Sei sulla pagina 1di 28

TYPES OF SULFIDE-RICH EPITHERMAL DEPOSITS AND THEIR AFFILIATION TO PORPHYRY SYSTEMS: LEPANTOVICTORIAFAR SOUTHEAST DEPOSITS, PHILIPPINES, AS EXAMPLES Jeffrey

W. Hedenquist1, Rene Juna R. Claveria and Gener P. Villafuerte2 1 Consulting Economic Geologist Canada 2 BA - Lepanto Consolidated Mining Company SUMMARY The Lepanto and Victoria epithermal deposits contain in excess of 8 Moz Au, associated with massive enargite-cemented breccia and quartz-carbonate-base metal veins, respectively. The two deposits are separated by <500 m, but the latter was not discovered until the former was nearly mined out. The sulfide assemblages of these adjacent deposits are characterized by high- and intermediate-sulfidation states, respectively. However, gold was introduced after enargite at Lepanto, associated with an assemblage that is similar to that of the Victoria ore. An even larger gold resource is present in similar to that of the Victoria ore. An even larger gold resource is present in the Far Southeast porphyry Cu-Au deposit, the top of which lies 200-400 m beneath and adjacent to these two epithermal deposits. The timing of ore deposition at FarSoutheast and Lepanto was about 1.4 to 1.3 Ma, whereas ore formed slightly later at Victoria, at 1.15 Ma. Although a genetic relationship has yet to be proven for these three deposits, their spatial and temporal affiliation highlights the potential for similar associations in other porphyry and/or epithermal districts. INTRODUCTION A variety of epithermal deposit types have long been recognized (Lindgren, 1933), and their possible affiliation with porphyry systems has been suggested (Sillitoe, 1975) and debated (Sillitoe, 1989). There is broad recognition and acceptance of two end-member types of epithermal system (Heald et al., 1987; White and Hedenquist, 1990; Sillitoe, 1993), although variations in style are noted (Bonham, 1986; Sillitoe, 1993). There has been much debate on whether or not these types are related genetically to each other, and to the underlying

magmatic engine, which in some cases has formed a porphyry deposit. We review the types of epithermal deposit from the basis of genetic affiliation, and note those with evidence for a spatial, and in some cases genetic relationship with porphyry deposits. In order to highlight these relationships, we discuss a case study of one district in Luzon, Philippines, that hosts two types of epithermal deposit located supradjacent to a large porphyry Cu-Au deposit. TYPES OF EPITHERMAL DEPOSIT The formation of one end-member type involves an early acid fluid that leaches the rock prior to deposition of high-sulfidation state Cu sulfides and gold by a less acid fluid. The other type of deposit forms from a near-neutral pH solution that deposits precious base metal ore in quartz veins, accompanied by alteration halos. These epithermal deposits have distinctly different

characteristics (Hedenquist et al., 2000), and are called high sulfidation and low sulfidation, respectively, to refer to the sulfidation state of their characteristic sulfide assemblages, an intrinsic feature of the deposit.

Characteristics Within the low-sulfidation (LS) type of deposit, widely different Ag:Au ratios have been noted, with some of the Ag-rich deposits having a high base-metal content. Although base-metal contents tend to increase with increasing depth in some deposits (Buchanan, 1981), the precious- and base-metal signature appears to be a fundamental characteristic of the deposit style (White et al., 1995; Simmons, 1995).

The Au-rich LS deposits are typically hosted by rhyolitic-dacitic rocks that have a bimodal character and formed in an extensional setting (Sillitoe, 1993; John et al., 1999). The deposits show crustiform vein textures in which chalcedony dominates, suggesting a relatively low-temperature and shallow depth of formation (Saunders, 1994). Adularia is a common gangue mineral, and illite typically forms an alteration halo. There is a broadly similar style that has a relation to alkalic volcanic host rocks (Bonham, 1986; Sillitoe, 1993). Sulfide

minerals are minor, and selenides (or tellurides, in the alkalic case) are common. The sulfides that occur record a low sulfidation state, including pyrite, pyrrhotite, arsenopyrite, and high-Fe sphalerite (Figure1; John et al., 1999). Type examples are listed in Table 1, and are best known from Nevada and California, USA, and Kyushu, Japan.

By contrast, Ag-rich deposits, where Ag:Au typically has a minimum value of 10 or 20:1, up to 100s:1, are hosted by andesite-rhyodacite volcanic rocks in arc settings, akin to the setting of most porphyry Cu deposits. Quartz veins are crystalline, massive and comb textured, and sericite is a common alteration mineral and gangue, whereas adularia is relatively uncommon. Sulfides are abundant, particularly in the base metal-rich variety. In addition to pyrite, Fepoor sphalerite, galena, chalcopyrite, and tennantite-tetrahedrite occur, along with Mn carbonates. This sulfide assemblage indicates an intermediate sulfidation state (Figure 1). Examples with type characteristics occur in Nevada and Colorado, USA, throughout Mexico and Per, and in many parts of the Philippines and Japan as well as in New Zealand and Romania (Table 1).

The high-sulfidation (HS) epithermal deposits, like the Ag base metal-rich style of the low-sulfidation type, also are hosted by andesite-rhyodacite volcanic rocks in an arc setting, and in many cases show a close spatial association with porphyry deposits (Sillitoe, 1999). The deposits form after leaching of host rocks by an acid fluid, the latter condensed from a magmatic vapor. The silicic leached core, commonly vuggy quartz, has a halo of quartz-alunite, and typically is underlain by roots of sericite or pyrophyllite. The early sulfide minerals are typically euhedral pyrite with enargite-luzonite, a high-sulfidation state assemblage. Where paragenetic studies have been conducted, gold is introduced after enargite deposition, along with fine pyrite, Fe-poor sphalerite, chalcopyrite, galena, and tennantite-tetrahedrite (Jannas et al., 1990; Hedenquist et al., 1998), essentially the same assemblage of the Ag-Au base metal-rich intermediate sulfidation style deposit.

Styles and Terminology Continued study on epithermal deposits is helping to identify fundamental distinctions in characteristics. This now allows classification that reflects basic variables that are related to formation, and which may be relevant to consider during exploration.

John et al. (1999) and John (2001) have examined the various types of epithermal deposit in Nevada. They note that in addition to the HS type of deposit, there are distinctly different characteristics of two styles of LS deposit. These distinctions follow those outlined above. In one style, typified by bonanza Au deposits in the Northern Nevada Rift, including Sleeper, Midas and Ivanhoe, the Au-rich, sulfide-poor veins have a very low sulfidation state assemblage (Figure 1). By contrast, within the Western Andesite Arc, Ag-rich deposits such as Comstock Lode and Tonopah have an intermediate sulfidation state assemblage. We refer to these two styles as end-member LS and intermediate sulfidation (IS), respectively, stressing their distinct tectonic settings, magmatic affiliations, metal complements, sulfide and gangue mineral assemblages, and in some cases, deposit form (Sillitoe, 1993).

We retain the HS terminology for deposits hosted by silicic and quartz-alunite altered bodies, with enargite-pyrite being the dominant sulfide assemblage, but we note that the precious metals in HS deposits are associated with a paragenetically late assemblage that has an intermediate sulfidation state, similar to that of IS deposits.

CAUSES OF VARIATION AMONG EPITHERMAL DEPOSITS Fluid Composition As noted above (Table 1), HS deposits form after the host rock has been leached by a very acid fluid. The salinity of the mineralizing fluid is known from only a few studies of fluid inclusions hosted by enargite, and ranges from 4 to 15 wt% NaCl equiv. However, gangue quartz contains inclusions that have salinities ranging from 1 to 40 wt% NaCl, and HS deposits are commonly

underlain by veins that contain hypersaline inclusion fluids (Arribas et al., 1995a; Hedenquist et al., 1995). This hypersaline fluid is related to an underlying porphyry system, formed by vapor separation (White, 1991; Arribas, 1995). By contrast, the intermediate salinity solution responsible for depositing the enargite may be an unseparated fluid of bulk magmatic composition, similar in composition and origin to the fluid that forms the sericitic overprint of porphyry deposits (Hedenquist et al., 1998).

The end-member LS deposits that are Au-rich typically have salinities <1 wt% NaCl, although the equivalent salinity is commonly reported to be higher, up to several wt%. The low freezing-point depressions that lead to the high equivalent salinity estimates are due to dissolved gases, largely CO2 but also including H2S (Hedenquist and Henley, 1985). Fluids of such composition account for the Au-rich but Ag and base metal-poor content of LS deposits, as gold is transported as a bisulfide complex whereas most of the silver and base metals are transported as chloride complexes (Henley, 1990).

The higher the salinity, the higher the concentration of dissolved Ag and base metals in solution (Henley, 1990). This accounts for the change from Au-Ag to Ag-Au to Ag-(Au)-base metal character of epithermal deposits, and the close match with increasing salinity (Hedenquist and Henley, 1985). This relationship is clearly seen in Mexican epithermal deposits, where the salinities are <1 wt% NaCl equiv for Au-Ag deposits, to 3-5 wt% NaCl for Ag-Au deposits, to 10-20+ wt% NaCl for Ag and base-metal rich deposits (Albinson et al., 2001).

If fluid composition, both salinity and gas content, is so critical in determining the metal complement of a deposit, just as fluid pH is critical in determining the alteration assemblages and zoning (Sillitoe, 1993), what controls this fundamental variable?

Tectonic and Magmatic Setting Low salinity (<1 wt% NaCl) but gassy fluids are typical of back-arc settings with bimodal volcanism (Hedenquist and Henley, 1985), consistent with the nature of the LS epithermal deposits formed in this setting. In these settings, large magma chambers are thought to lie relatively deep, possibly explaining the rarity of porphyry Cu deposits in such settings.

By contrast, volcanic arcs commonly host active geothermal systems with salinities that are an order of magnitude greater, up to about 2-3 wt% NaCl (Hedenquist and Henley, 1985). Geothermal systems of higher salinity are not well known, except for the demonstrably amagmatic systems that owe their high salinities, >20 wt% NaCl, to dissolution of evaporites (McKibben and Hardie, 1997). The high salinities that are recorded by fluid inclusions of Ag-rich IS deposits may reflect episodic incursions into low-salinity systems, based on evidence found at Fresnillo (Simmons, 1991). Thus, Ag base metal-rich brines many underlie many geothermal systems in arc settings, with tectonic or magmatic triggers causing periodic ascent and mineralization. This would explain also why such brines are not observed in active systems, as they are present for only a small portion of the life of the system due to density constraints. Porphyry deposits typically form within the roots of volcanic arcs, and thus there should be a supply of moderately saline fluid at sub-volcanic depths in these arcs.

Further evidence on the tectonic and magmatic relationship between fluid composition and epithermal style comes from igneous and volcanic rocks associated with epithermal deposits. John (2001) has shown that the oxidation state of the bimodal magma suites, determined from Fe-Ti oxide compositions, are 3-4 orders of magnitude more reduced in extensional settings associated with LS deposits, compared with the andesite arc magmas that host IS and HS deposits. This difference is similar to the oxidation state distinction based on sulfide mineral assemblages in the associated LS versus IS and HS deposits (Figure1). This may be the result of hydrothermal fluid simply equilibrating with

the volcanic and intrusive host rocks, or it may reflect a more intrinsic magmatic affiliation. If the latter is shown to be true, then reduced magmas may contribute reduced fluid to the hydrothermal systems that form LS deposits. By contrast, the more oxidized magmas that form during subduction processes close to the SO2-H2S gas buffer that is characteristic of porphyry Cu deposits (Burnham and Ohmoto, 1980) may control the higher oxidation states of fluids that form IS deposits. Generation of the very oxidized fluid that forms high sulfidation state sulfides in HS deposits may be a function mainly of the host rock a leached, silicic material having no buffer capacity to prevent a cooling fluid from evolving to enargite stability on cooling (Barton and Skinner, 1979). RELATIONSHIPS AMONG EPITHERMAL DEPOSIT TYPES: THE PORPHYRY CONNECTION Following on from the previous discussion, we have compiled a list of several HS and IS deposits (Table 2), noting the presence of related epithermal veins in a district, and the evidence for porphyry deposits. A similar but more thorough compilation (Sillitoe, 1999) was used to examine the HS-porphyry relationship. This followed an earlier argument (Sillitoe, 1983) that a spatial association reflected a genetic affiliation.

Based on the arguments of John et al. (1999) and John (2001), we conclude that end-member LS deposits form in a distinctly different tectonic environment and magmatic affiliation from IS and HS deposits. In addition, LS epithermal deposits are not known to occur in porphyry districts, perhaps because they are related to a more deeply seated magma chamber. For these reasons, we now focus only on the relationship between IS and HS epithermal deposits, and the evidence for a transition from these deposits to the porphyry environment, using a case study from the Philippines.

LEPANTO VICTORIA FAR SOUTHEAST, PHILIPPINES Introduction There are many examples of spatially associated epithermal and porphyry ore deposits (Table 2). Nowhere is this spatial association better seen than in the Mankayan mineral district in northern Luzon, where the Lepanto HS Cu-Au and Victoria IS Au-Ag-base metal deposit overlie the Far Southeast (FSE) porphyry Cu-Au ore body to the northwest and south, respectively (Figure 2). This is one of the richest mining districts in the Philippine archipelago in terms of economic value and abundance and diversity of hydrothermal ore deposits. Within an area of <25 km2, the district contains several porphyry Cu-Au and epithermal precious- and base-metal deposits, both HS and IS types (Sillitoe and Angeles, 1985; Hedenquist et al., 1998; Cuizon et al., 1998).

Lepanto was discovered in pre-Spanish time, as the massive enargite ore outcropped at the northwest end of the deposit (Figure 2). Large-scale mining commenced in 1936 and ceased in 1996. The FSE porphyry deposit was discovered in the 1980s during a drilling program that was investigating, among other indications, fragments of altered and mineralized porphyry within young volcanic outcrops. The Victoria deposit was discovered during the course of exploration to extend the Lepanto reserves. Drilling near the Nayak workings intersected a vein, and in 1995 horizontal drilling from the 900 m L of Lepanto cut eight veins with grades of 1.3 to 193 g/t Au. roduction began in 1997. The reserve of the IS veins may eventually compete in size with the total production and reserves of HS ore from Lepanto (Table 2).

Geology and Age of Deposits There are four main lithologic units in the district, three of which are intersected by a longitudinal section through the Lepanto-FSE-Victoria deposits (Figure 3). 1) The volcanic to epiclastic basement consists of several Late Cretaceous to middle Miocene sequences, and includes the Lepanto metavolcanic rocks and Balili volcaniclastic rocks. 2) A large Miocene tonalitic intrusion forms the western margin of the mineral district (Figure 1). 3) A Pliocene dacitic to

andesitic breccia and porphyry unit, the Imbanguila hornblende dacite, predates Cu-Au mineralization. (4) Unaltered Pleistocene dacite to andesite porphyritic lava domes and pyroclastic flow units, called the Bato hornblende-biotite dacite, are post-mineralization in age, although intrusion and volcanism was nearly continuous (Figure 4). A large part of the Lepanto ore body and most of the FSE ore body are hosted by basement metavolcanic or volcaniclastic rocks, whereas most of the presently known Victoria veins are hosted by the Imbanguila units (figures 5 and 6).

The ore deposits in the Mankayan district are spatially and temporally related to the Pliocene to Pleistocene event of intermediate-composition volcanism (Figure 2). The geologic and temporal relations at Lepanto-FSE indicate that mineralization occurred in the middle of this event, associated with quartz diorite porphyry dikes and bodies which are intersected in drill holes at depths of 800 to >1200 m below the present-day surface (Figure 3).

K-silicate alteration related to the FSE porphyry is centered on dikes (Concepcin and Cinco, 1989; Hedenquist et al., 1998), and formed at 1.41 0.05 Ma (n=6 K-Ar biotite dates; Figure 4). Leaching and quartz-alunite alteration occurred at 1.42 0.08 Ma (alunite, n=5). This alteration style formed over the top of the porphyry and extended NW >4 km along the basementpyroclastic rock contact, synchronous with potassic alteration. Quartz-illitesulfide veins with illite-chlorite halos that cut the porphyry system followed at 1.30 0.07 Ma (n=10). New K-Ar dating results (Sajona et al., 2000; LCMC, unpublished ages) indicates that illite-altered wall rock, dated at 1.50 0.14 Ma and indistinguishable from the FSE illite ages, was cut by the Victoria veins at 1.15 0.03 Ma (n=2).

The dating results suggest that the Victoria veins are younger than the principal sericite (illite) overprint at FSE, perhaps by as much as 50,000-100,000 years. Although the Lepanto enargite was deposited after much of the alunite, based on cross-cutting relationships, some alunite is intergrown with sulfides.

Hedenquist et al. (1998) argued that the enargite stage was related to the sericite overprint, based on the similarity of fluid inclusion compositions and trends. Where Lepanto-style enargite and Victoria-style carbonate-base metal veins are seen in the same location (e.g., at the base of the Lepanto deposit, and in Victoria at Zone 8, 1000 mL), the latter clearly cut the former, indicating that the Victoria ore was indeed the youngest event. Lepanto Ore Lepanto ore is closely associated with brecciated, massive or vuggy residual quartz and a halo of quartz-alunite. This silicic and quartz-alunite alteration zone mushrooms along the unconformity, and is exposed where this contact crops out in the vicinity of the mine (Figure 2). Gonzalez (1959) mapped this silicic and quartz-alunite alteration along a strike length of >4 km to the west and northwest of FSE, partially outlining the Lepanto ore body. This alteration style also crops out to the southeast and south, near the Guinaoang (Sillitoe and Angeles, 1985) and Palidan porphyry occurrences (Figure 2).

The leached silicic zones are best developed at the unconformity and in the dacite breccia. The quartz-alunite halo to the silicic zones that host ore include local occurrences of kaolinite, dickite, diaspore, pyrophyllite, and native sulfur. Within the metavolcanic sedimentary basement, below the unconformity, quartz-alunite is present immediately beneath or adjacent to the silicic zone. Beneath the unconformity, the advanced argillic zone grades downward or outward to pervasive chlorite alteration. By contrast, quartz-alunite is less well developed in the dacite, except at higher elevations distal from the FSE ore body (Garcia, 1991). In the dacite, there is a zone of kaolinite adjacent to the quartz-alunite zone that overlies the silicic core, and the kaolinite zone grades upward to illite-smectite. This last alteration type passes vertically and laterally to less altered and eventually fresh rock (Gonzalez, 1959; Garcia, 1991).

The ore interval rises in elevation from <700 to >1200 m as the unconformity rises to the northwest (Figure 3). Ore is dominated by veins of massive pyrite

10

and enargite occurring as open-space fillings, matrix or fragments in breccias and subsidiary faults, and as replacements (Gonzalez, 1959; Tejada, 1989). Where this fault intersects the unconformity, ore extends outward along the unconformity. Many subparallel veins, called the Branch veins, splay off to the west from the Lepanto fault. In the vicinity of the FSE porphyry, enargite ore is hosted entirely by the Imbanguila dacite brecca, in veins called the Easterlies (Garcia, 1991). Ore also occurs as stratabound lenses above and below the unconformity (Garcia, 1991).

The Lepanto ore is divided into early and late stages, postdating silicic and much of the quartz-alunite events (Figure 7). The high-sulfidation-state sulfosalts, enargite and luzonite, are the principal Cu minerals and occur with abundant euhedral pyrite (Stage I). Later fine-grained, anhedral pyrite is accompanied by tennantite-tetrahedrite, chalcopyrite, sphalerite, and galena, electrum (typically 900 fine), tellurides (including petzite, calaverite, hessite, and krennerite), selenides, and Bi- and Sn-bearing minerals. Gold ore is associated with tennantite-tetrahedrite and chalcopyrite, most of which appears

paragenetically later than enargite-luzonite (Gonzalez, 1956; Tejada, 1989; Claveria, 1998). Anhydrite, barite and, less commonly, alunite occur as gangue, followed by late quartz crystals and nacrite with minor kaolinite. Covellite occurs as a late alteration product of Cu-sulfide minerals. Enargite fluid inclusion studies indicate that the temperature and salinity decreased with increasing distance from the porphyry (4-2 wt% equiv., Th = 285-190C).

Victoria Ore The Victoria deposit consists of quartz-carbonate-sulfide veins with crustiform and banded textures and pinch and swell features common. Cymoid loops and ladder veins typify the extensional structures, which have a variable strike, from east-west and northeast to north northeast (Figure 5) and steep southeasterly dip, changing to a northwesterly dip to the southeast (Figure 6). Veins are relatively continuous and have been traced along strike for up to 600 m, with 39 g/t Au ore continuous over a 400-m vertical interval. The high-grade ore, >30

11

g/t, is more restricted, but still extends up to 250-m vertical intervals (Claveria et al., 1999, 2000). Fourteen zones of vein sets have been defined to date (Figure 3), with underground workings on the 900, 950 and 1000 mL. Silver grades range from an average of 76 g/t to 49 g/t from Zone 0 to Zone 8. In some zones, the Ag:Au ratio increases with decreasing elevation, although this pattern is not everywhere consistent.

The quartz veins contain abundant pyrite, sphalerite, galena and chalcopyrite. Sphalerite and galena increases in zones of high gold grades, whereas tetrahedrite is associated with massive chalcopyrite, locally accompanied by traces of bornite and chalcocite, the latter as a replacement (Claveria et al., 1999). Gold occurs in quartz and sulfides, particularly sphalerite.

Enargite replaced by tennantite-tetrahedrite occurs in east-west veins on the northern margin, and is probably related to Lepanto veins. The paragenesis of the Victoria veins indicates three stages of mineral deposition (Figure 7; Claveria et al., 1999; Sajona et al., 2000). The early quartz stage is accompanied by sulfides, including dark, Fe-rich sphalerite, and some gold. The middle stage constitutes Fe-poor sphalerite galena quartz, followed by rhodochrosite (with rhodonite) and late quartz. The decrease in Fe content of the sphalerite suggests progressively more reducing conditions with

paragenetic sequence, similar to that noted in the enargite to tennantitetetrahedrite sequence at the neighboring Lepanto deposit (Figure 1). This trend may reflect an increase in the degree of wall rock interaction.

The second stage was the dominant period of gold deposition, with gold content following both sulfide and carbonate abundance. Silver minerals occur as veinlets in sulfides, including acanthite, proustite, pyargyrite, tetrahedrite, and tennantite. Massive crustiform to botryoidal pyrite-chalcopyrite veins are late. The Mn carbonate occurrence is restricted to the north and western portions of the deposit, with decreasing amounts of carbonate at greater depths and to the southwest. This may indicate the northern area is proximal to the fluid source,

12

consistent with rhodochrosite and rhodonite at higher temperature in similar deposits elsewhere. The final stage of barren quartz and anhydrite cement brecciated vein material. Fluid inclusions in sphalerite and rhodochrosite have Th values of 200 to 250 C, with Tm data indicating salinities of 2 to 4 wt% NaCl, similar overall to the enargite data from Lepanto.

The alteration halos are narrower around Victoria veins compared with Lepanto veins. Silicification occurs adjacent to veins, followed by a sericite-clay (locally kaolinite) assemblage outward to a pyropylitic zone. Mapping at the surface prior to drilling the discovery drill hole identified narrow, sulfide-poor quartz (+gypsum) veins. The Imbanguila dacite pyroclastic rocks at the surface, at an elevation of about 1150 m, are argillically altered, with little evidence of silicification. Access to this area is presently limited. Far Southeast Ore FSE porphyry Cu and Au grades are concentric around the dikes and irregular intrusive bodies of melanocratic quartz diorite porphyry (Figure 3; Concepcin and Cinco, 1989). Grades are lower in the later leucocratic porphyry compared with the melanocratic unit, and decrease together with the alteration intensity from intrusive contacts inward toward the center of these later dikes. The lower grade of the later dikes is due to their intramineral timing of intrusion.

K-silicate alteration consists of a biotite-magnetiteK-feldspar assemblage and is associated with veins of vitreous, anhedral quartz. This alteration is partially to pervasively overprinted by alteration assemblages of chlorite plus hematite and/or sericite. Sillitoe and Gappe (1984) used the term sericite-clay-chlorite (SCC) to describe this assemblage, typical of porphyry deposits in the Philippines.

Definitive paragenetic evidence linking Cu sulfide minerals to the early veins of vitreous, anhedral quartz veins was not found (Hedenquist et al., 1998). However, there is petrographic evidence for Cu sulfides to be associated mainly

13

with a later event characterized by formation of euhedral quartz crystals with anhydrite (Hedenquist et al., 1998; Imai, 2000). Cathodoluminescent images show that the early anhedral quartz is overgrown by euhedral quartz (P. Redmond and J. Reynolds, pers. comm., 2000), the latter associated with sulfide deposition. Bleached halos of sericite, cm to m wide and including illite, accompany these euhedral quartz veins that also contain anhydrite-white micahematite-pyrite-chalcopyrite-bornite; these veins cut SCC alteration. Gold in the FSE deposit is present as free grains of electrum associated with chalcopyrite and bornite (Concepcin and Cinco, 1989), and locally is accompanied by BiTe-bearing tennantite (Imai, 2000).

Upward and outward from the core of economic porphyry mineralization the pervasive SCC assemblage grades from sericite-dominated with minor pyrophyllite locally to an assemblage in which pyrophyllite is abundant, variably accompanied by quartz, anhydrite, and kandite minerals (dickite, nacrite and kaolinite). This pervasively altered rock is overlain and, locally, cut by a zone of quartz-alunite that hosts the Lepanto ore, with variable anhydrite-diasporedickite-pyrophyllite. Genetic Relationships Based on studies at FSE-Lepanto, K-silicate and quartz-alunite assemblages had a coupled formation, and were associated with hypersaline liquid and lowsalinity vapor phases, respectively (Hedenquist et al., 1998). Phase separation occurred across the ductile-brittle transition, allowing vapor to ascend and form the silicic and quartz-alunite lithocap. The late sericite overprint was caused by a low salinity magmatic water, and this fluid deposited much of the Cu sulfide ore in both the FSE porphyry and Lepanto HS deposits, possibly remobilized from deeper protore. Exploration for either style of ore deposit should be based on the conclusion that they are genetically related to one another.

The Victoria deposit is closely related to the FSE porphyry and Lepanto HS deposits, possibly representing a late stage of ore deposition in fractures

14

opened subsequent to Lepanto formation. The alteration and ore mineral assemblages of Victoria are similar to those of the gold stage at Lepanto, suggesting that Victoria may be the late, lower sulfidation (Figure 1) equivalent of Lepanto, but formed in structures that did not intersect silicic altered rock. Alternatively, Victoria may have formed as the result of a later pulse of activity related to intrusion southeast of the FSE dikes. EXPLORATION IMPLICATIONS There is a clear genetic relationship between some HS epithermal prospects and underlying porphyry systems. In some cases, both environments host deposits that constitute ore bodies in their own right. As discussed elsewhere (Sillitoe, 1995, 1999), the quartz-alunite lithocaps that remain over or adjacent to porphyry Cu deposits constitute exploration targets for veins and disseminations of HS gold ore, although unless oxidized or particularly high grade, the ubiquitous and refractory Cu-As sulfides are a drawback. Conversely, where an advanced argillic lithocap outcrops, with or without HS ore, there is indication of a magmatic-hydrothermal system that may have been related to porphyry intrusion. In these cases, however, the depth to potential porphyry ore, unless relief or tilting are favorable, may limit the attractiveness of the discovery.

The common association of Au-Ag and/or Ag-base metal quartz veins on the periphery of HS occurrences and porphyry deposits has long been noted. Although a genetic relationship is not yet demonstrated between HS and IS ores, their similarity suggest that geological factors such as the nature of the host and the evolution of permeability may be factors that control which style of ore forms. Although few IS veins have these been economic to mine due to their typically small size, some (e.g., at Bingham) contain appreciable base metal ore. Other IS deposits (e.g., Victoria) can contain as much precious metal as the spatially associated HS ore body, the latter typically more prominent. For these reasons, old HS and porphyry districts, as well as ongoing prospects, should be examined carefully for their IS epithermal vein potential. The shallow

15

depth of their formation relative to a porphyry, and their typically free-milling Au ore relative to HS enargite, mean that these veins can have the position and metallurgy to be attractive if tonnage and grade can be found.

ACKNOWLEDGEMENTS We thank Mr. A.F. Disini, President of Lepanto Consolidated Mining Company, for permission to publish this paper, and the many Lepanto staff who contributed to the information presented here. Discussions with Antonio Arribas R., Marco Einaudi and Dick Sillitoe have contributed to the ideas presented here. REFERENCES
Albinson, T., Norman, D. I., Cole, D., and Chomiak, B., 2001, Fluid inclusion, gas analysis, and stable isotope characteristics of epithermal deposits in Mexico, in Albinson, T. and Nelson, C., Society of Economic Geologist Special Publication, in press. Arribas, A., Jr., 1995, Characteristics of high-sulfidation epithermal deposits, and their relation to magmatic fluid, in Thompson, J.F.H., ed., Magmas, fluids and ore deposits: Mineralogical Association of Canada Shortcourse Series, v. 23, p. 419-454. Arribas, A., Jr., Cunningham, C.G., Rytuba, J.J., Rye, R.O., Kelly, W.C., Podwysocki, M.H., McKee, E.H., and Tosdal, R.M., 1995a, Geology, geochronology, fluid inclusions, and isotope geochemistry of the Rodalquilar gold-alunite deposit, Spain: Economic Geology, v. 90, p. 795-822. Arribas, A., Jr., Hedenquist, J.W., Itaya, T., Okada, T., Concepcin, R.A., and Garcia, J.S., Jr., 1995b, Contemporaneous formation of adjacent porphyry and epithermal Cu-Au deposits over 300 ka in northern Luzon, Philippines: Geology, v. 23, p. 337-340. Barton, P.B., Jr., Bethke, 1977, Economic Geology, v. 72, p. Barton, P.B., Jr., and Skinner, B.J., 1979, Sulfide mineral stabilities, in Barnes, H.L., ed., 2nd edition, Geochemistry of hydrothermal ore deposits: New York, J. Wiley and Sons, p. 278-403. Bonham, H.F., Jr. 1986, Models for volcanic-hosted epithermal precious metal deposits: A review, in Proceedings Int. Volcanological Congress, Symposium 5, Hamilton, New Zealand, 1986: University of Auckland, Center Continuing Education, Auckland, New Zealand, p.13-17. Burnham, C.W., and Ohmoto, H., 1980, Late-stage processes in felsic magmatism. Mining Geology Special Issue, No. 8, p. 111. Buchanan, L.J., 1981, Precious metal deposits associated with volcanic environments in the Southwest. in Dickson, W.R. and Payne, W.D., eds., Relations of tectonics to ore deposits in the Southern Cordillera: Arizona Geological Society Digest 14, 237-262.

16

Claveria, R.J.R., 1998, A paragenetic study of the different sulfides and tellurides in the Lepanto enargite deposit, Mankayan, Benguet, Philippines: GEOCON 98, Manila, Philippines, p. 199-210. Claveria, R.J.R., Cuison, A.G., and Andam, B.V., 1999, The Victoria gold deposit in the Mankayan mineral district, Luzon, Philippines, in Australian Institute of Mining and Metallurgy, PacRim 99, Bali, Indonesia, 10-13 October, Proceedings, p. 73-80. Claveria, R.J.R., Villafuerte, G.P. and Francisco, D.G., 2000, Ore shoot development in the Lepanto Victoria gold ore shoot: GEOCON 2000, Manila, Philippines, in press. Concepcin, R.A., and Cinco, J.C., Jr., 1989, Geology of the Lepanto-Far Southeast gold-rich copper deposit [abs]: International Geological Congress, Washington, D.C., Proceedings, v. 1, p. 319-320, and preprint, 46 p Cooke, D. et al., 1996, Economic Geology, v. 91, p. Cuizon, A.L.G., Claveria, R.J.R., and Andam, B.V., 1998, The discovery of the Lepanto Victoria gold deposit, Mankayan, Benguet, Philippines: GEOCON 98, Manila, Philippines, p. 211-219. Garcia, J.S., Jr., 1991, Geology and mineralization characteristics of the Mankayan mineral district, Philippines, in Matsuhisa, Y., Aoki, M., and Hedenquist, J.W., eds., High-temperature acid fluids and associated alteration and mineralization: Geological Survey of Japan Report 277, p. 21-30. Gatter, 1999, Society of Economic Geologists, Guidebook, v. 31. Gonzalez, A.G., 1959, Geology and genesis of the Lepanto copper deposit, Mankayan, Mountain Province, Philippines: Unpub. Ph.D. dissert., Stanford Univ., 102 p. Harvey, B.A., Myers, S.A., and Klein, T., 1999, Yanacocha gold district, northern Peru, in Australian Institute of Mining and Metallurgy, PacRim 99, Bali, Indonesia, 10-13 October, Proceedings, p. 445-459. Heald, P., Foley, N.K., and Hayba, D.O., 1987, Comparative anatomy of volcanichosted epithermal deposits: Acid-sulfate and adularia-sericite types: Economic Geology, v. 82, p. 1-26. Hedenquist, J.W., and Henley, R.W., 1985, The importance of CO2 on freezing point measurements of fluid inclusions; evidence from active geothermal systems and implications for epithermal ore deposition: Economic Geology, v. 80, p. 1379-1406. Hedenquist, J.W., Matsuhisa, Y., Izawa, E., White, N.C., Giggenbach, W.F., and Aoki, M., 1994, Geology and geochemistry of high-sulfidation Cu-Au mineralization in the Nansatsu district, Japan: Economic Geology, v. 89, p. 1-30. Hedenquist, J.W., Arribas Jr., A., and Reynolds, T.J., 1998, Evolution of an intrusion-centered hydrothermal system: Far Southeast-Lepanto porphyryepithermal Cu-Au deposits, Philippines: Economic Geology: v. 93, p. 373404. Hedenquist, J.W., Arribas R., A., and Urien-Gonzalez, E., 2000, Exploration for epithermal gold deposits, in Gold in 2000, Society of Economic Geologists, Reviews in Economic Geology, v. 13, p. 245-277. Henley, R.W., 1990, Ore transport and deposition in epithermal ore environments, in Herbert, H.K. and Ho, S.E., eds., Stable isotopes and fluid processes in mineralization: University of Western Australia, Geology Department Publication 23, p. 51-69.

17

Hudson, D.M., 1993, The Comstock district, Nevada, in Lahre, M.M., Trexler, J.H., Jr., and Spinosa, C., eds, Crustal evolution of the Great Basin and Sierra Nevada: Cordilleran Rocky Mountain section, Geological Society of America Guidebook, p. 481-496. Imai, A., 2000, Mineral paragenesis, fluid inclusions and sulfur isotope systematics of the Lepanto Far Southeast porphyry Cu-Au deposit, Mankayan, Philippines: Resource Geology, v. 50, p. 151-168. Izawa, E., Urashima, Y., Ibaraki, K., Suzuki, R. Yokoyama, T., Kawasaki, K., Koga, A., and Taguchi, S., 1990, The Hishikari gold deposit: High-grade epithermal veins in Quaternary volcanic of southern Kyushu, Japan, in Hedenquist, J.W., White, N.C. and Siddeley, G., eds., Epithermal gold deposits of the Circum-Pacific: Journal of Geochemical Exploration, v. 36, p. 1-56. Jannas, R.R., Beane, R.E., Ahler, B.A., and Brosnahan, D.R., 1990, Gold and copper mineralization at the El indio deposit, Chile, in Hedenquist, J.W., White, N.C. and Siddeley, G., eds., Epithermal gold deposits of the CircumPacific: Journal of Geochemical Exploration, v. 36, p. 233-266. John, D.A., 2001, Miocene and early Pliocene epithermal gold-silver deposits in the northern Great Basin: Characteristics, distribution, and relationship to magmatism: Economic Geology, v. 96, in press. John, D.A., Garside, L.J., and Wallace, A.R., 1999, Magmatic and tectonic setting of late Ceonozic epithermal gold-silver deposits in northern Nevada, with an emphasis on the Pah Rah and Virginia ranges and the northern Nevada rift: Geological Society of Nevada, Special Publication no. 29, p. 65-158. Lindgren, W., 1933, Mineral deposits, 4th edition: New York, McGraw-Hill, 930 p. Losada-Caldern, A.J., and McPhail, D.C., 1996, Porphyry and high-sulfidation epithermal mineralization in the Nevados del Famitina mining district, Argentina, in Camus, F., Sillitoe, R.H., and Petersen, R., eds., Society of Economic Geology Special Publication 5, p. 91-117. McKibben, M.A., and Hardie, L.A., 1997, Ore-forming brines in active continental rifts, in Barnes, H.L., 3rd edition, Geochemistry of hydrothermal ore deposits: New York, John Wiley, p. 877-935. Saunders, J.A., 1994, Silica and gold textures in bonanza ores of the Sleeper deposit, Humboldt County, Nevada: Evidence for colloids and implications for epithermal ore-forming processes. Economic Geology, v. 89, p. 628 638. Sillitoe, R.H., 1975, Lead-silver, manganese, and native sulfur mineralization within a stratovolcano, El Queva, northwest Argentina: Economic Geology, v. 70, p. 1190-1201. Sillitoe, R.H., l983, Enargite-bearing massive sulfide deposits high in porphyry copper systems: Economic Geology, v. 78, p. 348-352. Sillitoe, R.H., 1989, Gold 88, Economic Geology Monograph. Sillitoe, R.H., 1993, Epithermal models: Genetic types, geometrical controls and shallow features, in Kirkham, R.V., Sinclair, W.D., Thorpe, R.I. and Duke, J.M., eds., Geological Association of Canada Special Paper 40, p. 403-417. Sillitoe, R.H., 1995, Exploration of porphyry copper lithocaps, in Mauk, J.L., and St. George, J.D., eds., PACRIM Congress 1995: Australasian Institute of Mining and Metallurgy, Publication Series No. 9/95, p. 527-532. Sillitoe, R.H., 1999, Styles of high-sulphidation gold, silver and copper mineralization in the porphyry and epithermal environments, in Australian Institute of Mining and Metallurgy, PacRim 99, Bali, Indonesia, 10-13

18

October, Proceedings, p. 29-44. Sillitoe, R.H., 2000, Enigmatic origins of giant gold deposits, in Geology and Ore Deposits: The Great Basin and Beyond, Geological Society of Nevada Symposium Proceedings, Reno, May 15-18, p. 1-18. Sillitoe, R.H., and Angeles, C.A., Jr., 1985, Geological characteristics and evolution of a gold-rich porphyry copper deposit at Guinaoang, Luzon, Philippines, in Asian Mining 85: London, Institution of Mining and Metallurgy, p. 15-26. Sillitoe, R.H., and Gappe, I.M. Jr., 1984, Philippine porphyry copper deposits: Geologic setting and characteristics: Bangkok, United Nations ESCAP, CCOP Technical Publication 14, 89 p. Simmons, S.F., 1991, Hydrologic implications of alteration and fluid inclusion studies in the Fresnillo district, Mexico: Evidence for a brine reservoir and a descending water table during the formation of hydrothermal Ag-Pb-Zn ore bodies: Economic Geology, v. 86, p. 1579-1601. Simmons, S.F., 1995, Magmatic contributions to low-sulfidation epithermal deposits, in Thompson, J.F.H., ed., Magmas, Fluids, and Ore Deposits: Mineralogical Association of Canada, Shortcourse 23, p. 455-477. Tan, 1991, Society of Economic Geologists Newsletter. Tejada, 1989 Vikre, P.G., 1998, Quartz-alunite alteration in the western part of the Virginia Range, Washoe and Storey Counties, Nevada: Economic Geology, v. 93, p. 338-346. White, N.C., 1991, High sulfidation epithermal gold deposits: Characteristics, and a model for their origin, in Matsuhisa, Y., Aoki, M. and Hedenquist, J.W., eds., Acid hydrothermal systems: Geological Survey of Japan Report 277, p. 920. White, N.C., and Hedenquist, J.W., 1990, Epithermal environments and styles of mineralization: variations and their causes, and guidelines for their exploration, in Hedenquist, J.W., White, N.C. and Siddeley, G., eds., Epithermal gold deposits of the Circum-Pacific: Journal of Geochemical Exploration, v. 36, p. 445-474. White, N.C., and Hedenquist, J.W., 1995, Epithermal gold deposits: Styles, characteristics and exploration: Society of Economic Geologists Newsletter, October, p. 1, 9-13. White, N.C., Leake, M.J., McCaughey, S.N., and Parris, B.W., 1995, Epithermal deposits of the southwest Pacific: Journal of Geochemical Exploration, v. 54, p. 87-136.

19

Table 1. Low-, intermediate- and high-sulfidation deposit characteristics


Low sulfidation Setting, related volcanic rocks Depth of formation Setting, typical host rock Bimodal rhyolitebasalt; extension 0-400 m Intermediate sulfidation Andesite-rhyodacite; arcs High sulfidation Andesite-rhyodacite, dominated by calc-alkalic magmas; arcs <100-1000 m >1000 m

300-800 m (rarely >1000 m) Domes; diatremes; . pyroclastic and sedimentary rocks Vein, breccia body, disseminated

Domes; pyroclastic and sedimentary rocks Vein, vein swarm, stockwork, disseminated Fine bands, combs, crustiform, breccia

Domes, central vent; pyroclastic and sedimentary rocks Disseminated, breccia, veinlet to* massive veins Vuggy quartz hosts replacement, to massive sulfide Silicic (vuggy), quartzalunite to pyrophyllitedickite-sericite Alunite, barite, kaolinite to anhydrite, dickite

Dome-diatreme. Porphyry, volcanic, sedimentary rocks Dissemination, veinlets, breccia

Deposit form

Ore textures

Coarse bands

Replacement

Alteration

Alunite-kaolinite blanket, clay halo

Clays, sericite, carbonates; roscoellite

Pyrophyllitesericite, quartzsericite Sericite, pyrophyllite

Gangue

Chalcedony-adulariaillite-calcite

Quartz-carbonaterhodonite-sericite adularia-barite- anhydritehematite- chlorite Pyrite-Au-Ag sulfides/sulfosalts, variable sphalerite, galena, chalcopyrite, tetrahedrite/tennantite

Sulfides

Cinnabar, stibnite; pyrite/marcasitearsenopyrite, Fe-rich sphalerite, pyrrhotite, Au-Ag selenides, Se sulfosalts, Au-Ag-As-Sb-Se(Te)-Hg-Tl Low Ag:Au (~1:1); <0.1-1% base metals Sinter, chalcedony blanket <1% NaCl, gas-rich, <220C McLaughlin, Sleeper, Midas, Ivanhoe, Hishikari (Round Mountain)

Enargite/luzonite, covellite, pyrite to later (deeper) tetrahedritetennantite, chalcopyrite, Fe-poor sphalerite

Bornite, digenite, chalcocite, covellite

Metals

Ag-Au-Pb-Zn, Ba, Mn, Se High Ag:Au (10:1-100s:1); 2-10 (20+)% base metals

Au-Ag, Cu leached (Hg overprint) to Cu-Au-AgBi-Te-Sn

Cu-Au

Notable features Fluids

Some IS veins adjacent to HS ore 3-5 and 10-20% NaCl, 220-280+C Comstock, Tonopah, Creede, Fresnillo, Pachuca, Guanajuato, Casapalca, Arcata, Orcopampa, Victoria, Baguio, Toyoha, Thames, Baia Mare

Steam-heated blanket to vuggy quartz host 4-15+ wt% NaCl

Overprinted on porphyry features Variable, typically hypersaline Bisbee, MM, Chuquicamata, to Far Southeast, Resck

Examples

Yanacocha, Pueblo Viejo, Pierina, La Coipa, Tambo, Pascua, Summitville, Kasuga, Rodalquilar, to El Indio, Lepanto, Chinkuashih, Goldfield, Lahca

Based on Lindgren, 1933; Buchanan, 1981; Heald et al., 1987; Sillitoe, 1993a, 1999; White et al., 1995; John et al., 1999; Albinson et al., 2000, Hedenquist et al., 2000. * Use of the term to refers to transition with increasing depth.

20

Table 2: Alternative nomenclature used for the two end-member epithermal environments and correspondence to active hydrothermal systems

Geothermal (dominated by neutral pH and reduced hypogene fluid)


Au-qtz veins in andesite & rhyolite Ag-Au, Ag, Au-Te, & Au-Se veins Base-metal veins with Au, Ag Hot spring Low sulfur Adularia-sericite Adularia-sericite Low sulfidation

Volcanic-hydrothermal (dominated by early acidic and oxidized hypogene fluid) Goldfield type, Au-alunite

Reference

Lindgren, 1933; Ransome, 1909 Giles and Nelson, 1982; Nakovnik, 1933, Ashley, 1982 Bonham, 1986 Heald et al., 1987 Berger and Henley, 1989 Hedenquist, 1987 Hedenquist et al., 2000 Sillitoe, 1995

Secondary quartzite; enargite-Au High sulfur Acid sulfate Alunite-kaolinite High sulfidation Intermediate sulfidation Barren quartz-alunite lithocap

21

Table 3. Association of HS and IS epithermal deposits, and affiliation with porphyry deposits Ore deposit Associated deposit or prospect
Victoria IS veins; >78 t Au (R), + Ag-Cu-Zn

Known porphyry

Relationships

References

Lepanto, Philippines, HS; 36.3 Mt @ 3.4 g/t Au, 10.8 g/t Ag, 2.9% Cu (P); 4.4 Mt @ 2.4 g/t Au, 1.7% Cu (R) Acupan-Antamok, Baguio, Philippines, IS; >700 t Au

Far Southeast; 650 Mt @ 0.65% Cu, 1.33 g/t Au (0.7% Cu eq cutoff) (r) Subjacent porphyry veins

HS, IS above and adjacent to porphyry, within 0.25 myr period

Hedenquist et al., 1998; Claveria et al., 1999 Cooke et al., 1996

Qtz-alunite lithocap around district, local Au values Chiufen-Wutanshan IS veins; 29 t Au (P)

Porphyry veins overprinted by IS veins, area capped by early qtzalunite zone IS veins 1 km W of HS veins, within continuous altered zone and same age as HS system Qtz veins 5 km N of HS bodies

Chinkuashih, Taiwan, HS; 20 Mt ore w/ >63 t Au, 183 t Ag, 119 kt Cu (P) Kasuga, Japan, HS; 4 Mt @ 2.8 g/t Au (P) Ladolam, Papua New Guinea, IS; 1190 t Au (r)

Proposed at depth from alteration patterns

Tan, 1991

Kago qtz veins, 0.6 t Au (P) Porphyry Cu overprinted by IS epithermal breccia Au-bearing veins on caldera margin Nasivi porphyry Cu prospect with qtzalunite-Au HS veins Resck; 109.4 Mt @ 0.96% Cu (O.8% Cu cutoff); skarn 36 Mt @ 2.2% Cu, 11.5 Mt @ 5% Zn Nevados del Famatina; 300 Mt @ 0.37% Cu, 0.3 g/t Au, 0.06% Mo (r) Tambo HS Au-barite; 37.2 Mt @ 4.2 g/t Au, 42 Mt @ 1 g/t Au (P). Ro del Medio Au-Ag veins Porphyry prospect drilled in late 1960s Au base metal veins at basement contact Porphyry prospects in district

Hedenquist et al., 1994 Sillitoe, 2000

Sector collapse and overprint of epithermal on low-grade porphyry Cu Porphyry prospect 3 km E, 0.4 myr older than epithermal veins

Emperor, Fiji, alkalic epithermal veins; 136 t Au (1993 P) Lahca, Hungary, HS; 3.1 Mt @ 0.63% Cu, 2 g/t Au (P), 5.5 Mt @ 1.4 g/t Au (0.5 g/t cutoff) (r)

Eaton and Setterfield, 1993

Replacement-veins 36.6 Mt @ 3.1-3.5% Zn, 1.2-2.1%Pb. Parfrd IS veins

Porphyry-skarn under HS body, replacement adjacent to porphyry. IS veins overprint HS sulfides 3 km SW Lahca, near adv arg zone Spatial and/or temporal evolution from porphyry to peripheral epithermal ores

Gatter et al., 1999

La Mejicana, Argentina, HS; 1 Mt @ 11 g/t Au, 80 g/t Ag, 3% Cu (P), 0.25 Mt @ 6.1 g/t Au, 64 g/t Ag, 1.08% Cu (R) El Indio, Chile, HS: 23.2 Mt @ 6.6 g/t Au, 50 g/t Ag, 0.2 Mt @ 209 g/t Au (P)

Losada-Caldern and McPhail, 1996

HS shallow breccia Au deposit 5 km to SE and qtz veins 5 km N of HS deposit

Jannas et al., 1990

Yanacocha, Per, HS; 128 t Au (1998 P), 1395 t Au (R, r) Chelopech, Bulgaria, HS; 52.1 Mt @ 3.3 g/t Au, 1.4% Cu (P) Rodalquilar, Spain, HS; 10 t Au (P) Comstock, US, IS; 260 t Au, 5980 t Ag P

Older porphyry deposits 15+ km E Porphyry Cu-Mo adjacent; sericitic roots to HS body Sericitic roots to HS body In district with porphyry prospects

Pyrophyllite on periphery of silicic lithocaps Deep drilling encountered andalusite-diaspore

Harvey et al., 1999 Sillitoe, 1999; personal observations Arribas et al., 1995a Hudson, 1993; Vikre, 1998

IS base-metal veins

Qtz-base metal IS veins within 2 km of HS body Pyrophyllite-alunite halos to veins, older than IS event?

In district with Hgrich lithocaps

HS, high sulfidation; IS, intermediate sulfidation; P, production; R, reserve; r, resource

22

FIGURE CAPTIONS

Figure 1: Oxidation state vs pH at 250C for alteration and sulfide minerals of epithermal interest (modified from Barton et al., 1977; John et al., 1999). Early leaching (vuggy quartz) and quartz-alunite form from a low pH, oxidized fluid (red), creating a barren advanced argillic lithocap that may host subsequent high-sulfidation (HS) state sulfide ore (orange), including enargite (en) with pyrite (py), e.g., the Lepanto Cu-Au deposit. Gold typically follows enargite, associated with intermediatesulfidation (IS) state minerals such as tennantite-tetrahedrite (tn), chalcopyrite (ccp) and sphalerite (spl) with low Fe content, and is sericite stable (blue). IS assemblages are also typical of Ag-Au base-metal vein deposits such as Comstock Lode, Nevada (John et al., 1999) and the Victoria Au-Ag-Cu-Pb-Zn deposit. By contrast, high-grade epithermal gold deposits such as Midas and Sleeper, Nevada, and Hishikari, Japan, are Ag- and base-metal poor. The sulfide assemblage of pyrite, pyrrhotite (po), arsenopyrite and high-Fe sphalerite indicates a very low-sulfidation (LS) state (green). cv-covellite, bn-bornite, di-digenite, ang-anglesite, gn-galena, hm-hematite, chl-chlorite (superimposed on magnetite stability).

23

Figure 2: Map of the Mankayan district, northern Luzon (inset), Philippines, showing the simplified geology, and location and type of known hydrothermal deposits. Outlines of the economically most important deposits, Far Southeast (FSE), Lepanto, Victoria and Guinaoang are shown projected to the surface. Modified from Garcia (1991). A-B-C shows line of section, Fig. 3.

Figure 3: Schematic northwest-southeast longitudinal section through the Lepanto HS and underlying Far Southeast (FSE) porphyry Cu-Au deposits, turning south (at B; Fig. 2) through the Victoria IS vein deposit, showing the geological units and extent of porphyry ore (from Concepcin and Cinco, 1989; Garcia, 1991, Cuison et al., 1998).

24

Figure 4: K-Ar ages for mineral separates from igneous and hydothermal minerals associated with the FSE, Lepanto, and Victoria deposits (Arribas et al., 1995a; Sajona et al., 2000; unpublished Lepanto Consolidated Mining Co. data). Samples arranged by lithologic or mineral groups; analytical uncertainty at 2 level contained within size of symbol, except where indicated by error bar. See Arribas et al. (1995a) and Hedenquist et al. (1998) for details on all but Victoria samples. Ages (2 sigma errors) and K concentrations for illite concentrates from Victoria samples: 5M P22 (quartz vein) 1.14 0.02 Ma, 5.53 wt%; 8M P1W (quartz vein) 1.16 0.02 Ma, 5.50 wt% (LCMC unpublished data; analyzed at BGR, Germany). 8K stope (altered wall rock) 1.5 0.14 Ma, 6.7 wt% (Sajona et al., 2000; analyzed at IGNS, New Zealand).

25

Figure 5: a) 1000 m L plan map of the Victoria veins, showing their arcuate trend, eastwest to the NE, to south southwest to the SW. b) 900 m L plan map, showing continued arcuate vein trends.

26

Figure 6: Northwest-southeast section (Fig. 5) through the Victoria veins; note change in direction of dip across deposit, creating an upward flaring pattern.

27

Figure 7: Paragenetic sequences of ore and related minerals for Lepanto and Victoria deposits (Claveria, 1998; Claveria et al., 1999).

28

Potrebbero piacerti anche