0.773 1.380 0.081 0.009 0.347 0.008 0.308 0.006 0.043 0.889 2.985 1.015 1.083 0.073 0.008 0.806 0.012 0.411 0.008 0.063 0.900 3.108 0.893 1.366 0.018 0.003 0.513 0.011 0.342 0.006 0.064 0.901 36.10 3.41 18.83 0.12 19.54 0.26 9.54 0.15 0.37 8.95 0.61 97.88 0.26 97.62 35.98 3.41 19.08 0.16 18.81 0.24 9.71 0.09 0.39 9.03 0.73 97.63 0.31 97.32 35.38 3.46 18.51 0.15 19.78 0.28 8.85 0.10 0.36 9.22 0.80 96.89 0.34 96.55 46.61 1.57 26.37 0.18 5.99 0.13 2.98 0.08 0.33 10.06 2.24 96.54 0.94 95.60 41.02 1.34 24.46 0.15 13.26 0.20 3.80 0.10 0.45 9.69 2.01 96.48 0.85 95.63 44.26 0.34 27.30 0.07 8.73 0.19 3.27 0.08 0.47 10.05 3.34 98.10 1.41 96.69 40.62 0.46 24.25 0.10 13.90 0.25 3.81 0.09 0.35 9.64 3.12 96.59 1.31 95.28 2.998 1.003 1.106 0.024 0.006 0.858 0.015 0.419 0.007 0.050 0.909 35.60 3.41 18.26 0.16 18.68 0.31 9.76 0.12 0.42 8.69 1.11 96.52 0.47 96.05 2.694 1.306 0.323 0.194 0.009 1.183 0.020 1.101 0.009 0.062 0.838 35.22 3.58 18.45 0.15 18.89 0.37 9.52 0.11 0.45 8.89 1.07 96.70 0.45 96.25 2.669 1.331 0.315 0.204 0.009 1.197 0.023 1.075 0.009 0.065 0.860 KONTAK AND CLARK

TABLE 2. Representative Electron Microprobe Analysis of Magmatic Biotite from the San Rafael and Antauta Granites, Southeast Peru

Compound or element

Antauta COCA-17 1

Antauta COCA-17 2

1754

SiO2 TiO2 Al2O3 Cr2O3 FeO(T) MnO MgO CaO Na2O K2O F Total F=O Total

36.34 3.28 19.72 0.10 16.86 0.22 9.73 0.13 0.86 8.89 0.56 96.69 0.24 96.45

35.73 3.15 20.17 0.13 17.81 0.28 9.79 0.11 0.85 8.84 0.88 97.74 0.37 97.37

Structural formula (11 oxygens) Si 2.699 Al(iv) 1.301 Al(vi) 0.426 Ti 0.183 Cr 0.006 Fe 1.047 Mn 0.013 Mg 1.077 Ca 0.010 Na 0.123 K 0.843

2.649 1.351 0.412 0.175 0.008 1.104 0.017 1.081 0.008 0.122 0.836

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1755

variation in major element contents, with Fe/(Fe + Mg) = 0.5 to 0.55, <1 wt percent F, and 3 to 5 wt percent TiO2. The micas, which contain 8 to 10 wt percent MgO, are distinctly magnesian compared with those in the majority of peraluminous granitoids, including those of Lachlan S-type. However, microscopically fresh biotite from a leucogranite, CC-156, deviates markedly in all plots and possesses strong but variable enrichment in Fe, Al, Si, and F (3.5 wt %). In the upper part of the San Rafael pluton, the majority of the plagioclase phenocrysts are normally zoned, most falling in the compositional range An2040 (Table 3, Fig. 10). Some, however, contain cores as calcic as An55. Although these cores are corroded and could represent restitic components, comparably Ca-rich compositions occur in the outer parts of phenocrysts in numerous samples from the 4200 m level of the San Rafael granite, where there is unambiguous megascopic evidence for commingling and mixing with mafic magmas (A.H. Clark, unpub. data). Such calcic spikes overgrew strong

50An
least altered SRG
tourmaline leucogranite (COCA-156)

minor alteration

moderate alteration

strong alteration

20 analyses

Ab

Or

FIG. 10. Compositions of feldspars from variably altered San Rafael and Quenamari granites plotted in part of the An-Ab-Or ternary. Plagioclase in the leucocratic granite (no. 156) approaches end-member albite. Translucent sanidine from a dike rock from the Antauta center (see text) is the Or75Ab25 phase in the first ternary plot. Note the progression toward more albitic plagioclase as the degree of alteration increases, culminating in end-member albite in the most altered rocks. 0361-0128/98/000/000-00 $6.00

internal corrosion surfaces recording complete dissolution, but sieve textures indicative of partial magmatic corrosion are present in groundmass and phenocrystic plagioclase throughout the intrusion (Fig. 5e, f), and core areas are commonly less calcic than rims (Table 3). As noted above, scattered alkali feldspar phenocrysts in the upper part of the San Rafael granite and in the Quenamari granite contain true rapakivi texture, i.e., overgrowth by oligoclase, but this feature is again more commonly observed below circa 4,300 m in the San Rafael granite. Microscopically unaltered plagioclase in the tourmaline leucogranite is markedly more sodic (An510) than that in the main intrusion (Fig. 10). Analyses of two separates of orthoclase perthite indicate that alkali feldspar in the least-altered granites has a bulk composition of Or74Ab23An3. Similar compositions are obtained both from separates and from single crystals of sanidine (Or75Ab25) from nearby intrusions of similar age and petrogenesis, i.e., Antauta and Santo Domingo (Fig. 10, Table 3). The K-rich domains of perthitic orthoclase range from Or89Ab11 to Or97Ab3 (Fig. 10). The abundances of Rb, Sr, Ba, Li, and Pb (Table 4, Fig. 11) in the separates from the San Rafael, Santo Domingo, and Antauta granites are also uniform, evidence for similar melt compositions (Ichenhower and London, 1996). The REE data for alkali feldspar separates (Fig. 11) have chondrite-normalized abundances, fractionation patterns [(La/Lu)N = 3480] and positive Eu anomalies (EuN/Eu* = 2025 for the San Rafael granite and 317 for Santo Domingo and Antauta), such as are typical of alkali feldspars in granites (Smith, 1974; Leeman and Phelps, 1981; Kontak and Martin, 1997). Again, similar compositions are shown by alkali feldspars from the San Rafael granite and other nearby intrusions (Fig. 11). The alkali feldspar is perthitic and exhibits strong (201) sodium peaks in X-ray powder diffraction patterns (Fig. 12a). Film perthite is cut by later flame perthite that locally coalesces into coarse bleb perthite (Fig. 5g, h). All alkali feldspar is turbid and incorporates areas characterized in thin section by pitted textures which, when viewed under high magnification, are seen to result from swarms of fluid inclusions. In contrast, the Antauta dike, interpreted as a small, quenched intrusion only locally affected by hydrothermal activity, exhibits translucent phenocrystic sanidine, thus providing a basis for assessment of ordering relationships during alkali metasomatism at San Rafael. The degree of Al/Si ordering of the alkali feldspar in the least altered granites in the wider San Rafael district, as summa rized in the (060) vs. (204) diagram of Stewart and Wright (1974; Fig. 12b), is variable. Whereas in the San Rafael granite proper the feldspars plot in a distinct field straddling the orthoclase tie-line, those from the Quenamari granite and the granitic dikes exhibit structures ranging from orthoclase to an intermediate position between orthoclase and sanidine, respectively, and stock- and dike-hosted samples are distinctly separate. In the altered samples, the feldspars invert to end-member K feldspar and plot at the boundary of the field (see Fig. 12b and discussion below). The ordering relationships in the San Rafael granite alkali feldspars and the extensive preservation of film perthite are strong evidence for rapid cooling and hence an epizonal emplacement level for the stock.

1755

1756

0361-0128/98/000/000-00 $6.00 TABLE 3. Representative Analyses of Feldspars in Unaltered Antauta and San Rafael Granites, Southeast Peru Granite, sample number, point, and phase Antauta COCA-17 13 Sanidine Sanidine Orthoclase Orthoclase Plagioclase Antauta COCA-17 14 San Rafael COCA-151 1 San Rafael COCA-151 5 San Rafael COCA-150 27 San Rafael COCA-151 2 Exsolved albite San Rafael COCA-151 3 Exsolved albite San Rafael COCA-151 4 Exsolved albite San Rafael COCA-151 25 Core sieve plagioclase San Rafael COCA-151 30 Rim sieve plagioclase San Rafael COCA-151 31 Rim sieve plagioclase 64.77 18.93 0.12 2.42 11.06 97.38 3.001 1.003 0.005 0.217 0.653 0.6 24.8 74.6 0.2 24.4 75.4 0.0 3.8 96.2 7.8 90.3 1.9 5.0 93.0 2.0 3.002 1.043 0.002 0.208 0.643 3.024 1.016 0.000 0.032 0.808 2.976 1.072 0.064 0.744 0.016 2.992 1.040 0.040 0.744 0.016 2.944 1.072 0.064 0.872 0.008 6.8 92.4 0.8 3.008 1.024 0.000 0.032 0.880 0.0 3.5 96.5 64.08 18.90 0.07 2.30 10.77 96.12 65.62 18.67 0.00 0.39 13.69 98.38 67.37 20.51 1.34 8.68 0.31 98.23 68.09 20.06 0.84 8.69 0.30 97.98 67.20 20.83 1.33 10.27 0.17 99.80 64.82 18.70 0.00 0.37 14.80 98.69 59.87 24.37 5.95 7.66 0.50 98.58 2.705 1.296 0.288 0.669 0.028 29.2 67.9 2.9 65.56 21.79 2.61 8.59 0.38 98.95 2.896 1.136 0.120 0.736 0.024 13.6 83.6 2.7 58.71 26.69 8.10 6.83 0.16 100.51 2.608 1.400 0.384 0.592 0.008 39.0 60.2 0.8 65.91 22.24 2.94 8.62 0.43 100.16 2.880 1.144 0.136 0.728 0.024 15.3 82.0 2.7 KONTAK AND CLARK

Compound or element

Antauta COCA-17 1

Antauta COCA-17 4

Plagioclase

Plagioclase

1756

SiO2 Al2O3 CaO Na2O K2O Total

61.26 24.04 5.28 8.23 0.47 99.52

57.97 25.68 7.70 6.81 0.34 98.55

Structural formula (8 oxygens) Si 2.736 Al 1.264 Ca 0.251 Na 0.711 K 0.026

2.628 1.371 0.374 0.598 0.019

End member (mol %) An 25.4 Ab 71.9 Or 2.7

37.7 60.3 2.0

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU TABLE 4. Trace Element Content of Alkali Feldspar Separates, Picotani Intrusive Suite Granites, Southeast Peru San Rafael Sample Mol % Or* Li Rb Sr Ba Pb COCA 407 73.1 19.8 407 346 2,554 153 4.55 7.65 0.79 2.74 0.52 3.63 0.65 0.04 0.22 0.04 0.11 0.01 0.08 0.008 COCA 408A 76.2 27.7 456 533 6,337 172 6.16 10.14 1.04 3.53 0.63 6.59 1.04 0.04 0.21 0.04 0.08 0.01 0.04 0.008 COCA 249 13.6 38.2 162 321 335 205 9.92 20.80 2.47 9.33 1.82 0.80 1.25 0.17 0.92 0.15 0.44 0.06 0.35 0.03 COCA 249 13.6 37.5 157 309 325 203 9.76 20.60 2.42 8.99 1.87 0.82 1.37 0.20 0.97 0.16 0.42 0.05 0.31 0.04 COCA 190 93.3 23.3 599 353 14,813 45.8 4.88 9.89 1.18 4.25 0.92 0.08 3.68 0.07 0.41 0.08 0.25 0.03 0.08 0.02 COCA 198 85.4 36.9 600 336 5,371 126 7.59 12.89 1.38 4.60 0.91 4.25 1.06 0.07 0.31 0.05 0.12 0.01 0.08 0.01 COCA 165B 87.8 42.6 723 302 3,880 129 4.61 8.35 0.89 3.06 0.54 3.53 0.53 0.04 0.25 0.03 0.08 0.009 0.015 0.008 Antauta COCA 15 70.0 251 485 352 2,518 179 7.52 13.70 1.50 5.09 0.88 1.22 0.90 0.06 0.36 0.05 0.21 0.03 0.15 0.02 SD 4 76.0 171 485 334 1,976 155 3.65 5.89 0.63 2.08 0.39 2.37 0.43 0.02 0.13 0.02 0.06 0.01 0.05 0.01 Santo Domingo SD 5 75.3 169 519 334 1,818 143 4.77 8.32 0.92 3.17 0.59 2.57 0.55 0.05 0.28 0.05 0.14 0.01 0.09 0.01 SD 3 86.3 28.9 509 509 6,841 186 5.41 7.82 0.98 3.30 0.69 0.82 1.31 0.05 0.27 0.05 0.12 0.02 0.07 0.015

1757

SD 6 89.7 44.8 690 579 6,612 85.2 10.60 14.00 1.95 6.72 1.29 3.64 1.88 0.10 0.52 0.09 0.28 0.03 0.16 0.03

Rare-earth elements La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

*Mol % Or determined on mineral separates by X-ray fluoresence

Petrology of Metasomatized Granite Outside of the chloritic alteration halos of the mineralized veins, much of the upper parts of the San Rafael granite and Quenamari granite appear unaltered in outcrop and hand specimen, because of the widespread preservation of apparently fresh biotite and the mafic appearance of the pinite
1000

pseudomorphs after cordierite (e.g., Fig. 4). The granites are not bleached and contain little megascopic secondary white mica. However, as emphasized by Kontak (1985) and Kontak and Clark (1988), intense alkali metasomatism and weaker hydrolytic alteration have affected much of the upper part of the San Rafael stock and all accessible outcrops of the Quenamari area stock, as well as the dikes associated with each.
1000

K-feldspar/Chondrite

San Rafael & Quenamari

K-feldspar/Chondrite

100

10

Or76 Or73 Or85 Or87 Or93 Or13 Or13

Santo Domingo & Antauta


100

10

Or90 Or76 Or75 Or70 Or86

407 408A 249 249(D)

165B 198 190

1
15 4 5 3 6

0.1
Li Sr Pb La Pr Sm Gd Dy Er Yb Rb Ba Ce Nd Eu Tb Ho Tm Lu

0.1
Li Sr Pb La Pr Sm Gd Dy Er Yb Rb Ba Ce Nd Eu Tb Ho Tm Lu

FIG. 11. Extended chondrite-normalized plots for trace- and rare-earth element data for alkali feldspar separates from granites at San Rafael (CC-407, 408A, 249, 165B), Quenamari (190, 198), Antauta (15), and Santo Domingo (4, 5, 3, 6). Sample 249D is a duplicate of 249. 0361-0128/98/000/000-00 $6.00

1757

1758

KONTAK AND CLARK


K (111) K K (130) (201)

a)
1

K (131) K (041) Sanidine

b)
42.0

(060)

QG stock SRG stock QG dikes

Orthoclase Perthite

Na (201)

degree of development of specific microscopic alteration fabrics. We distinguish between negligibly, weakly, moderately, and strongly altered facies, which possess approximate preservation of igneous mineral assemblages and textural relationships of, respectively, 90 to 95 percent, 80 to 95 percent, 50 to 80 percent, and <50 percent. This combined petrographic and mineralogic approach is inherently subjective, but it is, in our opinion, adequately robust. Petrography of alkali-metasomatized granite Although difficult to see in outcrop, the widespread alkali metasomatism in the San Rafael granite is clearly revealed by the use of standard staining methods for K and Ca-bearing feldspars. In Figure 13 a fresh granite (a) is contrasted with a cryptically altered sample (b, c) that reveals the extent of K metasomatism and plagioclase destruction only after being stained (cf. b, c).

Orthoclase Perthite

41.5 50.5

metasomatism SRG

Antauta dike

(204)

51.0

Na metasom.

c)
42.0
m ic

3 4 5 6

or

th

K metasom.

di

oc ne

la

(060)

Na metasom.

ro c

lin

se

2 cm

32

28

24

20

41.5 50.5

sa

2 degrees Cu K

(204)

FIG. 12. X-ray powder diffraction data for alkali feldspars from stocks and dike rocks in the San Rafael-Quenamari area (see Kontak et al., 1984, for details). Data are plotted in the orthoclase and microcline fields (i.e., degree of Al/Si ordering: diagram of Wright, 1968); compositions are also a function of Miller indices (i.e., increasing Or content toward the lower left of the plot). The Quenamari dike samples retain a less-ordered structure than those from the larger Quenamari and San Rafael intrusions. At San Rafael, the less-ordered samples occur at the roof of the stock, in the right-hand area of the outlined field, and the sanidine samples (diamonds) are from the Antauta dike. The metasomatism has caused the feldspars to invert to end-member K feldspar at the boundary of the field. The inferred trend with increasing alteration is from 1 to 6 (b). The corresponding diffractogram patterns are shown in (a).

ni

51.0

a
3 cm

Our studies of outcrop and underground exposures demonstrate that no circa 100 m3 volume of the San Rafael granite above the 4350 m level has escaped this alteration, which exhibits no control by megascopic fractures, including stage I veins and the mineralized lodes. The pervasive nature of the alkali exchange alteration precludes the preparation of maps or sections illustrating its extent and intensity, but overall it is evident that, in the upper 600 m of the San Rafael granite, unaltered or weakly altered granite is considerably less widespread within 500 m of the eastern and northeastern contacts than in the western half of the stock. The areal distribution of the hydrolytic alteration is less well established, but we tentatively conclude that it is most intense and pervasive within 20 to 30 m of the lodes. Despite the wide dispersion of the both alkali and hydrolytic alteration, most mineralogic features are heterogeneously developed at both hand-specimen and thin-section scales. In order to define the effects of metasomatism, therefore, samples have been assigned to alteration intensity groups on the basis of estimates, in stained hand specimens and in thin section, of the modal abundances of hydrothermal orthoclase, albite, chlorite, and sericite, coupled with the
0361-0128/98/000/000-00 $6.00

b
3 cm

c
FIG. 13. Slabs of fresh (a) and altered (b, c) San Rafael granite. (a) Fresh granite with calcic plagioclase (stained red) and coarse megacrystic and matrix phase K feldspar (stained yellow). Small, dark equant grains are cordierite and dark, platy grains are biotite. (b, c) Altered granite with stained side shown in (b) and unstained in (c). Note the markedly diminished amount of red color in this sample resulting from the albitization of the primary calcic plagioclase.

1758

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1759

The dominant alteration process in the granitoid rocks hosting the main San Rafael Lode system was the replacement both of perthitic orthoclase phenocrysts and of groundmass plagioclase by almost Na-free hydrothermal orthoclase, with remarkably little change in megascopic appearance. However, an increase in the turbidity of the igneous alkali feldspar at grain boundaries (Fig. 5g, h) and along intragrain fractures is evident at the hand-lens and microscopic scales in the more weakly K-metasomatized rocks. With increasing alteration, the turbidity extends throughout the grains, with concomitant development of abundant coarse perthite, fluid inclusions, and micropores. In the most strongly K-metasomatized rocks, pitted textures become pervasive within K feldspar grains, but perthite decreases. In rare cases, distinct microscopic domains of Ba-rich feldspar develop, initially along fractures and then invading the surrounding feldspar (Fig. 14). The coarser grains of barian feldspar exhibit rhombic, adularia-like forms (Fig. 14c, d). Accompanying this replacement is the generation of porosity, irregular in places (Fig. 14a, b), that assumes crudely rhombic outlines where the adularia grains are concentrated; such cavities host

Fe-rich chlorite euhedra (Fig. 14c). Hydrothermal biotite has not been observed. In contrast to the potassic alteration, Na metasomatism, erratically distributed in the main San Rafael granite but predominant in the Quenamari granite as well as in the Santo Domingo granite (A.H. Clark, unpub.), characteristically takes the form of replacive rims of white albite on cream alkali feldspar megacrysts (Fig. 4g, h, i), although groundmass plagioclase is also albitized. The rim development is a striking feature in outcrop, particularly where the altered phenocrysts exhibit rounded, ellipsoidal forms. Hence, some Na metasomatic domains are readily delimited although, without Ca staining, the albite overgrowths are difficult to distinguish from the much more locally developed true rapakivi textures. Because the outer surfaces of both the phenocrysts and the remnant K feldspar are rounded, it is possible that such textures are inherited from rapakivi-textured grains; they are not, however, observed in most areas of relict fresh granite in the upper part of the San Rafael stock. As in the K metasomatic domains, increased turbidity of magmatic K feldspar is apparent adjacent to albite rims.

a
25% FeO

100 m 100 m

b
wt. % BaO

20 100 m m

36% FeO

6.6% 6.1%

2.8%
Chl 29% FeO
d

6.3%

5.8% 5.4% 0% 20 100 m m

100 m 100 m

FIG. 14. Back-scattered electron images of K feldspar in altered San Rafael granite sample CC-190. (Note that the higher the average atomic number of an area the brighter the image.) These images illustrate (a, b) the pitted texture of orthoclase resulting from fluid-mediated alteration of magmatic K feldspar; (bright areas in a, c, d) the presence of celsian feldspar; (c, d) zonation and euhedral forms of adularia-like grains; and (c) the late development of chlorite euhedra (chl) occluding pores caused by formation of celsian felspar. Numbers in (d) are wt percent BaO in the K feldspar. 0361-0128/98/000/000-00 $6.00

1759

1760

KONTAK AND CLARK

Alteration of the quartzofeldspathic matrix of the granites results in the development of an allotriomorphic assemblage of subequant grains with sutured boundaries (Fig. 15a), contrasting markedly with the primary hypidiomorphic texture (cf. Fig. 5b, c). Staining shows that this matrix consists of widely variable proportions of pure orthoclase and albite, intergrown with quartz.

Petrography of hydrolytically altered granite Development of fine-grained white mica after feldspars is the most easily recognizable microscopic feature of hydrolytic alteration in many rocks. In K feldspar phenocrysts, muscovite is either disseminated throughout or concentrated along fractures (Fig. 15b). Replacement of plagioclase by

0.4 mm

1.0 mm

1.0 mm

1.0 mm

1.0 mm

1.0 mm

Kf

Tour

1.0 mm

1.0 mm

80 m

FIG. 15. Photomicrographs of altered San Rafael Granite. (a) Matrix showing extensive development of sutured grain boundaries and allotriomorphic texture. (b) White mica in fracture cutting orthoclase. (c) Fine-grained white mica replacing zoned plagioclase. (d) Incipient alteration of biotite along cleavage planes and formation of Ti-rich phases. (e) Extreme chloritization of biotite with formation of ragged edges and secondary minerals along cleavage traces. (f) Muscovite alteration of quartzofeldspathic matrix with formation of rosette textures. (g) Anhedral tourmaline (Tour) and abundant quartz replacing orthoclase (Kf). (h) Tourmaline (lower right) replacing fine-grained granitic matrix. Note that the tourmaline is much coarser than the matrix material. (i) Disseminated acicular cassiterite in metasomatic tourmaline in altered granite, adjacent to a vein exhibiting stages I and II mineral assemblages. 0361-0128/98/000/000-00 $6.00

1760

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1761

fine-grained muscovite, generally in concentric zones about the core area (Fig. 15c), is very common, but coarser muscovite rosettes form at more advanced stages. Secondary muscovite also develops within matrix feldspars and eventually dominates the assemblage, and both fine-grained and coarse rosettes of muscovite coexist with quartz (Fig. 15f). Igneous quartz, both phenocrystic and in the matrix, responds to alteration by development of sutured outlines and internal subdomains as grain-size reduction proceeds, and it becomes inundated with fluid inclusions. Secondary quartz is associated with tourmaline that replaces feldspars and in the matrix (Fig. 15g). Biotite records the effects of hydrogen metasomatism through progressive replacement by chlorite. The alteration apparently occurred at constant volume, the subhedral-to-euhedral forms being retained (Figs.15e, 16). Initial alteration along cleavage planes is associated with a change in color

from dark red-brown to pale orange-brown (Fig. 15d). As alteration continues, biotite is transformed to an Fe-rich chlorite. The alteration generates secondary phases, predominantly rutile, which nucleates along cleavage planes (Fig. 16a) and forms sagenite-like textures. Biotite is locally replaced by a mixture of muscovite and illite (Fig. 16b). Chlorite also replaces feldspar, as radiating or vermicular-textured grains, and is dispersed in the matrix. Beyond its intense development adjacent to some early stage I veinlets and breccia bodies, tourmaline occurs as subhedral to euhedral grains randomly dispersed within metasomatized granites, both within the matrix and replacing coarse K feldspar megacrysts (Fig. 15g). There is a strong association of quartz with coarser tourmaline (Fig. 15g, h). In rare cases, small, apparently systematically oriented, acicular cassiterite euhedra occur within metasomatic tourmaline grains (Fig. 15i). Although tourmaline is clearly a metasomatic phase in many areas, euhedral grains in the matrix of the tourmaline leucogranite (COCA156) are interpreted as magmatic. Whole-rock geochemistry of altered granites Hydrothermal alteration of the San Rafael granite and Quenamari granite, both alkali exchange and hydrolytic, was attended by significant metasomatism. The resulting compositional trends in general correlate satisfactorily with the petrographic subdivisions discussed in the previous section. The compositions of the altered rocks thus define arrays projecting away from the field delimited for the freshest San Rafael granite, with enrichment or depletion in the alkalies (Figs. 6a, 7a, c) and associated mobilization of Ca, Ba, Rb, and Sr (Fig. 6d, e, f). The K enrichment that affected large volumes of the San Rafael granite was associated with an increase in A/CNK ratio (Fig. 6b), which records the development of feldspar-destructive muscovite or its normative equivalent corundum (Fig. 7d), a feature not shown by the Na-metasomatized Quenamari granite. This difference in alkali enrichment in the two areas may reflect erosional level, in that the Quenamari area, from the abundance of fine-grained granitic dike rocks (Fig. 4) and Pb-Zn-Agrich mineralization, probably represents a higher section in the intrusive environment than San Rafael, and would therefore have cooled more abruptly. Ba, Sr, and Rb were mobile during alkali metasomatism. The positive correlation of K with both Rb (Fig. 6f: r = 0.80) and Ba (Fig. 6g: r = 0.70) reflects uptake of Rb in muscovite and of Ba in alkali feldspar. The poor correlation between Rb and Ba (r = 0.39) is further evidence that these elements are sequestered in different minerals. Depletion of both Ca and Sr is plausibly related to destruction of magmatic plagioclase (Fig. 6e). The positive correlation between Ba and Sr in the most altered rocks (Fig. 6d) also suggests that Sr is enriched in the samples with celsian-rich feldspar. However, rocks of varying alteration intensity from the upper 500 m of the main San Rafael stock do not have systematic differences in Zr or Nb contents (Figs. 6h, i, 7f), and these elements are inferred to have been relatively immobile during both alkali metasomatism and hydrolytic alteration. Tin is enriched (Table 2) in only about 20 percent of the analyzed altered rocks, attaining 61 ppm, and of these twothirds are enriched in K rather than Na (Fig. 6c). On this basis

Qtz Kf Alb Chl

Kf

Qtz
250 m

Rutile

Chl Il/Mus
100 m

FIG. 16. Back-scattered electron images of altered granite showing replacement of primary magmatic biotite by chlorite, rutile, and illite/muscovite. Alb = albite, Il = illite, Kf = K feldspar, Mus = muscovite, Qtz = quartz. 0361-0128/98/000/000-00 $6.00

1761

1762

KONTAK AND CLARK

we tentatively conclude that the apex of the largely K-metasomatized San Rafael granite experienced a significantly stronger, albeit erratic, enrichment in originally magmatic Sn than did the Quenamari intrusions. There was, in addition, an inconsistent enrichment of the base metals, and Zn ranges from 56 to 925 ppm and Pb from below detection to 675 ppm. In contrast, there was very little enrichment in Cu; maximum values do not exceed 40 ppm Cu despite the high Cu grades in the veins in the upper San Rafael granite. In Figure 8, REE spectra of selected altered rocks are arranged in order of increasing alteration intensity and are compared with the data for fresh San Rafael granite. In general, there is very little change in the overall abundances of REE in the metasomatized samples, which retain the strongly fractionated chondrite-normalized signature of the fresh granite. This is not unexpected, given that REE are largely sequestered in robust accessory mineral phases (e.g., monazite, zircon, xenotime, and apatite: Gromet and Silver, 1983; Bea, 1996). The Eu anomalies are similar to those in the fresh San Rafael granite, implying that the metasomatic fluid was reduced, thereby limiting the mobility of Eu; this reduced metasomatic fluid is in conformity with both the reduced composition of the granitic magma and the abundance of pyrrhotite in the veins. Samples with moderate alteration (Fig. 8c) exhibit a consistent depletion in LREE compared with both the fresh San Rafael granite and the more intensely altered samples (Fig. 8d). The heavy rare earth elements (HREE), in particular Yb and Lu, appear to deviate the most from the trend for fresh San Rafael granite, implying either greater mobility of these elements or their uptake in new mineral phases. Only one sample (CC-250) deviates markedly (Fig. 8d), and this for the LREE only, perhaps the result of a

reaction involving a LREE-rich accessory mineral such as monazite. Alteration mineral chemistry Plagioclase: Plagioclase progresses toward more sodic compositions in increasingly altered rocks, and albite (An1) characterizes the most altered samples (Fig. 10). Where albite is present, the grains commonly display a pitted texture. There is a clear hiatus from An13 to An2 that separates the moderately from the most strongly altered samples and corresponds closely to the peristerite gap (e.g., Moody et al., 1985). Alkali feldspar is Or90100 in altered rocks, but the most important feature is the erratic presence of barian feldspar. One hydrothermal orthoclase sample (CC-190) contains areas with up to 6.6 wt percent BaO (Table 5), and back-scatter electron imaging reveals a complex zonation with alternating zones rich and poor in Ba (Fig. 14c, d). Bulk analyses of alkali feldspar separates (Table 5) reveal both K and Na enrichment; compositions tend to Or93Ab6An1 and Or14Ab82An4, and reflect subsolidus reequilibration (Smith, 1974). Trace element analyses of separates show that Li, Rb, and Sr abundances are surprisingly uniform and comparable to those in fresh alkali feldspar (Fig. 11). In contrast, Ba shows extreme variability, attaining circa 15,000 ppm in the most K-rich sample (CC-190: Or93) and greatest depletion (325 ppm) in an albitized sample (CC-249: Ab87). Similar Ba-enrichment occurs in K feldspars in the Santo Domingo granite, where concentrations of 6,612 and 6,841 ppm are attained in rocks in which the host feldspar compositions, Or89 and Or86, indicate that postmagmatic processes have modified the primary magmatic chemistry. The REE contents and chondrite-normalized patterns for altered K feldspar (Fig. 11), along with their bulk

TABLE 5. Representative Analyses of Alkali Feldspar in Altered San Rafael-Quenamari Granites, Southeast Peru Granite, sample number, and point Compound or element SiO2 Al2O3 Na2O K2O CaO BaO Total Quenamari COCA-190 23 66.21 18.91 0.00 13.57 0.00 0.44 99.13 Quenamari COCA-190 25 60.31 20.19 0.80 12.25 0.00 6.16 99.72 2.880 1.136 0.072 0.744 0.000 0.112 7.8 80.2 0.0 12.1 Quenamari COCA-190 26 63.46 19.29 0.52 14.28 0.00 2.88 100.43 2.952 1.056 0.048 0.848 0.000 0.056 5.0 89.1 0.0 5.9 Quenamari COCA-190 31 60.22 19.64 0.87 12.39 0.00 6.61 99.74 2.896 1.112 0.080 0.760 0.000 0.128 8.3 78.5 0.0 13.2 San Rafael COCA-141 21 68.07 19.89 11.46 0.05 0.27 0.00 99.74 2.977 1.025 0.971 0.002 0.011 0.000 98.6 0.3 1.1 0.0 San Rafael COCA-151 3 68.09 20.06 8.69 0.30 0.84 0.00 98.46 2.992 1.040 0.744 0.016 0.040 0.000 93.0 2.0 5.0 0.0 San Rafael COCA-151 4 67.20 20.83 10.27 0.17 1.33 0.00 99.82 2.944 1.072 0.872 0.008 0.064 0.000 92.3 1.0 6.7 0.0 San Rafael COCA-165B 64.52 18.92 1.09 13.00 0.17 0.21 97.91 2.998 1.036 0.098 0.771 0.008 0.004 11.2 87.8 1.0 0.5 San Rafael COCA-249 67.50 19.13 7.33 1.85 0.72 0.04 96.57 3.035 1.013 0.639 0.106 0.035 0.001 81.9 13.6 4.5 0.0

Structural formula (8 oxygens) Si 3.032 Al 1.024 Na 0.000 K 0.792 Ca 0.000 Ba 0.008 End member (mol %) Ab 0.0 Or 99.0 An 0.0 Cs 1.0

Data for samples COCA-165B and 249 derived from bulk analyses of mineral separates analyzed by X-ray fluorescence (Kontak, 1985) Analyses of samples COCA-165B and 249 are bulk analyses of mineral separates determined by X-ray fluorescence (Kontak, 1985) 0361-0128/98/000/000-00 $6.00

1762

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1763

compositions expressed as mole percent Or, show that (1) there is an overall enrichment in REE in altered San Rafael granite samples, and two of the most altered samples (CC249, 190) show marked enrichment; (2) the chondrite-normalized patterns are nonetheless preserved during the alteration and enrichment; and (3) there is a notable decrease in EuN, which may reflect removal of Eu through the conversion of Eu2+ to Eu3+ during subsolidus fluid/rock interaction (Sverjensky, 1984). These observations would be consistent with transport of REE in saline, high-temperature fluids as preserved in the quartz megacrysts in the altered granites (discussed in detail below). Samples of alkali feldspar from variably altered granites were selected for X-ray diffraction study for comparison of their compositional and ordering relationships with those from relatively fresh granite. The data (Fig. 12) show that feldspar from metasomatized San Rafael granite differs markedly from that in relatively fresh rocks. In the former, the diffractograms reflect either Na or K metasomatism, as indicated by the increase (samples 4 and 5) or decrease (sample 6) in the intensity of albite reflections. Further, the K-rich domains in K feldspar subjected to either K or Na metasomatism have the composition Or100, as demonstrated by both Orville (1963) and Wright (1968), who contoured the (060) /(2 04) diagram using Na-K exchange experiments. Finally, it is evident that the dispersal of the metasomatized feldspar data along the base of the plot reflects a variable Si/Al ordering, similar to that shown by feldspar in the least-altered granites. Thus, although K-rich domains were developed during metasomatism via K-Na exchange, there was no accompanying migration of Al and Si within the feldspar structure. Biotite: Biotite in altered rocks does not record any systematic chemical changes compared with biotite in fresh samples, even in chloritized zones along cleavage traces. Chlorite was analyzed (Table 6) in three samples that represent contrasted modes of occurrence of this mineral, i.e., alteration zones in primary biotite (CC-151); domains of extreme Fe metasomatism adjacent to mineralized veins (CC-177); and within feldspar (CC-190). Electron microprobe analyses reveal two Fe/(Fe + Mg) populations in the ripidolite and daphnite fields (Fig. 17a, b). The most Fe-rich chlorite generally occurs in altered wall rock adjacent to stage II Sn ore, but similar compositions are also found in the latest chlorite overgrowths occluding secondary porosity in Ba-rich feldspar (Fig.14c). The remaining, more magnesian, chlorites replace primary biotite or are enclosed by feldspar. Minor Mn (to 1.5 wt % MnO) occurs in the chlorite. Tourmaline: Tourmaline exhibits strong optical zoning, and this zoning is reflected in both variable Fe (trace to 15 wt % FeOT) and Mg (trace to 10 wt % MgO), and minor variation in Na (trace to 2 wt % Na2O: Table 7). The compositional variations plotted in Figure 18 are for two tourmalines from altered San Rafael granite remote from a mineralized vein (CC-151,154), one from intensely altered granite adjacent to a mineralized vein (CC-177) and the other from the largest tourmaline leucogranite plug (CC-156). The data define a continuum between Fe-rich and Fe-poor compositions, the tourmaline from sample CC-177 occupying both extremes. The data also overlap with the compositions for tourmaline in stage I stockwork veins recorded by Palma (1981).
0361-0128/98/000/000-00 $6.00

Fe/(Fe+Mg)
a)
4 0.0
Corundophilite

0.5

1.0
Thuringite Chamosite Delessite
Pseudothuringite

Si
6

Sheridanite

Ripidolite Daphnite Brunsvigite

Clinochlore Pycnochlorite

Penninite

7
Diabantite Talc-chlorite

4
2+ 3+

12

Fe + Fe
b)
8
CC-151 CC-177 CC-190

Fe
4

0.4

0.6

0.8

1.0

Fe/(Fe+Mg)
FIG. 17. Compositions of chlorite from moderately (CC-151) and strongly (CC-190) altered granite, and from chlorite-tourmaline rock; CC-177) granite determined by electron microprobe analysis. (a) Binary plot of atomic proportions of Si versus Fe/(Fe + Mg) in classification diagram of Hey (1954). (b) Binary plot of atomic proportions of Fe versus Fe/(Fe + Mg). Note that in these diagrams the chlorite in the strongly altered sample is the most Fe rich; only minor amounts of such chlorite occur in the other samples, in which the Fe/(Fe + Mg) ratios are similar to those of fresh biotite (cf. Fig. 9).

Fluid Inclusions in Quartz Phenocrysts Petrography The fluids responsible for the extensive alteration of the granite, we propose, are represented by secondary inclusions in quartz phenocrysts (Fig. 19). These are locally extremely abundant, and occur both along healed fracture planes (Fig. 19a, d) and as isolated bodies. The inclusions are of irregular or negative-crystal shapes and generally 10 to 30 m in maximum diameter, although some attain dimensions of 75 to 300 m. They may be subdivided (Fig. 19) into the following categories: (i) aqueous-liquid rich, (ii) vapor rich (>8090% vapor), (iii) liquid-vapor- halite, (iv) liquid-vapor-sylvite, and

1763

1764

0361-0128/98/000/000-00 $6.00 TABLE 6. Representative Analyses of Chlorite in Altered San Rafael-Quenamari Granite, Southeast Peru Granite, sample number, and point San Rafael COCA-177 19 San Rafael COCA-177 7 Quenamari COCA-190 9 Quenamari COCA-190 10 Quenamari COCA-190 11 Quenamari COCA-190 12 Quenamari COCA-190 13 Quenamari COCA-190 34 San Rafael COCA-151 10 San Rafael COCA-151 15 San Rafael COCA-151 3 23.68 0.00 22.27 36.32 0.35 5.01 0.00 0.00 87.63 5.320 2.680 3.200 0.000 0.000 6.804 0.056 1.680 0.000 0.000 5.154 2.846 3.168 0.011 0.011 6.507 0.131 1.946 0.016 0.025 5.292 2.708 2.920 0.000 0.000 5.068 0.084 3.808 0.000 0.000 5.432 2.568 2.836 0.000 0.000 4.200 0.280 4.536 0.000 0.000 5.376 2.624 2.808 0.000 0.000 4.396 0.196 4.480 0.000 0.000 5.432 2.568 2.976 0.000 0.000 4.452 0.364 4.004 0.000 0.000 5.292 2.708 2.920 0.000 0.000 5.124 0.112 3.752 0.000 0.000 23.35 0.07 23.11 35.24 0.71 5.92 0.08 0.06 88.54 24.49 0.00 22.04 28.03 0.48 11.80 0.00 0.00 86.84 25.67 0.00 21.72 23.74 1.64 14.40 0.00 0.00 87.17 25.41 0.00 21.79 24.74 1.11 14.14 0.00 0.00 87.19 25.59 0.00 22.08 24.94 1.94 12.66 0.00 0.00 87.21 24.37 0.00 21.94 28.15 0.59 11.58 0.00 0.00 86.63 23.81 0.00 22.02 34.35 1.23 5.39 0.00 0.00 86.80 5.348 2.652 3.172 0.000 0.000 6.468 0.224 1.820 0.000 0.000 26.12 0.00 20.33 26.66 0.31 12.72 0.00 0.00 86.14 5.628 2.372 2.808 0.000 0.000 4.816 0.056 4.088 0.000 0.000 25.68 0.00 20.42 26.70 0.32 12.76 0.00 0.00 85.88 5.572 2.428 2.780 0.000 0.000 4.844 0.056 4.116 0.000 0.000 28.21 0.73 19.73 24.63 0.33 12.42 0.00 0.00 86.05 5.964 2.036 2.864 0.000 0.000 4.340 0.056 3.920 0.000 0.000 KONTAK AND CLARK

Compound or element

San Rafael COCA-177 17A

San Rafael COCA-177 18

1764

SiO2 TiO2 Al2O3 FeO MnO MgO CaO Na2O Total

21.94 0.00 21.05 40.39 0.84 1.28 0.00 0.00 85.50

21.27 0.00 20.07 39.70 1.33 1.76 0.00 0.00 84.13

Structural formula (28 oxygens) Si 5.236 Al(iv) 2.764 Al(vi) 3.144 Ti 0.000 Cr 0.000 Fe 8.036 Mn 0.168 Mg 0.448 Ca 0.000 Na 0.000

5.180 2.820 2.948 0.000 0.000 8.092 0.280 0.644 0.000 0.000

Sample COCA-190 is a zoned grain; columns 9 through 13 represent core to rim analyses

TABLE 7. Representative Analyses of Tourmaline in Altered San Rafael Granite, Southeast Peru

0361-0128/98/000/000-00 $6.00 35.76 0.00 34.62 13.21 0.00 0.97 0.00 1.59 86.15 6.832 1.168 6.616 0.000 2.100 0.000 0.280 0.000 0.588 6.860 1.140 6.280 0.000 0.728 0.000 1.988 0.056 0.728 6.804 1.196 6.084 0.028 0.084 0.000 2.744 0.084 0.952 6.916 1.084 6.252 0.056 2.380 0.000 0.308 0.028 0.672 6.692 1.308 6.924 0.000 0.084 0.000 0.168 0.000 0.616 6.692 1.308 6.924 0.000 0.084 0.000 0.168 0.000 0.616 6.664 1.336 7.204 0.000 0.140 0.000 1.540 0.000 0.420 6.664 1.336 7.204 0.000 0.140 0.000 1.540 0.000 0.420 6.580 1.420 6.364 0.084 2.072 0.000 0.392 0.028 0.812 6.216 1.784 6.644 0.056 1.988 0.028 0.224 0.028 0.784 6.412 1.588 6.504 0.056 1.932 0.000 0.392 0.028 0.784 6.664 1.336 6.140 0.112 1.512 0.000 1.204 0.056 0.780 37.22 0.00 34.14 4.76 0.00 7.19 0.26 2.07 85.64 37.76 0.30 34.20 0.47 0.00 10.15 0.48 2.69 86.05 35.67 0.30 32.12 14.56 0.00 1.07 0.17 1.76 85.65 37.14 0.00 38.74 0.47 0.00 6.88 0.00 1.73 84.96 37.14 0.00 38.74 0.47 0.00 6.88 0.00 1.73 84.96 36.96 0.00 40.26 0.85 0.00 5.75 0.00 1.16 84.98 36.96 0.00 40.26 0.85 0.00 5.75 0.00 1.16 84.98 34.57 0.68 34.70 13.10 0.14 1.39 0.18 2.20 86.96 32.67 0.43 37.47 12.48 0.33 0.88 0.22 2.18 86.66 34.08 0.45 36.44 12.31 0.13 1.41 0.19 2.19 87.20 35.54 0.87 33.84 9.61 0.00 4.35 0.31 2.13 86.65 35.47 1.06 32.39 8.87 0.00 5.59 0.62 2.18 86.18 6.692 1.308 5.888 0.140 1.400 0.000 1.568 0.140 0.812

Compound Sample number and point or COCA-177 COCA-177 COCA-177 COCA-177 COCA-177 COCA-177 COCA-177 COCA-177 COCA-177 COCA-177 COCA-177 COCA-156 COCA-156 COCA-156 COCA-151 COCA-151 element 1 2 3 4 5 6 7 8 8 13 13 51 52 53 20 21

SiO2 TiO2 Al2O3 FeO MnO MgO CaO Na2O Total

35.78 0.00 36.16 12.13 0.00 0.74 0.24 1.13 86.18

36.36 0.00 35.95 11.49 0.00 0.69 0.00 1.24 85.73

37.29 0.00 32.62 4.29 0.00 8.21 0.61 2.39 85.41

Structural formula (28 oxygens) Si 6.776 6.860 Al(iv) 1.224 1.140 Al(vi) 6.840 6.868 Ti 0.000 0.000 Fe 1.932 1.820 Mn 0.000 0.000 Mg 0.196 0.196 Ca 0.056 0.000 Na 0.420 0.448

6.888 1.112 5.972 0.000 0.672 0.000 2.268 0.112 0.868

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1765

a)

b)

c)

Ca
0.1 0.2

Na

Fe

0.3

0.4

0.6

0.8

1.0

0.0 0.0 0.2 0.4 0.6

Fe/(Fe+Mg)
0.8 1.0

CC-151 CC-154 CC-156 CC-177

1765

FIG. 18. Compositions of tourmaline from negligibly (CC-154), moderately (CC-151) and strongly altered (chlorite-tourmaline rock, CC-177) granites and the tourmaline leucogranite (CC-156), determined by electron microprobe analysis. Results are plotted in various binary element plots (atomic proportions). Note that the extreme end-member compositions are represented by the most altered rock (CC-177) and that the tourmaline in the leucogranite sample is extremely Fe rich, whereas that in the other granite samples has an Fe/(Fe + Mg) ratio similar to that of fresh biotite (cf. Fig. 9).

(v) liquid-vapor-multiple solids, which, from optical properties and scanning electron microscopy (SEM) data include, in addition to halides, Fe-rich mica, hematite, and sulfides. The simple liquid-vapor inclusions possess a large range in salinity, but it is not possible to distinguish high- and low-salinity (i.e., 023 wt % NaCl equiv) types on the basis of petrography. Petrographic evidence that fluid unmixing occurred within the system (e.g., Bodnar et al., 1985b) includes the occurrence of discrete fluid inclusion assemblages (i.e., FIA of Goldstein and Reynolds, 1994) of the appropriate inclusion types (e.g., Fig. 19c) along healed fractures, and the coexistence of vapor-rich inclusions with inclusion types iii and iv. Moreover, because phenocrysts containing the greatest number of

1766

KONTAK AND CLARK

300 m

200 m

100 m 20 m H

d g
H 100 m B

100 m 30 m

30 m

f
H

30 m H

15 m

30 m

FIG. 19. Photomicrographs of fluid inclusions in phenocrystic quartz of the San Rafael granite; all photos taken in planepolarized transmitted light. (a) Typical area of quartz phenocryst in San Rafael granite showing abundant healed fracture planes decorated with fluid inclusions. (b) Typical area of quartz phenocryst inundated with liquid- and vapor-rich fluid inclusions. Two areas are enlarged to illustrate the contrasting nature of the fluid inclusion populations. (c) Healed fracture plane inundated with negative-crystalshaped, vapor-rich fluid inclusions. Inset shows part of the area enlarged. (d) Multiple healed fracture planes in quartz that are decorated with abundant fluid inclusions of liquid-vapor and liquid-vapor-solid types. Two areas (g, j) are enlarged in Figures 19g and j. (e) Pair of multisolid fluid inclusions with halite (H), birefringent (B) phase and other undetermined solids. Image is a mosaic with these inclusions forming part of a plane containing numerous such inclusions. (f) Pair of equant liquid-vapor fluid inclusions; (g) Multisolid, liquid-rich, negative-shaped fluid inclusions with halite and other unidentified solids; (h) Healed fracture plane with two-phase liquid-vapor fluid inclusions of negative crystal shape; (i) Healed fracture plane with two-phase liquid-vapor fluid inclusions of irregular shape; (j) vapor-rich, halite-bearing fluid inclusion; and (k) Plane of liquid-vapor halite fluid inclusions that are in close proximity to plane of liquid-vapor fluid inclusions (large inclusion on the right). Image is mosaic to show the various inclusion populations represented. 0361-0128/98/000/000-00 $6.00

1766

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1767

vapor-rich inclusions also host numerous hypersaline (liquidvapor-halite-sylvite) inclusions, commonly along the same fracture planes, we infer that unmixing occurred within the upper part of the stock. Microthermometric relationships Following reconnaissance examination of approximately 45 samples of granitic rocks from outcrop and underground exposures, eight surface samples (Fig. 1c) were selected for detailed microthermometric study. These samples represent both the main San Rafael granite and Quenamari granite stocks, as well as several of the flanking dikes, and embrace all categories of pervasive alteration intensity. The data for heating and freezing experiments are summarized in Figure 20. Ice was the last phase to melt in all two-phase (liquid-vapor) aqueous inclusions, but in many cases the melting of hydrohalite could also be observed. First-melting or eutectic temperatures range from 65 to 0.5C (Fig. 20a), and distinct populations occur at 20C, 30 to 45C, and 50C. As emphasized by Kontak and Clark (1997a), these data show that the fluid must contain solutes in addition to NaCl.

Nonsaline fluids (i.e., 0 wt % NaCl equiv) are also very widely preserved. The lack of clathrate formation or freezing of the vapor phase during cooling runs indicates that condensed gases such as CO2, if present, occur in minor concentrations. Fluid salinities, estimated from the last melting of ice and dissolution of daughter minerals (Bodnar, 1994; Bodnar and Vityk, 1994), range from circa 0 to 65 wt percent NaCl equiv, but with two distinct groupings that correspond to liquidvapor and liquid-vapor-halite inclusion types (Fig. 20b). The histogram indicates that the liquid-vapor inclusions exhibit a continuum in salinity, but L/V ratios confirm that variable salinities commonly occur within apparently coherent populations of such inclusions. This behavior is observed for liquid-vapor type inclusions with salinities below circa 10 wt percent equivalent NaCl; thus, the last melting of ice varies from 3 to 7C in such assemblages, indicating a range of 4 to 7 wt percent NaCl. For hypersaline inclusions, the NaCl/(NaCl + KCl) ratio varies from 0.47 to 0.63, as determined from projection of the dissolution temperatures of halite and sylvite into the NaCl-KCl-H2O ternary diagram. Inclusion homogenization occurs at 205 to 540C (Fig. 20c)

a)
Frequency

60 50 40 30 20 10 0 -70 -60 -50 -40 -30 -20 -10 0

b)
Mg, Fe

200 160 120 80 40 0 0 10 20 30 40 50 60 70

Na, K Ca

first melting

( C)

Frequency

Salinity (wt. % eq. NaCl)

c)
Frequency

100 80 60 40 20 0 200 300

d)
Th(Halite ) Th(L) Th(V)

80
Th(L) Th(Halite) Th(V)

(wt. % eq. NaCl)

60
I

Salinity

40 20
IV II III

400
o

500

600

0 100

200

300

400
o

500

homogenization

( C)

homogenization

( C)

FIG. 20. Summary of microthermometric data for secondary fluid inclusions hosted by phenocrystic quartz in the San Rafael granite, and a comparison with primary inclusions in quartz and cassiterite in the San Rafael Lode. (a) First-melting or eutectic temperatures. The cations shown are the suggested dominant species in solution accounting for the first-melting temperatures. (b) Salinity vs. wt percent NaCl equiv. (c) Final homogenization temperature for inclusions grouped according to the nature of homogenization, namely expansion of the liquid phase (Thliquid), expansion of the vapor phase (Thvapor), and dissolution of halite (Thhalite); (d) Comparative plot of salinity vs. homogenization temperature for all inclusion types hosted by quartz phenocrysts and lode minerals fields for stages I to IV (from Palma (1981) and A.H. Clark (unpub. data), as discussed in the text). 0361-0128/98/000/000-00 $6.00

1767

1768

KONTAK AND CLARK


o T(H)-T(L)=100 C

Pressure (fluid) kbars

L(50) + H L

L(40) + H L

40

through dissolution of solid phases (i.e., halite and sylvite), liquid expansion, and vapor expansion. Whereas homogenization by the first two processes occurs throughout this wide T range, homogenization by vapor expansion is restricted to 370 to 480C, except for a few cases at 305C. In view of the possibility of error in establishing the homogenization of vapor-rich inclusions (Sterner, 1992), microthermometry was undertaken only on inclusions with the appropriate shape (i.e., reentrants, as discussed by Roedder (1984, p. 256)). The fact that the highest value for Th-V is similar to that for Th-L suggests that the observed measurements are close to the true trapping temperature. For inclusions containing halite, final homogenization occurred by either dissolution of halite or expansion of the liquid phase after halite had dissolved (Fig. 21a). The relationship between salinity and homogenization temperature is shown in Figure 20d. Here, the data are compared with the fields delimited for the lode mineralization for stages I, II, III, and IV by Palma (1981) and A.H. Clark (unpub. data). There is clearly a correspondence and overlap in the data for fluid inclusions in the quartz phenocrysts and those hosted by vein minerals. We emphasize the large range in salinity for fluids homogenizing within the temperature range 300 to 500C, which strongly implies fluid mixing. In contrast, there is a relatively limited range in salinity for inclusions homogenizing below 300C. Solute compositions The compositions of the fluids trapped in the inclusions were herein estimated from the final melting temperatures of ice and hydrohalite, the dissolution of solid phases, analysis of decrepitates, and scanning electron microscopic analysis of opened inclusions. The first two methods reveal a nearly continuous range of salinity from 0 to 65 wt percent NaCl equiv, excluding a small gap in the 25 to 35 wt percent interval (Fig. 20b). Although much of the solute is inferred to be NaCl, the low first-melting temperature, combined with melting of hydrohalite and ice (Oakes et al., 1990), indicates that CaCl2, although usually less than 10 wt percent, locally accounts for up to 40 wt percent of the solute (Kontak and Clark, 1997a). Phase relations in the system NaCl-MgCl2-H2O imply that MgCl2 is also present (Crawford, 1981). In the few cases where dissolution of both sylvite and halite was observed, the bulk NaCl/KCl/H2O ratio (wt %) is estimated from the NaClKCl-H2O ternary to be 35:25:40; the maximum NaCl content is 50 wt percent. Thus, the fluid component of some inclusions is inferred to contain significant quantities of both CaCl2 and KCl. Fluid composition was also estimated by analysis of decrepitate mounds (Figs. 22, 23; cf. Haynes et al., 1988). Although no attempt has been made to quantify the data in terms of the bulk fluid chemistry, the results of the analyses can be used to characterize the composition of the fluid in the quartz-hosted inclusions. Using both point- and raster-type analyses, the detected cations, in order of decreasing abundance, are Na, K, Ca, Fe, Mn, and Mg. Traces of Sr and P occur, but Ba was not detected in any analysis. The abundance of cations other than Na and K may account for the massive nature of some mounds (Fig. 22), because more equant shapes develop where the alkalies dominate (D.J.
0361-0128/98/000/000-00 $6.00

a)
T ( Halite) homogenization ( oC)

600

Th(H)>T(L)
500
(55 wt. %) (50 wt. %)

400

3 2
300

A
(35 wt. %)

200

Th(H)<T(L)
200 300 400 500 600

T ( L) homogenization (C)

b)
20 0C
C 150
0 C 25
5.0

30

0 C
0 35
gr a Bea u voir g r a n

1:
(45 wt. %) (40 wt. %)
4.0

1
0 C
3.0

45

C 0

satu H 2O satura

2.0

g ted gr

1.0

3 2
200 400

ni

600

Temperature (C)
FIG. 21. (a) Binary plot of homogenization temperatures for three-phase liquid-vapor-halite inclusions. Circled numbers 1, 2, 3 represent the three different homogenization behaviors for liquid-vapor-halite inclusions. The area A refers to that part of the diagram for which there are no homogenization data (i.e., no Thliquid at <280C), the significance of which is discussed in the text. The horizontal dashed lines with numbers refer to wt percent NaCl inferred from dissolution temperature of halite (Bodnar and Vityk, 1994), whereas the inclined dashed line is a reference line for inclusions displaying homogenization of halite at 100C above Thliquid, a relationship that has implications for the minimum pressure of entrapment (see text for discussion). (b) A pressure versus temperature plot for the system H2O-NaCl (after Bodnar, 1994) showing (1) lines of constant liquid-vapor homogenization temperature or iso-Th lines for fluid inclusions containing 40 wt percent NaCl; (2) the liquidi for two relevant compositions (i.e., liquid[40, 50] + halite); and (3) the 40 wt percent NaCl vapor-pressure curve for a 40 wt percent NaCl system (liquid[40] + vapor). The circled numbers 1, 2, 3 represent different homogenization behavior for liquid-vapor-halite inclusions, as in (a). The horizontal dashed line refers to the inferred maximum pressure for fluid entrapment on the basis of the difference in homogenization of halite and liquid-vapor in three phase liquid-vapor-halite inclusions (i.e., T(H) T(L) = 5075C). The horizontal solid line with arrow indicates path of isobaric cooling for the fluids. Also shown are the solidi for two relevant granite systems, one a volatile-rich (F, B) system (Beauvoir granite; Pichavant et al., 1987), and the other the water-saturated haplogranite (Tuttle and Bowen, 1958). Note that the inclusion data are consistent with exsolution of a fluid from a volatile-rich Beauvoir-type granite at low pressures with subsequent isobaric cooling.

e i te

te

800

1768

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1769

10 m
FIG. 22. Back-scattered electron image for fluid inclusion decrepitate mound, released from a quartz phenocryst in the San Rafael granite. The variation in brightness reflects compositional variability.

Kontak, unpub. data). In most cases the Na/K ratio is much greater than unity, but K dominates in several mounds (Fig. 23a). There is a decoupling of K from Ca and Fe (Fig. 23a, b) and also from Mn and Mg (not shown), whereas (4) Na is generally coupled with Fe (Fig. 23b) and to a lesser extent Ca (Fig. 23a, c), although Fe appears to be decoupled from Na in the most saline fluids (Fig. 23e). Finally, the existence of a Ca-dominated fluid is confirmed, consistent with some lowtemperature first-melting and the hydrohalite-ice melting relationships. The high concentrations of Fe in some fluid inclusions is also demonstrated by SEM analysis (A.J. Anderson, pers. commun., 1987) of several opened inclusions in quartz phenocrysts in a single sample of weakly altered granite, which revealed numerous Fe chloride daughter crystals and lesser numbers of Fe-Mn chloride crystals. Because of the dispersed areal distribution of our observations on fluid inclusions in the quartz phenocrysts, it is difficult to correlate specific fluid cation contents with alteration facies. However, Fe-rich fluids may be more prevalent in proximity to the major lodes within the San Rafael granite, in conformity with the Fe metasomatism recorded by their chloritized envelopes.

a)

Na

d)
60

wt. % Na

40 20 0 0 10 20 30 40 50

Ca

K Na

wt. % K

b)

e)
wt. % Fe

60 40 20 0 0 10 20 30 40 50

Fe

K Na

wt. % Na

c)

f)
wt. % Ca

8 6 4 2 0 0 10 20 30 40 50

Fe

Ca

wt. % Na
FIG. 23. Binary and ternary element plots for analyses of decrepitate mounds (results normalized to 100 wt %) released from fluid inclusions hosted by quartz phenocrysts from the San Rafael granite. 0361-0128/98/000/000-00 $6.00

1769

1770

KONTAK AND CLARK

Temperature-salinity relationships and correlation with mineralization stages The data for quartz phenocryst-hosted fluid inclusions may be grouped (Fig. 20d) into three domains: a high-salinity trend (>35 wt % NaCl equiv) with Th = 260 to 530C; a hightemperature (300480C), variable-salinity (323 wt % NaCl equiv) group; and a lower-salinity (<10 wt % NaCl equiv) and lower-temperature (200 to 300C) group. Although the highsalinity inclusions reflect saturation or near-saturation of the fluid with respect to halite, it is significant that homogenization of these inclusions occurs either by dissolution of halite or expansion of the L phase (Fig. 21a). From Figure 20d, it is evident that the data for stage I barren quartz-tourmaline veins are similar to those for the high-temperature, highsalinity quartz phenocryst-hosted inclusions of this study. The main mineralizing stages, II and III, correspond to the majority of the fluid inclusion data in this study, although numerous inclusions in quartz phenocrysts homogenize at higher temperatures than those recorded by Palma (1981) and A.H. Clark (unpub. data) for the lodes (420C). Stage IV of the lode paragenesis formed from fluids with uniformly low salinities (5 wt % NaCl equiv) and temperatures (170300C: Palma, 1981), overlapping with a large proportion of the data collected for quartz phenocryst-hosted inclusions in this study. Pressure of fluid trapping The pressure of entrapment for the earliest-stage fluids preserved in the quartz phenocrysts is estimated to have varied from 150 to 615 bars (Kontak and Clark, 1997a), on the basis of microthermometric observations on inclusions that we infer record fluid unmixing in the uppermost levels of the stock and for which the temperatures of halite dissolution and homogenization to liquid are available, and by applying the method of Bodnar and Vityk (1994; Fig. 24). We emphasize that this does not imply that higher fluid pressures did not occur, but that the fluids exhibiting appropriate homogenization behavior and recording these higher pressures have not specifically been encountered. In fact, the homogenization behavior of types iii and iv inclusions, where Th(H) exceeds Th(L) by 50 to 100C, and rarely up to 200C, indicates that pressures as high as 2 kbars must have been attained, as discussed below. Given the high-level setting of the San Rafael granite, such elevated pressures plausibly record fluid overpressuring caused by the tensile strength of the overlying rock, preventing fluid escape during second boiling. That fluid pressure fluctuated at a later stage is implied by the multiphase hydrothermal brecciation textures characteristic of the stage II assemblages in the veins (e.g., Palma, 1981). As proposed by Cline and Bodnar (1994), and also discussed by Bodnar (1994), the three different modes of homogenization of the halite-bearing inclusions are also consistent with fluctuations in fluid pressure combined with cooling of the fluid (Fig. 21a), as modeled for a 40 wt percent NaCl fluid in Figure 21b. The higher-salinity fluids represented in Figure 21a can also be modeled as for the 40 wt percent fluid (e.g., Bodnar and Vityk, 1994); the L + H L plane was displaced to higher temperatures in P-T space with increasing salinity. The retrograde evolution of similar high-salinity
0361-0128/98/000/000-00 $6.00

1500
30 40 50

Pressure (bars)

wt. % NaCl
1000

60 70

500

CP
0

L+

V+

100

300

500

700

Temperature (C)
FIG. 24. Plot of pressure vs. temperature with vapor-pressure curves for H2O-NaCl solutions with salinities of 30 to 70 wt percent (from Bodnar and Vityk, 1994). Also shown are the liquid-vapor curve for H2O (0), the critical point (CP) and the three phase (liquid + vapor + halite: L + V + H) curve in the H2O-NaCl system. Plotted are homogenization data for fluid inclusions from San Rafael that represent trapping in a two phase field (i.e., fluid unmixing demonstrated). See Bodnar and Vityk (1994) for details of this plot.

fluids has been discussed for the Capitan pluton, New Mexico (Ratajeski and Campbell, 1994; Campbell et al., 1995). In Figure 21a, the three P-T regimes of entrapment, representing different fluid types in terms of bulk density, are distinguished on the basis of the mode of homogenization: by vapor-phase disappearance, by simultaneous halite dissolution and inclusion homogenization, and by halite dissolution. That homogenization occurs by all three processes is consistent with a model involving fluid evolution through a combination of cooling and overpressuring, given the shallow crustal setting. The inclusions that homogenized at temperatures above circa 500C must have been trapped at no more than 0.5 to 1 kbar for salinities of circa 40 wt percent NaCl equiv, because the iso-Th lines intersect the solidus of a representative evolved granite at this point (e.g., Beauvoir granite in Fig. 21b). Higher pressures and by inference higher temperatures are, we argue, inconsistent with the clearly secondary context of the fluids in a rock that had attained brittle conditions that permitted development of microcracks along which the fluids infiltrated. Cooling of this fluid with concomitant overpressuring would lead to trapping of inclusion types 2 and 3. The absence of data in area A of Figure 21a indicates that when fluids of appropriate salinity (i.e., 35 wt % NaCl) were circulating, overpressures could not have existed, as occurred for fluid type 3, or the inclusions would exhibit homogenization via Th (H) > Th (L). This is an additional argument favoring a generally low- pressure environment during mineralization, but with overpressuring accounting for inclusions (fluid 3) with Th(H) > Th(L). The generally lower pressures of formation we infer are in agreement with the low, epizonal pressures estimated from fluid inclusion data for vein mineralization in both the San Rafael and Quenamari sectors of the district. Hannington

1770

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1771

(1983) estimated that the pressures attending vein formation at Quenamari (Veta Nazareth) may have been as low as 180 bars, whereas Palma (1981) concluded that the main lode stage (stage II) at San Rafael formed at 165 bars. These estimates are in keeping with the shallow level of emplacement of the San Rafael granite and Quenamari granite inferred from the petrologic features and the intensely brecciated nature of the veins. An alternative model would involve quenching, perhaps through dilution with a cool fluid, as discussed above to account for the salinity-Th correlations in Figure 20d. Given the structural features of the mineralized veins, a model involving both pressure fluctuation and cooling is preferred. In view of the temporal hiatus between initial retrograde boiling and ore formation, it is plausible that the transition from highly saline stage I fluids to the more dilute, cooler fluids that generated stages II and III coincided with a change from dominantly lithostatic to dominantly hydrostatic conditions. Ore-Genetic Modeling The large-scale hydrothermal alteration system we document herein introduces a new perspective to the ore-genetic model for this world-class lithophile-metal deposit. We propose that the Sn-Cu mineralization was generated through an exceptional conjunction of magmatic, regional and local tectonic, and hydrothermal processes, many of which are broadly shared by other major sulfide-cassiterite centers (Smirnov, 1937), a diversified but coherent clan that now dominates prospective hardrock cassiterite mineralization globally (Clark, 1995; Clark et al., 1993, 1995). Such common, if not invariable, characteristics include parental peraluminous, reduced, granitoid magmas, generally of high-temperature Lachlan, rather than Himalayan type, that exhibit evidence of commingling and mixing with mantle-derived, oxidized, mafic melts, and that widely include K-rich lamprophyric members. Subordinate volumes of strongly fractionated, commonly F- and/or B-enriched melts may have been closely linked to ore formation, but their role remains uncertain. Second, chalcopyrite, iron sulfides and, less consistently, sphalerite are abundant, representing integral constituents of the ores. Third, the ore-forming fluids are of both magmatic and nonmagmatic, most commonly meteoric, origin, and, finally, ore deposition significantly postdated the initiation of retrograde boiling in an epizonal magma chamber, and it was constrained by structures imposed by regional tectonism (Clark et al., 2000). We regard these varied factors as ultimately responsible for the extreme shallow-crustal concentration of Sn, although the dilational jogs on the main San Rafael Lode structure (Fig. 2) undoubtedly controlled the development of the enormous orebodies that dominate the deposit. Despite the close association of hydrolytic wall-rock alteration with both Sn and Cu ore-zones, it is clear that the San Rafael Lode system differs in almost all salient features from the widespread low-grade, greisen-bordered, sulfide-poor, cassiterite (-wolframite) vein swarms that have been the subject of most investigations of lithophile-metal mineralization. In contrast, as was initially emphasized by Palma and Clark (1982), the San Rafael deposit is remarkably similar to the classic Cornwall-type Sn-Cu-polymetallic lodes, particularly those of the Camborne-Redruth district (Clark et al., 1995).
0361-0128/98/000/000-00 $6.00

San Rafael does not conform to the tin porphyry model of Sillitoe et al. (1975), in that cassiterite deposition was entirely focused within throughgoing brittle shear zones, and retrograde boiling processes did not immediately give rise to the mineralization. The major goals of this contribution were to determine whether the P-T-X evolution of the dispersed fluids documented herein corresponded to that of the Sn-Cu lodes, and thereby to evaluate the extent to which the structurally localized ore formation process is reflected throughout a considerably greater area by alteration assemblages and, specifically, fluid inclusions trapped in the granitic host rocks. Significance of cryptic pervasive alteration The upper 700 m of the San Rafael stock widely preserves coarsely phenocrystic megascopic textures and even mineral associations, e.g., poorly ordered alkali feldspars exhibiting film perthite lamellae, of unambiguous magmatic origin (Brown and Parsons, 1989). However, the majority of surface and underground exposures of both stocks and dikes, outside of the well-defined alteration envelopes of the veins, record moderate to intense alkali metasomatism and/or hydrolytic alteration. The presence of pervasive, alkali-dominant alteration within rare-element (Sn-W-Nb-Ta)mineralized felsic systems has been documented globally (e.g., Pollard, 1983; Witt, 1984, Pollard and Taylor 1986; Charoy and Pollard, 1989), and there is a consensus that alkali redistribution occurred during circulation of hydrothermal fluids, a process manifested by a vertical zonation of alkalis in intrusive cupolas that contain Na-rich apices and K-rich cores (Orville, 1963; Lagache and Weisbrod, 1977; Pichavant, 1983; Lagache, 1984; Webster, 1995). Such metasomatism may be dispersed or concentrated along veins, and the geochemical and mineralogic similarities of these alteration facies reveal a cogenetic origin (Pollard, 1983; Witt, 1988), the former recording entrapment of late magmatic or orthomagmatic fluids, and the latter the involvement of hydrofracturing (Pollard and Taylor, 1986; Titley, 1994). In addition, this early alkali metasomatism is almost ubiquitously succeeded by hydrolytic alteration. The nature and distribution of the early alkali metasomatism in the San Rafael granite strongly support an origin through the subsolidus infiltration of fluid rather than the in situ retention of a late-stage magmatic brine, such as is inferred for the essentially closed systems represented by the tourmaline orbicules in the Seagull batholith (Samson and Sinclair, 1992) or the cassiterite pipes of the Zaaiplaats deposit (Pollard et al., 1989, 1991), the former barren and the latter high grade, but modest in scale. The alteration in the San Rafael granite is variable in both intensity and distribution, fluid infiltration having been controlled by parameters such as fluid/rock ratio and permeability (cf. Ferry, 1979). The lack of a relationship between the alteration and the veinlets and the absence of vein-selvage alkali alteration, such as is observed in porphyry deposits, indicate that fluid migration was not a consequence of fracture-controlled permeability. The selective replacement and incremental progression of the alteration normally associated with porphyry deposits would not have provided an efficient mechanism for formation of a large and high-grade tin deposit (Heinrich, 1990), even if the temperatures were not excessive for cassiterite

1771

1772

KONTAK AND CLARK

saturation. Whereas comparative relationships among the various granitic stocks of the Picotani intrusive suite (Kontak and Clark, 1988) strongly suggest that the K metasomatism exhibited by the San Rafael granite is an inherently favorable feature from the exploration standpoint, no significant areal zonation in alteration intensity exists with respect to the major lode concentration. The Al/Si ordering relations in the alkali feldspar provide insight into the conditions of the alkali metasomatism. The absence of microcline, normally the dominant K-rich alkali feldspar in similar settings (e.g., Martin and Bowden, 1981), is consistent with temperatures above the orthoclase-microcline inversion at circa 450C (Smith, 1974). In contrast, the compositions of the Na and K feldspar phases, inferred from both microprobe (Fig. 10) and X-ray powder data, suggest that equilibration persisted to temperatures as low as 300C. There was evidently insufficient strain energy associated with this lower-temperature alteration to promote ordering in the feldspars (Brown and Parsons, 1989). Magmatic versus nonmagmatic fluid components and integration with paragenetic stages The secondary fluid inclusions hosted by the quartz phenocrysts of the San Rafael granite provide evidence of the nature of the metasomatic fluid that infiltrated the stock. The following three distinct fluid populations have been recognized: a spectrum of hypersaline Na-K fluids of 35 to 65 wt percent NaCl equiv that homogenize between 280 and 540C; moderately saline fluids of 10 to 20 wt percent NaCl equiv that homogenize at 300 to 470C; and a low-salinity fluid of 0 to 10 wt percent NaCl equiv that homogenizes at 200 to 350C. Also present is a low-density, vapor-rich inclusion population for which limited thermometric data indicate a salinity of range of 4.2 to 19.4 wt percent NaCl equiv (n = 10) and homogenization temperatures of 370 to 480C overall, most falling in the 380to 400C range. The Th values and salinities of these inclusions overlap with the high-temperature limit of the vein stage II fluids. Petrographic observations strongly support the coexistence of all types of inclusions along individual healed fracture planes, providing evidence for fluid unmixing. The question that must therefore be addressed is whether the spectrum of fluid inclusion composition and temperature, in both the pervasive alteration and lode environments, may be explained through the evolution of a single magmatic brine or, alternatively, requires the involvement of two or more hydrothermal fluids of disparate origin. In the magmatic-hydrothermal model, associations of highand low-salinity aqueous fluids could record either unmixing of an original moderate-salinity fluid or, at low pressure, direct exsolution of associated brine and vapor from the magma (e.g., Bodnar et al., 1995a). Although these processes may have contributed to the fluid inclusion populations at San Rafael, it is significant that, in both the quartz phenocrysts and veins, a significant proportion of the liquid-vapor inclusions that homogenize at 200 to 300C yielded ice-melting temperatures of 0 to 0.5C, evidence for negligible salinity. The various potential evolutionary paths of a magmatic brine would not, according to phase equilibrium data for the system H2O-NaCl (Bodnar et al., 1985a), yield fluids with less than 2
0361-0128/98/000/000-00 $6.00

to 3 wt percent NaCl. We therefore advocate the involvement of nonmagmatic, most plausibly meteoric, waters with salinities in the range 0 to 0.7 wt percent NaCl equiv, in agreement with many other studies of epizonal magmatic-hydrothermal environments (e.g., Norton, 1982; Dong and Zhou, 1996; Hezarkhani and Williams-Jones, 1998). Indeed, the inferred overall trajectory of hydrothermal evolution in the upper section of the San Rafael stock was similar in almost all salient respects to that defined in many magmatic-hydrothermal centers, including the great majority of porphyry copper deposits, in which evidence of late-stage meteoric water incursion is almost ubiquitous. As discussed previously, and illustrated in Figure 20d, there is an excellent correlation between the fluid inclusion data from this study and those for the vein stages. Interpretation of the links between the dispersed fluids and those channeled by the major lode structures must, however, take into account the apparent circa 2 m.y. hiatus between mineralization stages I and II. The formation of stage I quartz-tourmaline veins is plausibly related to the high-salinity, high-temperature fluids that have a signature typical of retrograde boiling and presumably coexisted with the vapor-rich fluids. These inclusions, therefore, preserve a record of an early period of fluid infiltration into the San Rafael granite and its surrounding ring dikes. The introduction of the moderately saline (25 wt % NaCl equiv) fluids may have begun late in stage I but was, we infer, largely coincident with stages II and III mineralization, at which time mixing with a more dilute fluid yielded the observed trends in temperature-salinity space (Fig. 20d). The separation of these fluid events in time, but not space, is supported in part by the apparent hiatus in the data of Figure 20d, a realistic inference given the substantial amount of data collected from numerous samples. We note, however, that we cannot discriminate between the stage II and III fluids and potential candidates for the parental fluid of appropriate composition that unmixed to give rise to the highly saline fluids of stage I. The terminal stage IV fluids, dominantly of lowsalinity nature, indicate that the magmatic contribution to the hydrothermal system had waned by this time. Implications of pervasive metasomatism for Sn and Cu mineralization The present study of quartz phenocryst-hosted secondary inclusions has demonstrated that the San Rafael stock was penetrated by a wide range of hydrothermal fluids that strikingly mimic the Th-salinity relationships of those that deposited quartz, cassiterite, and chalcopyrite in the major lodes and very probably paralleled the overall evolutionary history of the mineralization. Paradoxically, however, there is no economically significant disseminated or veinlet-controlled Sn or Cu mineralization in either the San Rafael or Quenamari intrusive centers. This is, in our opinion, a problematic aspect of the database. The lack of dispersed cassiterite is particularly striking where the initiation of semipervasive hydrolytic alteration at temperatures below circa 450C paralleled the fluid evolutionary event that generated the enormous stage II orebodies of the San Rafael Lode (Fig. 2). In contrast, widespread development of sericite in the East Kemptville leucogranite, Nova Scotia, was associated with the introduction of several hundred parts per million Sn through much of

1772

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1773

the intrusion (Kontak et al., 1995). The absence of such dispersal at San Rafael could reflect one of two factors: the fluids entering the San Rafael granite were depleted in Sn and Cu, or the dispersed fluids were different in source and/or composition from those related to lode mineralization. We consider the first to be the more probable in view of our fluid inclusion studies. Late fluid ingress to the upper part of the San Rafael granite was overwhelmingly controlled by the major, evolving shear zones, and particularly that which hosts the San Rafael Lode, channeling both the rise of magmatic Sn- and Cu-bearing brines and the incursion of cool, dilute ground waters, thereby providing a constrained environment in which abrupt mixing of the two fluids took place. There are several potential corollaries to this model. First, because the Th-salinity relationships of the resulting mixed hydrothermal fluids involved in stages II and III were similar and well within the range in which chalcopyrite is precipitated, e.g., porphyry Cu deposits (Hezarkhani and WilliamsJones, 1998), the absence of chalcopyrite and other sulfides in stage II assemblages may indicate that neither Cu nor reduced S was present in significant amounts in either the hightemperature brine or the meteoric water at this time. This absence, in turn, predicates a late-stage process in which copper and sulfur became available, perhaps involving the mixing of residual granitic magmas with newly introduced, oxidized, mafic melts, either lamprophyric or basaltic. It is also possible that Sn was largely derived from the mafic melts, particularly the minette (A.H. Clark, unpub. data), as has been proposed for the major South Crofty lode system, Cornwall (Clark et al., 1993; Chen et al., 1996). Second, the barren nature of the stage I veins, stockworks and breccias reflects two, interrelated factors: in this widespread early episode, the magmatic brines rising through the upper part of the San Rafael granite stock did not encounter large volumes of meteoric water; and this absence of meteoric water precluded both the large-scale quenching and salinity reduction that were prerequisites for cassiterite and chalcopyrite deposition. Significance of age relationships Stage I barren veining was broadly contemporaneous with the cooling of the upper kilometer of the San Rafael granite through the Ar-retention temperature of biotite and probably took place less than 1 m.y. after granite intrusion at 24.65 0.2 Ma (Clark et al., 2000). In contrast, the economic lode was emplaced at circa 21.9 to 22.7 Ma; the 40Ar/39Ar plateau dates for stage II hydrothermal muscovites and stage III adularias (low sanidine) were unresolvable within 2 error. By this stage, the upper parts of the San Rafael granite and Quenamari granite stocks had cooled to at least 230C. Clark et al. (2000) propose that such delays in the initiation of ore deposition, a feature also shown by the Cornubian Sn-Cu lodes and the Pasto Bueno W-Cu-Ag veins, northern Peru, may be characteristic of world-class lithophile-metal mineralization, constituting a significant departure from, for example, porphyry Cu deposits in which retrograde boiling is immediately responsible for stockwork development and ore metal deposition. Thus, in Cornubia, it is clear that the sparse deposition of cassiterite and chalcopyrite in early veins associated with retrograde boiling was not the result of excessive fluid temperatures or inappropriate Cl contents, and the hydrothermal
0361-0128/98/000/000-00 $6.00

fluids possessed Th and salinity ranges similar to those which generated the later major Sn-Cu lodes (Chen, 1994; Chen et al., 1996). At San Rafael, cassiterite and chalcopyrite mineralization was confined to brittle shear zones that formed in response to regional tectonism rather than to cooling stresses related to intrusion. Similarly, at Pasto Bueno, wolframite deposition was initiated, some 2 m.y. after retrograde boiling and pervasive development of greisen in the host upper Miocene Consuzo stock, within dilatant structures that developed during the initial uplift and unroofing of the contiguous Cordillera Blanca batholith (Clark et al., 2000; A.H. Clark, unpub. data). It is therefore proposed that intrusion-ore formation hiatuses may play a major role in the concentration of the lithophile and associated base and precious metals. This may be evidence for the accumulation of volumes of metal-enriched dense brines in the interiors of the parental granitoid plutons, where they may coexist with small bodies of specialized magma, such as would be represented by the tourmaline leucogranite at San Rafael. Upward expulsion of the ore metal-rich brines would, in this model, be caused by regional tectonic activity, involving uplift, exhumation and, probably, renewed incursion of granitic and/or mafic melts. Conclusions Despite the absence of dispersed, porphyry-style Sn or Cu mineralization at San Rafael, the semipervasive hydrothermal alteration of the upper 700 m of the host stock records an evolution closely linked to that of the early barren and later economic veins and lodes. On a district scale, as recognized by Kontak and Clark (1988), there is a correlation between the predominant alkali metasomatic facies and intensity of Sn-Cu mineralization, and enrichment in K, as at San Rafael proper, was apparently more favorable than that in Na, as at Quenamari and Santo Domingo. Intrusions, such as the Antauta granite, that preserve magmatic sanidine are not known to be associated with significant lithophile or basemetal mineralization. Precipitation of cassiterite and, subsequently, chalcopyrite at San Rafael was almost entirely confined to the major shear zones, because only therein did the rising metal-rich brines encounter pristine meteoric waters. Development of chlorite in the lode envelopes may have promoted the deposition of, particularly, cassiterite through fluid neutralization, but this process is unlikely to have contributed more than marginally to the extreme ore grades exhibited by the Veta San Rafael. Whereas essentially undiluted high-temperature magmatic fluids permeated much of the apex of the San Rafael stock before, and in the initial stages of, development of the northnorthweststriking fractures, migration of the lower-temperature, mixed fluids was largely channeled along these structures and, we infer, only spent, Sn- and Cu-depleted effluent penetrated the surrounding granite. The most remarkable observation made in the present study is that, although the fluids that generated alkali and hydrogen metasomatism throughout much of the apex of the San Rafael granite did not cause the deposition of disseminated cassiterite or chalcopyrite, their trajectory in Th-salinity space mimicked that of the fluids responsible for the lode mineralization. The microthermometric research documented

1773

1774

KONTAK AND CLARK Bea, F., 1996, Residence of REE, Y, Th and U in granites and crustal protoliths: Implications for the chemistry of crustal melts: Journal of Petrology, v. 37, p. 521552. Berger, G.W., and York, D., 1981, Geothermometry from 40Ar/39Ar dating experiments: Geochimica et Cosmochimica Acta, v. 45, p. 795811. Bodnar, R.J., 1994, Synthetic fluid inclusionsXII. The system H2O-NaCl. Experimental determination of the halite liquidus and isochores for a 40 wt % NaCl solution: Geochimica et Cosmochimica Acta, v. 58, p. 10531063. Bodnar, R.J., and Vityk, M.O., 1994, Interpretation of microthermometric data for H2O-NaCl fluid inclusions, in De Vivo, B., and Frezzotti, M.L., eds., Fluid Inclusions in Minerals: Methods and Applications, Short Course of the Working Group (IMA) Inclusions in Minerals, p. 117130. Bodnar, R.J., Burnham, C.W., and Sterner, S.M., 1995a, Synthetic fluid inclusions in natural quartz. III. Determination of phase equilibrium properties in the system H2O-NaCl to 1000oC and 1500 bars: Geochimica et Cosmochimica Acta, v. 49, p. 18611873. Bodnar, R.J., Reynolds, T.J., and Kuehn, C.A., 1985b, Fluid-inclusion systematics in epithermal systems, in Berger, B.R., and Bethke, P.M., eds., Geology and Geochemistry of Epithermal Systems: Reviews in Economic Geology, v. 2, p. 7398. Brown, W.L., and Parsons, I., 1989, Alkali feldspars: Ordering rates, phase transformations and behaviour diagrams for igneous rocks: Mineralogical Magazine, v. 53, p. 2542. Campbell, A., Banks, D.A., Phillips, R.S., and Yardley, B.W.D., 1995, Geochemistry of the Th-U-REE mineralizing magmatic fluids, Capitan Mountains, New Mexico: ECONOMIC GEOLOGY, v. 90, p. 12711287. Chappell, B.W., and White, A.J.R., 1974, Two contrasting granite types: Pacific Geology, v. 8, p. 173174. Charoy, B., and Pollard, P. J., 1989, Albite-rich, silica-depleted metasomatic rocks at Emuford, northeast Queensland: Mineralogical, geochemical and fluid inclusion constraints on hydrothermal evolution and tin mineralization: ECONOMIC GEOLOGY, v. 84, p. 18501874. Chen, Y., 1994, Geochronological relationships in the Cornubian metallogenetic province, southwest England: unpublished Ph.D. dissertation, Kingston, Ontario, Queens University, 621 p. Chen, Y., Clark, A.H., Farrar, E., Wasteneys, H.A.H.P., and Hodgson, M.J., 1996, Evolution and genesis of the South Crofty tin deposit, Cornwall, southwest England [extended abs.]: International Geological Congress, 30th, Beijing, China, 1996, Abstracts, v. 1, p. 408. Clark, A.H., 1995, Nature or Nurture, in Clark, A.H., ed., Giant Ore DepositsII, Controls on the scale of orogenic magmatichydrothermal mineralization: 2nd Giant Ore Deposits Workshop, Kingston, Ontario, 1995, QMinEx, Queens University, Proceedings, p. viix. Clark, A.H., Kontak, D.J., and Farrar, E., 1983a, Evolution of the central Andean magmatic arcs [abs.]: American Geophysical Union Transactions, v. 64, p. 326. Clark, A.H., Palma, V.V., Archibald, D.A., Farrar, E., Arenas, F., M.J., and Robertson, R.C.R., 1983b, Occurrence and age of tin mineralization in the Cordillera Oriental, southern Peru: ECONOMIC GEOLOGY, v. 78, p. 514520. Clark, A.H., Kontak, D.J., and Farrar, E.,1984, A comparative study of the metallogenetic and geochronological relationships in the northern part of the central Andean tin belt, SE Peru and NW Bolivia, in Janelidze, T.V., and Tvalchralidze, A.G., eds.: IAGOD Symposium, 6th, Tbilisi, USSR, 1982, Proceedings, p. 267279. Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., Arenas, M.J., France, L.J., McBride, S.L., Woodman, P.L., Wasteneys, H.A., Sandeman, H.A., and Archibald, D.A., 1990a, Geologic and geochronologic constraints on the metallogenic evolution of the Andes of southeastern Peru: ECONOMIC GEOLOGY, v. 85, p. 15201583. Clark, A.H., Kontak, D.J., and Farrar, E., 1990b, The San Judas Tadeo W(Mo, Au) deposit: Permian lithophile mineralization in southeastern Peru: ECONOMIC GEOLOGY, v. 85, 16511668. Clark, A.H., Chen, Y., Farrar, E., Wasteneys, H.A.H.P., Stimac, J.A., Hodgson, M.J., Willis-Richards, J., and Bromley, A.V., 1993, The Cornubian SnCu (-As, W) metallogenetic province: Product of a 30 m.y. history of discrete and concomitant anatectic, intrusive and hydrothermal events: Proceedings of the Ussher Society, v. 8, p. 112116. Clark, A.H., Chen, Y., Sandeman, H.A., Farrar, E., Hodgson, M.J., Kontak, D.J., and Arenas, M.J., 1995, Two giants and their lairs: The South Crofty, Cornwall, and San Rafael, Peru, tin (copper) deposits: Proceedings of the Ussher Society, v. 8, p. 457458. Clark, A.H., Chen, Y., Grant, J.W., Kontak, D.J., Wasteneys, H.A., Sandeman, H.A., Farrar, E., and Archibald, D.E., 2000, Delayed inception of ore deposition in major lithophile-metal vein systems: the San Rafael tin and

herein focused on the largest, least problematic, fluid inclusions hosted by quartz phenocrysts in samples from the restricted outcrop (Figs. 1c, 2a) of the San Rafael granite, in none of which cassiterite occurs as either disseminated grains or in veinlets. Nonetheless, our experience suggests that the evolutionary history of the hydrothermal system that generated the world-class San Rafael Lode could probably be established through study of a small number of quartz phenocrysts from many sites in the exposed depth interval of the stock and its surrounding dikes. This observation has unambiguous implications for mineral exploration in this and other comparable provinces. Acknowledgments This study, a contribution to the Queens University Central Andean Metallogenetic Project (QCAMP), is an outgrowth of the senior authors doctoral research. Funding was provided by Natural Sciences and Engineering Research Council of Canada grants to A.H. Clark and Edward Farrar. Field work was supported logistically by Minsur S.A., Lima, through the good offices of Fausto Zavaleta Cruzado and, latterly, Lucio Pareja Chvez. Underground studies were unselfishly assisted by former mine manager, Adolfo Mdico Bao, and by numerous mine geologists, particularly Jlver Alvarez Romero and Pastor Luque Mlaga. QCAMP research in southern Peru would not have been possible without the enthusiastic support and encyclopedic knowledge of Mario Arenas Figueroa, consulting geologist. Geochronologic data were determined by Edward Farrar and Yanshao Chen, at Queens University, and by Hardolph Wasteneys, Royal Ontario Museum, Toronto, in both cases supported by NSERC. Our indebtedness to Vicente Palmas pioneering contributions at the outset of this research is clearly evident from our text. Alan Anderson is thanked for carrying out SEM scans of daughter minerals. Discussions with consulting geologists David Kerr and Robert Mason have been particularly helpful. Bob McKay, Dalhousie University, is thanked for his assistance with the electron microprobe analysis, while Jim Reynolds provided advice on the interpretation of fluid inclusion relationships, determined at the Nova Scotia Department of Natural Resources. Joan Charbonneau expertly typed the manuscript, and Tom Ullrich assisted with the figures. Nigel Grant, Jos Perell, Jean Cline, Roger Taylor, and an anonymous referee, as well as the indefatigable editor, promptly and constructively reviewed the original manuscript. Minsur, S.A., has given permission for publication of this manuscript, which is prepared in honor of Adolfo Mdico, murdered at the San Rafael mine by members of the Sendero Luminoso, December 15th, 1986. July 31, August 21, 2002
REFERENCES
Arenas, M.J., 1980, El distrito minero San Rafael: Estao en el Per: Boletn de la Sociedad Geolgica del Per, no. 66, p. 111. 1999, Exploracin y geologa de la mina San Rafael, Puno: Minera, Instituto de Ingenieros de Minas del Per, Lima, no. 260, May, 1031. Bateman, P.W., 1982, Mineralogy, vein textures and fluid inclusion studies of the Condoriquea and Santo Domingo tin-base metal deposits, SE Peru, with a comparison to the San Rafael tin-copper deposit: Unpublished B.Sc. Honors thesis, Kingston, Ontario, Queens University, 158 p.

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

1774

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU Pasto Bueno tungsten deposits, Peru: 2000, Geological Society of America Abstracts with Programs, v. 32, no. 7, Reno, Nevada, p. A-279. Cline, J.S., and Bodnar, R.J., 1994, Direct evolution of brine from a crystallizing silicic melt at the Questa, New Mexico, molybdenum deposit: ECONOMIC GEOLOGY, v. 89, p. 17801802. Cox, S.F., Knackstedt, M.A., and Braun, J., 2001, Principles of structural control on permeability and fluid flow in hydrothermal systems, in Richards, J.P., and Tosdal, R.M., eds., Structural Controls on Ore Genesis: Reviews in Economic Geology, v. 14, p. 124. Crawford, M.L., 1981, Phase equilibria in aqueous fluid inclusions, in Hollister L.S., and Crawford, M.L., eds., Mineralogical Association of Canada Short Course Notes, p. 75100. Dines, H.G., 1956, The metalliferous mining region of south-west England: Geological Survey of Great Britain Memoir, 2 vols., 795 p. Dong, G.Y., and Zhou, T., 1996, Zoning in the Carboniferous-Lower Permian Cracow epithermal vein system, central Queensland, Australia: Mineralium Deposita, v. 31, p. 210224. Eugster, H.P., and Wilson, G.A., 1985, Transport and deposition of ore-forming elements in hydrothermal systems associated with granites, in Halls, C., et al., eds., High Heat Production (HHP) Granites, Hydrothermal Circulation and Ore Genesis, St. Austell, England: Institution of Mining and Metallurgy, London, p. 8798. Farrar, E., Yamamura, B.K., Clark, A.H., and Taipe, J., 1990, 40Ar/39Ar ages of magmatism and tungsten- polymetallic mineralization, Palca 11, Choquene district, southeastern Peru: ECONOMIC GEOLOGY, v. 85, p. 16691676. Ferry, J.M., 1979, Reaction mechanisms, physical conditions, and mass transfer during hydrothermal alteration of mica and feldspar in granitic rocks from south-central Maine, USA: Contributions to Mineralogy and Petrology, v. 68, p. 125139. Fryer, B.J., 1977, Rare earth evidence in iron formations for changing Precambrian oxidation states: Geochimica et Cosmochimica Acta, v. 41, p. 361367. Goldstein, R.H., and Reynolds, T.J., 1994, Systematics of fluid inclusions in diagenetic minerals: Society for Sedimentary Geology Short Course 31, 199 p. Gromet, L.P., and Silver, L.T., 1983, Rare earth element distribution among minerals in a granodiorite and their petrogenetic implications: Geochimica et Cosmochimica Acta, v. 47, p. 925940. Halter, W., Williams-Jones, A., and Kontak, D.J., 1996, The role of greisenization in cassiterite precipitation at the East Kemptville tin deposit, Nova Scotia: ECONOMIC GEOLOGY, v. 91, p. 368385. Hannington, M., 1983, Mineralogy and paragenesis of Veta Nazareth, SE Peru: unpublished B.Sc. Honors thesis, Kingston, Ontario, Queens University, 87 p. Harrison, T.M., Duncan, I., and McDougall, I., 1985, Diffusion of 40Ar in biotite: Temperature, pressure and compositional effects: Geochimica et Cosmochimica Acta, v. 49, p. 24612468. Haynes, F.M., Sterner, S.M., and Bodnar, R.J., 1988, Synthetic fluid inclusions in natural quartzIV. Chemical analysis of fluid inclusions by SEM/EDA: Evaluation of method: Geochimica et Cosmochimica Acta, v. 52, p. 969977. Heinrich, C.A., 1990, The chemistry of hydrothermal tin(-tungsten) ore deposition: ECONOMIC GEOLOGY, v. 85, p. 457481. Hey, M.H., 1954, A new review of the chlorites: Mineralogical Magazine, v. 30, p. 277292. Hezarkhani, A., and Williams-Jones, A.E., 1998, Controls of alteration and mineralization in the Sungun porphyry copper deposit, Iran: Evidence from fluid inclusions and stable isotopes. ECONOMIC GEOLOGY, v. 93, p. 651670. Ichenhower, J., and London, D., 1996, Experimental partitioning of Rb, Cs, Sr, and Ba between alkali feldspar and peraluminous melt: American Mineralogist, v. 81, p. 719734. Jenner, G.A., Longerich, H., Jackson, S., and Fryer, B.J., 1990, ICP-MSA powerful tool for high-precision trace-element analysis in earth sciences: Evidence from analysis of selected U.S. Geological Survey reference samples: Chemical Geology, v. 83, p.133148. Kontak, D.J., 1985, The magmatic and metallogenetic evolution of a cratonorogen interface: The Cordillera de Carabaya, central Andes, SE Peru: Unpublished Ph.D. dissertation, Kingston, Ontario, Queens University, 714 p. Kontak, D.J., 1998, A study of fluid inclusions in sulphide and non-sulphide mineral phases from a carbonate-hosted Zn-Pb deposit, Gays River, Nova Scotia. ECONOMIC GEOLOGY, v. 93, p. 793817. Kontak, D.J., and Clark, A.H, 1988, Exploration criteria for tin and tungsten mineralization in the Cordillera Oriental of southeastern Peru, in Taylor, R.P., and Strong, D.F., eds., Recent Advances in the Geology of Granite-Related

1775

Mineral Deposits: Canadian Institute of Mining and Metallurgy Special Volume 39, p. 157169. Kontak, D.J., and Clark, A.H., 1997a, The roles of magmatic and meteoric fluids in the formation of the giant, high-grade, San Rafael Sn deposit, SE Peru: Geological Association of Canada-Mineralogical Association of Canada, Abstracts with Program, v. 20, p. A80. 1997b, The Minastira perlauminous granite, Puno, southeastern Peru: A quenched, hypabyssal intrusion recording magma commingling and mixing: Mineralogical Magazine, v. 61, p. 743764. Kontak, D.J., and Martin, R.F., 1997, Alkali feldspar in the peraluminous South Mountain batholith, Nova Scotia: Trace-element data: Canadian Mineralogist, v. 35, p. 959977. Kontak, D.J., Clark, A.H., and Farrar, E., 1984, The influence of fluid and rock composition and tectono-thermal processes on Al-Si distribution in alkali feldspars in granitoid rocks: Bulletin de Minralogie, v. 107, p. 387400. Kontak, D.J., Clark, A.H., Farrar, E., Pearce, T.H., Strong, D.F., and Baadsgaard, H., 1986, Petrogenesis of a Neogene shoshonite suite, Cerro Moromoroni, Puno, southeastern Peru: Canadian Mineralogist, v. 24, p. 117135. Kontak, D.J., Clark, A.H., Farrar, E., Archibald, D.A, and Baadsgaard, H., 1987, Geochronological data for Tertiary granites of the southeast Peru segment of the central Andean tin belt: ECONOMIC GEOLOGY, v. 82, p. 16111618. Kontak, D.J., Farrar, E., Clark, A.H., and Archibald, D.A., 1990, Eocene tectono-thermal rejuvenation of an upper Paleozoic-lower Mesozoic terrane in the Cordillera de Carabaya, Puno, southeastern Peru, revealed by K-Ar and 40Ar/39Ar dating: Journal of South American Earth Sciences, v. 3, p. 231246. Kontak, D.J., Clark, A.H., Halter, W., and Williams-Jones, A., 1995, Metal concentration versus dispersal in the magmatic-hydrothermal environment: A case study contrasting low-grade (East Kemptville, Nova Scotia) and high-grade (San Rafael, Peru) Sn-base metal deposits, in Clark, A.H., ed., Giant Ore DepositsII. Controls on the Scale of Orogenic MagmaticHydrothermal Mineralization: Giant Ore Deposits Workshop, 2nd, Kingston, Ontario, 1995, QMinEx, Queens University, Proceedings, p. 347413. Lagache, M., 1984, The exchange equilibrium distribution of alkali and alkaline-earth elements between feldspars and hydrothermal solutions, in Brown, W.L., ed., Feldspars and Feldspathoids: Structures, Properties and Occurrences: NATO ASI Series, Boston, D. Reidel Publishing, p. 247280 Lagache, M., and Weisbrod, A., 1977, The system, two alkali feldspars-KClNaCl-H2O at moderate to high temperatures and low pressures: Contributions to Mineralogy and Petrology, v. 62, p. 77101. Laubacher, G., 1978, Gologie de la Cordillre Orientale et de lAltiplano au nord et au nord-ouest du Lac Titicaca (Prou): ORSTOM, Traveaux et Documents, v. 95, 217 p. Leeman,W.P., and Phelps, D.W., 1981, Partitioning of rare earths and other trace elements between sanidine and coexisting volcanic glass: Journal of Geophysical Research, v. 86, p. 10,19310,199. Lentz, D.R., ed., 1994, Alteration and alteration processes associated with ore-forming systems: Geological Association of CanadaMineralogical Association of Canada Short Course Notes, v. 11, 467 p. Martin, R.F., and Bowden, P., 1981, Peraluminous granites produced by rock-fluid interaction in the Ririwai nonorogenic ring-complex, Nigeria: Mineralogical evidence: Canadian Mineralogist, v. 19, p. 6582. Moody, J.B., Jenkins, J.E., and Meyer, D., 1985, An experimental investigation of the albitization of plagioclase: Canadian Mineralogist, v. 23, p. 583598. Muecke, G.K., and Clarke, D.B., 1981, Geochemical evolution of the South Mountain batholith, Nova Scotia: Rare- earth element evidence: Canadian Mineralogist, 19, p. 133162. Norton, D.L., 1982, Fluid and heat transport phenomena typical of copperbearing pluton environments, in Titley, S.R., ed., Advances in Geology of Porphyry Copper Deposits of Southwestern North America: Tuscon, University of Arizona Press, p. 5972. Oakes, C.S., Bodnar, R.J., and Simonson, J.M., 1990, The system NaClCaCl2-H2O. I. The ice liquidus at 1 atm total pressure: Geochimica et Cosmochimica Acta, v. 54, p. 603610. Orville, P.M., 1963, Alkali ion exchange between vapor and feldspar phases: American Journal of Science, v. 261, p. 201237. Palma, V.V., 1981, The San Rafael tin-copper deposit, SE Peru: Unpublished M.Sc. thesis, Kingston, Ontario, Queens University, 235 p. Palma, V.V., and Clark, A.H., 1982, The San Rafael tin-copper lode system, Puno, SE Peru: A Cornwall-type deposit in the central Andean tin belt [abs.]: Geological Association of CanadaMineralogical Association of Canada Program with Abstracts, v. 6, p. 71.

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

1775

1776

KONTAK AND CLARK Sterner, S.M. 1992: Homogenization of fluid inclusions to the vapor phase: The apparent homogenization phenomenon: ECONOMIC GEOLOGY, v. 87, 16161623. Stewart, D.B., and Wright, T.L., 1974, Al/Si order and symmetry of natural alkali feldspars, and the relationship of strained cell parameters to bulk composition: Bulletin de la Socit Franaise de Minralogie et Cristallographie, v. 97, p. 356377. Stimac, J.A., Clark, A.H., Chen, Y., and Garcia, S., 1995, Enclaves and their bearing on the origin of the Cornubian batholith, southwest England: Mineralogical Magazine, v. 59, p. 277296. Sverjensky, D.A. 1984. Europium redox equilibria in aqueous solution: Earth and Planetary Science Letters, 67, p. 7078. Sylvester, P.J., 1998, Post-collisional strongly peraluminous granites: Lithos, v. 45, p. 2944. Tischendorf, G., 1977, Geochemical and petrographic characteristics of silicic magmatic rocks associated with rare-element mineralization, in Burnol, L., and Tischendorf, G., eds., Metallization Associated with Acid Magmatism: Geological Survey of Czechoslovakia, v. 2, p. 4198. Tischendorf, G., Geisler, M., Gerstenberger, H., Budzinski, H., and Vogler, P., 1987, Geochemistry of Variscan granites of the Westerzgebirge-Vogtland regionAn example of tin deposit-generating granites: Chemie der Erde, v. 46, p. 213235. Titley, S.R., 1994, Evolutionary habits of hydrothermal and supergene alteration in intrusion-centred ore systems, southwestern North America, in Lentz, D.R., ed., Alteration and Alteration Processes Associated with OreForming systems: Geological Association of Canada Short Course Notes 11, p. 237260. Turneaure, F.S., 1971, The Bolivian tin-silver province: ECONOMIC GEOLOGY, v. 66, p. 215225. Tuttle, O.F., and Bowen, N.L., 1958, Origin of granite in light of experimental studies in the system NaAlSi3O8- KAlSi3O8-SiO2-H2O: Geological Society of America Memoir 74, 153 p. Webster, J.D., 1995, Evolution of Cl-bearing fluids from chlorine-enriched mineralizing granitic magmas and implications for metal transport, in Clark, A.H., ed., Giant Ore Deposits II, QMinEx, Kingston, Ontario, Department of Geological Sciences, Queens University, p. 221261. White, A.J.R., and Chappell, B.W., 1983, Granitoid types and their distribution in the Lachlan fold belt, southeastern Australia, in Roddick, J.A., ed., Circum-Pacific Plutonic Terranes: Geological Society of America Memoir 159, p. 2134. Wilson, A.D., 1960, The microdetermination of ferrous iron in silicate minerals by a volumetric and colourimetric method: Analyst, v. 85, p. 823827. Witt, W. K., 1988, Evolution of high-temperature hydrothermal fluids associated with greisenization and feldspathic alteration of a tin-mineralized granite, northeast Queensland: ECONOMIC GEOLOGY, v. 83, p. 310334. Wright, T.L., 1968, X-ray and optical study of alkali feldsparII. An X-ray method of determining the composition and structural state from measurement of 2 values for three reflections: American Mineralogist, v. 53, p. 88104. Yamamura, B.K., 1991, The Palca 11 tungsten deposit and associated granitic rocks, Choquene district, Puno, SE Peru: Unpublished M.Sc. thesis, Kingston, Ontario, Queens University, 262 p. Zweng, P.L., and Clark, A.H., 1995, Hypogene evolution of the Toquepala porphyry copper-molybdenum deposit, Moquegua, southeastern Peru, in F.W. Pierce and J.B. Bolm, eds., Porphyry Copper Deposits of the American Cordillera: Arizona Geological Society Digest 20, p. 566612.

Patino-Douce, A.E., 1999, What do experiments tell us about the relative contributions of crust and mantle to the origin of granitic magmas?, in Castro, A., Fernndez, C., and Vigneresse, J.L., eds., Understanding granites: Integrating new and classical techniques: Geological Society of London Special Publication 168, p. 5575. 1983, (Na, K) exchange between alkali feldspars and aqueous solutions containing borate and fluoride anions, experimental results at P = 1 kbar: NATO Advanced Study Institute on Feldspars, 3rd, Rennes, p. 102. Pichavant, M., Boher, M., Stenger, J.-F., Aissa, M., and Charoy, B., 1987, Relations de phase des granites de Beauvoir 1 et 3 kbar en conditions de saturation en H2O: Gologie de la France, no. 23, p. 7786. Pollard, P.J., 1983, Magmatic and post-magmatic processes in the formation of rocks associated with rare-element deposits: Institution of Mining and Metallurgy Transactions, v. 92, Section B, p. B1B9. Pollard, P.J., and Taylor, R.G., 1986, Progressive evolution of alteration and tin mineralization: Controls by interstitial permeability and fracture-related tapping of magmatic fluid reservoirs in tin granites: ECONOMIC GEOLOGY, v. 81, p. 17951800. Pollard, P.J., Taylor, R.G., and Tate, N.M., 1989, Textural evidence for quartz and feldspar dissolution as a mechanism of formation of Maggs pipe, Zaaiplaats tin mine, South Africa: Mineralium Deposita, v. 24, p. 210218. Pollard, P.J., Andrew, A.S., and Taylor, R.G., 1991, Fluid inclusion and stable isotope evidence for interaction between granites and magmatic hydrothermal fluids during formation of disseminated and pipe-style mineralization at the Zaaiplaats tin mine: ECONOMIC GEOLOGY, v. 86, p. 121141. Powell, T., Salmon, S., Clark, A.H., and Shail, R.K., 1999, Emplacement styles within the Lands End granite, west Cornwall: Geoscience in SouthWest England, v. 9, p. 333339. Ratajeski, K., and Campbell, A.R., 1994, Distribution of fluid inclusions in igneous quartz of the Capitan pluton, New Mexico, USA: Geochimica et Cosmochimica Acta, v. 58, p. 11611174. Roedder, E., 1984, Fluid Inclusions: Reviews in Mineralogy, v. 12, 644 p. Samson, I., and Sinclair, W.D., 1992, Magmatic hydrothermal fluids and the origin of quartz-tourmaline orbicules in the Seagull batholith, Yukon Territory: Canadian Mineralogist, v. 30, p. 937954. Sandeman, H.A.I., 1995, Lithostratigraphy, petrology and geochronology of the Crucero Supergroup, Puno, SE Peru: Implications for the Cenozoic geodynamic evolution of the southern Peruvian Andes: Unpublished Ph.D. dissertation, Kingston, Ontario, Queens University, 382 p. Sandeman, H.A., and Clark, A.H., 1993, Mingling and mixing of minette and S-type, anatectic rhyodacitic magmas, Crucero Supergroup, Puno, SE Peru: Geological Association of CanadaMineralogical Association of Canada Program with Abstracts, v. 18, p. A-92. Sandeman, H.A., Clark, A.H., and Farrar, E., 1995, An integrated tectonomagmatic model for the evolution of the southern Peruvian Andes (1320S) since 55 Ma: International Geology Review, v. 37, p. 10391073. Sandeman, H.A., Clark, A.H., Farrar, E., and Arroyo, G., 1997, Lithostratigraphy, petrology and 40Ar/39Ar geochronology of the Crucero Supergroup, Puno Department, SE Peru: Journal of South American Earth Sciences, v. 10, p. 223245. Shaw, D.M., 1968, A review of K/Rb fractionation trends by covariance analysis: Geochimica et Cosmochimica Acta, v. 32, p. 573601. Sillitoe, R.H., Halls, C., and Grant, J.N., 1975, Porphyry tin deposits in Bolivia: ECONOMIC GEOLOGY, v. 70, p. 913927. Smirnov, S.S., 1937, Some outlines concerning the sulphide-cassiterite deposits: Academy of Sciences, USSR, B, Series Geology, v. 5, p. 853862 [in Russian]. Smith, J.V., 1974, Feldspar Minerals. 1. Crystal Structure and Physical Properties: New York, Springer-Verlag.

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

1776

ALTERATION & HIGH-GRADE Sn MINERALIZATION, SAN RAFAEL, PERU

1777

APPENDIX Analytical Procedures Whole-rock samples were analyzed at Queens University for major and minor (Si, Ti, Al, Fe, Mg, Ca, Na, K, Mn, P) and trace (Rb, Sr, Ba, Nb, V, Cr, Zr, Y, Sn) element content using a Siemens Model VRS X-ray vacuum spectrometer. Additional analyses for trace elements (Cu, Pb, Zn, Ni) were carried out on an Instrumentation Laboratory Model 251 atomic absorption spectrophotometer, and ferrous iron determinations were obtained using a modified titrametric method of Wilson (1960). Full details of sample preparation and analyses, along with standards, duplicates, precision, and accuracy are given in Kontak (1985). Rare-earth element abundances were determined at Memorial University using the thin-film X-ray fluorescence method of Fryer (1977). Analysis of silicate minerals was carried out on polished thin sections with a JEOL 733 Superprobe at Dalhousie University (Halifax, Nova Scotia) in the energy dispersive mode, and using the following operating conditions: 1 to 3 m beam diameter but defocused to 10 m for feldspars; 10 nA beam current; and 15 kV accelerating voltage. F in biotite was determined using wavelength-dispersive spectrometry on the same instrument. Back-scattered and secondary electron images were collected for examination of textural and mineralogic details of the samples, as discussed in the text. Major and trace element compositions of high quality mineral separates (biotite, alkali feldspars) were obtained using the same X-ray fluoresence analytical procedures as for the whole rocks. Trace elements (Li, Mo, Cs, Hf, Ta, Bi, Th, U, rare earth elements) were measured for alkali feldspar separates by solution chemistry inductively coupled plasma-mass spectrometer (Memorial University, Newfoundland) using the method of Jenner et al. (1990). Fluid inclusion thermometric measurements were obtained using a U.S. Geological Surveytype gas-flow heating/cooling stage housed at the Nova Scotia Department Natural Resources, Halifax (see Kontak [1998] for full details of operating conditions). Precipitate or decrepitate mounds generated from artificial thermal decrepitation of fluid inclusions (Haynes et al., 1988) were also analyzed in the energy dispersive spectrometer mode using the JEOL 733 Superprobe with full imaging facilities, using the operating conditions noted above. Alkali feldspars were analyzed for their Al/Si ordering/disorder relationships using a Picker X-ray diffractometer with Ni-filtered Cu radiation ( = 1.5418 A; 40 kV; 18 mA). A single scan was made for each sample from a 2 angle of 13o to 70o at a goniometer speed of 1o per minute. A silicon internal standard was used to check for and correct any peak shift. The order/disorder relationships and bulk compositions ofthe phases were determined with reference to the (060)-(204) diagram of (Wright, 1968; Stewart and Wright, 1974). Further details of the methodology employed, including estimates of precision and accuracy, are given in Kontak (1985).

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

1777

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

1778

Vous aimerez peut-être aussi