You are on page 1of 29

ProExplo Congreso, Lima, Perú, 24-28 April 2001

Types of sulfide-rich epithermal deposits, and their affiliation to


porphyry systems:

Lepanto–Victoria–Far Southeast deposits, Philippines, as examples

Jeffrey W. Hedenquist
99 Fifth Avenue, Suite 420, Ottawa, Ontario, K1S 5P5, Canada
Hedenquist@aol.com, fax 1-613-230-9292

Rene Juna R. Claveria and Gener P. Villafuerte


Lepanto Consolidated Mining Company
BA – Lepanto Building, 8747 Paseo de Roxas, 1226 Makati City, Philippines

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 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
Far Southeast 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
is related to porphyry deposits. We review the types of epithermal deposit on a genetic
basis, and note those with evidence for a spatial, and in some cases genetic
relationship to 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 type of epithermal deposit, widely different Ag:Au ratios
have been noted, with some of the Ag-rich deposits having a high base-metal content
(Table 1). 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).

2
The Au-rich low-sulfidation (LS) deposits are typically hosted by rhyolitic to
dacitic rocks that commonly have a bimodal association with basalts in an extensional
setting (Sillitoe, 1993; John et al., 1999). The Ag:Au ratio is relatively low, typically
close to 1:1. 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 (Fig. 1;
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, the majority of epithermal deposits have higher Ag:Au ratios and are
hosted by andesite-rhyodacite volcanic rocks in arc settings, akin to the setting of most
porphyry Cu deposits. The Ag:Au ratio typically has a minimum value of 10:1, and can
range up to >100:1, and sulfides are relatively abundant, particularly in the base metal-
rich variety. There is a wide variety of characteristics, much wider than in the case of
the bonanza LS deposits. Quartz veins are crystalline, massive and comb textured, and
sericite is a common alteration mineral and gangue, whereas adularia is relatively
uncommon; Mn carbonates are common, particularly in the sulfide-rich examples. In
addition to pyrite, Fe-poor sphalerite, galena, chalcopyrite and tennantite-tetrahedrite
occur, in some deposits forming massive veins. This sulfide assemblage indicates an
intermediate sulfidation state (Fig. 1), although wide variations in sulfide assemblage
occur. Examples with type characteristics occur in Nevada and Colorado, USA,
throughout Mexico and Perú, and in many parts of the Philippines and Japan as well as
in New Zealand and eastern Europe, particularly 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). These 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

3
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. There are many approaches to classifying epithermal
deposits (Table 2), ranging from form to alteration assemblage (White and Hedenquist,
1990). The metal content and sulfide plus gangue assemblage 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) 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 near the Northern
Nevada Rift, including Sleeper, Midas, Ivanhoe and Mule Canyon, the Au-rich, sulfide-
poor veins have a very low sulfidation state assemblage (Fig. 1). By contrast, within the
Western Andesite Arc, relatively 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 (Table 2; Hedenquist et
al., 2000), stressing their distinct tectonic settings, magmatic affiliations, metal
complements, sulfide and gangue mineral assemblages, and in some cases, deposit
form (Table 1; 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. However,
the majority of the precious metal content of HS deposits is typically associated with a
paragenetically late assemblage that has an intermediate sulfidation state (Jannas et
al., 1990; Hedenquist et al., 1998), with this assemblage similar to that of IS deposits
(Table 1).

4
Causes of variation among epithermal deposits

Fluid composition
As noted above, 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 (Table 1).
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 led 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?

5
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 (John, 2001).
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 systems that owe their high salinities, >20 wt% NaCl, to a demonstrably amagmatic
source, the dissolution of evaporites (McKibben and Hardie, 1997), although they are
driven by rift-related magmatism. The high salinities in the epithermal environment that
are recorded by fluid inclusions of Ag-rich IS deposits may reflect episodic incursions of
brine into low-salinity systems, based on evidence found at Fresnillo (Simmons, 1991)
and similar deposits (Albinson et al., 2001). Thus, Ag ± base metal-rich brines many
underlie many geothermal systems in arc settings, with tectonic or magmatic triggers
causing periodic brine ascent and mineralization. Such stable-stratified double-diffusive
systems (Fournier, 1987) would explain also why such brines are seldom 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 would be one 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 in Nevada, 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 (Fig. 1). This may be the result of hydrothermal fluid
simply equilibrating with the volcanic and intrusive host rocks, or it may reflect a more

6
intrinsic magmatic affiliation (John, 2001). If the latter is 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 rock – having no
buffering 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 3), 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 3). Nowhere is this spatial association better seen than in the
Mankayan mineral district in northern Luzon, where the Lepanto HS Cu-Au and Victoria

7
IS Au-Ag-base metal deposit overlie the Far Southeast (FSE) porphyry Cu-Au ore body
to the northwest and south, respectively (Fig. 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; Concepción and Cinco, 1989; 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 (Fig. 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 exploration to extend the Lepanto reserves. Drilling near the
Nayak workings (Fig. 2) 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 (Fig. 3). Production
began in 1997. The reserve of the IS veins should eventually compete in size with the
total production and reserves of HS ore from Lepanto (Table 3).

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 (Fig. 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
(Fig. 2). 3) A Pliocene dacitic to andesitic breccia and porphyry unit, the Imbanguila
hornblende dacite, predates Cu-Au mineralization; and (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 (Fig. 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
(Fig. 5).

8
The ore deposits in the Mankayan district are spatially and temporally related to
the Pliocene to Pleistocene event of intermediate-composition volcanism (Fig. 4). 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
(Fig. 3).
K-silicate alteration related to the FSE porphyry is centered on dikes
(Concepción and Cinco, 1989; Hedenquist et al., 1998), and formed at 1.41 ± 0.05 Ma
(n=6 K-Ar biotite dates; Fig. 4). A dacite breccia also erupted at this time (hornblende,
1.43 ± 0.21 Ma). 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 basement-pyroclastic rock contact, synchronous with the K-silicate
alteration stage. Quartz-illite-sulfide veins with illite-chlorite halos that cut the porphyry
system followed at 1.30 ± 0.07 Ma (n=10). New K-Ar dating indicate that illite-altered
wall rock, dated at 1.50 ± 0.14 Ma (Sajona et al., 2000) and indistinguishable from the
FSE illite ages, was cut by the Victoria veins at 1.15 ± 0.03 Ma (Ar-Ar results on vein
illite, n=2; LCMC, unpublished ages, Fig. 4). This younger event corresponds to a K-Ar
date on fresh dacite collected from the surface over Victoria (biotite, 1.18 ± 0.08 Ma).
The dating results suggest that the Victoria veins are younger than the principal
muscovite (sericite) 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 muscovite overprint of the underlying
porphyry, 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. By contrast, quartz veins with sphalerite and gold, similar to those of
Victoria, cut enargite ore and advanced argillic halos in the root zone of the Lepanto
deposit (Cuison et al., 1998; Claveria et al., 1999a). Thus, the paragenetic and dating
results are consistent, and indicate that Victoria ores were deposited slightly later than
Lepanto ores, and that there is some spatial overlap between the two.

9
Lepanto ore
Lepanto ore is closely associated with brecciated, massive or vuggy residual
quartz with alunite. This silicic alteration zone mushrooms along the unconformity, and
is exposed where this contact crops out in the vicinity of the mine (Fig. 2). Gonzalez
(1959) mapped the quartz-alunite zone outcropping at the base of the Imbanguila dacite
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 (Fig. 2).
The leached silicic zones are best developed at the unconformity and in the
dacite breccia. The alteration halo around the silicic zones that host ore include local
occurrences of alunite, 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 (Fig. 3). Ore is dominated by veins of massive pyrite 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 sub-parallel veins,
called the Branch veins, splay off to the west from the Lepanto fault (Fig. 6). 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 the silicic
alteration event (Fig. 7). The high-sulfidation-state sulfosalts, enargite and luzonite, are
the principal Cu minerals and occur with abundant euhedral pyrite (early stage). Later

10
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 (Fig. 7; 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-
190°C).

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 (Figs. 5 and 6) and steep southeasterly dip, changing to a
northwesterly dip to the southeast (Fig. 5). Veins are relatively continuous and have
been traced along strike for up to 600 m, with 3-9 g/t Au ore continuous over a 400-m
vertical interval. The high-grade ore, >30 g/t, is more restricted, but still extends up to
250-m vertical intervals (Claveria et al., 1999a,b). Fourteen zones of vein sets have
been defined to date (Fig. 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 increase 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., 1999a). Gold occurs in quartz
and sulfides, particularly sphalerite. Enargite replaced by tennantite-tetrahedrite occurs
in east-west veins on the northern margin of the deposit, and is probably related to the
paragenetically earlier Lepanto veins.

11
The paragenesis of the Victoria veins indicates three stages of mineral deposition
(Fig. 7; Claveria et al., 1999a; 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 early, high Fe content of the sphalerite indicates
relatively reduced conditions, whereas the Fe-poor sphalerite during the second stage
indicates somewhat more oxidized conditions, similar to that of the tennantite-
tetrahedrite sequence associated with gold in the neighboring Lepanto deposit (Fig. 1).
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, consistent with rhodochrosite and rhodonite formed
at higher temperature in similar deposits elsewhere. The final stage, largely barren of
metals, consists of quartz and anhydrite that cement brecciated vein material. Fluid
inclusions in sphalerite and rhodochrosite have Th values of 200 to 250 C and salinities
of 2 to 4 wt% NaCl equivalent, similar to fluid inclusion data from Lepanto enargite.
The alteration halos are narrower around Victoria veins compared to Lepanto
veins. Silicification occurs adjacent to veins, followed by a muscovite-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 dikes and irregular
intrusive bodies of melanocratic quartz diorite porphyry (Fig. 3; Concepción 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

12
contacts inward toward the center of these later dikes. The lower grade of the later
dikes may be caused by their intrusion during hydrothermal activity.
K-silicate alteration consists of a biotite-magnetite±K-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
muscovite. 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 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 muscovite, cm to m wide
and including illite, accompany these euhedral quartz veins that also contain anhydrite-
muscovite-hematite-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 (Concepción and Cinco, 1989), and locally is accompanied by Bi-Te-bearing
tennantite (Imai, 2000).
Upward and outward from the core of economic porphyry mineralization the
pervasive SCC assemblage grades from muscovite-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 silicic zone with local alunite that hosts the
Lepanto ore, with variable anhydrite-diaspore-dickite-pyrophyllite.

Genetic relationships
Based on studies at FSE-Lepanto, K-silicate and quartz-alunite assemblages
were coupled, associated with hypersaline liquid and low-salinity vapor phases,
respectively (Hedenquist et al., 1998). Phase separation occurred across the ductile-
brittle transition, allowing vapor to ascend and form the silicic lithocap with alunite. The
late muscovite overprint was caused by a moderate salinity magmatic water, and this

13
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.
The Victoria deposit is closely related to the FSE porphyry and Lepanto HS
deposits, and may represent a late stage of ore deposition in fractures opened
subsequent to Lepanto formation. The alteration and ore mineral assemblages of
Victoria are similar to those of the gold stage at Lepanto (Fig. 7), and also to the later
porphyry event. This observation could be explained if Victoria was the late, lower
sulfidation (Fig. 1) equivalent of Lepanto, formed during the waning stages of the
hydrothermal system (Cuison et al., 1998; Claveria et al.., 1999a), where fluid ascended
along structures that did not intersect silicic altered rock. This trend to lower sulfidation
state may simply reflect an increase in the degree of wall rock interaction with time.
However, the dating suggests that the Victoria hydrothermal system was associated
with a slightly later pulse of magmatic activity south of the FSE center. Thus, the
relationship between the epithermal deposits may be traced to magmatic evolution
rather than simply to hydrothermal evolution.

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 lithology-controlled, massive silicic lithocaps with hypogene alunite that remain over
or adjacent to porphyry Cu deposits constitute exploration targets for HS epithermal
gold veins and disseminated replacements. In such situations, there needs to be
evidence that gold was introduced subsequent to leaching and lithocap development,
e.g., the presence of Cu sulfides or barite. However, unless oxidized or particularly high
grade, the refractory Cu-As sulfides that are ubiquitous with gold ore are a drawback.
Conversely, the presence of a silicic lithocap, with or without HS ore, is evidence that a
magmatic-hydrothermal system and intrusion exists at depth, and this may be related to
a related porphyry system. In these cases, however, the depth to potential porphyry
ore, unless there is favorable relief or tilting, may limit the attractiveness of the

14
discovery. In addition, it is essential to distinguish hypogene lithocaps from advanced
argillic blankets of steam-heated or supergene origin (Sillitoe, 1993; Hedenquist et al.,
2000), as the latter may have no relationship to porphyry systems.
Gold-Ag and/or Ag-base metal quartz veins with IS characteristics are common
on the periphery of HS occurrences and porphyry deposits, and their presence has long
been noted (Sillitoe, 1975, 1993). Although a genetic relationship has not yet been
demonstrated between HS and IS ores, their overall paragenetic similarity is striking.
This suggests that geological factors, such as the nature of the host and the evolution of
permeability, may control which style of ore forms. Intermediate sulfidation veins
around the world have been the most economic style of epithermal deposit throughout
history. Despite this fact, relatively few IS veins on the margins of porphyry deposits
have been economic to mine due to their typically small size, although there are notable
exceptions (e.g., at Bingham). Other IS deposits (e.g., Victoria) can contain as much
precious metal in veins as the spatially associated HS ore body, the latter typically more
prominent due to their resistant lithocap hosts. In addition, it must be appreciated that
the precious-metal content of most HS deposits is essentially IS in nature (e.g., El
Indio). Finally, it is notable that HS deposits hosted by carbonate rocks are known to
grade out to rich Zn-Pb-Ag ores (e.g., East Tintic, Cerro de Pasco, San Gregorio;
Fontboté and Bendezú, 1999). For these reasons, old HS and porphyry districts, as
well as ongoing prospects, should be examined carefully for their IS epithermal vein
potential at distances of several hundred meters to a few km (Fig. 6). The shallow
depth of their formation relative to porphyry deposits (Fig. 3), and their typically free-
milling Au ore compared with enargite of HS deposits, 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 present and past Lepanto staff who contributed to
the information presented here. Antonio Arribas R., Marco Einaudi, David John, Dick
Sillitoe and Noel White are among those who have contributed significantly to the ideas
presented here. Noel White critically reviewed the manuscript.

15
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., Concepción, 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.

Ashley, R.P, 1982, Occurrence model for enargite-gold deposits, U.S. Geol. Survey Open-File Report 82-
795, p. 144-147.

Barton, P.B., Jr., Bethke, P.M., and Roedder, E., 1977, Environment of ore deposition in the Creede
mining district, San Juan mountains, Colorado: Part III. Progress toward interpretation of the chemistry of
the ore-forming fluid for the OH vein: Economic Geology, v. 72, p. 1-24.

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.

Berger, B.R., and Henley, R.W., 1989, Advances in understanding of epithermal gold-silver deposits, with
special reference to the western United States. Econ. Geol. Monograph 6, p. 405-423.

Bonham, H.F., Jr.1986, Models for volcanic-hosted epithermal precious metal deposits: a review, in
Proceedings International 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. 1–11.

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.

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., 1999a, 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., 1999b, Ore shoot development in the Lepanto
Victoria gold ore shoot: GEOCON 2000, Manila, Philippines, CD release.

Concepción, 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

16
Cooke, D.R., McPhail, D.C., and Bloom, M.S., 1996, Epithermal gold mineralization, Acupan, Baguio
district, Philippines: Geology, mineralization, alteration, and the thermochemical environment of ore
deposition: Economic Geology, v. 91, p. 243-272.

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.

Fontboté, L., and Bendezú, R., 1999, The carbonate-hosted Zn-Pb San Gregorio deposit, Colquijirca
district, central Peru, as part of a high-sulfidation epithermal system, in Stanley, C.J. et al., eds., Mineral
deposits: Processes to processing, Rotterdam, Balkema, p. 495-498.

Fournier, R.O., 1987, Conceptual models of brine evolution in magmatic-hydrothermal systems, in


Decker, R.W., Wright, T.L., and Stauffer, P.H., eds., Volcanism in Hawaii: U.S. Geological Survey
Professional Paper, 1350, p. 1487-1506.

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, I., Molnár, F., Foldessy, J., Zelenka, T., Kiss, J., and Szebényi, G., 1999, High- and low-sulfidation
epithermal mineralization of the Mátra mountains, northeast Hungary, in Molnár, F., Lexa, J., and
Hedenquist, J.W., eds., Epithermal mineralization of the western Carpathians: Society of Economic
Geologists, Guidebook, v. 31, p. 155-179.

Giles, D.L., and Nelson, C.E., 1982, Epithermal lode gold deposits of the circum-Pacific, in Transactions,
rd
3 Circum Pacific Energy and Mineral Resources Conference, Honolulu, August, 1982: Tulsa, Oklahoma,
American Association of Petroleum Geologists, p. 273-278.

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 volcanic-hosted 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 porphyry-epithermal Cu-Au deposits, Philippines: Economic
Geology: v. 93, p. 373-404.

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 Circum-Pacific: 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, submitted.

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.
th
Lindgren, W., 1933, Mineral deposits, 4 edition: New York, McGraw-Hill, 930 p.

Losada-Calderón, 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.
rd
McKibben, M.A., and Hardie, L.A., 1997, Ore-forming brines in active continental rifts, in Barnes, H.L., 3
edition, Geochemistry of hydrothermal ore deposits: New York, John Wiley, p. 877-935.

Nakovnik, N.I., 1933, New data about the so called secondary quartzites. Problemy sovetskoy geologii,
N6, p.228-242.

Ransome, F.L., 1909, The geology and ore deposits of Goldfield, Nevada: U.S. Geological Survey,
Professional Paper 66, 258 p.

Sajona, F.J., Claveria, R.J.R., Izawa, E., Motomura, Y., Sakakibara, H., Imai, A., and Watanabe, K., 2000,
Victoria’s secret: implications of a carbonate-base metal low sulfidation gold deposit to mineralization in
the Mankayan district: GEOCON 2000, Manila, Philippines, CD release.

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 deposits in western Pacific island arcs: The magmatic connection, in Keays,
R.R., Ramsey, W.R.H., and Groves, D.I., eds., the geology of gold deposits: The perspective in 1988:
Economic Geology Monograph 6, p. 274-291.

18
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 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, L.P., 1991, The Chinkuashih gold-copper deposits, Taiwan: Society of Economic Geologists
Newsletter, number 7, p. 1, 22-24.

Tejada, M.L.G., 1989, Characteristics of paragenesis of luzonite in the Lepanto copper-gold deposit,
Mankayan, Benguet, Philippines: Unpublished M.Sc. thesis, National Institute of Geological Sciences,
University of the Philippines, 89 p.

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. 9-20.

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 epithermal deposit characteristics

Low sulfidation Intermediate sulfidation High sulfidation

Setting, related Bimodal rhyolite- Andesite-rhyodacite; arcs Andesite-rhyodacite, dominated by calc-alkalic


volcanic rocks basalt; extension magmas; arcs

Depth of 0-400 m 300-800 m <100-1000 m >1000 m


formation (rarely >1000 m)

Setting, typical Domes; pyroclastic Domes; diatremes; . Domes, central vent; Dome-diatreme.
host rock and sedimentary pyroclastic and pyroclastic and Porphyry, volcanic,
rocks sedimentary rocks sedimentary rocks sedimentary rocks

Deposit form Vein, vein swarm, Vein, breccia body, Disseminated, breccia, Dissemination,
stockwork, disseminated veinlet down to massive veinlets, breccia
disseminated veins

Ore textures Fine bands, combs, Coarse bands Vuggy quartz hosts Replacement
crustiform, breccia replacement, down to
massive sulfide

Alteration Alunite-kaolinite Clays, sericite, Silicic (vuggy), quartz- Pyrophyllite-


blanket, clay halo carbonates; roscoellite alunite, to pyrophyllite- sericite, quartz-
dickite-sericite roots sericite

Gangue Chalcedony-adularia- Quartz-carbonate- Alunite, barite, kaolinite Sericite,


illite-calcite rhodonite-sericite± to deeper anhydrite, pyrophyllite
adularia-barite- anhydrite- dickite
hematite- chlorite

Sulfides Cinnabar, stibnite; Pyrite-Au-Ag Enargite/luzonite, Bornite, digenite,


pyrite/marcasite- sulfides/sulfosalts, covellite, pyrite to later chalcocite,
arsenopyrite, Fe-rich variable sphalerite, (deeper) tetrahedrite- covellite; enargite-
sphalerite, pyrrhotite, galena, chalcopyrite, tennantite, chalcopyrite, luzonite
Au-Ag selenides, Se tetrahedrite/tennantite Fe-poor sphalerite
sulfosalts,

Metals Au-Ag-As-Sb-Se- Ag-Au-Pb-Zn, Ba, Mn, Se Au-Ag, Cu leached (Hg Cu(-Au)


(Te)-Hg-Tl High Ag:Au (10:1-100s:1); overprint) to deeper Cu-
Low Ag:Au (~1:1); 2-10 (20+)% base metals Au-Ag-Bi-Te-Sn; locally
<0.1-1% base metals bonanza Au grades

Notable Sinter, chalcedony Some IS veins adjacent to Steam-heated blanket Overprinted on


features blanket HS ore over vuggy quartz host porphyry features

Fluids <1% NaCl, gas-rich, 3-5 and 10-20% NaCl, 4-15+ wt% NaCl Variable, typically
<220°C 220-280+°C hypersaline

Examples McLaughlin, Sleeper, Comstock, Tonopah, Yanacocha, Pueblo Bisbee, MM,


Midas, Ivanhoe, Creede, Fresnillo, Viejo, Pierina, La Coipa, Chuquicamata,
Hishikari (Round Pachuca, Casapalca, El Indio-Tambo, Pascua- down to Far
Mountain) Arcata, Orcopampa, Lama, Summitville, Southeast, Resck
Victoria, Baguio, Toyoha, Goldfield, Rodalquilar,
Thames, Baia Mare Chelopech, Lahóca,
Lepanto, Chinkuashih

Based on Lindgren, 1933; Buchanan, 1981; Heald et al., 1987; Sillitoe, 1993, 1999; White et al., 1995;
John et al., 1999; Albinson et al., 2000, Hedenquist et al., 2000.

20
Table 2: Alternative nomenclature used for the two end-member epithermal environments and
correspondence to active hydrothermal systems

Geothermal Volcanic-hydrothermal Reference


(dominated by neutral pH and (dominated by early acidic and
reduced hypogene fluid) oxidized hypogene fluid)

Au-qtz veins in andesite & rhyolite Goldfield type, Au-alunite Lindgren, 1933; Ransome,
Ag-Au, Ag, Au-Te, & Au-Se veins 1909
Base-metal veins with Au, Ag

Hot spring Secondary quartzite; enargite-Au Giles and Nelson, 1982;


Nakovnik, 1933, Ashley, 1982

Low sulfur High sulfur Bonham, 1986

Adularia-sericite Acid sulfate Heald et al., 1987

Adularia-sericite Alunite-kaolinite Berger and Henley, 1989

Low sulfidation High sulfidation Hedenquist, 1987

Intermediate sulfidation Hedenquist et al., 2000

Barren quartz-alunite lithocap Sillitoe, 1995

21
Table 3. Association of HS and IS epithermal deposits, and affiliation with porphyry deposits
Ore deposit Associated Roots, known Relationships References
deposit/prospect porphyry deposit

Lepanto, Philippines, Victoria IS veins; Far Southeast; 650 Mt HS, IS above and adjacent Hedenquist et
HS; 36.3 Mt @ 3.4 g/t >78 t Au (R), + @ 0.65% Cu, 1.33 g/t to porphyry, within 0.25 myr al., 1998;
Au, 10.8 g/t Ag, 2.9% Ag-Cu-Zn Au (0.7% Cu eq cutoff) period Claveria et al.,
Cu (P); 4.4 Mt @ 2.4 g/t (r) 1999
Au, 1.7% Cu (R)

Acupan-Antamok, Qtz-alunite Porphyry veins at depth, Porphyry veins overprinted by Cooke et al.,
Baguio, Philippines, IS; lithocap around older IS veins, area capped by 1996
>700 t Au district, local Au early qtz-alunite zone

Chinkuashih, Taiwan, Chiufen- Proposed at depth from IS veins 1 km W of HS veins, Tan, 1991
HS; 20 Mt w/ >63 t Au, Wutanshan IS alteration patterns within continuous altered
183 t Ag, 119 kt Cu (P) veins; 29 t Au (P) zone, same age as HS

Kasuga, Japan, HS; 4 Kago qtz veins, Roots of base-metal Qtz veins 5 km N of HS Hedenquist et
Mt @ 2.8 g/t Au (P) 0.6 t Au (P) veins, sericite bodies al., 1994

Ladolam, PNG, IS; Porphyry Cu Sector collapse and overprint Sillitoe, 2000
1190 t Au (r) overprinted by IS of epithermal on low-grade
epithermal breccia porphyry Cu

Emperor, Fiji, alkalic Au-bearing veins Nasivi porphyry Cu Porphyry prospect 3 km E, Eaton and
epithermal veins; 136 t on caldera margin prospect with qtz- 0.4 myr older than epithermal Setterfield,
Au (1993 P) alunite-Au HS veins veins 1993

Lahóca, Hungary, HS; Replacement- Resck; 109.4 Mt @ Porphyry-skarn under HS Gatter et al.,
3.1 Mt @ 0.63% Cu, 2 veins 36.6 Mt @ 0.96% Cu (O.8% Cu body, replacement adjacent to 1999
g/t Au (P), 5.5 Mt @ 1.4 3.1-3.5% Zn, 1.2- cutoff); skarn 36 Mt @ porphyry. IS veins overprint
g/t Au (0.5 g/t cutoff) (r) 2.1%Pb. 2.2% Cu, 11.5 Mt @ 5% HS sulfides 3 km SW Lahóca,
Paráfürdö IS veins Zn near adv arg

La Mejicana, Argentina, Nevados del Famatina; Spatial and/or temporal Losada-


HS; 1 Mt @ 11 g/t Au, 300 Mt @ 0.37% Cu, evolution from porphyry to Calderón and
80 g/t Ag, 3% Cu (P), 0.3 g/t Au, 0.06% Mo (r) peripheral epithermal ores McPhail, 1996
0.25 Mt @ 6.1 g/t Au,
64 g/t Ag, 1.08% Cu (R)

El Indio, Chile, HS: 23.2 Tambo HS Au- Porphyry prospects in HS shallow breccia Au Jannas et al.,
Mt @ 6.6 g/t Au, 50 g/t barite; 37.2 Mt @ district, no known deposit 5 km to SE and Río 1990
Ag, 0.2 Mt @ 209 g/t 4.2 g/t Au, 42 Mt association del Medio Au-Ag veins 5 km
Au (P) @ 1 g/t Au (P) N of HS deposit

Yanacocha, Perú, HS; Porphyry prospect Older porphyry deposits Pyrophyllite on periphery of Harvey et al.,
128 t Au (1998 P), 1395 drilled in late 15+ km E silicic lithocaps 1999
t Au (R, r) 1960s

Chelopech, Bulgaria, Au – base metal Porphyry Cu-Mo Deep drilling encountered Sillitoe, 1999;
HS; 52.1 Mt @ 3.3 g/t veins at basement adjacent; sericitic roots andalusite-diaspore personal
Au, 1.4% Cu (P) contact to HS body observations

Rodalquilar, Spain, HS; IS base-metal Sericitic roots to HS Qtz-base metal IS veins Arribas et al.,
10 t Au (P) veins body within 2 km of HS body 1995a

Comstock, US, IS; 260 t In district with Hg- Porphyry prospects in Pyrophyllite-alunite halos to Hudson, 1993;
Au, 5980 t Ag P rich lithocaps district, older veins, older than IS event? Vikre, 1998

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

22
Fig. 1: Oxidation state vs pH at 250°C for alteration and sulfide minerals of epithermal interest (modified
from Barton et al., 1977; John et al., 1999). Early leaching is caused by a low pH, oxidized fluid (red),
creating a barren, advanced argillic zone of vuggy quartz with alunite. This lithocap may host subsequent
high-sulfidation (HS) state sulfide ore (orange), including enargite (en) with pyrite (py). Gold typically
follows enargite, associated with intermediate-sulfidation (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
Fig. 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). Extent of silicic and quartz-alunite altered outcrop from Gonzalez (1959).

24
Fig. 3: Schematic northwest-southeast longitudinal section, looking northeast through the Lepanto HS
deposit, with the underlying Far Southeast (FSE) porphyry Cu-Au deposit projected from the east. Line of
section turns south (at B; Fig. 6) through the Victoria IS vein deposit (from Concepción and Cinco, 1989;
Garcia, 1991, Cuison et al., 1998). The solid red veins of Victoria have been proven by drilling and
underground workings, whereas the dashed lines show veins that are inferred only on the basis of drilling.
Most veins are shown as terminating for lack of information, not necessarily due to pinching.

25
Fig. 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; all symbols or error bars ±2 sigma errors). Ages (±2 sigma errors) and K
concentrations for illite concentrates from Victoria samples collected on the 900 mL: 5M P22 (quartz vein)
1.14 ± 0.02 Ma, 5.53 wt%; 8M P1W (quartz vein with base-metal sulfides) 1.16 ± 0.02 Ma, 5.50 wt%
(LCMC unpublished Ar-Ar dates; 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).

26
Fig. 5: Plan maps of Victoria vein distribution and host rocks on the 900 and 1000 m levels, plus a
northwest-southeast section through the veins along A-A’.

27
Fig. 6: Detailed spatial relationship between the Lepanto ore shoots in yellow (composite projection to the
surface of >3.0 wt% Cu equivalent ore, largely from 700 to 1000 m elevation), the FSE outline at 0 m
elevation (>0.7 wt% Cu eq.), and the Victoria vein distribution at 1000 m elevation. Also shown are
surface and near-surface outcrops of the Nayak Au-Ag quartz veins. From unpublished Lepanto
Consolidated Mining Company files. Note the arcuate trend of veins, from east-west in the north, roughly
parallel to the Lepanto Easterly ore shoots, to north-south in the direction of the Nayak vein outcrops.

28
Fig. 7: Paragenetic sequences of ore and related minerals for Lepanto and Victoria deposits (Claveria,
1998; Claveria et al., 1999a). Lepanto ore was deposited subsequent to an early stage of leaching, which
generated the vuggy quartz ore host. Note the similarity in sulfide assemblage between the late, Au-rich
stage of Lepanto and the main-stage Victoria ores.

29

You might also like