You are on page 1of 36

In Magmas, Fluids, and Ore Deposits, Ed.: J.FM.

Thompson,
Mineralogical Association of Canada Short Course Vol. 23 (1995)

Chapter 19

CHARACTERISTICS OF HIGH-SULFIDATION EPITHERMAL


DEPOSITS, AND THEIR RELATION TO MAGMATIC FLUID
Antonio Arribas Jr.
Mineral Resources Department, Geological Survey ofJapan,
1-1-3 Higashi, Tsukuba 305, Japan

INTRODUCTION

A consequence of the increased exploration for


gold deposits during the late 1970s and early
1980s was the revision of the classification of
epithermal deposits in order to account for the
variations observed in styles of mineralization and
inferred genetic environments. Among the
numerous classifications that followed, one group
of deposits clearly showed a common set of
features; this deposit type is characterized by the
presence of minerals diagnostic of highsulfidation states (e.g., enargite and luzonite) and
acidic hydrothermal conditions (e.g., alunite,
kaolinite, pyrophyllite). The terms enargite-gold
(Ashley 1982), Goldfield-type (Bethke 1984, after
Ransome 1909), high-sulfur (Bonham 1984,
1986), quartz-alunite Au (Berger 1986), acidsuifate (Heald et al. 1987), and alunite-kaolinite
(Berger & Henley 1989) were applied to this
group in reference to some of its mineralogical or
inferred geochemical attributes. The term highsulfidation (HS) (Hedenquist 1987) is now widely
used; the term was proposed originally to refer to
a fundamental genetic aspect, the relatively
oxidized state of sulfur contained in the
hydrothermal system (i.e., initially S02-rich). This
aspect is significant because it links HS deposits
with one of the two main types of terrestrial
magma-related hydrothermal systems (Henley &
Ellis 1983), those associated with andesitic
volcanoes whose surface manifestation includes
high-temperature fumaroles and acid sulfatechloride hot springs and crater lakes. By contrast,
low-sulfidation deposits form from neutral-pH,
reduced (H2S-rich) hydrothermal fluids similar to
those encountered in geothermal systems (Henley
& Ellis 1983), with surface manifestation

including silica sinter-depositing hot springs and


steam-heated acid-sulfate alteration.
The main objective of this review is to
summarize the characteristics of HS mineralization formed primarily within the epithermal
environment, though recognizing the potential for
HS conditions to occur at greater depths. Earlier
studies have argued for a magmatic fluid
component in HS deposits (e.g., Sillitoe 1983,
1989, 1991a; Hayba et al. 1985; Henley 1991;
White 1991; Rye 1993; Hedenquist et al. 1994a),
and the identification and characterization of HS
deposits has contributed to a re-evaluation of the
role of magmatic fluids in other types of
hydrothermal systems (Hedenquist & Lowertstern
1994; Simmons this volume; de Ronde this
volume). In this context, particular attention is
given to the characteristics that are helpful in
determining the nature of the magmatic contribution to the hydrothermal system through time
and space. This review considers features of many
of the deposits listed in Table 1, with locations
shown in Figure 1, but is based on a selection of
fourteen deposits for which the results of detailed
geological and geochemical studies are available
(Tables 2, and 3). For simplification, bibliographic references are not given in the text for
general deposit features; these references may be
fduh* in Table 1. For regional studies of HS
deposits, particularly with respect to other types of
magmatic-hydrothermai base- and precious-metal
deposits, the reader is referred to reviews by
Heald et al. (1987), Bonham (1989), Sillitoe
(1989, 1991a), Berger & Bonham (1990), Camus
(1990), White & Hedenquist (1990), Mitchell &
Leach (1991), Mitchell (1992), and White et al.
(1995).

419

A. Arribas, Jr.
Table 1. Principal high-sulfldation deposits or documented prospects ordered geographically
Nin
Fig. 1
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50

Deposit

References

Asia & Australasia


Dobroyde, Australia
Whiter al. (1995)
Rhyolite Creek, Australia
Raetz & Partington (1988)
Temora, Australia
Thompson et al. (1986)
Peak Hill, Australia
Cordery (1986), Harbon (1988), Masterman (1994)
ML Kasi, Fiji
Turner (1986)
Wafi River, Papua New Guinea
Leach & Erceg (1990), Erceg et al. (1991)
Nena, Papua New Guinea
Asami & Britten (1980), Hall et al. (1990)
Motomboto, Indonesia
Perelld (1994)
Nalesbitan, Philippines
Sillitoe et al. (1990)
Lepanto, Philippines
Gonzalez (1959), Garcia (1991), Arribas et al. (1995b)
Chinkuashih, Taiwan
Huang (1955), Hwang & Meyer (1982), Tan et al. (1993)
Zijinshan, China
Zhang et al. (1994)
Seongsan & Ogmaesan, South Korea
Yoon (1994)
Nansatsu (Iwato, Akeshi & Kasuga), Japan
Izawa & Cunningham (1989), Hedenquist et al. (1994a)
Yoji, Japan
Yui&Matsueda(1994)
Teine, Japan
Ito (1969)
Akaiwa, Japan
Akamatsu & Yui (1992), Akamatsu (1993)
Mitsumori-Nukeishi, Japan
Aoki & Watanabe (1995)
North & Central America
Northwestern Vancouver Island, Canada
Panteleyev & Koyanagi (1994)
Goldfield, Nevada
Ransome (1907,1909), Ashley (1974), Vikre (1989)
Paradise Peak, Nevada
John et al. (1991), Sillitoe & Lorson (1994)
Summitville, Colorado
Steven & Ratte" (1960), Stoffrcgen (1987), Rye (1993)
Red Mtn-Lake City, Colorado
Bove et al. (1990), Rye (1993)
Red Mtn-Silverton, Colorado
Burbank (1941), Fisher and Leedy (1973)
Mulatos, Mexico
Staude(1994)
Pueblo Viejo, Dominican Republic
Muntean et al. (1990), Russell & Kesler (1991)
South America
Julcani, Peru
Petersen et al. (1977), Deen (1990), Rye (1993)
Castrovirreyna, Peru
Vidal & Cedillo (1988)
Ccarhuarso, Peru
Vidal ef a/. (1989)
San Juan de Lucanas, Peru
Vidal & Cedillo (1988)
Cerro de Pasco, Peru
Graton & Bowditch (1936), Einaudi (1977)
Colquijirca, Peru
Vidal etal. (1984)
Sucuitambo, Peru
Vidal & Cedillo (1988)
Laurani, Bolivia
Murillo et al. (1993)
Choquelimpie, Chile
GiOpper et al. (1991)
Guanaco, Chile
Puig et al. (1988), Cuitifio et al. (1988)
El Hueso, Chile
Sillitoe (1991a)
Esperanza, Chile
Vila (1991), Moscoso et al. (1993), Cuitifio et al. (1994)
La Coipa, Chile
Oviedo et al. (1991), Cecioni & Dick (1992)
Nevada & Sancarron, Chile
Siddeley & Araneda (1990)
El Indio-Tambo, Chile
Siddeley & Araneda (1986), Jannas et al. (1990)
La Mejicana-Nevados del Famatina, Argentina
Losada-Calderon & McPhail (1994)
Europe
Rodalquilar, Spain
SSnger-von Oepen et al. (1989), Arribas et al. (1995a)
Furtei-Serrenti, Sardinia
Ruggieri (1993a,b)
Spahievo, Bulgaria
Velinovrtc/.(1990)
Chelopech, Bulgaria
Bogdanov (1982,1986)
Western Srednogorie region, Bulgaria
Bogdanov (1982), Velinov & Kanazirski (1990)
Bor, Yugoslavia
Jankovic et al. (1980), Jankovic (1982)
Lahoca, Hungary
Baksa (1975,1986), First (1993)
Enasen, Sweden
HaUberg(1994)

High-sulfidation Epithermal Deposits

WT

C^IU-10
Western
i f l P J5V-9 V
*lcjk /S'fy g /
Pacific

--TV?
Figure 1. Worldwide distribution of high-<sulfidatfori deposits and principal documented prospects. The main highsulfidation metallogenic provinces are indicated. See Table 1 for deposit names and selected references.
OPENING REMARKS ON GENETIC
ENVIRONMENT

Based on detailed research of the Summitville


Au-Cu-Ag deposit, Stoffregen (1987) demonstrated that a nearly ubiquitous feature of HS
deposits, fracture-controlled vuggy silica rock
(intensely leached volcanic rock consisting
dominantly of quartz; Fig. 2) is the product of
very acidic conditions (pH <2 at T= ~250 C) that
occur within a sulfate-rich hydrothermal fluid
formed by absorption of magmatic vapor. In
addition to S0 2 disproportionate to H 2 S0 4 ,
significant concentration of HC1 from the
magmatic vapor contributes to the acidic
conditions necessary for alumina to be soluble,
leading to vuggy silica alteration (Hedenquist et
al 1994a,b). Neutralization of the acidic solution
by reaction with the wallrock results in a sequence
of alteration zones, outward from the
hydrothermal conduit, which is indicative of
decreasing acidity and is defined by the presence
of alunite, kaolinite, illite, and montmorillonite
chlorite (Steven & Rattd 1960; Fig. 2).
This same alteration sequence, without the
vuggy silica zone but with enargite-bearing ores,
was documented in the Butte polymetallic deposit
(Meyer et al. 1968) and in the roots of the

advanced argillic zones that commonly cap


porphyry copper systems {e.g., Sillitoe 1973; Corn
1975; Gustafson & Hunt 1975; Koukharsky &
Mirre 1976; Wallace 1979). Indeed, several of the
deposits considered in this review are underlain by
porphyry-type mineralization (Table 2). This
advanced argillic assemblage is also typical Qf
that associated with acidic crater lakes atop active
volcanoes (Christenson & Wood 1993; Delmelle
& Bernard 1994; Rowe 1994; Hedenquist this
volume).
The implications of a genetic relation between
porphyry and epithermal mineralization, e.g., with
respect to the origin of metals or the nature of the
fluid inclusions in HS deposits, are discussed
below. The observation made here is that an
alunite-enargite assemblage records a similar
geochemical environment, whether forming an
epithermal deposit or as part of the alteration
zoning of an orebody formed at greater depths.
High-sulfidation deposits form in a position
intermediate between intrusions and the surface;
therefore, they may be located close to a porphyry
copper deposit or in a near-surface environment,
such as the roots of an acid crater lake.
Comprehensive genetic models for HS
deposits have been proposed only recently (e.g.,
Berger & Henley 1989; Sillitoe 1989; White 1991;

421

A. Arribas, Jr.
Table 2. Main geological characteristics of 14 selected high-sulfidation epithermal deposits
Deposit/district,
location

Age
(Ma)

Motomboto,
Indonesia

1.9

Nalesbitan,
Philippines
Lepanto,
Philippines

Pliocene

Chinkuashih,
Taiwan

1.3-1.0

1.5-1.2

Zijinshan,
China

-94

Nansatsu,
Japan

5-3.5

Summitville,
Colorado

22.5

Metals,
(tonnes)1

Local volcanic
setting

Principal host
rocks

Genetically
related rocks

Time
between host
rock & deposit

Deposit form

<1.0 m.y.

Hbx , veins, dis in


VS

None observed

N/A

Hbx, veinlets

Ands/dac vol,
Miocene + older
volx + metavol

Qtz-diorite
porphyry

<analyt. error
(0.1 m.y.)

Dome complex

Dae vole
Miocene sed

Dacite domes
and flows

Dome along
caldera margin?

Jurassic granite,
Cretaceous dac
porpyhry+pyr
Ands pyr, flows +
volx

Not reported

Vertical breccias,
veins, stratabound
replacements
Veins or "ledges",
hbx, dis and stk
surrounding veins
Veins, hbx, stk

Cu, Au, Ag
60,000 t Cu. 4 i
Au, 180tAg(c)
Au
15 t Au (c)

Central-vent
volcano

Dae dome, ands/dac/rhy Dioritic, qtzdioritic stocks


flows, pyr and volx

Small centralvent volcano

Ands pyr + flows

Cu, Au, Ag
900,000 i Cu,
1201 Au(c)
Au, Cu, Ag
92 t Au, 183 t Ag,
120,000 tCu(p)
Cu, Au
>10 t Au (c)

Diatreme
complex

Small volcanos
Au
18 t Au (p) + 18 t in a caldera?
Au reserves
Au, Cu, Ag
Dome along
17lAu
preexisting
caldera margin

7.0 m.y.(?)
(poorly-dated)

HorWende ands <0.5 my


(Middle Voles)

Qu-lalite porphyry

Qtz-monzonite
porphyry

<analyt. error
(0.5 m.y.)

Dis in stratabound
VS/MS bodies,
veins, hbx
"Ledges" with
veins, hbx + dis
inVS

21

Au (Ag, Cu)
Domes along
130 t Au, t 43 Ag, preexisting ring
37,000 Cu (p)
fracture

Miocene andesite

Andesite

<analyt. error
(0.4 m.y.)

"Ledges" with
veins, hbx + dis
in MS

Paradise Peak,
Nevada

19-18

Composite welded tuff,


volx + ands flows

And/dacvd

<analyt. error
(1.0 m.y.)

Pueblo Viejo,
Dominican Rep.

-130

Au, Ag, Hg
Within or close
47 t Au, 1255 Ag to a central-vent
457 l Hg (p)
volcano
Au, Ag
Maar-diatreme
>600 t Au (p;
complex
Sillitoe. 1993)
Ag, Cu, Pb, Au,
Dome complex
W, Bi, Zn
around a central
diatreme
Au, Ag, Cu
Stratovo!cano(?)
-140 t Au,
in earlier caldera
-1,100 t A g ( c )
Dome complex(?)
Cu, Au, Ag
>10-15tAu(c)

Maar sed + basaltic


vol (spilite)

CAbimodal
N/A
(Rhy + basalt)
volcanic suite
Dac/rhyodacitic <analyt. error
porphyry
(0.1 m.y.)

Stratabound bodies
commonly with
hbx
Mushroom-shaped
bodies with stk +
dis

Goldfield,
Nevada

Julcani,
Peru

9.8

El Indio,
Chile

13-8

La Mejicana & Nevados del Famatina,


Argentina
Rodalquilar,
Spain

4.0-3.6

11-10

Au
10 l Au (p)

Caldera margin

Dactorhyodacitic
domes and tuffs

Veins

Dac, rhy pyr;


dac + ands vol

CA vol

N/A

Veins + stk

Paleozoic seds +
granites. Pliocene
intrusive dacite
Ands to rhy pyr flows,
collapse bxs + domes

Dac/rhyodaoic
porphyry
stocks
Ands flows
+ dykes

<1.2 m.y.

Veins; also hbx at


N. del Famtina

<analyt. error
(0.7 m.y.)

Veins, hbx, dis in


VS

Abbreviations used:
CA calc-alkaline, MS = massive silica, VS - vuggy silica, ands - andesitic, bre - b r e c c t a ^ 5 n L 2 r f c k 'stk
Abbrev.at.ons
usea: w
disseminations, hbx - hydrothermal vein breccia or breccia pipes, pyr = pyroclastics, qtz - quartz, rhy - rhyol.t,c, sed - segmentary rock, stk
= stockwork, vol volcanic rock (unspecified), volx, volcaniclastics
IVIVUillVIWUVd
' (p) produced, (c) estimated total contained 1 Approximate number, quoted from paper or estimated from figures: 150 m for Paradise
Peak is for individual orebodies

Giggenbach 1992a; Rye 1993; Hedenquist et al.


1994a). However, the basic genetic controls, as we
understand them now, were formulated almost
ninety years ago by Ransome (1907) following his
classic study of the Goldfield Au-Ag-Cu deposit.
In his own words "the [ore depositing] solutions
were essentially emanations from a solidifying
body ofdacitic magma" and " . . the initially acid
emanations would be neutralized and modified in
their ascent through fissured rock . .by the
distance and kind of rock traversed, the quantity
and character of admixed surface-derived waters,

and the pressure and temperature gradients". This


concept formed the basis for Ransome's "direct
volcanic hypothesis", though it was quickly
abandoned in favor of a "simultaneous solfatarism
and oxidation" model (Ransome 1909). The
change in genetic interpretation has more than
anecdotal value because it illustrates the source of
a not-uncommon misconception on the environment of mineralization of epithermal deposits.
The crucial aspect is identification of the
origin of alunite or acid-sulfate alteration, which
can be generated by different mechanisms in three

High-sulfidation Epithermal Deposits


Table 2 (continued)
Deposit/district,
location

Control on mineralization

Vertical extent of epith


ore(m)2

Relation to
porphyry system

References

Motomboto,
Indonesia

Contact between dome and


volcanic rock, steep fault

250

Porphyry Cu-Au
prospects nearby, age
within 1.0 m.y.

Perell6(l94)

Nalesbitan,
Philippines
Lepanio,
Philippines

Steep strike-slip fault

ISO

Proposed,
none known

Sillitoerta/. (1990)

Major steep + minor faults,


diatreme contact, unconformity, permeable layers

500

Above + adjacent
same age porphyry
Cu-Au deposit

Garcia (1991),
Arribasrta/. (1995b)

Chinkuashih,
Taiwan

Steep normal faults +


their intersections,
bedding planes

800

None known

Huang (1955),
Tan etal. (1993)

Zijinshan,
China

600(7)

None known

Ren era/. (1992),


Zhang etal. (1994)

Nansatsu,
Japan

Steep strike-slip fault


zones + contact of
volcanic vent
Steep fractures + permeable
pyroclastic layers

<150

None known

Izawa & Cunningham (1989),


Hedenquist etal. (1994a)

Summitville,
Colorado

Steep radial fractures +


dome contact

250

Intrusion-centered
sericitic, low grade
stk mineralization

Steven & Ratti (1960), Menhert


etal (1973), Stoffregen (1987),
Rye (1993) Gray & Coolbaugh
(1994)

Goldfield,
Nevada

Moderately + shallow
dipping faults & fissures

400

None known

Paradise Peak,
Nevada

Steep faults and permeable


pyroclastic layers

<150

Ransome (1909). Ashley (1974),


Ashley & Silberman (1976),
Vikre (1989, written commun.
1995)
John et al. (1991),
Sillitoe & Lorson (1994)

Pueblo Viejo,
Dominican Rep.

Diatreme ring fault +


permeable layers

400(7)

Julcani,

Steep fractures

600

Sericitic, stk Au
mineralization (East
Zone)
None known

None known

Pem
El Indio,
Chile

Steep normal faults

La Mejicana & Nevados del Famatina,


Argentina
Rodalquilar,
Spain

Local faults

Caldera ring faults +


normal local faults

>300

<150

Porphyry Cu-Mo
mineralization
nearby

Russell &Kesler (1991),


Mumeanero/. (1990)
Petersen etal. (1977),
Noble & Silberman (1984),
Deen(1990)
Siddeley & Araneda (1986),
Jannasai(1990)

Losada-Calderon & McPhail


HS ore at Nevado del
Famatina is a part of a (1994), Losada-Calderon et al.
(1994)
porphyry Cu prospect
Arribas etal. (1995a)
None known

principal geologic environments (Bethke 1984;


Rye et al. 1992): (1) by the disproportionation of
magmatic SOj to H2S04 and H2S following
absorption
by
groundwater
(magmatichydrothermal), (2) by atmospheric oxidation of
H2S in the vadose zone over the water table,
associated with fumarolic discharge of vapor
released by deeper boiling fluids (steam-heated),
and (3) by atmospheric oxidation of sulfides
during weathering (supergene). Magmatichydrothermal alunite occurs with minerals such as
diaspore, pyrophyllite, kaolinite, dickite, and

zunyite, which are typical of hypogene (T = 200350 C) acidic conditions (advanced argillic
assemblage; Meyer & Hemley 1967). This type of
alunite is characteristic of HS deposits, but it may
also appear in areas of advanced argillic alteration
void of ore mineralization (e.g., Iwao 1962; Hall
1978). Alunite in steam-heated environments
forms with kaolinite and interlayered illitesmectite at about 100 to 160 C where fumarolic
vapor condenses above the boiling zone of
neutral-pH, H2S-rich fluid, typical of geothermal
systems that form low-sulfidation deposits.

423

A. Arribas, Jr.
Vuggy silica
Quartz alunite
PropylWc

Chlorite-rich
rock

100 m

Montmorillonite-rich
rock

Figure 2. Cross-section of alteration zones characteristic of high-sulfidation. deposits,. as observed at the


Summitville Au-Cu deposit, Colorado. Diagramat left (simplified from Steven & Ratte 1960) shows schematic
outward zonation from a subvertical mineralized body, shown at right (from Stoffregren 1987).
Because of the relatively shallow and dynamic
environment of mineralization, overprinting
among the three types of acid-sulfate alteration
(including supergene) is possible; however, the
spatial relation of each type of alunite to ore is
different, and correct identification is important
for exploration (Rye el al. 1992; White &
Hedenquist 1995).
DISTRIBUTION, A G E AND ECONOMIC
SIGNIFICANCE

In
common
with
other
magmatichydrothermal deposits (e.g., porphyry copper
deposits), HS deposits coincide worldwide with
plutonic-volcanic arcs. This association is best
observed in the Cenozoic deposits of the CircumPacific and-the Balkan belt of southeastern Europe
(Fig. 1). These deposits occur in two main
settings: in island arcs and at continental margins.
The tectonic regime during formation of the
deposits seems to be dominantly extensional
(Sillitoe 1993). Some deposits (e.g., Goldfield,
Rodalquilar, Summitville) formed in intracontinental regions during periods of extension
that followed regional compression and subduction by several m.y.
Tertiary HS deposits predominate, and only a
few deposits are Mesozoic (e.g., Pueblo Viejo,
Zijinshan), Paleozoic (e.g., Temora and others in
southeastern Australia), or PreCambrian (the early

Proterozoic Enasen Au deposit located in the


Baltic shield of central Sweden; Fig. 1). The
youngest deposits are Pleistocene (<1.6 Ma) and
occur in the central western Pacific (Kelly,
Lepanto, and Chinkuashih). The concentration of
deposits in young volcanic areas is mainly a
reflection of the fact that older HS deposits are
more likely to be eroded.
Gold, copper, and variable amounts of silver
are the main products of HS deposits (Table 2).
Gold (Nalesbitan, Rodalquilar), occasionally with
silica by-product (Nansatsu), is the only economic
metal in the smaller deposits. No copper is
produced at Paradise Peak and Pueblo Viejo.
Mercury is produced at Paradise Peak, and the
Julcani district has been a source of a remarkable
polymetallic assemblage consisting of Ag, Cu, Pb,
Au, W, Bi, and Zn (Table 2). The six largest
deposits or districts (Chinkuashih, El Indio,
Goldfield, La Coipa, Lepanto, and Pueblo Viejo)
each contains more than about 100 tonnes of gold.
The economic potential of this type of
mineralization is clear in regions such as the
Chilean Andes (Sillitoe 1991a).
VOLCANIC SETTING AND ASSOCIATED
IGNEOUS ROCKS

The high-sulfidation deposits considered in


Table 2 occur within intermediate-composition
volcanic rock sequences having ages broadly

High-sulfidation Epithermal Deposits


Figure 3. K 2 0 versus Si02 variation diagram
for rocks thought to be genetically related to
high-sulfidation deposits. The samples from
12 deposits or districts ( = 140) define a
small compositional field, which contrasts
sharply with the large field defined by
volcanic rocks associated with lowsulfidation or intrusion-related Au deposits
(>100 samples from 16 districts; Sillitoe
O
1991b, 1993; Muller & Groves 1993). The
degree of alteration of the rock samples and
precision of the analytical data are largely
unknown; however, according to the
individual data sources, most of the samples
are unaltered or very weakly altered. Circles
indicate average values for each highsulfidation deposit or district: Ch =
Chinkuashih, Cq = Choquelimpie, Go =
50
70
80
60
Goldfield, In = El Indio, Ju = Julcani, La =
SCv, (wt%)
Laurani, Le = Lepanto, Mo = Motomboto,
Na = Nansatsu, PP = Paradise Peak, Ro = Rodalquilar, Su = Summitville. Compositional fields after Keith et al.
(1991). See Appendix for references and information on data plotted.
CM

similar to that of mineralization. Where abundant


radiometric ages are available, the age of the host
rocks and the age of mineralization are within
analytical precision; where a difference is
indicated, it is typically less than ~1.0 m.y. (Table
2). A common spatial association exists between
the deposits and shallow, typically porphyritic
intrusions. These intrusions are interpreted to be
the roots of volcanic domes or the feeders of
central-vent volcanoes or maar-diatreme complexes, the three main volcanic settings for HS
deposits (Table 2). Some deposits are hosted
entirely within a single dome (Summitville), or
within a dome complex (Julcani). In most cases
the mineralization extends from the subvolcanic
intrusion into country rocks, such as the Main
Vein Cu-Au-Ag deposit and associated breccia
deposits in the Penshan area of the Chinkuashih
district. Some deposits, however, do not show any
(known) spatial association with subvolcanic
intrusions thought to be genetically related to
mineralization (e.g., Nalesbitan, Nansatsu). In the
Rodalquilar Au deposit, dykes and small
intrusions of hornblende andesite which are
interpreted to be temporally related to the
mineralization represent only a fraction of the
altered and mineralized area exposed at the
present depth of erosion; a larger intrusive body is

thought to exist at depth (Arribas et al. 1995a).


The main control on location of mineralization at
Rodalquilar is the structural margin of two nested,
resurgent calderas. With the exception of
Rodalquilar, the role of calderas in the formation
of HS deposits seems to be limited to facilitating
the emplacement of late intrusive magma along
preexisting caldera ring-fractures (Rytuba et al.
1990).
The magmas thought to be genetically related
to HS deposits have a remarkably limited
compositional variation. The ranges of wt.% K 2 0
and SiC2 for twelve deposits overlap greatly and
show a dominance of calc-alkaline andesitic and
dacitic compositions, with subordinate rhyolite
(Fig. 3). Intermediate calcic volcanic rocks are
limited to porphyritic intrusions in the Lepanto
and Motomboto Cu-Au-Ag districts, and
intermediate-to-felsic alkali-calcic rocks are
characteristic of the Summitville and Laurani
districts (Fig. 3). Interestingly, no deposits have
been discovered in association with alkaline or
mafic magmas, even though these magmas can be
genetically related to low-sulfidation and
intrusion-related Au deposits (Sillitoe 1991b,
1993; Muller & Groves 1993; Richards this
volume). The data shown in Figure 3 suggest a
relation exists between magma composition and
425

A. Arribas, Jr.
Table 3. Main alteration and mineralization characteristics of 14 selected high-sulfidation
epithermal deposits

Deposit

Lateral alteration zoning


(outward from mineralized bodies)

Venical alteration
zoning
(shallow to deep)

Ore
mincratizaUon
Princial ore minerals

in:

Ag/Au

Motomboto

VS-qtz-alu-*qtz-kao-
kao-sme-Mll-chl

VS/MS-qtz-alu-*qtzkao-*ill-kao-*cbl

Py, ena-luz, mar, sph, gal, lentet, are, cpy, arg, nat.Au, tell

Silica core

Nalesbitan

Silicified Hbx-*qtz-kaoalu-ill-sme-chl-cal

Silicified Hbx-*qtz-kaoalu- ill-sme-chl-cal

Py, chalc.qtz, ceo, bor, cov,


ena, tell

Silica core

Very
low Ag

Lepanto

VS/MS-* tz-alu-kao-*
kao-qtz-ill-*chl-ill

MSA'S-* AA-SER-
(K-silicate in subjacent
FSE porphyry copper)

Silica core

Chinkuashih

VS/MS-*qtz-alu-kao-
ill-chl-kao

Silica core

Zijinshan

VS/MS-*qtz-dic-alu-qtzdic-ser-qtz-scr

Ena-luz, py, ten-let, cpy, py, ele.


sph, gal, mar, sele, tell, Snbearing sulf
py, ena-luz, fam, ten-let, nat.Au,
ele, bar, naLHg, tell, sph, gal,
cpy, geo, bou
VS/MS-* qtz-dic-lu- qtz- py, dig, ena, cov, mol, naLAu
cpy, bor, let-ten, gal, sph
dic-ser-qtz-ser

Silica core

N/A

Nansatsu

VS/MS-* alu-dic-pyo-*
ill-kao-sme-*
PRO

VS/MS-*alu-*dic-serpy-ser-chl-PRO

Silica core

<2

Summitville

VS(MSh*qtz-alu--
qtz-kao-*kao-iU-*
sme-chl

VS(MS)-qtz-kaoalu-*qtz-kao-*SER

ena-luz. py, ele, nat.Au, arg,


pyr, cpy, bor, sph, gal, cas, sta,
mol, can
py, ena-luz, cov, mar, naLS,
nat.Au, sph, gal, bar, cpy, ten

Silica core

2-20

Goldfield

MS{VS)-*qtz-alu-kao-*
iII-sme-
PRO

MS>VS-*qtz-alukao-*qtz-kaopyo

Silica core

<1

Paradise Peak

Vertical (due to deposit


style): MS(VS)-*
qtz-alu-kao- sme-chl

MS(VSHqtz-alu-kao
(SER in faulted, deeperC?)
East Zone deposit)

Silica core

10-30

Pueblo Viejo

Complex + overprinted

Early: Kao-py-qtz-*
alu-py-qtz
Late: MS-pyo-dia

Julcani

Pre-ore:VS/MS-qrz-alu-

Ellndio
La Mejicana,
Nevadosdel
Famatina
Rodalquilar

py, sph, ena, nal.Au, nat.S, bar,


len-tet, fam, gal, bar, stb, ele,
sele, tetl, Bi- Pb- Ag- sulf
py, wol, cas, nat.Au, ena, luz,

InAA +
MS zones
Veins

470

tet, ten, cpy, gal, sph, bar, sid,


Pb- Bi- Ag-bearing sulf

kao~*qtz-kao; Syn-ore:
qtz-pyo-py-qtz- kao-py-<
qtz-scr-py qtz-kao-sme
Cu stage veins-* kao-aluser-qlz; Au stage
veins-* ser-kao-pyo-qtz

VS/MS-*qtz-alu-lcao-
qtz-kao-ill-* ill-sme-chl

py, fam, ten-let, bis, got,


naLAu, ena-luz, bar, tell, sph,
cov
bar, stb, bis, nal.Au, mar, py,
nat.S, cin, sph, gal, cpy, ars,
let, arg, cov, fam

35-45

Alu-kao-*qtz-ser- (K-silicate in N. del Famatina


porphyry copper)
VS/MS-qtz-alu-kao-*
qtz-kao-ser-qtz-ser-py

Ena, py, tel, nat.Au, ten, cpy,


gal, sph, hue, bor, dig, emp,
ceo, mar, bar,
py, ena, cpy, sph, ten-let, cov,
ceo, fam, luz, naLAu, gal, mol,
ele, tell, col Sn-Bi-Pb-Ag-sulf
Py, naLAu, ena, tell, cas, col.
cov, dig, bor, gal, sph, Bi- sulf

Veins
Silica core

10-20

Silica core

<1

Abbreviations used: AA - advanced argillic, Hbx - hydrotherma! breccia, MS = massive silica, PRO Propylitic, SER =
sericitic, VS - vuggy silica, VS (MS) vuggy silica dominant, alu - alunite, ars = arsenoyrite, bar = barite, bis =
bismuthinite, bor = bomite, bou = boumonite, cal - calcite, ceo = chalcocite, chal.qtz = chalcedony or chalccdonic quartz,
chl chlorite, cin cinnabar, can = canfieldite, cas cassiterite, col colusite, cov = covellite, cpy = chalcopyrite, die =
dickite, dig = digenite, ele electrum, emp emplectite, fam - famatinite (stibioluzonite), gal - galena, geo geocronite,
gol = goldfieldite, hue = hubnerite, ill - illite, kao e kaolinite, luz - luzonite, mar = marcasite, mol = molybdenite, nat.Au =
native gold, natS - native sulfur, nat.Te - native tellurium, oro orpimcnt, py = pyrite, pyo = pyrophyllite, qtz quartz,
rea = realgar, sele - selenides, ser sericite, sid = siderite, sme - smectite, sph = sphalerite, sta stannite, stb stibnite,
sulf- sulfides or sulfosalts, tell * tellurides, ten ~ tennantite, tet - tetrahedrite, tou tourmaline, wol - wolframite
' Based on fluid-inclusion (flinc) or geological (geol) evidence; blank where not specified.
2
Boiling (Hbx) -= boiling due to abrupt pressure reduction associated with hydrotherma! brecciation

426

High-sulfidation Epithermal Deposits

Table 3 (continued)

Deposit

N and order of
main mineral
events

Inferred
mineralization
depth (m)1

Supergene oxidation/
secondary Au
enrichment?

Inferred oreforming
mechanism

References

Motomboto

2
AA-> sulfide-Au

Unknown

Unknown

Irregular to 100 m

Perello(1994)

Nalesbitan

2
Py-qtz-Cu-Au

300-500
flinc

Boiling (Hbx)2

Complete to 130
m; yes

Sillitoe el al. (1990)

Lepanto

3
AA-CuAu
-Au
2
AA-Cu-Au +
late bar-Au
SeveraK?),
Qtz-ser-dic-*qtzalu-dic- sulfide

Unknown

Mixing/cooling

Not important

Garcia (1991), Claveria &


Hedenquist(1994)

Chinkuashih
Zijinshan

500
Unknown

Mixing/boiling

150-300
flinc

Mixing/cooling

Nansatsu

2
AA-*Cu-Au

Summitville

3
400-500
AA-CuAu-+
flinc + geol
bar-base-metal-Au
5+
100-300+
AA-ena-Augeol
ten-py-Bi-Tc-
3
Unknown
AA-Au-Ag-+
Hg
2
Shallow; lacusAlu-kao-py-Au- trine sediments
MS-pyo-dia-Au
preserved
Several
200-300
AA+VS-tou
breccias-* sulfide
veins (main, late)
2
AA+Cu- Au

Goldfield

Paradise Peak

Pueblo Viejo
Julcani

El Indio

La Mejicana,
Nevadosdel
Famatina
Rodalquilar

2
AA-silica-py-Au

Unknown

200-300
flinc + geol

Huang (1955),
Tan el al. (1993)
Important in upper Ren el al. (1992),
Zhang et al. (1994)
250 nv, yes

Widespread to 100
nv, yes but may
be steam-heated
Cu-Au by
Irregular to 100 m;
mixing, bar-Au
may be partly
by oxidation
steam-heated
Mixing/cooling; Widespread to 80
oxidation
nv, supergene
alunite(-lOMa)
Boiling, with
Widespread to 250
Hbx in early Au- m; supergene aluAg stage
nite (100.5 Ma)
Sulfidation +
Widespread to
boiling
-100 m
Mixing + boiling

Boiling (Hbx)2 + Widespread to 80


mixing/cooling
m; supergene
alunite (4-3 Ma)

DEPOSIT FORM AffD CONTROL:


CLASSIFICATIONS

High-sulfidation deposits display a wide


variety of styles of mineralization that includes
veins, hydrothermal breccia bodies, stockworks,

Steven & Ratte (I960,)


Stoffregen (1987), Gray &
Coolbaugh(1994)
Ransome(1907, 1909).
Ashley (1974), Vikre (1989,
written comm., 1995)
John eial. (1991),
Sillitoe & Lorson (1994)
Keslerai(1981)
Russell AKesler (1991).
Munteane/a/. (1990)
Deen (1990), Rye (1993)

Siddeley & Araneda (1986),


Jannasrtai(1990)

Mixing/cooling

development of the oxidized and reactive


magmatic vapor plume that is thought crucial to
the formation of HS deposits.

Izawa & Cunningham (1989),


Hedenquisl el al. (1994a)

Losada-Calderon & McPbail.


(1994), Brodtkorb & Paar
(1993)
SSInger-von Oepen et al.
(1989)Arribasfl/.
(1995a)

and disseminations or replacements. This


variation in the structure of the orebodies is
complemented with variations in other deposit
features, including ore and alteration mineralogy,
paragenesis, and metal ratios (Tables 2, 3). In
addition, some deposits present complex relations
which may be composite, e.g., between high- and
low-sulfidation mineralization (e.g., quartz-Austage veins at El Indio, and some of the veins at
Julcani). Definition of styles of HS mineralization

427

A. Arribas, Jr.
is difficult, but useful for discussion of the
differences among deposits and design of
exploration strategies. In this context, White
(1991) distinguished three end-member styles of
HS deposits, named after deposits of the CircumPacific: Temora, El Indio, and Nansatsu. Irregular
bodies of disseminated, silicified ores dominate in
the Temora-style. Cavity-filling veins with
sericitic and clay-rich haloes are characteristic of
El Indio-style Au deposits. A large group of
deposits falls into White's (1991) Nansatsu-style,
which is characterized by wallrock-alteration
zoning similar to that shown in Figure 2, and by
the occurrence of enargite-bearing ores within a
silica core consisting of vuggy or massive silica
rock (Table 3). Mineralization in this style of
deposit forms irregular stratabound bodies (e.g.,
Nansatsu, Lepanto) or subvertical vein-like
masses or "ledges" {e.g., Chinkuashih, Goldfield,
Lepanto, Rodalquilar, Summitville). These
deposits contain breccia bodies, veins, stockworks
of small veins, and disseminated ores that replace
or impregnate intensely altered country rock.
Ericksen & Cunningham (1993) distinguished two
styles of HS deposits in the Andean province: Agand Au-rich polymetallic base-metal veins, and
low-grade vuggy silica and breccias; the two types
are broadly comparable with El Indio- and
Nansatsu-styles, respectively.
Local subvertical faults and fractures are the
dominant control on HS mineralization and they
are present in most deposits (Table 2). Other
examples of structural controls observed in some
districts among the fourteen selected include:
moderately to shallow-dipping faults (Goldfield),
caldera ring and radial faults (Rodalquilar), the
dilational jog of a strike-slip fault (Nalesbitan),
diatreme ring-faults (Lepanto, Pueblo Viejo), the
contact between a dome or volcanic conduit and
country rock (Motomboto, the Missionary
orebody at Summitville), and a lithologic
unconformity (Pueblo Viejo, Lepanto). In three of
the fourteen deposits, the principal control is
lithological (maar sediments at Pueblo Viejo, and
interbedded pyroclastic layers at Paradise Peak
and Nansatsu; Table 2).
A unique combination of the structural and
lithological controls characteristic of HS deposits
is exhibited by the Lepanto Cu-Au-Ag deposit.
/1">0

The deposit is 3 km long and consists of a main


zone of breccia and replacement mineralization
along the Lepanto Fault (Fig. 4A). Multiple veins
associated with smaller diagonal faults branch
from the main zone and extend into both the
hanging wall and footwall (Garcia 1991). The
characteristic mushroom-shaped cross-section of
many of the orebodies at Lepanto is related tc the
intersection of the steeply dipping Lepanto fault
and branch veins with the unconformity at the
base of Imbanguila dacite (Fig. 4B). Lithologic
variations in the host rocks also played an
important role in the formation of the deposit, as
shown by lenses of stratiform enargite-luzonite
ore which resulted from replacement of detrital
layers within volcaniclastic and sedimentary
basement units (Garcia 1991).
ALTERATION MINERALOGY AND ZONING

As mentioned above, the lateral alteration


zoning that is characteristic of HS deposits
reflects the reaction and neutralization of hightemperature acidic fluids with wallrock. The
innermost zone of vuggy or massive silica
alteration commonly has sharp boundaries with a
zone that may contain quartz, alunite, kaolinite,
dickite, pyrophyllite, diaspore, and zunyite . This
advanced argillic assemblage grades into a second
envelope of argillic alteration, composed of
minerals such as quartz, kaolinite, illite, sericite,
and smectite, and an outermost halo of propylitic
alteration, with chlorite, illite, smectite, and
carbonate (Fig. 2, Table 3). The width of each
zone varies widely; for example, vuggy silica and
advanced argillically altered rock form narrow
(<70 cm) vein selvages at Julcani (Deen 1990),
but form wide (>50 m) rock bodies at Summitville
or Lepanto (Figs. 2 and 4). Late-stage, cavityfilling planar veins at Julcani and El Indio may
extend outside the zone of alunite-kaolinite. In the
majority of HS deposits, however, most of the ore
is contained within the silica core, inside the
advanced argillic envelope (Table 3).

In Russian and eastern European terminology these rocks are


commonly termed 'metasomatic quartzites', with more specific
names such as porous quartzites, diaspore quartzites, alunite
quartzites, and dickite quartzites (e.g., Velinov et al. 1990).

High-sulfidation Epithermal Deposits

h>B
Lepanto fault

1.56-1.17 I [
1.45-1.22 E%3
1.2-0.9
(ill
22-1.8
Cre.-Mto.

)n/mineralizatJon

Ore deposits and lithology

Age (Ma)

High suttkJation Cu-Au ore


Porphyry Cu-Au ore
Post-mineralization cover

t3
rIM

Ouartz-oloiUe porphyry
Dacte porphyry & pyroclastics

Basement rocks "

m
Stratabound f ^ ,

m
raaftHi

Jea level

-700m

JSWfl

pgg
BSa
RS|9
^ ^
f=g

Vuggy silica/
massive silica
Advanced
argiHic
ArgiHic

1 Propyiitic

Figure 4. Longitudinal (A) and transverse (B) cross-sections of the Lepanto-FSE Cu-Au-Ag deposits (Philippines),
showing structural and Uthologic controls on formation of the high-sulfidation and porphyry-type ores (simplified
from Garcia 1991). Potassium-argon dating of country rocks and alteration minerals associated with the porphyry and
high-sulfidation deposits indicates that hydrothermal Cu-Au mineralization took place in the middle of a Pliocene to
Pleistocene event of dacitic-andesitic magmatism (Arribas et al. 1995b). Note the overall spatial overlap of the
magmatic and hydrothermal "plumbing" systems (i,e., volcanic vents of Pliocene dacite, quartz diorite intrusions,
porphyry deposit, and deeper parts of epithermal mineralization).
The zones of alteration with increasing depth
typically grade from a shallow silicic zone
through advanced argillic, argillic, argiHic/
sericitic, into a sericitic or phyllic zone with
quartz, sericite, and pyrite. This alteration
sequence occurs over a vertical interval that
ranges from a few hundred meters to more than
1000 m, and has been best documented by deep
drillholes in the deposits of smaller size, in which
the vertical span of mineralization is less than
about 300 m (e.g., Rodalquilar, Summitville; Fig.
5B). At Lepanto, sericitic alteration at depths of
400 to 500 m below the epithermal deposit gives
way, laterally towards the south, to K-silicate
alteration of the FSE porphyry Cu-Au deposit.
Porphyry-type stockwork mineralization at
Paradise Peak is contained within the sericitic ores
of the East Zone deposit which, according to
Sillitoe & Lorson (1994), formed underneath the
main HS orebodies in the area. A quartz-sericitepyrite zone with trace amounts of chalcopyrite and
molybdenite surrounds an intrusion of monzonite
porphyry 300 m below the HS deposit at
Summitville (Gray & Coolbaugh 1994).

The lateral and vertical alteration zones


described above correspond to a generalized
model. They are useful in exploration because
they help in understanding the genetic environment of a deposit and provide spatial "markers"
within the extinct hydrothermal
system.
Experimental data on the relative stability of
minerals such as alunite, kaolinite, pyrophyllite,
and diaspore (Hemley et al. 1969, 1980), coupled
with the temperature ranges noted for these and
other related acid minerals in active systems
(Reyes 1990; Reyes et al. 1993), also provide
information that contributes to definition of the
paleoconduits in extinct systems.
If studied in detailed, several superimposed
and crosscutting stages of pervasive as well as
fracture (conduit)-related mineralization may be
recognized in the majority of deposits. These are
the expected result of variations, during the course
of mineralization, in temperature, pressure, and
composition of the hydrothermal fluid and the
degree of wallrock interaction. Detailed field and
petrographic studies at the Monte Negro orebody
in the Pueblo Viejo deposit have resulted in

429

A. Arribas, Jr.

Rodalqultar caldg--r?
margin

^<^^"''

^KmSmm
iominacakieraV
ftt^ls)/
^Un
s, margin
N
Mm^7
rffflsRSm
^^V^
\
jHodafottf/ar
i Vuggy silica
"^ (||l
,
r = j Advanced argilltc
**V.
_
ip^lttjr
TO3 Argillic
^^^5i2Sii5S>liS*i^^^^^\.._
f;||flp-::f
LA*I Serkatic
p l l Propylitic
GOOD Intense supergene actd-sulfate overprint

Au-{Cu-TeSn) Nghsutfidation deposits

400 .

lilHi

0 ,

-400

Lower limit of
sulfide oxidation

:ij|HrA I
X
vf* %AA A A
A /

JCSJSSw

A A

ft. A A X.
A A A * A A A J
, A A A * A 4* * ,
A A A /
[A
A A *

Pb-Zn-{Cu-/
quartz veins

Elevation

Figure 5. Generalized surface alteration map (A) and cross-section (B) of the Rodalquilar
HS deposit in the Rodalquilar and Lomilla calderas, southeastern Spain (from Arribas et
al. 1995a). The boundaries shown between alteration zones are irregular and gradational.
identification of two stages of mineralization,
interpreted to correspond to two distinct magmatic
pulses (Muntean et al. 1990). During the first
stage (responsible for ~60% of the Au in the
deposit), shallow kaolinite-quartz-pyrite and deep
alunite-quartz-pyrite-quartz zones were developed, with gold mineralization in association
with disseminated pyrite in the wallrock; during
the second stage (responsible for about 40% of the
Au), an extensive zone of siliciflcation with pyrite
sphalerite enargite veins formed at shallow
levels, above a zone of pyrophyllite-diaspore
alteration (Muntean et al. 1990).
O R E AND GANGUE MINERALOGY, AND
TIMING O F MINERALIZATION

White et al. (1995) and White & Hedenquist


(1995) presented detailed discussions on various
aspects of epithermal gold mineralization on the
basis of observations from a large number of
deposits around the Pacific; their conclusions with
respect to ore and gangue mineralogy in HS
deposits are included here, in addition to the

430

particular features of the deposits listed in Table


3. Pyrite and enargite (and its low-temperature
dimorph luzonite) are the dominant sulfides in HS
deposits; pyrite is abundant but the amount of
enargite and luzonite is variable. Common ore
minerals, listed by decreasing abundance from
variable to very minor, include tennantitetetrahedrite, covellite, native gold and argentian
gold (electrum), marcasite, chalcopyrite, sphalerite, and galena. Famatinite is locally abundant
in some deposits (Goldfield, La Mejicana). Sparse
ore minerals include bornite, cassiterite, cinnabar,
molybdenite, orpiment, realgar, stibnite, and
wolframite (the last locally important at Julcani).
Other minerals present in minor amounts in
several deposits include Pb-, Ag-Pb, Bi- and Snbearing sulfosalts (Table 3).
Fine-grained quartz is the dominant gangue in
HS deposits. Other common but minor gangue
minerals include barite, kaolinite, alunite,
pyrophyllite, diaspore, and Ca-,Sr-, Pb- and REEbearing phosphate-sulfate mineral(s) such as
svanbergite-woodhouseite or crandallite (Stoffregen & Alpers 1987). For example, high-grade

High-sulfidation Epithermal Deposits

vein specimens from Chinkuashih, Goldfield, and


La Mejicana have spectacular intergrowths of ore
minerals with kaolinite, alunite, or pyrophyllite.
This observation implies that ore formation
occurred under moderately acidic to acidic
conditions, which are inconsistent with transport
of Au as Au(HS)2" (Seward 1973). Recent
studies of Au solubility in high-temperature acid
sulfide solutions have resulted in identification of
AuHS as one of the principal gold complexes in
HS mineralization (Bening & Seward 1994), the
other possibility being AuCl2~(e.g., Hedenquist
et al. 1994a).
The number and order of mineralizing events
provide critical information for reconstruction of
the hydrothermal system that results in HS
mineralization. A minimum of two stages of
alteration/mineralization has been recognized in
most deposits on the basis of crosscutting
relations (Table 3). The most common evolution
is from an early leaching and alteration stage to a
later ore-forming stage. Vuggy silica rock and the
advanced argillic assemblage with disseminated
pyrite form typically early-stage acidic alteration,
and are followed by Cu Au Ag deposition.
Detailed studies in some districts (e.g., El Indio,
Lepanto), however, have resulted in identification
of two metal stages, an early Cu-rich, Au-poor
stage, dominated by enargite-luzonite, and a late
Au-rich, Cu-poor stage, associated with
intermediate-sulfidation-state sulfides such as
tennantite-tetrahedrite and chalcopyrite, and
tellurides. The transition from quartz-alunitepyrite alteration to enargite-pyrite and finally to
tennantite-tetrahedrite, the last typically without
sulfate (alunite) but with quartz-sericite gangue
and wallrock alteration, indicates a fluid
progressively more reduced and less acid. At
Summitville and Chinkuashih (also Tambo and
Furtei-Serrenti; Table 1), a late stage of baritegold has been documented.
CHARACTERISTICS AND SOURCES OF
HYDROTHERMAL FLUIDS

Results of recent detailed fluid-inclusion and


stable-isotopic studies reveal much about the
composition, temperature and sources of
hydrothermal fluids in HS deposits. Combination

of these data with geological and mineralogical


observations mentioned above allows the nature
of the altering and ore-forming fluids to be
determined. The framework for the interpretation
has benefited from information on the composition and fluxes of volcanic discharges and active
magmatic-hydrothermal systems (Hedenquist &
Lowenstern 1994; Giggenbach this volume;
Hedenquist this volume).
Fluid-inclusion Evidence
Suitable hosts for fluid-inclusion studies are
scarce in HS deposits, as the gangue minerals are
typically fine-grained and even millimeter-size
hydrothermal quartz crystals are usually late stage
and vug-filling. Satisfactory results are obtained
on secondary fluid-inclusions in igneous quartz
phenocrysts from altered wallrocks; although
lacking temporal information, these inclusions
seem to provide a representative cross-section of
the fluids involved. The most reliable data on the
ore-forming fluids are obtained through infrared
microscopy directly on ore minerals, such as
enargite (Deen 1990; Mancano & Campbell
1995).
The temperatures and salinities estimated for
HS deposits define a wide range, from 90 to 480
C and <1 to 45 equiv. wt.% NaCl, respectively
(Table 4). There is no systematic difference in
salinity among Au-, and Ag- or base-metal-rich
deposits, in contrast to that noted for lowsulfidation Au versus Ag deposits (Hedenquist &
Henley 1985). Large variations in both
temperature and salinity also occur within a single
deposit; these reflect the dynamic environment,
with high- and low-temperature and high- and
low-salinity fluids interacting during the course of
mineralization. Four broad groups of hydrothermal fluids are recognized here on the basis of
the estimated temperatures and interpretations
given by most workers. The temperature
boundaries chosen for each group are only
indicative, as significant variations exist among
and within deposits; each group, however,
provides relevant information on various genetic
aspects.
Group 1. Higher temperature (e.g., >300 C)
fluids of variable salinity, which have been
documented in several deposits and are generally
431

A. Arribas, Jr.

Table 4. Summary of fluid-inclusion microthermometric data for high-sulfidation deposits

Deposit

Host-mineral
studied

Temperature
(C)'

Associated
Salinity
(equiv wt.9E- NaCI) alteration

Motomboto, Indonesia
Nalesbitan, Philippines
Lepanto, Philippines

Barite
Quartz
Enargite

150-180
220-260
170-290

0.2^.5

Chinkuashih, Taiwan

Quartz, barite,
al unite
Quartz (no details
reported)

180-330

0.2-12

160-300
220-380
100-160
(300-420)
130-250
-270
250-310
190-240
210-330

2-22
3-19
0-5
(3-20)
<1
up to 30

180-280

2-18

(300-390)
-100
230-480+
210-280
(370-410)
180-210
300-380
(up to 450)
160-280
360-450
230-330
220-250
330-380
230-260
170-350

(up to 9)

190-280
140-180
(>300)
200-460
160-340
230-480

0.1-4
0.1-2.7
(up to 27)
1-37
0.3-12
^47

AA
Ser

Quartz, quartz
phenoe

170-300
220-450

2-30
2-45

Ser

Quartz, barite,
quartz phenoe

190-320
90-140
(390-500)

0.4-23
0.4-1.6
(32-45)

Zijinshan, China

Nansatsu. Japan

.^

Akaiwa, Japan
Mitsumori-Nukcishi, Japan

Summitville, Colorado

Goldfield, Nevada

Paradise Peak, Nevada

Julcani, Peru

Ccarhuaraso, Peru
Colquijirca, Peru
Can-Can (La Coipa),
Chile
El Indio, Chile

La Mejicana (LM) and


Nevados Famatina (NF),
Argentina
Rodalquilar, Spain

. .
Funei-Serrenti, Italy

Quartz
Diaspore
Quartz, barite,
quartz phenoe
Quartz phenoe
Barite
Quartz phenoe
Quartz, barite
Quartz, barite
Quartz
Quartz phenoe
Quartz phenoe
Wol, ena, quartz
Siderite
Quartz phenoe
Quartz phenoe
Sphalerite, quartz
hubnerite
Quartz phenoe
N/A

<1

A A/si 1
AA/sil
AA/sil

AA
Ser
Sil
AA/sil
AA/sil

Ser
AA/sil

0.5-14

AA
AA/sil

5-18
0.2-8

AA/sil
AA/sil

<3
30-35

AA/sil

5-24
38-46
9-20
6-9
7-18
4-11
<l-40

AA/sil

Ser

AA/sil
AA/sil
AA/sil
dominant
AA/Ser

AA + ser

AA/sil
AA/sil

Abbreviations used: AA = advanced argillic, ena = enargite, phenoe = phenocrysts, ser =


sericitic, sil = silicic: wol = wolframite: see Table 3 for paleodeplh estimations
1
temperatures are rounded to the nearest 10; brackets used to indicate high-iemperature
inclusions typically interpreted as having formed early or being anomalous

interpreted as "anomalous" or unrelated to ore and


are associated with early stages of alteration.
Two-phase entrapment may explain some of the
unusually high homogenization temperatures (7*),
particularly considering the shallow mineralization depth inferred for many of the deposits
(Table 3). However, most workers agree that such
All

entrapment cannot account for all the high Th


values. The consistent presence of these fluids in
several deposits indicates a high temperature
gradient, and implies the presence of a shallowdepth intrusion, and possibly lithostatic confining
pressures. On the basis of fluid-inclusion, as well
-34
as isotopic (8 Ssuifate-suifide) temperatures (see

High-sulfidation Epithermal Deposits


Table 4. (continued)
Deposit
Motomboto, Indonesia
Nalesbitan, Philippines
Lepanto, Philippines

Comments

References
Perell6(1994)
Sillitoe et al. (1990)
Mancano & Campbell (1995),
Garcia (1991)

Reconnaisance study in late-stage barite


Reconnaissance study; liquid CO2 observed
Sampled interval 3 km long by 0.5 km high ; cooling fluids
away from subjacent porphyry Cu-Au deposit, where
Th >450C & salinity up to 54 eq wt.% NaCl
Poorly-documented samples along a 450-m vertical interval;
the higher T|,s in samples at -750 m depth; CO2 observed
Associated with main stage Cu
Deep alteration zone (>600 m depth)
Associated with late, shallow silica-Au
Associated with early silica and quartz-dickite
Late, vug-filling quartz
Qtz in breccia, saline liquid and low-salinity vapor coexist
Vein quartz -400 m below Kasuga deposit
Coarse-grained diasporc
Not (known) Au or Cu mineralization, but high salinity
fluids
Liquid-rich; salinity >6 eq w..1o NaCl only in vuggy silica
associated with Cu mineralization; CO2 observed
Liquid- and vapor-rich inclusions; also polyphase inclusions
Late barite-Au assemblage
True Th is interpreted to be 250-290C
Hydrostatic and near-lithostatic pressures suggested

Bruha & Noble (1983), R.


Stoffregen (written
commun., 1994)
Cunningham (1985)
Bruha & Noble (1983)
Vikrc(1989)

Paradise Peak, Nevada

Late, vug-filling crystals in hydrotherma! breccia;


From stockwork Au East Zone deposit; CO2 observed

John et al. (1991)


Sillitoe &Lorson (1994)

Julcani, Peru

Quartz-alunitepyrite
Pre-ore tourmaline breccia dykes, lithostatic pressures likely.
Main-stage ore fluids, also inner veins, liquid-rich inclusions
Late-stage ore fluids, also in outer veins; P correction applied
Quartz-alunitepyrite
Quartz-alunitepyrite
Two generations identified; both may be very saline. Evidence
for P above hydrostatic and higher salinities at depth
Copper and gold stages
Late stage
Interpreted as early, with vapor-rich inclusions, CO2 observed
LM & NF: includes liquid-, vapor-rich and polyphase inclusions
NF: complete transition from porphyry-type fluids in Ksilicate stage (SOO-oW^C, up to 67 eq wt% NaCl)
through sercitic to epithermal fluids in HS (AA) stage;
vapor-rich inclusions typically less saline
Vertical temperature and salinity gradient: high-temperature
brines coexist with low -aiinity vapor inclusions;
hydrostatic and near-lithostatic pressures suggested
Includes high + low-salinity fluids (22-23, <6 eq wt% NaCl)
Late stage

Bruha & Noble (1983)


Shelnutt& Noble (1985)
Deen(1990)
Deen(1990)
Bruha & Noble (1983)
Bruha & Noble (1983)
Townley(1991)

Chinkuashih, Taiwan
Zijinshan, China

Nansatsu, Japan
Akaiwa, Japan
Mitsumori-Nukcishi, Japan
Summitville. Colorado -

Goldfield, Nevada

Ccarhuaraso, Peru
Colquijirca, Peru
Can-Can (La Coipa),
Chile
El Indio, Chile
La Mejicana (LM) and
Nevados Famatina (NF).
Argentina
Rodalquilar, Spain
Furtci-Serrenti, Italy

below), pressures above hydrostatic have been


suggested for several deposits, including Julcani
(Shelnutt & Noble 1985), Goldfield (Vikre 1989),
Summitville (Rye 1993), and Rodalquilar (Arribas
etal. 1995a).
Group 2. Intermediate-temperature fluids
(e.g., 180-330 C), with salinities variable from

Folinshce el al. (1972), Yen


(1976), Tan el al. (1993)
Zhang et al. (1994)

Hedenquist et al. (1994a)


Akamatsu & Yui (1992)
Aoki&Watanabe(1995)

Jannase/tz/. (1990)
Losada-Caldcr6n & McPhail
(1994)

Sanger-von Ocpen et al. (1989),


Arribas <</. (1995a)
Ruggieri (1993b)

<1 to ~18 equiv. wt.% NaCl. With the possible


exception of deposits for which only the late-stage
minerals have been studied, these typically liquidrich inclusions are found in all deposits. Mainstage ore fluids are contained within this group.
The temperatures measured in fluid inclusions in
enargite at Lepanto (Mancano & Campbell 1955)

433

A. Arribas, Jr.
and Julcani (Deen 1990) are broadly similar, but
their salinities are distinctly different (0.2-4.5
equiv. wt.% NaCl versus 8-18 equiv. wt.% NaCl,
respectively), providing constraints on the role of
a saline magmatic liquid (versus low-salinity
vapor) in the generation of HS deposits.
Group 3. Lower temperature (e.g., 90-180
C), dilute (typically <5 equiv. wt.% NaCl)
liquids; these have been documented in a few
deposits associated with late-stage (e.g., Aubarite) mineralization. The late-stage ore fluids at
Julcani are hotter (220-250 C; Deen 1990) and
slightly more saline (6-9 equiv. wt.% NaCl), than
these averages, but no correlation among the late
stages in different deposits is attempted here.
Group 4. "Sericitic" fluids. As mentioned
above, sericitic (quartz-sericite-pyrite) is the most
common alteration assemblage observed below
the ore zone in some HS deposits. Although
detailed documentation is lacking for many
deposits, higher temperatures and higher salinity
fluid-inclusions seem to characterize the sericitic
zone with respect to the shallower zones of
alteration (Table 4). For example, at Rodalquilar
(Arribas et al. 1995a), documentation of temperature and salinity along a >600-m vertical
interval (extending 500 m below the ore zone; Fig.
6) shows a gradient which correlates with the
change in dominant alteration, from silicic and
advanced argillic (T = 170-300 C, salinity = 2-15
equiv. wt.% NaCl at the elevation of the orebody)
to sericitic (7"= 220-450 C, salinity = 2-45 equiv.
wt.% NaCl) assemblages.
The transition from advanced argillic alteration,
through quartz-sericite-pyrite, to K-silicate
alteration and typical porphyry-type hightemperature (600+ C) and high-salinity (up to 67
equiv. wt.% NaCl) fluids of magmatic origin is
displayed, among the examples reviewed, at the
Lepanto-FSE and La Mejicana-Nevados del
Famatina epithermal-porphyry copper systems.
The cooler and less saline inclusion fluids
documented in the ore zone of the HS deposits are
interpreted to reflect mixing of magmatic and
meteoric fluids in an environment shallower than
that of porphyry mineralization. Furthermore, in
common with porphyry-type deposits, hightemperature, vapor-rich, low-salinity fluid
inclusions coexist with high-temperature, liquid434

Temperature (C)
200

400 -

300

400

500

HjO + 5 wt% NaCl


(hydrostatic)

Elevation of
Onto deposits

200
(hydrostatic)

200

(lithostatic)

B1
s -|400

>

1I

3
- 600

1
I

Q.

-200

o
800

-400

Figure 6. Elevation versus temperature diagram


showing the range (horizontal line) and average
(vertical line) of fluid-inclusion homogenization
temperatures measured in the Rodalquilar Au deposit,
Spain. Also shown are the temperatures calculated, on
the basis of 834S suir,dMuifatt for four coexisting alunitepyrite samples (large filled circles), reference boilingpoint curves, and vertical spans of the alteration zones
mentioned in the text. Estimated salinities of fluid
inclusions in the shallow advanced argillic/silicic zone
and deep sericitic zone range between 2 to 30 equiv.
wt.% NaCl and 2 to 45 equiv. wt.% NaCl, respectively
(modified from Arribas et al. 1995a).

rich hypersaline inclusions (i.e., with Groups 1


and 4, above). These fluids may be the result of
boiling of a high-temperature liquid, or they may
reflect immiscible vapor and hypersaline liquid
derived directly from shallow-emplaced magma
(Rye 1993; Hedenquist & Lowenstern 1994;
Shinohara 1994; Hedenquist this volume).
Sulfur-isotope Evidence
The abundance of coexisting hydrothermal
sulfides and sulfates, in addition to the possibility

High-sulfidation Epithermal Deposits


.Sulfides

Sulfates

T &V - 534SJS

Lepanto
Chinkuashlh
i

Nansatsu
Summltville
Goldfield

Pueblo Viejo

I
I

Julcani

A 3 4 SH 2 S-SO4

Temp. (C)* H 2 S/S0 4


220 - 420

2-6

220 - 270

200 - 240

200-390

200 - 350

180-260

210-270

220-330

El Indto
Rodalqullar

-H
-10

1
0

1
1
10
&*S (%., CDT)

h20

'(mineral pairs)

30

Figure 7. Range of 834S (per mil) values for sulfides and sulfates from nine highsulfidation deposits. Also shown are the values calculated for 834S for total sulfur in the
hydrothermal system (triangles), H2S/S04, and the range of temperatures determined
from sulfide-sulfate mineral pairs. Solid triangles indicate deposits in which 834SS was
calculated on the basis of isotopic analyses of samples of unaltered whole rock
genetically related to mineralization. See Appendix for references and information on
data plotted.
of measuring 34 S/ S in host rock and genetically
related igneous rock (Sasaki et al. 1979), allows
sulfur-isotope studies to provide information on
the composition, temperature, and sulfur sources
of the hydrothermal fluids. The results of detailed
studies in nine HS districts show a remarkable
consistency (Fig. 7). In agreement with the
observations in active volcanic-hydrothermal
systems (e.g., Kiyosu & Kurahashi 1983), sulfide
and sulfate minerals are mainly in isotopic
equilibrium, and, therefore, their overall S/ S
depends on the temperature of mineralization and
the 34S/32S of total sulfur in the hydrothermal
system. Only the data for alunite from the
Campana vein in El Indio (Fig. 7) are different. If
the measured El Indio alunites are not steamheated or supergene (unlikely as they contain finegrained pyrite; Jannas et al. 1990), the most likely
explanation is a "magmatic-steam" (Rye et al.
1992) origin, in which the 834S of alunite is close
to the composition of total sulfur in the system
(e.g., Alunite Ridge in Marysvale; Cunningham et
al. 1984; Rye et al. 1992) . Combined with the

8 S values of pyrite and enargite from the same


vein, these values indicate drastic changes in
H 2 S/S0 4 during the course of mineralization
(similar to those for the Red Mountain alunite
deposit; Bove et al. 1990; Rye 1993).
The main conclusions of the sulfur-isotope
studies in HS deposits are: (1) sulfur in the
deposits is magmatic, but the magmatic sulfur is
overall heavier than mantle values (from 534S = 2
0/

0/

2 'oo at Summitville, to 9 2 'oo at Rodalquilar;


Fig. 7). This is not surprising given the most
common geological setting of the deposits;
isotopically heavy igneous sulfur is common in
volcanic arc environments (e.g., Ueda & Sakai
1984). (2) A simple mass-balance calculation
done in several deposits using the S/ S values
of the igneous rocks and the average 34S/32S
values of sulfides and sulfates indicates that
H 2 S/S0 4 in the hydrothermal fluids was generally
about 4 2 (Fig. 7; Rye et al. 1992; Hedenquist et
al. 1994a; Arribas et al. 1995a). This is a
minimum value for ore-forming fluids because it
applies mainly to the early stage of hydrothermal

435

A. Arribas, Jr.

alteration, which is characterized by a sulfate-rich


alunite-pyrite assemblage. (3) Isotopic equilibrium between sulfide and sulfate in the
hydrothermal solutions results, in a majority of the
deposits, in reliable temperatures calculated on the
basis on A SH2s-so4 (Fig- 7). Pyrite-alunite
mineral pairs were used most commonly, and
where sampling with depth is available, they show
a thermal gradient: e.g., 220 to 330 C over 200-m
elevation at Rodalquilar (Arribas et al. 1995a),
200 to 390 C over -900 m at Summitville (Rye
1993); 220 to 420 C over 500 m at Lepanto
(Hedenquist and Garcia 1990; J.W. Hedenquist,
unpub. data). Other mineral pairs used with
consistent results include pyrite-barite (Vikre
1989; Deen 1990), sphalerite-barite (Vennemann
et al. 1993), and pyrite-gypsum (Vikre 1989). The
range of isotopic temperatures is consistent with
temperatures estimated from fluid inclusions and
alteration mineralogy (e.g., Hemley et al. 1980;
Reyes 1990; Reyes et al. 1993). The range is also
consistent with formation of alunite at
temperatures below ~400 C, when S0 2 gas starts
to disproportionate in the hydrothermal solution
(Sakai & Matsubaya 1977; Bethke 1984).
Oxygen- and Hydrogen-isotope Evidence
In terms of oxygen and hydrogen isotopic
composition, the fluids that form HS deposits are
arguably some of the better documented and
understood in ore-deposit studies. This situation
contrasts sharply with that of a decade ago, at
which time no data were available to corroborate
the affinity-suggested between fluids in active
volcanic-hydrothermal systems and HS deposits
(e.g., Heald et al. 1987; Hedenquist 1987). Stableisotope studies of HS deposits are particularly
illuminating because of: (1) the abundance and
variety of oxygen- and hydrogen-bearing minerals
(e.g., alunite, illite, kaolinite), (2) the development
of analytical procedures for complete stableisotope analysis of alunite, including 8 l 8 O s o 4 and
61 0 O H that help to distinguish the various types
of alunite and associated acid-sulfate alteration
(Rye et al. 1992; Wasserman et al. 1992), (3)
fewer limitations on the interpretation of the
isotopic data because of the relatively young age
of mineralization of most HS deposits and general

4 -> r

lack of post-depositional effects that disturb the


stable-isotope systematics, and (4) the availability
of detailed information on the isotopic
composition of fluids in active geothermal and
volcanic-hydrothermal systems, which allows
fluids estimated in HS deposits to be compared
with those in their active equivalents.
Some limitations still exist. These may be
independent of obvious factors such as sampling
or mineral-preparation procedures (fundamental
for achieving representative and reliable results),
analytical imprecision, and natural variations, as
observed in active systems (e.g., Aoki 1991, 1992;
Rowe 1994). Important limitations that must be
taken into account for optimum use of the stableisotope data are related to (1) the choice of
temperature of mineral formation for calculation
of the fluid isotopic composition, (2) the lack of
mineral-water fractionation factors for some
minerals (e.g., pyrophyllite), and (3) the
disagreement among fractionation constants
proposed for even common minerals such as illite
(see Dilles et al. 1992, for a discussion) and
kaolinite. For example, at 200 C there is a
difference of 20 Aw between the D/H fractionation constants for kaolinite - water as given
by Marumo et al. (1980) on the basis of samples
of minerals and water from active systems, and by
Liu & Epstein (1984) on the basis of experimental
results. For these reasons, discussion of the
sources of water during acidic alteration in the
deposits considered here is based on the average
of the data collected for alunite, for which
fractionation factors are well-known (Stoffregen
et al. 1994). The magmatic-hydrothermal alunite
typical of-HS deposits gives good results because
it is relatively coarse-grained (post-mineral D-H
exchange is not a problem; Stoffregen et al. 1994)
and commonly is closely associated with ore, thus
recording equilibrium conditions of a fluid closer
in composition to the ascending mineralizing
solution than the kaolinite or illite from outer
alteration zones.
Oxygen and hydrogen isotopic compositions
of water in HS deposits are clearly consistent with
mixing between a high-temperature magmatic
fluid of 8 18 0 = 9 l/oo and 8D = -30 20^oo and
meteoric groundwaters (Fig. 8). In part because of

High-sulfidation Epithermal Deposits

0-20-

Alunite alteration stg.

O
Q

Ore mineralization stg.


Alteration/
mineralization

Subduction-related
volcanic vapor
Arc + crustal
felsic magmas

-40-

Acidic fluids In high] sulfldation deposits


.20

W W

Active systems
(Giggenbach, 1992b)

so
6D(%.)

Volcanic
Geothermal

-100

-20

i
-15

1
-10

1
-5

1
0
6

18

'r
5

10

15

20

0 (%o, SMOW)

Figure 8. Summary diagram showing variation in oxygen- and hydrogen-isotope composition of hydrothermal
fluids in high-sulfidation deposits. The average isotopic composition for the main stages of acidic alteration
(squares) and ore-mineralization (circles) fluids are shown. Where possible, only alunite data were used for the
alteration stage (SD and 8l8OSOi,); &I800H is not used because hydroxyl oxygen requilibrates with the hydrothermal
fluid during cooling (Rye et al. 1992). Tie-lines between data points connect samples from the same deposit. Inset
shows the isotopic composition of fields defined by waters from active geothermal systems and high-temperature
fumarole condensates in subduction-related andesitic volcanoes (from Giggenbach 1992b). Go = Goldfield, Ju =
Julcani, Le= Lepanto, Nansatsu district: Ka = Kasuga, Iw = Iwato, NF = Nevados del Famatina, PV = Pueblo
Veijo, Ro = Rodalquilar, RM = Red Mountain, Lake City, Colorado, Su = Summitville. The approximate
compositions of groundwaters suggested for several deposits are indicated by the intials parallel to the meteoric
water line. See Appendix for references and information on data plotted.
the very light isotopic composition of local
meteoric water, this meteoric-magmatic watermixing trend is displayed particularly well by the
three stages of alteration/mineralization at Julcani
(Deen 1990; Rye 1993): from a magmatic-waterdominated early stage of (alunite) acid-sulfate
alteration (Ju; Fig. 8), through main ore-stage
fluid-inclusion waters (Jut and JU2), to meteoricwater-dominated late ore-stage fluid-inclusion
waters (Ju3). In addition to Julcani, the ore fluids
at Summitville (Rye et al 1990; Rye 1993) and
Rodalquilar (Arribas et al. 1995a) also have lower
5 1 8 0 values than those of acidic alteration fluids,
indicating greater dilution by groundwater (Fig.
8). The extent of an O-shift in the groundwater
component due to water-rock interaction, as
typically seen in some neutral-pH geothermal
systems, is not known, but such a shift is not
indicated by the Julcani data.
The overall oxygen- and hydrogen-isotope

relations are identical to those of volcanichydrothermal and geothermal systems associated


with subduction-related volcanism (Giggenbach
1992b; Fig. 8, inset). The similarity is even closer
between the composition of acidic alteration fluids
(large shaded field, Fig. 8) and the vapor
condensates from high-temperature fiimaroles of
andesitic volcanoes (dark shaded field, Fig. 8,
inset), such as Nevado del Ruiz, Satsuma
Iwojima, or White Island, the last documented to
have a geochemical environment similar to that of
HS mineralization (Hedenquist et al. 1993).
The origin of the D-enriched magmatic (endmember) fluid of HS deposits has been interpreted
in two ways. Most workers conclude that the
acidic fluid in HS deposits is derived from
absorption of magmatic vapors outgassing from
arc volcanoes or felsic magmas in crustal settings
{e.g., Hedenquist & Aoki 1991; Matsuhisa 1992;
Giggenbach 1992a; Vennemann et al. 1993;

437

A. Arribas, Jr.

Hedenquist et al. 1994a; Arribas et al. 1995a).


The O and D enrichment of the volcanic vapors
with respect to their parent magmas (Fig. 8) is a
consequence of fractionation during degassing
from the melt (Taylor 1986, 1992; Matsuhisa
1992). In an alternative interpretation proposed by
Deen (1990) and Rye (1993), the enrichment in
deuterium is the result of reaction in a low
water/rock environment between magmatic fluid
of 8D ~-80/Oo (calculated on the basis of the D/H
of igneous biotite) and wallrock at magmatic
temperatures down to T = -400 C (Rye 1993),
beneath the ore zone. This interpretation requires
that there have been chemical and isotopic
equilibrium between the magmatic fluid and
country rock; however, this would likely result in
neutralization of the magmatic fluid, a condition
that would not favor the extreme acidity required
for formation of vuggy silica at shallow depths
(Giggenbach 1992a).
SOURCES OF METALS

The origin of metals in HS deposits is more


speculative. There is consensus among most
workers (e.g., Sillitoe 1989, 1991a; Henley 1991;
White 1991; Rye 1993; Hedenquist et al. 1994a)
that the bulk of the ore-forming components is
contributed by magmatic fluid, either directly by a
magmatic vapor or hypersaline liquid that is
incorporated into the hydrothermal system, or
indirectly by remobilization of metals from a
porphyry-type protore. The study of radiogenic
isotopes so far has not provided conclusive
evidence. The reason for this is a consequence of
the common intimate association between the
deposits and host rocks; all are indistinguishable,
e.g., for Pb-isotope systematics. Lead-isotope
studies at Summitville (Doe et al. 1979) and
Rodalquilar (Arribas et al. 1995a) have shown
that the Pb in the deposits is igneous, but these
studies provide no information about the
processes by which the hydrothermal system
acquired the metal {i.e., leaching of the host rocks
or derivation from a crystallizing magma). By
contrast, in low-sulfidation gold deposits it is not
uncommon to detect a component of basementrock Pb (e.g., Doe et al. 1979; Arribas & Tosdal

A1<2

1994). Detailed and comprehensive radiogenic


isotope studies are needed, particularly in districts
where a majority of the host or country rocks are
expected to have isotopically distinct signatures
from those of the genetically related igneous
rocks.
If magmatic fluid is accepted as the dominant
metal source in porphyry copper deposits, then the
best evidence for a magmatic fluid source in HS
deposits comes, perhaps, from consideration of
the common genetic connection between the two
deposit types (Sillitoe 1988, 1989, 1991a, this
volume; Henley 1991; Mitchell 1992). In this
context, Arribas et al (1995b) have argued for a
common magmatic metal source for the FSELepanto porphyry-epithermal HS copper-gold
system on the basis of the intimate spatial and
temporal relations (Fig. 4) and similar metal
associations (i.e., Au/Ag values and high Te, Sn,
and Bi).
POST-DEPOSITIONAL EFFECTS

In deposits that have not suffered deformation


and metamorphism (those in Table 1, aside from
the Paleozoic deposits in southeastern Australia
and the Proterozoic Enasen deposit in Sweden),
the principal post-depositional processes affecting
HS deposits are mineral deposition from nearsurface steam-heated or supergene waters. Both
can result in acid-sulfate alteration which may
mask hypogene alteration zones and ore. On the
basis of geological and mineralogical criteria, KAr dating, and stable-isotope analyses of alunite,
identification of the origin of acid-sulfate minerals
overprinting should be straightforward (e.g.,
Hayba et al. 1985; Rye et al. 1992; White &
Hedenquist 1995). At Paradise Peak, a drop of the
paleowater table during a late stage led to
overprinting of steam-heated acid leaching with
Hg mineralization and abundant cristobalite
(Sillitoe & Lorson 1994). Atmospheric oxidation
of sulfides during weathering in at least three HS
deposits (Goldfield, Rodalquilar, and Summitville; Table 3) led to development of surficial
supergene acid-sulfate blankets. Supergene alunite
at Rodalquilar forms white to yellow,
porcelaneous veins consisting of fine-grained

High-sulfidation Epithermal Deposits

(Jannase/a/. 1990).
A feature of HS deposits which has been
noted commonly is a change in the hydrothermal
system with time towards a fluid that is less
reactive and less oxidized. To explain this
observation, Berger & Henley (1989) suggested
that precious-metal mineralization in HS deposits
is introduced by later incursion of low-sulfidation
geothermal-type fluids into previously formed HS
alteration zones of magmatic origin. This model
GENETIC MODELS
requires coincidence of two distinctively different
hydrothermal fluids along the same plumbing
Genetic models for HS deposits have been
system. These unusual conditions seem to have
presented recently by various workers. A selection
occurred at the Kelly mine, Philippines (Comsti et
that emphasizes the variations in conceptual
al. 1990; Aoki et al. 1993) and Masupa Ria,
models includes Sillitoe (1989), Berger & Henley
Indonesia (Thompson et al. 1994), where two
(1989), Giggenbach (1992a), Rye (1993),
distinct, overlapping hydrothermal assemblages
Hedenquist et al. (1994a, which incorporates
were identified. In these two deposits the known
White 1991), and Sillitoe (this volume). In
precious-metal mineralization occurred during the
common among these models is an evolution from
low-sulfidation stage. However, mineralization in
a main stage of acidic wallrock alteration (early) HS deposits does not occur under the reduced
to a main ore-forming stage (late). The differences
conditions of low-sulfidation-type geothermal
among the models mentioned above are relatively
fluids (see Fig. 3 in Hedenquist this volume).
minor and lie primarily in the details of the oreThe initial requirement for formation of a HS
forming stage. Most of these models consider the deposit is emplacement of an oxidized, typically
origin of HS deposits in the broad context of
intermediate calc-alkaline magma to within a few
metallic deposits formed within magmatickilometers of the paleosurface. Rather than a
hydrothermal systems {i.e., including lowspatially and temporally isolated magmatic event,
sulfidation deposits, porphyry-type deposits, the subvolcanic intrusions seems to represent an
replacement skarn deposits, and even submarine
"aborted" late-stage feeder of a comagmatic
volcanogenic massive sulfide deposits). This
central-vent volcano or a maar - diatreme or flowaspect is important from a genetic perspective, as
dome complex. Within the ascending, crysit is unwise to place strict limits between these tallizing, and cooling magma, a hydrothermal
ore-forming environments (Hedenquist &
fluid phase exsolves and concentrates ore-forming
Lowenstern 1994). Similarly, the two stages
metals and volatiles. Because of magma
described here should be considered as successive convection and concentration of the magmatic
phases of an evolving hydrothermal system, in
fluid phase in the roof of the intrusion, the total
most cases, without sharp boundaries between
amount of water and metals available for a
them. For example, much of the precious-metal
potential suprajacent hydrothermal system far
mineralization at Pueblo Viejo was directly
exceeds that of the comparatively small stock that
associated with acidic alteration (Muntean et al.
reaches the shallow depth (Lowenstern 1994;
1990). In contrast, crosscutting relations at Julcani = Shinohara & Kazahaya this volume). At the
are unambiguous; the early stage of acid-sulfate
confining pressure of a few kilometers depth (500alteration is separated from the base- and
1000 bars), and given the amount of water in the
precious-metal veins by emplacement of
type of magmas associated with HS deposits, the
tourmaline-pyrite-quartz-altered breccia dykes
magmatic fluid will separate into two aqueous
(Shelnutt & Noble 1985; Deen 1990). El Indio
phases, a low-salinity vapor and a hypersaline
also shows clear crosscutting relations between
liquid (Fournier 1987; Hedenquist & Lowenstern
early enargite veins and later gold-quartz veins
1994; Shinohara 1994), with chloride-complexed
(commonly <50 \ixn) pseudocubic alunite with
subordinate jarosite, kaolinite, and hydrated
amorphous silica (Arribas et al. 1995a). In the
western United States, formation of supergene
alunite at ~10 Ma at Goldfield and Paradise Peak
(Table 3) seems to be related to regional tectonic
episodes (e.g., Arehart et al. 1992). For an account
of supergene effects in deposits of the Chilean
Andes, see Sillitoe (1991a).

439

A. Arribas, Jr.
ORE DEPOSITION

ALTERATION

B2

Heated ground- ..
waters

Magmatic vapors
(incl., S 0 2 . H a )

Magmatic
brine

Absorption of
high-P vapor

f
Vuggy silica
Alunite
Kaolinite
Sericite
K-silicate

Magmatic
brine

7*
j Heated
groundwater

Possible
Cu{Au)

Figure 9. Model showing the two main stages of evolution of HS deposits. A: Early stage of advanced argillic
alteration dominated by magmatic vapor. B, and B2: Two genetic hypotheses proposed for the stage of ore
formation. B, = absorption of high-pressure vapor by entrainment in meteoric water cell at depth to explain lowsalinity, mixed magmatic-meteoric ore fluid (Hedenquist this volume). B2 = ascending metal-bearing magmatic
brine with shallow cooler meteoric waters to explain high-salinity, mixed magmatic-meteoric ore fluid (White
1991; Rye 1993; Hedenquist et al. 1994a).
metals strongly partitioned into the high-density
liquid (Hemley et al 1992; Hedenquist this
volume).
At this early intrusive stage, several modes of
magma degassing may occur which will lead to
different
styles of magmatic-hydrothermal
systems with or without associated mineralization
(Giggenbach 1992a). To form the styles of
alteration and the spatial distribution of alteration
zones characteristic of HS deposits, degassing
must be very efficient, with oxidized hightemperature magmatic vapor reaching shallow
depths with little reaction with rock or dilution by
groundwaters at greater depths (Fig. 9A). Dilution
with groundwaters is unlikely because the high
temperatures surrounding the cooling magma
cause meteoric water cells to be displaced from
the magma core (Fig. 9A). In addition to the
relatively low pressure at the depth of intrusion,
effective degassing will be favored by the
structural factors characteristic of HS deposits,
such as fractured volcanic domes or roots of
domes, caldera or diatreme faults, volcanic vent
contacts, and active faults with a dilational
component.
As the high-temperature magmatic vapor

440

reaches shallow depths of less than a kilometer, it


may be absorbed by groundwater if it does not
discharge as a fumarole. The acidity of this
groundwater-absorbed vapor condensate increases
as the liquid cools, first at temperatures below
~400 C by disproportionate of S0 2 to form
H 2 S0 4 and H2S (Day & Allen 1925; Sakai &
Matsubaya 1977), then by progressive dissociation of H 2 S0 4 and HCl at lower temperatures
(<300 C). Reaction of the increasingly acidic
liquid with wallrock results in the upward
alteration sequence of sericite-kaolinite-
alunite-^vuggy silica (Fig. 9A); the residual
vuggy silica rock results from complete leaching
of the rock components, except silica, by a
hydrothermai solution with a pH <2 and
temperatures probably <250 C (Stoffregen 1987).
The extremely acidic conditions may even lead to
formation of dissolution cavities in which the only
remnant of the host rock is a basal sedimentary
layer of quartz phenocrysts (e.g., Rodalquilar;
Arribas et al 1995a).
For the quartz-alunite-pyrite assemblage of
the advanced argillic zone, the stable-isotope
evidence is consistent with magmatic vapor being
absorbed by meteoric waters, with the latter

High-sulfidation Epithermal Deposits

constituting a relatively small part of the mixture


(generally <l/3; Fig. 8). The fluid-inclusion
evidence, by contrast, is inconclusive because of
the lack of temporal information. Nevertheless,
high-temperature, high-salinity inclusion fluids
have been interpreted to form early in most HS
deposits (e.g., Bruha & Noble 1983; Ruggieri
1993b; Arribas et al. 1995a). These fluids may be
restricted to greater depths, as demonstrated at
Rodalquilar and in other deposits where highsalinity inclusion fluid is associated with the deep
sericitic alteration (Table 4). This latter observation suggests an episodic ascent of high-salinity
magmatic liquid from the greater depths of the
hydrothermal system, where the hypersaline liquid
tends to stay because of its high density (Fig. 9A).
These magmatic brines may be more closely
related to the K-silicate alteration (and, in places,
porphyry mineralization) that envelopes the
intrusion (Fig. 9A; Sillitoe 1989).
The conditions during the main stage of ore
formation are not yet as well-understood, and this
reflects the much more variable geochemical
environment in comparison with that associated
with acidic alteration. During the ore stage, the
hydrothermal liquid may be less dominated by a
magmatic vapor phase and its associated "sulfurgas buffer" (Giggenbach 1987). The presence of
this S02-H2S buffer is the reason that the early
stage of alteration is so oxidized, as reflected by
the alunite-pyrite assemblage (Whitney 1988;
Giggenbach 1992a). Instead, conditions during the
ore stage fluctuate within a range of redox
potential that is reflected by enargite-pyrite
alunite and enargite-tennantite-chalcopyrite associations, which are relatively high to intermediate
sulfidation-state assemblages, respectively (see
Fig. 3 in Hedenquist, this volume). In the Lepanto
(Claveria & Hedenquist 1994) and El Indio
(Jannas et al. 1990) deposits, these two
assemblages are related to Cu-rich and Au-rich
mineralization, respectively, with the latter being
of later stage in both cases. The more reduced
conditions are a likely consequence of increased
water-rock interaction, and, to some extent,
increased dilution of the oxidized magmatic fluid
by meteoric water; this trend is also consistent
with the isotopic composition of waters in the
main ore stage of various deposits (Fig. 8). No

discrimination, however, can be made between a


meteoric-water component that is incorporated at
deep or shallow levels within the hydrothermal
system. Importantly, salinities during the main ore
stage can be low (e.g., Lepanto and El Indio, <4
equiv. wt.% NaCl; Table 4) or moderate to high
(Julcani, up to 18 equiv. wt.% NaCl; Zijinshan, up
to 22 equiv. wt.% NaCl; Table 4).
Assessment of a Model
No single model adequately explains all of
these various observations, and several hypotheses
have been proposed, each reflecting an emphasis
on individual deposits or different interpretations
of the fluid-inclusion and stable-isotope data. A
basic understanding of this ore-forming environment may be gained by considering the principal
end-member fluid components and ore-forming
processes. The spectrum of characteristics
displayed by HS deposits may be then analyzed in
the context of such a genetic framework.
Four fluid regimes have been identified in the
HS environment; evidence for all is present in the
early stage of HS alteration, and three of them are
critical to formation of porphyry systems (e.g.,
Henley & McNabb 1978; Sillitoe 1989). These
end-members are: (1) a metal-rich hypersaline
magmatic liquid which tends to remain in the
vicinity of the intrusion, but may ascend (or be
driven) to shallower depths if the ambient
temperature is low enough (<400 C) for the
mechanical strength of the rock to increase
sufficiently to result in brittle fracturing (Foumier
1992), (2) a low-salinity magmatic vapor whose
metal-transporting capacity decreases sharply with
decreasing pressure (Hedenquist this volume), (3)
heated meteoric or connate water in deep
convection cells that collapse inward and
downward as the intrusive stock progressively
solidifies and cools, and (4) shallow and cool
meteoric groundwater.
Two
main
end-member
ore-forming
hypotheses are considered. In the "volatile
transport" hypothesis (Fig. 9B,), the magmatic
hypersaline liquid may remain at depth throughout
the evolution of the hydrothermal system, and the
low-salinity vapors are responsible for mineralization (Sillitoe 1989; Vennemann et al. 1993);
deep meteoric water entrainment of high-pressure
441

A. Arribas, Jr.
vapor is required for transport of sufficient
amounts of metals (Hedenquist this volume;
Sillitoe this volume). These conditions are
consistent with the low salinity of the Lepanto and
El Indio fluid-inclusion data. Mineral deposition
in this case may be caused by mixing with cooler
groundwater or by boiling, possibly resulting from
the abrupt pressure reduction associated with
hydrothermal brecciation.
In the "hypersaline liquid transport"
hypothesis (Fig. 9B2), following waning of the
magmatic vapor plume responsible for early
alteration, the lithostatic-pressured system fractures and the metal-bearing hypersaline liquid
ascends into the porous leached zone (Deen 1990;
White 1991; Rye 1993; Hedenquist et al. 1994a).
The dominant ore-forming mechanism in this case
is mixing of the metal-bearing hypersaline liquid
with cooler groundwaters at the site of deposition,
not at depth in the meteoric water convection cell.
This hypothesis has been proposed to explain the
high salinities recorded by inclusion fluids in
several deposits (e.g., Julcani).
A part of the ore-forming components may
originate from leaching of wallrock, but both
hypotheses agree on a dominantly magmatic
source for metals, with an increase in the meteoric
water component with time. The principal
difference between the two hypotheses is in the
nature of the magmatic phase responsible for
transporting the metals into the epithermal
environment, and in the site of meteoric water
dilution. A potential contributor to ore formation
in HS deposits involves remobilization of the
metals by a meteoric-water-dominated hydrothermal system from a subjacent K-silicate
assemblage and porphyry-type protore, such as
that which may have formed close to the intrusion
(e.g., Brimhall 1980). This mechanism, however,
has not been suggested as the main ore-forming
process in any of the deposits reviewed in this
study.
The three models for formation of HS ores,
assimilated here from the literature, are not
mutually exclusive; on the contrary, they may
occur in the same HS deposit as the magmatichydrothermal system evolves, with complexities
arising from multiple intrusions, variations in
depth of emplacement, and changes in the local
447

tectonic and hydrodynamic environment. None of


the three models satisfies the overall evidence. For
example, if metals were supplied only by a dense,
high-salinity liquid, a relation would be expected
among estimated salinities, metal associations,
and ore grade or metal abundances of the various
deposits. Such seems not to be the case. Similarly,
if alteration and mineralization were solely the
result of interaction between groundwater and
low- and high-pressure vapor, respectively, high
salinities should not be as common as they are
unless they are explained by local boiling of dilute
to moderately saline meteoric or seawaterdominated fluids.
SYNTHESIS

Gold, Cu, and Ag (and in a few exceptional


cases also Hg, W, Bi, Pb, and Zn) are produced
from HS deposits. As a source of Au, and because
their mode of occurrence and the potential to
overlie porphyry-type mineralization have been
widely recognized only within the past 10 to 15
years, HS deposits represent a valuable
exploration target that has been overlooked in
some regions. Most known HS deposits are young
in age, Tertiary and even Quaternary. Highsulfidation deposits form dominantly in
subduction-related
plutonic-volcanic
arcs,
commonly during crustal extension. The deposits
form at a depth intermediate between the surface
and shallow (few kilometers depth) intermediatecomposition intrusions.
The intimate relationship among HS deposits,
volcanic host rocks, and oxidized magmatic fluid
derived from a degassing intrusion is supported by
the following observations: (1) the volcanic rocks
hosting HS deposits were erupted immediately
prior to mineralization, (2) the ore-forming
hydrothermal system commonly follows the same
plumbing as that of the magmatic system (i.e.,
mineralization spatially associated with domes or
volcanic conduits), (3) the isotopic composition of
hypogene sulfides (e.g., enargite and pyrite) and
sulfates (e.g., alunite and barite) commonly can be
modelled from the 34S/32S of sulfur in igneous
rocks thought to be genetically related, by
equilibrium fractionation between H2S and SO4 in
solution at T -200-400 C, and (4) on the basis of

High-sulfidation Epithermal Deposits


oxygen and hydrogen isotopic ratios, the waters
involved in formation of HS deposits are identical
to waters in active volcanic-hydrothermal systems, in which the same HS geochemical
environment has been documented.
Ore formation in some HS deposits may
accompany acidic alteration, and recent studies of
the hydrothermal geochemistry of Au provide
preliminary evidence that this element may be
transported in HS and low-sulfidation systems as
different hydrosulfide complexes (AuHS0 and
Au(HS) 2 , respectively; Bening & Seward 1994;
Seward 1973). On the other hand, the presence of
moderate to high salinities in many HS deposits,
the intimate association with porphyry coppertype deposits, and the assumptions of the most
recent genetic models (transport of Au and Cu by
either hypersaline liquid or high-pressure vapor)
indicate that chloride complexes must also be
considered for metal transport.
Most HS deposits evolve from an early period
of acidic wallrock alteration to a late period of
precious- and base-metal mineralization. Acidic
alteration is characterized by advanced argillic
assemblages and porous (leached) rock, and the
hydrothermal fluid responsible for this alteration
is dominated by high-temperature magmatic vapor
containing S0 2 , H2S, and HC1. Less reactive and
oxidized fluids are typically responsible for ore
mineralization. Factors such as multiple intrusions
and opening or closing of fractures (conduits)
result in variations in the temperature, pressure,
and composition of the ascending solutions.
Combined with the shallow environment of
mineralization, these conditions lead to a variety
of deposit styles (mainly replacements, breccias,
and veins) that usually occupy a limited vertical
span of <300 to 500 m (except for >800 m at the
giant Chinkuashih deposit). The geological,
mineralogical, and
geochemical
evidence,
particularly the association between the orebodies
and the lateral and vertical zones of alteration,
illustrates the basic genetic condition of HS
deposits, that a magmatic fluid interacts extensive
ly with country rock and groundwaters on its
relatively short path to the earth's surface.

ACKNOWLEDGMENTS
Valuable insight on various aspects related to
this exciting ore-forming environment was gained
through discussions and field work with M. Aoki,
A. Arribas Sr., C. G. Cunningham, J. Hedenquist,
W.C. Kelly, R. O. Rye, J. J. Rytuba,and T. A.
Steven. Earlier versions of this manuscript
benefited from constructive reviews by Phil
Bethke, Andrew Campbell, Anne Thompson, John
Thompson, Peter Vikre, Noel White, and Jeff
Hedenquist, who also provided abundant
documentation on HS deposits worldwide.
REFERENCES
AKAMATSU, K. (1993): Acid Hydrothermal
Alteration at Otaru City, Hokkaido. M.S. thesis,
Hokkaido Univ., Sapporo, Japan (in Japanese)
AKAMATSU, K. & YUI, S. (1992): Acid sulfate
alteration at Akaiwa, near Otaru, southwestern
Hokkaido. In Hydrothermal Ore Deposits and Wall
Rock Alteration in Southwestern Hokkaido (H.
Matsueda, S. Yui. & K. Kurosawa, eds.). Soc.
Resource Geology, Tokyo, 29th Intemat. Geol.
Congress Field Trip Guide Book d, 17-23.
AOKI, M. (1991): Mineralogical features and genesis
of alunite solid solution in high temperature
magmatic-hydrothermal systems. Geol. Surv.
Japan Report 277, 35-37
AOKI, M. (1992): Magmatic fluid discharging to the
surface from the Osorezan geothermal system,
northern Honshu, Japan. Geol. Surv. Japan Report
279, 16-21
AOKI, M. COMSTI, E.C., LAZO, F.B. &
MATSUHISA, Y. (1993): Advanced argillic
alteration and geochemistry of alunite in an
evolving hydrothermal system at Baguio, northern
Luzon, Philippines. Resource Geol. 43, 155-164.
AOKI, M. & WATANABE, Y. (1995): Gold mineralization in an evolving magmatic hydrothermal
system at Mitsumori-Nukeishi area, Minamikayabe, southern Hokkaido. In Report of Regional
Geological Survey, General Evaluation of
Geological Structure, 1994 fiscal year, Ministry of

443

A. Arribas, Jr.

International Trade and Industry, Tokyo, Japan (in


Japanase).

Metallogeny of Copper Deposits (G.H. Friedrich,


ed.). Springer-Verlag, 280-290.

AREHART, G.B., KESLER, S E., O'NEIL, J.R. &


FOLAND, K.A. (1992): Evidence for the
supergene origin of alunite in sediment-hosted
micron gold deposits, Nevada. Econ. Geol. 87,
263-270.

BENING, L.G. & SEWARD, T.M. (1994): The


solubility of gold as AuHS in high temperature
hydrosulfide solutions. Beih. z. Eur. J. Mineral. 61,24.

ARRIBAS, A., JR. & TOSDAL, R.M. (1994): Isotopic


composition of Pb in ore deposits of the Betic
Cordillera, Spain: Origin and relationship to other
deposits in southern Europe. Econ. Geol. 89, 10741093.
ARRIBAS, A., JR., CUNNINGHAM, C.G., RYTUBA,
J.J., RYE, R.O., KELLY, W.C., McKEE, E.H.,
PODWYSOCKY, M.H. & TOSDAL, R.M.
(1995a): Geology, geochronology, fluid inclusions,
and isotope geochemistry of the Rodalquilar Aualunite deposit, Spain. Econ. Geol. 90 (accepted).
ARRIBAS, A., JR., HEDENQUIST, J.W., ITAYA, T.,
OKADA, T., CONCEPClON, R.A. & GARCIA,
J.S. (1995b): Contemporaneous formation of
adjacent porphyry and epithermal Cu-Au deposits
over 300 ka in northern Luzon, Philippines.
Geology 23 (accepted).
ASAMI, N. & BRITTEN, R.M. (1980): The porphyry
copper deposits at the Frieda River prospect, Papua
New Guinea. Resource Geol. Special Issue 8, 117130.
ASHLEY, R.P. (1974): Goldfield mining district.
Nevada Bur. Mines Geol. Report. 19,49-66.
ASHLEY, R.P. (1982): Occurrence model for enargitegold deposits. U.S. Geol. Surv. Open-file Report
82-795, 144-147.
ASHLEY, R.P. & SILBERMAN, M.L. (1976): Direct
dating of mineralization at Goldfield, Nevada, by
potassium-argon and fission-track methods. Econ.
Geol. 71,904-924.
BAKSA, C. (1975): New data on the enargite-luzonitepyrite massive sulfide deposits, North from
Lahoca-Hill, Recsk. FcHdtani Kazldny, Bull, of the
Hungarian Geol. Soc. 105, 58-74 (in Hungarian)
BAKSA, C. (1986): Genetic aspects of the Recsk
mineralized complex, Hungary. In Geology and

AAA

BERGER, B.R. (1986): Descriptive model of epithermal quartz-alunite Au. In Mineral Deposit
Models (D.P. Cox and D.A. Singer, eds.). U.S.
Geol. Surv. Bull. 1693, p. 158.
BERGER, B.R. & BONHAM, H.F., JR. (1990):
Epithermal gold-silver deposits in the western
United States: time-space products of evolving
plutonic, volcanic and tectonic environments J.
Geochem. Explor. 36, 103-142.
BERGER, B.R. & HENLEY, R.W. (1989): Advances
in the understanding of epithermal gold-silver
deposits, with special reference to the western
United States. Econ. Geol. Monogr. 6, 405423.
BETHKE, P.M. (1984): Controls on base and precious
metal mineralization in deeper epithermal
environments. U.S. Geol. Surv. Open-file Report
84-890.
BOGDANOV, B. (1982): Bulgaria. In Mineral
Deposits of Europe, Vol. 4/5, Southwest and
Eastern Europe (F. W. Dunning, W. Mykura & D.
Slater, eds.). Instit. Mining Metall., London, 215232.
BOGDANOV, B. (1986): Copper ore deposits in
Bulgaria.
In
Geotectonic
Evolution
and
Metallogeny of the Mediterranean Area and
Western Asia (W.E. Petrascheck & S. Jankovic,
eds.). Akad. Wiss. Schrifteneihe Erdwissenschaftlichen Kommissionen 8,103-112.
BONHAM, H.F., JR. (1984): Three major types of
epithermal precious metal deposits. Geol. Soc. Am.
Abstr. Programs 16,449.
BONHAM, H.F., JR. (1986): Models for volcanichosted epithermal precious metal deposits: a
review. In Proceedings Internat. Volcanological
Congress, Symposium 5, Hamilton, New Zealand
1986. Univ. Auckland, Centre Continuing
Education, Auckland, New Zealand, 13-17.

High-sulfidation Epithermal Deposits

BONHAM, H.F., JR. (1989): Bulk mineable gold


deposits of the western United States. Econ. Geol.
Monogr. 6, 193-207.

CLAVERIA, R.J.R. & HEDENQUIST, J.W. (1994):


Paragenesis of Au and related minerals in the
Lepanto Cu-Au deposit. Resource Geol. 44, 267.

BOVE, D., RYE, R.O. & HON, K. (1990): Evolution


of the Red Mountain alunite deposits, Lake City,
Colorado. U.S. Geol. Surv. Open-file Report 90235.

COMSTI, M.E.C., VILLONES, R.I., DE JESUS, C.V.,


NATIVIDAD, A.R., ROLLAN, L.A. & DUROY,
A.C. (1990): Mineralization at the Kelly gold
mine, Baguio district, Philippines: fluid-inclusion
and wall-rock alteration studies. J. Geochem.
Explor. 35, 341-362.

BRIMHALL, G.H., JR. (1980): Deep hypogene


oxidation of porphyry copper potassium-silicate
protore at Butte, Montana: A theoretical evaluation
of copper remobilization hypothesis. Econ. Geol.
75, 384-409.
BRODTKORB, M.K. & PAAR, W.H. (1993): New
data on the ore mineralogy of the Upulungos mine,
La Mejicana district, Sierra de Famatina,
Argentina. In Current Research in Geology
Applied to Ore Deposits (P. Fenoll Hach-AF, J.
Torres-Ruiz & F. Gervilla, eds.). Proceedings 2nd
Biennial Soc. Geol. Applied to Mineral Deposits
Meeting, Granada, Spain, 9-11 September, 1993,
57-59.
BRUHA, D.J & NOBLE, D.C. (1983): Hypogene
quartz-alunite pyrite alteration formed by
moderate saline, ascendant hydrothermal solutions.
Geol. Soc. Am. Abstr. Programs 15-5, 325.
BURBANK, W.S. (1941): Structural controls of ore
deposition in the Red Mountain, Sneffels and
Telluride districts of the San Juan Mountain,
Colorado. Sci. Soc. Colorado Proceedings 14-5,
141-261.
CAMUS, F. (1990): The geology of hydrothermal gold
deposits in Chile. J. Geochem. Explor. 36, 197232.
CECIONI, A.J. & DICK, L.A. (1992): Geolog'a del
yacimiento epitermal de oro y plata Can Can,
Franja de Maricunga, Precordillera de Copiapo,
Chile. Revista Geol. Chile 19.1, 3-17.
CHEN, J.C. & HUH, C.A. (1982): Geochemistry of
dacites from Chinkuashih area, northeastern
Taiwan. Geol. Soc. China [Taiwan] Proceedings
25,67-81.
CHRISTENSON, B.W. & WOOD, C.P. (1993):
Evolution of a vent-hosted hydrothermal system
beneath Ruapehu Crater Lake, New Zealand. Bull.
Volcanol. 55, 547-565.

CORDERY, G. (1986): Epithermal alteration zonation


at Peak Hill. Geol. Soc. Australia Abstr. 18, 1-8.
CORN, R.M. (1975): Alteration-mineralization zoning,
Red Mountain, Arizona. Econ. Geol. 70, 14371447.
CUITINO, L., D'AZ, S. & PUIG, A. (1988): Aspectos
de la mineralogia, geoqu'mica, y geotermometr'a
de los yacimientos epitermales Guanaco y Cachinal
de la Sierra, Antofagasta, Chile. Cong. Geol.
Chileno, 5th, Santiago, 1988, Actas 1, B273-B298.
CUITINO, L., MOSCOSO, R. & MAKSAEV, V.
(1994): Aspectos mineralogicos y termometricos
del prospecto Esperanza-Cerros Bravos. Maricunga, III Region, Chile. Cong. Geol. Chileno, 7th,
Concepcion, 1994, Chili, Actas 1, 771-775CUNNINGHAM, C.G. (1985): Characteristics of
boiling-water-table and carbon dioxide models for
epithermal gold deposition. In Geologic Characteristics of Sediment- and Volcanic-hosted
Disseminated Gold Deposits - Search for an
Occurrence Model (E.W. Tooker, ed.). U.S. Geol.
Surv. Bull. 1646,43-46.
CUNNINGHAM, C.G., RYE, R.O., STEVEN, T.A. &
MEHNERT, H.H. (1984): Origins and exploration
significance of replacement and vein-type alunite
deposits in the Marysvale volcanic field, west
central Utah. Econ. Geol. 79, 50-71.
DAY, A.L. & ALLEN, E.T. (1925): The volcanic
activity and hot springs of Lassen Peak. Carnegie
Inst. Washington Publication 360.
DEEN, J.A. (1990): Hydrothermal Ore Deposition
Related to High-level Igneous Activity: A Stableisotopic Study of the Julcani Mining District, Peru.
Ph.D. thesis, Univ. Colorado, Boulder, Colorado.
DELMELLE, P. & BERNARD, A. (1994): Geo-

445

A. Arribas, Jr.

chemistry, mineralogy, and chemical modeling of


the acid crater lake of Kawajh Ijen volcano,
Indonesia, Geochim. Cosmochim. Acta 58, 24452460.
DILLES, J.H., SOLOMON, G.C., TAYLOR, H.P., JR.
& EfNAUDI, M.T. (1992): Oxygen and hydrogen
isotope characteristics of hydrothermal alteration at
the Ann-Mason porphyry copper deposits,
Yerington, Nevada. Econ. Geol. 87,44-63.
DOE, B.R., STEVEN, T.A., DELEVAUX, M.H.,
STACEY, J.S., LIPMAN, P.W. & FISHER, F.S.
(1979): Genesis of ore deposits in the San Juan
volcanic field, southwest Colorado - lead isotope
evidence. Econ. Geol. 74, 1-26.
EINAUDI, M.T. (1977): Environment of ore deposition
at Cerro de Pasco, Peru. Econ. Geol. 72, 893-924.
ERCEG, M.M., CRAIGHEAD, G.A., HALFPENNY,
R. & LEWIS, P.J. (1991): The exploration history,
geology and metallurgy of a high sulphidation
epithermal gold deposit at Wafi River, Papua New
Guinea. In PNG Geology, Exploration and
Metallurgy Conference, Proceedings (R. Rogerson,
ed.). Rabaul, 1991, Australasian Instit. Mining and
Metallurgy, Parkville, 58-64.
ERICKSEN, G.E. & CUNNINGHAM, C.G. (1993):
Precious-metal deposits in the Neogene-Quaternary
volcanic complex of the Central Andes. In
Investigaciones de Metales Preciosos en el
Complejo Volcanico Neogeno-Cuaternario de los
Andes Centrales. Servicio Geoldgico de Bolivia,
La Paz, Bolivia, 1-16.
FIRST, D.M.X1993): Precious metal and Cu-Sn sulphosalt mineralogy of the Lahoca acid-sulphate
deposit, Recsk, Hungary. In Field Conference on
Plate Tectonic Aspects of Alpine Metallogeny in
the Carpatho-Balkan Region, Hungarian Geol.
Surv., Budapest, Hungary, May 1993, p. 35.

135, 323-335.
FOURNIER, R.O. (1987): Conceptual models of brine
evolution in magmatic-hydrothermal systems. U. S.
Geol. Surv. Prof. Paper 1350, 1487-1506.
FOURNIER, R.O. (1992): The influences of depth of
burial and the brittle-plastic transition on the
evolution of magmatic fluids. Geol. Surv. Japan
Report 279, 57-59.
GARCIA, J.S. (1991): Geology and mineralization
characteristics of the Mankayan mineral district,
Philippines. Geol. Surv. Japan Report 111, 21-30.
GIGGENBACH, W.F. (1987): Redox processes
governing the chemistry of fumarolic gas
discharges from White Island, New Zealand.
Applied Geochem. 2, 143-161.
GIGGENBACH, W.F. (1992a): Magma degassing and
mineral deposition in hydrothermal systems along
convergent plate boundaries. Econ. Geol. 87, 19271944.
GIGGENBACH, W.F. (1992b): Isotopic shifts in
waters from geothermal and volcanic systems
along convergent plate boundries. Earth Planet.
Sci.Letl. 113,495-510.
GONZALEZ, A.G. (1959): Geology and Genesis of the
Lepanto Copper Deposit, Mankayan, Mountain
Province, Philippines. Ph.D. thesis, Stanford Univ.,
Stanford, California.
GRATON, L.C. & BOWDITCH, S. (1936): Alkaline
and acid solutions in hypogene zoning at Cerro de
Pasco. Econ. Geol. 31, 651-698.
GRAY, J.E.& COOLBAUGH,M. F. (1994): Geology
and geochemistry of Summitville, Colorado: An
epithermal acid-sulfate deposit in a volcanic dome.
Econ. Geol. 89 (accepted).

FISHER, F.S. & LEEDY, W.P. (1973): Geochemical


characteristics of mineralized breccia pipes in the
Red Mountain district, San Juan Mountains,
Colorado. U.S. Geol. Surv. Bull. 1381.

GROPPER, H., CALVO, M., CRESPO, H., BISSO,


C.R., CUADRA, W.A., DUNKERLEY, P.M. &
AGUIRRE, E. (1991): The epithermal gold-silver
deposit of Choquelimpie, Northern Chile. Econ.
Geol. 86, 1206-1221.

FOLINSBEE,
R.E.,
KIRKLAND,
K.,
NEKOLAICHUK, A. & SMEJKAL, V. (1972):
Chinkuashih-a gold-pyrite-enargite-barite hydrothermal deposit in Taiwan. Geol. Soc. Am. Memoir

GUSTAFSON, L.B. & HUNT, J.D. (1975): The


porphyry copper deposit at El Salvador, Chile.
Econ Geol. 70, 857-912

High-sulfidation Epithermal Deposits


HALL, R.B. (1978): World nonbauxite aluminum
resources-Alunite. U. S. Geol. Surv. Prof. Paper
1076-A.
HALL, R.J., BRITTEN, R.M. & HENRY, D.D. (1990):
Frieda River copper-gold deposits. In Geology of
the Mineral Deposits of Australia and Papua New
Guinea 2 (F.E. Hughes, ed.). Australasian Inst.
Mining Metall. Monogr. 14, 1709-1716.
HALLBERG, A. (1994): The Enasen gold deposit,
central Sweden. 1. A paleoproterozoic highsulfidation epithermal mineralization. Mineral.
Deposita 29, 150-162.
HARBON, P. (1988): Peak Hill project, Peak Hill, New
South Wales. In Bicentennial Gold 88 Core Shed
Guidebook (M.S. Bloom & P.J.Parrington, eds.).
Geol. Soc. Australia, p. 30.
HAYBA, D.O., BETHKE, P.M., HEALD, P. &
FOLEY, N.K. (1985): Geologic, mineralogic, and
geochemical characteristics of volcanic-hosted
epithermal precious-metal deposits. Reviews Econ.
Geology 2, 129-167.
HEALD, P., FOLEY, N. K. & HAYBA, D.O. (1987):
Comparative
anatomy
of volcanic-hosted
epithermal deposits: acid-sulfate and adulariasericite types. Econ. Geol. 82, 1-26.
HEDENQUIST, J.W. (1987): Mineralization associated
with volcanic-related hydrothermal systems in the
Circum-Pacific basin. In Transactions of the Fourth
Circum-Pacific Energy and Mineral Resources
Conference (M.K. Horn, ed.). August, 1986,
Singapore. Am. Assoc. Petroleum Geol., Tulsa,
Oklahoma, 513-524
HEDENQUIST, J.W. & AOKI, M. (1991): Meteoric
interaction with magmatic discharges in Japan and
the significance for mineralization. Geology 19,
1041-1044.
HEDENQUIST J.W., AOKI, M. & SHINOHARA, H.
(1994b): Flux of volatiles and ore-forming metals
from the magmatic-hydrothermal system of
Satsuma Iwojima volcano. Geology 22, 585-588.
HEDENQUIST, J.W. & GARCIA J.S., JR. (1990):
Sulfur isotope systematics in the Lepanto mining
district, northern Luzon, Philippines. Mining Geol.
40, p. 67.

HEDENQUIST, J.W. & HENLEY, R.W. (1985): The


importance of COz on freezing point measurements
of fluid inclusions; evidence from active
geothermal systems and implications for
epithermal ore deposition. Econ. Geol. 80, 13791406.
HEDENQUIST, J.W. & LOWENSTERN, J.B. (1994):
The role of magmas in the formation of
hydrothermal ore deposits. Nature 370, 519-527.
HEDENQUIST, J.W., MATSUHISA, Y., IZAWA, E.,
WHITE, N.C., GIGGENBACH, W.F. & AOKI, M.
(1994a): Geology and geochemistry of highsulfidation Cu-Au mineralization in the Nansatsu
district, Japan. Econ. Geol. 89, 1-30.
HEDENQUIST,
J.W.,
SIMMONS,
S.F.,
GIGGENBACH, W.F. & ELDRIDGE, C.S.
(1993): White Island, New Zealand, volcanichydrothermal system represents the geochemical
environment of high-sulfidation Cu and Au ore
deposition. Geology 21, 731-734.
HEMLEY, J.J., CYGAN, G.L., FEIN, J.B.,
ROBINSON, G.R. & D'ANGELO, W.M. (1992):
Hydrothermal ore-forming processes in the light of
studies in rock-buffered systems: I. Iron-copperzinc-lead sulfide solubility relations. Econ. Geol.
87, 1-22.
HEMLEY, J.J., HOSTETLER, P.B., GUDE, A.J. &
MOUNTJOY, W.T. (1969): Some stability relations of alunite. Econ. Geol. 64, 599-612.
HEMLEY, J.J., MONTOYA, J.W., MARINENKO,
J.W. & LUCE, R.W. (1980): Equilibria in the
system Al203-Si02.H20 and some general
implications for alteration/mineralization processes. Econ. Geol. 75, 210-228.
HENLEY, R.W. (1991): Epithermal gold deposits in
volcanic terranes. In Gold Metallogeny and
Exploration (R.P. Foster, ed.). Blackie & Son Ltd.,
London, UK, 133-164.
HENLEY, R.W. & ELLIS, A.J. (1983): Geothermal
systems, ancient and modem: A geochemical
review. Earth Sci. Reviews 19, 1-50.
HENLEY, R.W. & McNABB, A. (1978): Magmatic
vapor plumes and ground-water interaction in
porphyry copper emplacement Econ Geol. 73,1-20.

447

A. Arribas, Jr.

HUANG, C.K. (1955): Gold-copper deposits of the


Chinkuashih mine, Taiwan, with special reference
to the mineralogy. Acta Geol. Taiwanica 7, 1-20.
HWANG, J.Y. & MEYER, H.O.A. (1982): The mineral
chemistry and genesis of the Chinkuashih ore
deposits, Taiwan. Geol. Soc. China [Taiwan]
Proceedings 25, 88-101.
ITO, T. (1969): Geology and ore deposits of the Teine
mine in Hokkaido. Mining Instil. Hokkaido J. 25,
30-37.
IWAO, S. (1962): Silica and alunite deposits of the
Ugusu mine; a geochemical consideration on an
extinct geothermal area in Japan. Japanese J. Geol.
Geography33, 131-144.
IZAWA, E. & CUNNINGHAM, C.G. (1989): Hydrothermal breccia pipes and gold mineralization in
the Iwashita ore body, Iwato Deposit, Kyushu,
Japan. Econ. Geol. 84, 715-724.
JANKOVIC, S. (1982): Yugoslavia. In Mineral
Deposits of Europe, Vol. 4/5, Southwest and
Eastern Europe (F. W. Dunning, W. Mykura & D.
Slater.eds.). Instit. Mining Metall., London, 143202.
JANKOVIC, S., TERZIC, M., ALEKSIC, D.,
KARAMATA, S., SPASOV, T., JOVANOVIC,
M., MILICIC, M., MISKOVIC, V., GRUBIC, A.
& ANTONIJEVIC, I. (1980): Metallogenic features of copper deposits in the volcano-intrusive
complexes of the Bor district, Yugoslavia. In
European Copper Deposits (S. Jankovic & R.H.
Sillitoe, eds.). Fac. Mining Geol., Belgrade Univ.,
Belgrade, Soc. Geol. Applied to Mineral Deposits
Special Pub. 1,42-49.
JANNAS, R.R., BEANE, R.E., AHLER, B.A. &
BROSNAHAN, D.R. (1990): Gold and copper
mineralization at the El Indio Deposit, Chile. J.
Geochem. Explor. 36, 233-266.
JENSEN, M.L., ASHLEY, R.P. & ALBERS, J.P.
(1971): Primary and secondary sulfates at
Goldfield, Nevada. Econ. Geol. 66, 618-626.
JIMENEZ, N., LIZECA, J.L., MURILLO, F.,
SANJINES, O., BARRERA, L. & FLORES, O.
(1993): Marco geologico del distrito minero de
Laurani. In Investigaciones de Metales Preciosos
en el Complejo Volcanico Neogeno-Cuaternario de
AA<f

los Andes Centrales. Servicio Geologico de


Bolivia, La Paz, Bolivia, 123-130.
JOHN, D.A., NASH, J.T., CLARK, C.W. &
WULFTANGE, W.H. (1991): Geology, hydrothermal alteration, and mineralization at the
Paradise Peak gold-silver-mercury deposit, Nye
County, Nevada. In Geology and Ore Deposits of
the Great Basin. Symposium Proceedings (G.L.
Raines, RE. Lisle, R.W. Schafer & W.H.
Wilkinson, eds.). Geol. Soc. Nevada, Reno, 10201050.
KEITH, S.B. LAUX, D.P., MAUGHAN, J., SCHAWB,
K., RUFF, S., SWAN, M.M., ABBOTT, E. &
FRIBERG, S. (1991): Magma series and
metallogeny: A case study from Nevada and
environs. In Geology and Ore Deposits of the
Great Basin, Field Trip Guidebook and
Compendium 1 (R.H. Buffa & A.R. Coyner, eds.).
Geol. Soc. Nevada, Reno, 404-493.
KESLER, S.E., RUSSELL, N., SEAWARD, M
RIVERA, J., MCCURDY, K, CUMING, G.L. &
SUTTER, J.F. (1981): Geology and geochemistry
of sulfide mineralization underlying the Pueblo
Viejo gold-silver oxide deposit, Dominican
Republic. Econ. Geol. 76, 1096-1117.
KIYOSU, Y. & KURAHASHI, M. (1983): Origin of
sulfur species in acid sulfate-chloride thermal
waters, northeastern Japan. Geochim. Cosmochim.
Acta 47, 1237-1245.
KOUKHARSKY, M. & MIRRE J.C. (1976): Mi Vida
prospect: A porphyry copper-type deposit in
northwestern Argentina. Econ. Geol. 71, 849D863.
LEACH T.M. & ERCEG M. (1990): The Wafi River
high sulfidation epithermal Au deposit, Papua New
Guinea. Pacific Rim Congress, Gold Coast,
Australia, 1990, Proceedings 2,451-456.
LIU, K. & EPSTEIN, S. (1984): The hydrogen isotope
fractionation between kaolinite and water. Isotope
Geosci. 2, 335-350.
L O S A D A - C A L D E R O N , A. & McPHAIL, D. (1994):
The Nevados del Famatina mining district:
porphyry- and epithermal-style mineralization, La
Rioja Prov., Argentina. 7th Congreso Geologico
Chileno, Conception, Chile, Actas 2, 1585-1589.
LOSADA-CALDER6N,

A.J., McBRIDE, S.L. &

High-sulfidation Epithermal Deposits

BLOOM, M.S. (1994): The geology and " W A r


geochronology of magmatic activity and related
mineralization in the Nevados del Famatina mining
district, La Rioja province, Argentina. J. South
American Earth Sci. 7, 9-24.
LOWENSTERN, J.B. (1994): Dissolved volatile
concentrations in an ore-forming magma. Geology
22, 893-896.
MAKSAEV, J.V., MOSCOSO, D.R., MPODOZIS,
M.C. & NASI, P.C. (1984): Las unidades
volcanicas y plut6nicas del Cenozoico Superior en
la Alta Cordillera del Norte Chico (29-31S):
Geologia, alteraci6n hidrotermal y mineralizacion.
Revista Geol. Chile 21,11-51.
MANCANO, D.P. & CAMPBELL, A.R. (1995):
Microthermometry
of enargite-hosted
fluid
inclusions from the Lepanto, Philippines, highsulfidation
Cu-Au deposit using
infrared
microscopy. Geochim. Cosmochim. Acta
59
(accepted).

(The Graton-Sales Volume) 1933-1967 2, Am.


Inst. Mining Metall. Petroleum Engineers, New
York, 1373-1416.
MITCHELL, A.H.G. (1992): Andesitic arcs, epithermal
gold and porphyry-type mineralization in the
western Pacific and eastern Europe. Trans. Inst.
Mining Metall. 101, B125-B138.
MITCHELL, A.H.G. & LEACH, T.M. (1991):
Epithermal Gold in the Philippines: Island Arc
Metallogenesis, Geothermal Systems and Geology.
Academic Press Inc., London, UK, 457 p.
MOSCOSO, R., MAKSAEV, V., CUITINO, L., DIAZ,
F.,
KOEPPEN,
R.P.,
TOSDAL,
R.M.,
CUNNINGHAM, C.G., MCKEE, E.H. &
RYTUBA, J.J. (1993): El complejo volcanico
Cerros Bravos, regi6n de Maricunga, Chile:
Geologia, alteracion hidrotermal, y mineralizacion.
In Investigaciones de Metales Preciosos en el
Complejo Volcnico Neogeno-Cuatemario de los
Andes Centrales. Servicio Geologico de Bolivia,
La Paz, Bolivia, 131-165.

MARUMO, K., NAGASAWA, K. & KURODA, Y.


(1980): Mineralogy and hydrogen isotope
geochemistry of clay minerals in the Ohnuma
geothermal area, northeastern Japan. Earth Planet.
Sci. Lett. 47,255-262.

MULLER D. & GROVES, D.I. (1993): Direct and


indirect associations between potassic igneous
rocks, shoshonites and gold-copper deposits. Ore
Geol. Reviews 8, 383-406.

MASTERMAN, G.J. (1994): High Sulphidation GoldCopper Mineralisation at Peak Hill, New South
Wales: Origin, Alteration Zoning and Deformation
History. B.Sc. thesis, Earth Sci. Department, Univ.
Melbourne, Australia.

MUNTEAN, J.L., KESLER, S.E., RUSSELL, N. &


POLANCO, J. (1990): Evolution of the Monte
Negro acid-sulfate Au-Ag deposit, Pueblo Viejo,
Domincan Republic: Important factors in grade
development. Econ. Geol. 85, 1738-1758.

MATSUHISA, Y. (1992): Origin of magmatic waters


in subduction zones: stable isotope contraints.
Geol. Surv. Japan Report 279, 104-109.

MURILLO, F., SANJINES, O., BARRERA, L.,


JIMENEZ, N., LIZECA, J.L., FLORES, O.,
HOFFSTRA, A.H., HARDYMAN, R.M. &
NASH, T.J. (1993): Geologia, alteracion, y
mineralizaci6n del depbsito mineral tipo sulfatodcido de Laurani, Altiplano Norte de Bolivia. In
Investigaciones de Metales Preciosos en el
Complejo Volcanico Neogeno-Cuaternario de los
Andes Centrales. Servicio Geologico de Bolivia,
La Paz, Bolivia, 123-130.

MENHERT, H.H., LIPMAN, P.W. & STEVEN, T.A.


(1973): Age of mineralization at Summitville,
Colorado, as indicated by K-Ar dating of alunite.
Econ. Geol. 68, 399-401.
MEYER, C. & HEMLEY, J.J. (1967): Wall rock
alteration. In Geochemistry of Hydrothermal Ore
Deposits (H.L. Barnes, ed.). Holt, Rinehart &
Winston, New York, 166-235.
MEYER, C , SHEA, E.P., GODDARD JR., C.C. &
Staff (1968): Ore deposits at Butte, Montana. In
Ore Deposits of the United States (J.D. Ridge, ed.),

NOBLE, D.C. & SILBERMAN, M.L. (1984):


Evoluci6n volcanica e hidrotermal y cronologia de
K-Ar del distrito minero de Julcani, Peru. Soc.
Geol. Peru, Vol. Jubilar, 60thAnniv. 5, 1-35.
OVIEDO, L., FUSTER, N., TSCHISCHOW, N.,

449

A. Arribas, Jr.

RIBBA, L., ZUCCONE, A., GREZ, E. &


AGUILAR, A. (1991): General geology of La
Coipa precious metal deposit, Atacama, Chile.
Econ. Geol. 86, 1287-1300.
PANTELEYEV, A. & KOYANAGI, V.M. (1994):
Advanced argillic alteration in bonanza volcanic
rocks, northern Vancouver Island-lithologic and
permeability controls. In Geological Fieldwork
1993 (B. Grant & J.M. Newell, eds.). Brit. Col.
Ministry of Energy Mines Petroleum Resources
Paper 1994-1, 101-110.
PERELL6, J.A. (1994): Geology, porphyry Cu-Au,
and epithermal Cu-Au-Ag mineralization of the
Tombulilato district, North Sulawesi, Indonesia. J.
Geochem. Explor. 50,221 -256.
PETERSEN, U., NOBLE, D.C., ARENAS, M. J. &
GOODELL, P.C. (1977): Geology of the Julcani
mining district, Peru. Econ. Geol. 72, 931-949.
PUIG, A., DIAZ, S. & CUITINO, L. (1988): Sistemas
hidrotermales asociados a calderas en el arco
volcanico Pale6geno de la region de Antofagasta,
Chile: distritos El Guanaco, Cachinal de la Sierra y
El Soldado. Revista Geol. Chile 15.1, 57-82.
RAETZ, M.C. & PARRINGTON, P.J. (1988): Rhyolite
Creek, Victoria, A Lower Paleozoic epithermal
gold prospect. Bicentennial Gold 88, Geol. Soc.
Australia, Abstracts 23, 291-294.
RANSOME, F.L. (1907): The association of alunite
with gold in the Goldfield district, Nevada. Econ.
Geol. 2, 667-692.
RANSOME, F.L. (1909): The geology and ore deposits
of Goldfield, Nevada: U. S. Geol. Surv. Prof.
Paper 66.
REN, Q., ZHANG, C , YANG, R., XIE, X. & XU., Z.
(1992): Hydrothermal systems related to
epithermal gold deposits in Mesozoic volcanic
areas in eastern China. In Proceedings Intemat
Symposium on Water-Rock Interaction, 7th, July
1992, Park City, Utah, 1609-1611.
REYES, A.G. (1990): Petrology of Philippine
geothermal systems and the application of
alteration mineralogy to their assessment. J.

4sn

Volcano!. Geotherm. Research 43, 279-309.


REYES, A.G., GIGGENBACH, W.F., SALERAS,
J.R.M., SALONGA, N.D. & VERGARA, M.C.
(1993): Petrology and geochemistry of Alto Peak,
a vapor-cored hydrothermal system, Leyte
Province, Philippines. Geothermics 22, 479-519.
ROWE, G.L. (1994): Oxygen, hydrogen, and sulfur
isotope systematics of the crater lake system of
Poas Volcano, Costa Rica. Geochem. J. 28, 263287.
RUGGIERI, G. (1993a): Ore genesis in epithermal
environment: the examples of Furtei (Sardinia),
Frassine and La Campigliola (Southern Tuscany)
gold mineralization. Plinius Supplemento italiano
al'European J. Mineral. 9, 134-139.
RUGGIERI, G. (1993b): Minerogenesi in Ambiente
Epitermale: Gli Esempi Delle Mineralizzazioni
Aurifere di Furtei (Sardegna), Frassine e La
Campigliola (Toscana Meridionale). Ph.D. thesis,
Univ. Firenze, Firenze, Italy.
RUSSELL, N. & KESLER, S.E. (1991): Geology of
the maar-diatreme complex hosting precious metal
mineralization at Pueblo Viejo, Dominican
Republic. Geol. Soc. Am. Special Paper 262, 203216.
RYE, R.O. (1993): The evolution of magmatic fluids in
the epithermal environment: The stable isotope
perspective. Econ. Geol. 88, 733-753.
RYE R.O., BETHKE, P.M. & WASSERMAN, M.D.
(1992): The stable isotope geochemistry of acidsulfate alteration. Econ. Geol. 87, 225-262
RYE, R.O., STOFFREGEN, R. & BETHKE, P.M.
(1990): Stable isotope systematics and magmatic
and hydrothermal processes in the Summitville,
CO gold deposit. U.S. Geol. Surv. Open-file Report
90-626.
RYTUBA, J.J., ARRIBAS, A. JR., CUNNINGHAM,
C.G., MCKEE, E.H., PODWYSOCKI, M.H.,
SMITH, J.G., KELLY, W.C. & ARRIBAS, A.
(1990): Mineralized and unmineralized calderas in
Spain; Part II, evolution of the Rodalquilar caldera
complex and associated gold-alunite deposits,
Mineral. Deposita 25 (Suppl.), S29-S35.

High-sulfidation Epithermal Deposits


SAKAI, T. & MATSUBAYA, 0. (1977): Stable
isotopic studies of Japanese geothermal systems.
GeothermicsS, 97-124.

SILLITOE, R.H. (1989): Gold deposits in the western


Pacific Island arcs: The magmatic connection.
Econ. Geol. Monogr. 6, 274-291.

SANGER-VON OEPEN, P., FRIEDRICH, G. &


VOGT, J.H. (1989): Fluid evolution, wallrock
alteration, and ore mineralization associated with
the Rodalquilar epithermal gold-deposit in
southeast Spain. Mineral. Deposila 24, 235-243.

SILLITOE, R.H. (1991a): Gold metallogeny of Chile an introduction. Econ. Geol. 86, 1187-1205.

SASAKI, A., ARIKAWA, Y. & FOLINSBEE, R.E.


(1979): Kiba reagent method of sulfur extraction
applied to isotope work. Geol. Surv. Japan Bull.
30,241-245.
SEWARD, T.M. (1973): Thio complexes of gold and
the transport of gold in hydrothermal ore solutions.
Geochim. Cosmochim, Acta 37, 379-399.
SHELNUTT, J.P. & NOBLE, D.C. (1985): Premineralization radial dikes of tourmalinized
fluidization breccia, Julcani district, Peru. Econ.
Geol. 80, 1622-1632.
SHTNOHARA, H. (1994): Exsolution of immiscible
vapor and liquid phases from a crystallizing silicate
melt: Implications for chlorine and metal transport.
Geochim. Cosmochim. Acta 58, 5215-5221.
SIDDELEY, G. & ARANEDA, R. (1986): The El
Indio-El Tambo gold deposits, Chile. In Gold '86
(A.J. Macdonald, ed.). Konsult International,
Willowdale, Ontario, 445-456.
SIDDELEY, G. & ARANEDA, R. (1990): Gold-silver
occurrences of the El Indio belt, Chile. Earth. Sci.
Ser., Circum-Pacific Council Energy Mineral
Resources 11, 273-284.
SILLITOE, R.H. (1973): The tops and bottoms of
porphyry copper deposits. Econ. Geol. 68, 799815.
SILLITOE, R.H. (1983): Enargite-bearing massive
sulfide deposits high in porphyry copper systems.
Econ. Geol. 78, 348-352.
SILLITOE, R.H. (1988): Gold and silver deposits in
porphyry systems. In Proceedings Bulk Mineable
Precious Metal Deposits of the Western United
States, Symposium (R.W. Schafer, J.J. Cooper &
P.G. Vikre, eds.). Reno, Nevada 1987, Geol. Soc.
Nevada, Reno, Nevada, 233-257.

SILLITOE, R.H. (1991b): Intrusion-related gold


deposits. In Gold Metallogeny and Exploration
(R.P. Foster, ed.). Blackie & Son Ltd., London,
UK, 165-209.
SILLITOE, R.H. (1993): Giant and bonanza gold
deposits in the epithermal environment. In Giant
Ore Deposits (B.H. Whiting, C. J. Hodgson & R.
Mason, eds.). Soc. Econ. Geol. Special Publication
2,125-156.
SILLITOE, R.H., ANGELES, C.A., COMIA, G.M.,
ANTIOQUIA, E.C. & ABEYA, R.E. (1990): An
acid-sulphate type lode gold deposit at Nalesbitan,
Luzon, Philippines. J. Geochem. Explor. 35, 387412.
SILLITOE, R.H. & LORSON, R.C. (1994): Epithermal
gold-silver-mercury deposits at Paradise Peak,
Nevada: Ore controls, porphyry gold association,
detachment faulting and supergene oxidation.
Econ. Geol. 89, 1228-1248.
STAUDE, J.-M.G. (1994): Acid sulfate gold systems of
the Mulatos district, Sonora: northern Mexico's
largest gold system. Geol Soc. Am. Abstr.
Programs 26, A42.
STEVEN, T.A. & RATTE J.C. (1960): Geology and
ore deposits of the Summitville district, San Juan
Mountains, Colorado. U. S. Geol. Surv. Prof.
Paper 343.
STOFFREGEN, R.E. (1987): Genesis of acid-sulfate
alteration and Au-Cu-Ag mineralization at
Summitville, Colorado. Econ. Geol. 82, 15751591.
STOFFREGEN, R.E. & ALPERS, C.N. (1987):
Woodhousite and svanbergite in hydrothermal ore
deposits: products of apatite destruction during
advanced argillic alteration. Can. Mineral. 25,201211.
STOFFREGEN, R.E., RYE, R.O. & WASSERMAN,
D.M. (1994): Experimental studies of alunite I:

451

A. Arribas, Jr.
I8

0- I 6 0 and D-H fractionation factors between


alunite and water at 250-450C. Geochim.
Cosmochim. Acta 58, 903-916.
TAN, L.P., YU, B.S. & KUO, C.L. (1993):
Geochemical zonations of the Chinkuashih goldcopper deposits, Taiwan. Resource Geol. Special
Issue 16, 95-106.
TAYLOR, B.E. (1986): Magmatic volatiles: isotopic
variation of C, H, and S. In Stable Isotopes in High
Temperature Geological Processes (J.W. Valley,
H.P. Taylor, Jr. & J.R. O'Neil, eds.). Reviews
Mineral. 16, 185-226.
TAYLOR, B.E. (1992): Degassing of H 2 0 from
rhyolite magma during eruption and shallow
intrusion, and the isotopic composition of
magmatic water in hydrothermal systems. Geol.
Surv. Japan Report 279, 190-194.
THOMPSON, J.F.H., ABIDIN, H.Z., BOTH, R.A.,
MARTOSUROYO, S., RAFFERTY, W.J. &
THOMPSON, A.J.B. (1994): Alteration and
epithermal mineralization in the Masupa Ria
volcanic center, Central Kalimantan, Indonesia. J.
Geochem. Explor. 50, 429-456.
THOMPSON, J.F.H., LESSMAN, J. & THOMPSON,
A.J.B. (1986): The Temora gold-silver deposit: A
newly recognized style of high
sulfur
mineralization in the Lower Paleozoic of Australia.
Econ. Geol. 81, 732-738.
TOWNLEY, B. (1991): Evolucidn y Zonacidn de la
Mineralizacidn de Au y Ag, en el Sistema
Epitermal de Can Can, III Regidn. B.S. thesis,
Univ. de^Chile.
TURNER, S. (1986): Fluid inclusion, alteration and ore
mineral studies of an epithermal vein system:
Mount Kasi, Vanua Leva, Fiji. In Proceedings
Intemat. Volcanological Congress, Symposium 5,
Hamilton, New Zealand 1986. Univ. Auckland,
Centre Continuing Education, Auckland, New
Zealand, 87-94.
UEDA, A. & SAKAI, H. (1984): Sulfur isotope study
of Quaternary volcanic rock from the Japanese
Island Arc. Geochim. Cosmochim. Acta 48, 18371848.
VELINOV, I. & KANAZIRSKI, M. (1990): For-

452

mational nature and physico-chemical analysis of


mineral parageneses in the metasomatic zones of
acid leaching in the western Srednogorie. Geol.
Balcanica 20.3, 59-71.
VELINOV, I., KANAZIRSKI, M. & KUNOV, A.
(1990): Formational nature and physico-chemical
conditions of formation of metasomatites in the
Spahievo ore field (Eastern Rhodopes, Bulgaria).
Geologica Balcanica 20.4,49-62.
VENNEMANN, T.W., MUNTEAN, J.L., KESLER,
S.E., O'NEIL, J.R., VALLEY, J.W. & RUSSELL,
N. (1993): Stable isotope evidence for magmatic
fluids in the Pueblo Viejo epithermal acid sulfate
Au-Ag deposit, Dominican Republic. Econ. Geol.
88,55-71.
VIDAL, C.E. & CEDILLO, E. (1988): Los yacimientos
de enargita-alunita en el Peru. Soc. Geol. Peru Bol.
78, 109-120.
VIDAL, C , MAYTA, O., NOBLE, D.C. & MCKEE,
E.H/ (1984): Sobre la evolution de soluciones
hidrotermales
desde
el centro
volcanico
Marcapunta en Colquijirca-Pasco. Soc. Geol.Peru,
Vol. Jubilar Homenaje Dr. G. Petersen, Fasc. 10,
1-14.
VIDAL, C. E., NOBLE, D.C, MCKEE, E.H.,
BENAVIDES, J.E. & H.C. DE LOS RIOS, M.
(1989): Hydrothermally altered and mineralized
Late Pliocene-Quaternary central volcanoes in the
Andes of southern Peru. Internat. Geol. Congress,
28th, Washington, 3, p. 297.
VIKRE, P.G. (1989): Ledge formation at the Sandstorm
and Kendall gold mines, Goldfield, Nevada. Econ.
Geol. 84,2115-2138.
VILA, T. (1991): Epithermal silver-gold mineralization
at the Esperanza area, Maricunga belt, high Andes
of northern Chile. Revista Geol. Chile 8, 37-54.
WALLACE, A.B. (1979): Possible signatures of buried
porphyry-copper deposits in middle to late Tertiary
volcanic rocks of western Nevada. In Proceedings
Intemat. Assoc, on the Genesis of Ore Deposits,
5th Symposium. Nevada Bur. Mines Geol. Report
33,69-76.
WASSERMAN, M. D., RYE, R. O., BETHKE, P. M.
& ARRIBAS, A., JR. (1992): Methods for

High-sulfidation Epithermal Deposits


separation and total stable isotope analysis of
alunite. U.S. Geol. Surv. Open-file Report 92-9.
WHITE, N.C. (1991): High sulfidation epithermal gold
deposits: Characteristics, and a model for their
origin. Geol. Surv. Japan Report 227, 9-20.
WHITE, N.C. & HEDENQUIST, J.W. (1990):
Epithermal environments and styles of mineralization: variations and their causes, and guidelines
for exploration. J. Geochem. Explor. 36,445-474.
WHITE, N.C. & HEDENQUIST, J.W. (1995): Epithermal gold deposits: styles, characteristics and
exploration. Soc. Econ. Geol. Newsletter 21,
(accepted).
WHITE, N.C, LEAKE, M.J., McCAUGHEY, S.N. &
PARRIS, B.W. (1995): Epithermal deposits of the
southwest Pacific. J. Geochem. Explor. (accepted).
WHITNEY, J.A. (1988): Composition and activity of
sulfurous species in quenched magmatic gases
associated with pyrrhotite-bearing silicic systems.
Econ. Geol. 83, 86-92.
YEN, C.C. (1976): Trapping temperature and pressure
of the fluid inclsuions in the gangue minerals of
gold-silver-copper deposits at Chinkuashih,
Taiwan. Geol. Soc. China [Taiwan] Proceedings
19, 127-133.
YOON, C.H. (1994): Gold content variations in the
acid-sulfate alteration zone of the Seongsan and
Ogmaesan clay deposits in Naenam area, Korea.
Resource Geol. 44, 277.
YUI, S. & MATSUEDA, H. (1994): Several mineral
deposits in Saku-Machi, Nagano prefecture.
Resource Geol. 44, 305.
ZHANG, D. LI, D., ZHAO, Y., CHEN, J., LI, Z. &
ZHANG, K. (1994): The Zijinshan deposit: the
first example of quartz-alunite type epithermal
deposit in the continent of China. Resource Geol.
44, 93-99.

APPENDIX 1
Summary of data and references used to compile
Figures 3, 7, and 8.
Figure 3
K 2 0 versus Si0 2 variation diagram. The name
of lithologic units analyzed, number of samples
(w), and data sources are given: Chinkuashih,
dacite n - 18 (Chen & Huh 1982); Choquelimpie,
Choquelimpie volcanic complex (5 units), n = 20
(Gropper et al. 1991; chemical data for the
feldspar porphyries genetically related to
mineralization are not available); Goldfield,
rhyodacite n = 6 (Ransome 1909; Ashley, unpub.
analyses in Sillitoe 1993); El Indio, Cerro de las
Tortolas Formation, n = 15 (Maksaev et al. 1984;
in Sillitoe 1993); Julcani, dacite and rhyodacite, n
= 10 (Petersen et al. 1977); Laurani, Laurani
volcanic and intrusive rocks, n = 10 (Jimenez et
al. 1993); Lepanto, Imbanguila dacite and least
altered quartz diorite porphyry, n = 4 (A. Arribas,
unpub. data); Motomboto, porphyritic intrusions,
n = 10 (Perello 1994, and written comm. 1995);
Nansatsu, Upper Formation and hornblende
andesite in Middle Volcanic rock, n = 2 (E. Izawa,
written comm. 1995); Paradise Peak, average of
Younger andesites, n = 31 (John et al. 1991);
Rodalquilar, hornblende andesite, dacite tuff, and
rhyolite domes, n - 1 (Arribas et al. 1995a);
Summitville, Fisher quartz latite, n = 7 (Steven &
Ratte 1960).
Figure 7
Range of 534S (/oo) values. Given below are
the number of measurements for sulfides ("H2S),
sulfates (so4) sulfide-sulfate mineral pairs
(A34S)> and references: Lepanto, H2S = 52, s 0 4
= 38 (Hedenquist & Garcia 1990; J. Hedenquist &
M. Aoki, unpub. data); Chinkuashih, wH2S = 4>
n S04 = 2, A34S = 2 (Folinsbee et al. 1972);
Nansatsu, H 2 S = 6, n S o 4 = 9 (Hedenquist et al.
1994a); Summitville, H2S = >11, n S 0 4 = 17,
"A34S = 7 (Rye et al. 1990); Goldfield, H2s = 16,
n S04 = 16, A34S = 7 (Jensen et al. 1971; Vikre
1989); Pueblo Viejo, H2s = 19, n S04 = 7, n^s =
4 (Vennemann et al. 1993); Julcani, H2S = l g3

453

A. Arribas, Jr.
"so 4 = 55, A34S = 7 (Deen 1990); El Indio, nH?s =
U> "SO4 = 3 (Jannas et al. 1990), Rodalquilar,
H2S = 44, nS04 = 1 1 , A 3 4 S - 4 (Arribas et al.
1995a). Temperatures for Chinkuashih were
calculated using the 4 S/ S data from Folinsbee et
al. (1972) and more recent fractionation equations.
Sulfide-sulfate mineral temperatures higher than
350 C were documented only at depth at
Summitville (T = 390 C, -900 m below the
present surface; Rye et al. 1990) and Lepanto (7" =
420 C at the 700-m level, immediately above the
FSE porphyry copper deposit; Hedenquist &
Garcia 1990). On the basis of phase equilibria, the
sulfide/sulfate values for the Pueblo Viejo stage 1
and stage 2 mineralization were estimated by
Muntean et al. (1990) to be about 3 and 35,
respectively.
Figure 8
8D
versus
8 O
variation
diagram.
Explanation: Go = Goldfield, hypogene alunite, n
= 1 (Rye et al. 1992); Ju = Julcani, alunite (/? = 6),
Jut = average of main-stage ore fluids in
wolframite, enargite, tetrahedrite, and galena fluid
inclusions, Ju2 = average of main-stage ore fluids
in sphalerite and chalcopyrite, Ju3 = late-stage ore
fluids in barite, siderite, and botroidal pyrite (Deen
1990); Le = Lepanto, alunite, n = 2 (Y. Matsuhisa
& J. Hedenquist, unpub. data); Nansatsu district:
Ka = Kasuga, alunite n = 1, Iw = Iwato, alunite n
1 ft

= 2, 6 O values of residual vuggy silica


associated with ore in both depositsfall between
Ka and Iw (Hedenquist et al. 1994a); NF =

454

Nevados del Famatina, stage V alunite-kaolinite, n


= 1 (Losada-Calderon & McPhail 1994); the
average 8D and 8 , 8 0 values for La Mejicana {n =
9) are similar to NF; K-silicate and quartz-sericite
at Nevados del Famatina have 8 O between 4 and
10/oo, reflecting a larger magmatic component
(Losada-Calderon & McPhail 1994); PV = Pueblo
Viejo: PV] = stage 1 alunite and kaolinite, PV2 =
stage 2 pyrophyllite (Fig. 9 in Vennemann et al.
1993); Ro = Rodalquilar, alunite, n = 10,
chalcedonic ore, n = 6 (Arribas et al. 1995a); RM
= Red Mountain, Lake City, Colorado, alunite, n =
12 (Bove et al. 1990; Rye 1993); Su =
Summitville, alunite, average of n = 10 (8D) and n
= 16 (8 18 0) (Rye et ah 1992), ore fluids from Rye
(1993). The main ore stage.at Rodalquilar (stage
2) is based on 8 1 8 0 of chalcedonic quartz; 8D are
not available for this stage but present-day
groundwaters, alunite, kaolinite, and illite fluids in
the deposits have a limited range of 8D values,
suggesting significant variations are unlikely
(Arribas et al. 1995a). Stage 2 (pyrophyllite)
fluids for Pueblo Viejo involve several
assumptions with respect to the choice of
fractionation factors for oxygen and hydrogen.
The data for stage 2 at Rodalquilar and Pueblo
Viejo should be viewed as approximate. Data for a
single alunite for Goldfield (Rye et al. 1992)
i o

suggest that mixing of


enriched magmatic fluid
waters may result in
hydrothermal acid-sulfate
1989).

a the 8D- and 8 Owith isotopically light


D- and
O-depleted
fluids (see also Vikre

You might also like