2011-epitermal (hs-2;henley and berger) 2

Upload: evafragozo

Post on 07-Aug-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    1/12

    Magmatic-vapor expansion and the formation of high-suldation gold deposits:Chemical controls on alteration and mineralization

    Richard W. Henley a, Byron R. Berger b,⁎a Research School of Earth Sciences, Australian National University, Canberra, Australiab U.S. Geological Survey, Federal Center MS 964, Denver, CO 80225-0046, United States

    a b s t r a c ta r t i c l e i n f o

     Article history:

    Received 14 July 2010Received in revised form 8 November 2010

    Accepted 8 November 2010

    Available online 23 November 2010

    Keywords:

    High-suldation

    Gold

    Magmatic vapor

    Enargite

    Silica-alunite

    Sulfosalt melt

    Tennantite

    Solfatara

    Large bulk-tonnage high-suldation gold deposits, such as Yanacocha, Peru, are the surface expression of 

    structurally-controlled lode gold deposits, such as El Indio, Chile. Both formed in active andesite–dacite

    volcanic terranes. Fluid inclusion, stable isotope and geologic data show that lode deposits formed within

    1500 m of the paleo-surface as a consequence of the expansion of low-salinity, low-density magmatic vapor

    with very limited, if any, groundwater mixing. They are characterized by an initial  ‘Sulfate’ Stage of advanced

    argillic wallrock alteration± alunite commonly with intense silicication followed by a   ‘Sulde’  Stage   —  a

    succession of discrete sulde–sulfosalt veins that may be ore grade in gold and silver. Fluid inclusions in

    quartz formed during wallrock alteration have homogenization temperatures between 100 and over 500 °C

    and preserve a record of a vapor-rich environment.

    Recent data for El Indio and similar deposits show that at the commencement of the Sulde Stage,

    ‘condensation’   of Cu–As–S sulfosalt melts with trace concentrations of Sb, Te, Bi, Ag and Au occurred at

    N600 °C following pyrite deposition. Euhedral quartz crystals were simultaneously deposited from the vapor

    phase during crystallization of the vapor-saturated melt occurs to Fe-tennantite with progressive

    non-equilibrium fractionation of heavy metals between melt-vapor and solid. Vugs containing a range of 

    suldes, sulfosalts and gold record the changing composition of the vapor.

    Published   uid inclusion and mineralogical data are reviewed in the context of geological relationships to

    establish boundary conditions through which to trace the expansion of magmatic vapor from source tosurface and consequentalteration andmineralization. Initially heat loss from the vapor is high resulting in the

    formation of acid condensate permeating through the wallrock. This Sulfate Stage alteration effectively

    isolates the expansion of magmatic vapor in subsurface fracture arrays from any external contemporary

    hydrothermal activity. Subsequent fracturing is localized by the embrittled wallrock to provide high-

    permeability fracture arrays that constrain vapor expansion with minimization of heat loss. The Sulde Stage

    vein sequence is then a consequence of destabilization of metal-vapor species in response to depressurization

    and decrease in vapor density.

    The geology, mineralogy,  uid inclusion and stable isotope data and geothermometry for high-suldation,

    bulk-tonnage and lode deposits are quite different from those for epithermal gold–silver deposits such as

    McLaughlin, California that formed near-surface in groundwater-dominated hydrothermal systems where

    magmatic   uid has been diluted to less than about 30%. High sul dation gold deposits are better termed

    ‘Solfataric Gold Deposits’   to emphasize this distinction. The magmatic-vapor expansion hypothesis also

    applies to the phenomenology of acidic geothermal systems in active volcanic systems and equivalent

    magmatic-vapor discharges on the  anks of submarine volcanoes.

    Published by Elsevier B.V.

    1. Introduction

    High-suldation gold deposits range in scale from small, high-grade

    lode deposits to very large,low-grade deposits.Theyhave in commonan

    envelope of pre-ore advanced argillic alteration assemblages  –  Sulfate

    Stage   –   that are commonly, but not exclusively, characterized by

    oxidized sulfuras sulfate in alunite(KAl3(SO4)2(OH)6),anda subsequent 

    contrasting sequence of ore-stage sulde assemblages  – Sulde Stage –

    that typically include pyrite, a range of sulfosalt minerals (including

    enargite and tennantite), gold, and silver (Arribas, 1995; Berger and

    Henley, 2011).

    Since the late 1990s, a number of studies of high-suldation gold–

    silver deposits have consistently shown that the temperatures recorded

    by uid inclusions in quartz in the Sulfate Stage alteration assemblages

    Ore Geology Reviews 39 (2011) 63–74

    ⁎   Corresponding author.

    E-mail address: [email protected] (Byron Berger).

    0169-1368/$   –  see front matter. Published by Elsevier B.V.

    doi:10.1016/j.oregeorev.2010.11.003

    Contents lists available at  ScienceDirect

    Ore Geology Reviews

     j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / o r e g e o r ev

    http://dx.doi.org/10.1016/j.oregeorev.2010.11.003http://dx.doi.org/10.1016/j.oregeorev.2010.11.003http://dx.doi.org/10.1016/j.oregeorev.2010.11.003mailto:[email protected]://dx.doi.org/10.1016/j.oregeorev.2010.11.003http://www.sciencedirect.com/science/journal/01691368http://www.sciencedirect.com/science/journal/01691368http://dx.doi.org/10.1016/j.oregeorev.2010.11.003mailto:[email protected]://dx.doi.org/10.1016/j.oregeorev.2010.11.003

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    2/12

    range from about 100 to over 500 °C and that vapor-rich inclusions are

    abundant (see  Section 2.1.1). In turn recent FESEM data for sulfosalt

    assemblages from El Indio, Summitville and Lepanto (Mavrogenes et al.,

    2010) strongly suggest near-magmatic temperatures for the initial

    copper-rich ore stage of high-suldation deposits (see  Section 2.1.2).

    These data show that, following initial pyrite deposition, arsenic-rich

    sulfosalt melts withtrace contentsof other heavy metals (tin,silver,gold,

    bismuth, selenium, and tellurium) were formed from a vapor phase. The

    high-temperature melts partially crystallized   rst to enargite and

    Fe-tennantite and distinctive euhedral quartz crystals (Fig. 1a–c)

    whose oxygen isotope composition (Mavrogenes and Henley, unpub-

    lished data) is consistent withthat demonstratedfor magmaticuid atEl

    Fig. 1.   Euhedral quartz and Fe-tennantite (Cu10.6, Fe1.4As3.9Sb0.1S13; Tn  0–5%, Fe 0.8 to 2.2 at.%) inll a void and associated fractures in earlier pyrite, b) euhedral quartz and

    Fe-tennantite (Cu10.6Fe1.1As4S13) inll a void in concentrically banded and fractured pyrite, c) euhedral quartz  lls in a void in earlier pyrite and with fractional crystallization of 

    Fe-tennantite aroundquartzfrom Tn 4.0 to 77% (bright zones), d) curved interfaces between low-antimony Fe-tennantite (Tn 3.4%)and fractionated Fe-tennantite (Tn 4.0–77%).Note

    the vug-rich rims of the low Tn  phase, the distribution of vugs (black) and euhedral quartz, and the minor Sb-enrichment adjacent to one vug, e) dissolution of early pyrite by

    Fe-tennantite (Cu10FeAs4S13) with transient development of a metal-decient chalcopyrite corona and isolated bornite and with quartz inlling void space, f) emplectite (CuBiS2)

    growing in an isolated vug within enargite. Abbreviations: Qz, quartz, Fe–Tn, Fe-tennantite (numbers are atomic percent Sb/(As+Sb)), ccp, chalcopyrite, Py, pyrite, v, vug. (Sample

    location data: El Indio; EI1300, Enargite vein, DDH1300, 490.5 m, EI4216F; DDH4216, 27.7 m; Lepanto; L4=3.29.3 of  Mancano (1994, Fig. 8); Red Mountain, CO6, Underground

    sample, Longfellow Mine, Red Mountain, Silverton, Colorado). Images: a), b) and e) are reected light images, c) and d) are ASB FESEM images and f) is a BSE in-lens detector FESEM

    image.

    64   R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    3/12

    Indio ( Jannas,1995).Fractionation of heavymetals (Sb, Bi,Ag, Te,and Au)

    occurred as the residual melt crystallized in the presence of vapor (Fig. 1

    c–d) anddissolved earlier pyrite(Fig. 1e) andenargite. Metal transportin

    the vapor phase is recorded through the preservation of micron-scale

    vugs containing a wide range of discrete suldes and sulfosalts including

    tennantite and tetrahedrite, chalcocite, bismuthinite and emplectite

    (CuBiS2) (Fig. 1f) as well as quartz, uorapatite and gold (Mavrogenes et

    al., 2010).

    By contrast the homogenization data for 

    uid inclusions inenargite at Lepanto, Philippines, suggested a much lower temperature

    range (less than 310 °C) for the ore stage (Mancano, 1994; Mancano

    and Campbell, 1995). Although, as determined by  Hedenquist et al.

    (1998, p. 386), no primary uid inclusions were found in the Lepanto

    enargite or quartz samples, the acceptance of these   uid inclusion

    homogenization data (b310 °C) as recording the temperature of 

    mineralization forced the conclusion that substantial system-scale

    cooling had occurred between the alteration and ore stages

    (Hedenquist et al. (1998, Fig. 19). In the absence of independent

    temperature estimates for the Sulde Stage, this hypothesis was

    subsequently adsorbed into a number of generic models for the

    formation of high-suldation gold deposits (e.g.   Heinrich, 2005;

    Muntean and Einaudi, 2001). However a number of fundamental

    dif culties with the geology, physics and chemistry attend these

    models, not least of which is accounting for how the uid in such large

    scale hydrothermal systems lost some 75% of its heat between its

    evolution at magmatic temperatures (NN750 °C) and the assumed

    temperature of mineralization (b310 °C). Each of these models

    imposes critical assumptions about pressure as well as temperature

    in the depositionalregime of high suldationmineralsystems that are

    not sustained by critical review of available geochemical and,

    crucially, geologic data. Neither can these models accommodate the

    new high temperature data for the Sulde Stage so that a fresh

    approach is necessary.

    In this paper, in order to set boundary conditions, werst reviewthe

    interpretation of available   uid inclusion data in the context of the

    geology and geochemistry of high-suldationdeposits.We then provide

    a new hypothesis for the origin of these important gold deposits based

    on the   expansion   of a magmatic-vapor phase under low-pressureconditions (see Section 3) that is fully consistent with the geological

    relationships (Berger and Henley, 2011) and stable isotope data and

    with the active volcanic context in which these deposits formed.

    2. The environment of formation of high-suldation lode-gold

    deposits

    The advanced argillic alteration and ore stages in high-suldation

    lode deposits track active faults and branch fractures through

    intrinsically low-permeability host rocks such as marine sediment

    sequences or ash-ow tuffs. These fracture arrays crosscut host rocks

    that commonly contain evidence of previous or concurrent weak,

    near-neutral pH groundwater-related alteration indicating that largerscale groundwater-dominated geothermal systems may have operated

    contemporaneously with the formation of the high-suldation

    mineralization. Based on   eld data,   Berger and Henley (in press)

    suggested that the aggressive silica–alunite1 alteration localized by

    syn-hydrothermalfault arrays effectively isolated magmaticuidowin

    fractures from any external hydrothermal system—a discretization that

    characterizes acidic geothermal systems in active volcanic settings today

    (e.g., Ruaya et al., 1992) (see Section 5.2).

    Sulfur isotope data consistently emphasize the predominance of 

    magma-sourced  uid in both the alteration and mineralization stages

    (e.g., Rye, 1993, 2005) of these deposits. Other stable isotope and  uid

    inclusion data are alsoconsistent witha high-level volcanicenvironment

    and thepredominance of a low-salinity (b5 wt.% NaCleq) magmaticuid

    (Bethkeet al., 2005;Jannas et al., 1999; Rye et al., 1992). Theinvolvement

    of groundwater has been investigated in a number of 

     18

    O–

    D studies, butits proportion in the ore setting is generally much less than 25% and, in

    some cases, is close to zero (Bethke et al., 2005; Deyell et al., 2004, 2005;

    Hedenquist et al., 1994,1998;Jannas et al., 1999;Rye, 1993; Vennemann

    et al., 1993). This is the reverse of the relative proportions of magmatic

    uid and groundwater in high-level adularia–sericite (so-called   ‘low

    suldation’   gold–silver) and analogous submarine massive sulde

    systems. In high-suldation deposits, evidence for the involvement of 

    groundwater (and seawater in coastal settings) is strongest for

    distinctive alunite veins and barite that post-date the sulfosalt-gold

    stage in many deposits and record waning thermal activity.

     2.1. Constraints on temperature

    In order to set boundary constraints on, for example,   uidtemperature and pressure, for the modeling of the physics and chemistry

    of the common and characteristic silica–alunite advanced argillic and

    sulde stages of high-suldation gold deposits, it is   rst necessary to

    carefully review the interpretation of   uid inclusion and other data

    within their geologic context. The dif culties in interpreting   uid

    inclusion data for deposits formed at shallow paleo-depths (see

    Section 2.2) are well known and primarily relate to uncertainties with

    respect to the timing of trapping relative to primary mineral deposition,

    the extent of cooling and involvement of groundwater (e.g., Foley et al.,

    1989) and as discussed below,phase changes, in thehost mineral,during

    the cooling stages of alteration and mineralization.

     2.1.1. Silica–alunite stage temperatures

    There is general agreement that the silica–alunite stage of intenseleaching and alteration results from the   in situ   formation of acidic

    condensate derived from magmatic vapor (Arribas, 1995; Chouinard

    et al., 2005; Hedenquist et al., 1994, 1998; Sillitoe and Hedenquist,

    2003; Stoffregen, 1987; Vennemann et al., 1993). At Chinkuashih,

    Wang et al. (1999)  noted an abundance of vapor-rich inclusions in

    Sulfate Stage quartz samples. At Furtei, Sardinia, Ruggieri et al. (1997)

    also comment on the abundance of vapor-rich inclusions, and report a

    population of distinctive high-salinity, high-temperature inclusions

    (390°–500 °C) as wellas lower temperature populations. At Summitville,

    Bethke et al. (2005)  similarly report an abundance of vapor-rich

    inclusions with homogenization temperatures greater than 500 °C.

    Hedenquist et al. (1994)   also reported an abundance of vapor-rich

    inclusions in alteration assemblages in the Nansatsudistrict, Japan, as did

     Jannas et al. (1999) in alunite in the Campaña fracture-vein system at ElIndio. Nevertheless, homogenization temperature data for vapor-rich

    inclusions2 have been largely ignored in the interpretation of these

    deposits, even though their abundance conrms a vapor-rich

    environment and a wide range of temperatures and temperature

    gradients within the alteration zones during their development.

    The mineralogy of the Sulfate Stage has been extensively

    documented (e.g.,  Heald et al., 1987; Hedenquist et al., 1994, 1998;

     Jannas, 1995; Stoffregen, 1987). As well as alunite and a range of silica

    polymorphs including quartz and chalcedony, it comprises the

    minerals that are typical of advanced argillic alteration; namely,1 Silica–alunite. We are concerned in this paper with the reactions between acidiccondensate and rock during alteration and subsequent cooling in the Sulfate Stage. As

    a convenience, we shall use the generic term   “silica–alunite”  to refer to the  range  of 

    advanced argillic alteration assemblages (Meyer and Hemley, 1967) that are

    encountered in the high-suldation mineral districts whether or not alunite is a

    product.

    2 Homogenization temperature data for vapor-rich inclusions provide minimum

    trapping temperatures for the inclusions.

    65R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    4/12

    kaolinite, pyrophyllite, diaspore, andalusite and disseminated pyrite

    and in rare instances, corundum (e.g., Famatina, Argentina,  Losada-

    Calderon and Bloom, 1990). Alunite itself is stable up to at least 450 °C

    (Stoffregen et al., 1994) in accord with temperatures as high as 500 °C

    determined through   uid inclusion measurements in silica–alunite

    alteration zones in many deposits despite the frequent assumption

    that the upper stability limit of alunite is 300 °C (e.g.,  Muntean and

    Einaudi, 2001). The alteration assemblage or mineralogical proportions

    can change with depth. At El Indio, for example, pyrophyllitepredominates over alunite below 4000 m asl ( Jannas et al., 1999) and

    at Chinkuashih, alunite simply disappears at deep levels (Sakazaki et al.,

    1964).

    Hemley et al. (1980)   experimentally determined the phase

    relationships of kaolinite, pyrophyllite, diaspore, andalusite, and

    corundum as a function of temperature, pressure and the activity of 

    silica in solution (Fig. 2). During advanced argillic alteration, these

    phases may be substituted by alunite (Hemley et al., 1969) in the

    presence of alkalis derived from volcanic glass or feldspar in the host

    rock, and high sulfate and hydrogen ion concentrations derived from

    ef cient partitioning of highly soluble SO2 and HCl from the vapor into

    the condensate fraction. When a strongly acidic condensate reacts

    with volcaniclastic or sedimentary rocks, it results in high silica

    activity in the liquid phase (Africano et al., 2002; Fournier, 1985) such

    that metastable polymorphs of silica precipitate with alunite or

    pyrophyllite, or with kaolinite (Fig. 2). In addition silica is added to the

    wallrock during alteration since the solubility of quartz in the vapor

    phase in the main fracture is, at 500 bars and 600 °C, about 1000 mg/kg

    (Morey and Hesselgesser, 1951; Fournier and Potter, 1982). When a

    small condensate fraction is formed in response to heat loss, partitioning

    of silica to the liquid condensate leads immediately to very high silica

    concentrations that drive very rapid silica deposition (Fig. 2). During

    cooling of the alteration assemblage in the presence of condensate

    trapped in pore space, as well as later with the inltration of 

    groundwater, progressive recrystallization of metastable silica occurs

    through a sequence of polymorphs and ultimately to quartz (Africano

    and Bernard, 2000; Okamoto et al., 2010). This sequence of processes

    results in massive, highly silicied rocks, including   “vuggy silica”

    textured rocks of the type described by Stoffregen (1987) at Summit-

    ville, Jannas et al. (1999) at El Indio and Wang (2010) at Chinkuashih.

    Consequently, the trapping of inclusion populations takes place over an

    extended period so thata wide range of temperatures arerecorded(e.g.,

    Bethke et al., 2005, p. 290). Furthermore, since alteration progressively

    reduces permeability, the altered rock becomes both a barrier to the

    ingress of groundwater and,at thesame time, an effectiveinsulator with

    respect to the dispersion of heat away from the 

    uid phase that ischanneled through the principal controlling fractures (Berger and

    Henley, 2011). The temperature gradients within such insulating

    wallrock environments are consequently very large, an attribute that is

    reected in the wide range of   uid inclusion homogenization data

    (N500 to 100 °C), from alteration assemblages in which silica

    recrystallized.

    During cooling, redistribution and recrystallization of silica in the

    altered wallrock occur where available water penetrates or wets the

    surface of the primary and successive silica precipitates.

    Fluid inclusions in quartz in this evolving assemblage do not therefore

    necessarily record the presence of a continuum of liquid water or a

    boiling-point depth relationship through which to convert

    homogenization and freezing data to an estimate of depth. Rather

    the pore pressure during silica recrystallization is more reliably

    related to lithostatic pressure (see   Section 2.2). Importantly, since

    most of the homogenization temperature data record trapping during

    cooling fom   N500 °C the primary deposition temperature of the host

    assemblage was necessarily hotter.

     2.1.2. Sul de Stage temperatures

    It is demonstrable that it is Sulfate Stage embrittlement of the

    wallrock that localizes subsequent fracturing and consequent Sulde

    Stage lode mineralization (Berger and Henley, 2011). The latter is

    associated with little recognizable wallrock alteration other than, for

    example, isolated replacement of early alunite by illite–sericite at El

    Indio ( Jannas, 1995). In most deposits, pyrite (±solid solution gold

    and silver; e.g., Deditius et al., 2009) is the rst sulde to precipitate in

    lodes and is followed by enargite and Fe-tennantite (±minor

    chalcopyrite and base-metal suldes) assemblages that also containdistinctive euhedral quartz crystals (Fig. 1) (Mavrogenes et al., 2010;

    Mavrogenes and Henley, in preparation). Later stages of sulde–

    sulfosalt deposition are characterized by more complex textures and

    mineral suites that include some tellurides, selenides and bismuth

    minerals as well as ore-grade silver and gold (e.g.,   Claveria, 2001;

     Jannas, 1995). The textural and compositional data (Fig. 1) also show

    that high-suldation sulde–sulfosalt assemblages do not show

    equilibrium relationships nor are they analogous to the sulde

    assemblages encountered in epithermal adularia–sericite type

    deposits.

    Sulde stage mineral assemblages are not amenable to standard

    methods of transmitted light  uid inclusion analysis so that infrared

    techniques have been progressively developed for the measurement

    of homogenization temperatures and salinities in enargite andsometimes pyrite.   Moritz (2006) has cautioned that many infrared

    determinations of inclusion   uid salinity and homogenization

    temperatures may be in error and has since published new data for

    enargite sampled from a number of deposits in Europe and Peru

    (Moritz and Benkhelfa, 2009). These liquid-phase inclusion data

    dene a range of salinities from very low to less than 5 wt.% NaCleqand have a range of homogenization temperatures from about

    150–300 °C. Kouzmanov et al. (2010) have reported a similar range

    of   uid inclusion homogenization temperatures for well-formed

    enargite crystals in drusy  late-stage  cavities in minor post-porphyry

    veins from the Rosia Poieni porphyry copper deposit, Romania.

    The recrystallization of sulfosalts to complex sulfosalt assemblages

    together with a wide variety of tellurides (e.g., goldeldite) and native

    gold, during cooling to below about 300 °C, has been previously

    Fig. 2.  Stability relations in the system Al2O3–SiO2–H2O at PH2O=250 bars with the

    addition of silica activity-temperature relations for the common polymorphs of quartz

    to temperatures well above 300 °C (Okamoto et al., 2010). Deposition of silica occurs

    initially as metastable silica polymorphs that subsequently recrystallize to quartz

    during cooling of the alteration assemblage with development of pseudo-secondary

    uid inclusions with a wide range of homogenization temperatures from ~100 to

    N500 °C. Point MH  shows, for illustration, the concentration of silica when a liquid

    condensate fraction of 10% separatesfrom thevaporat 550 °C based on thevaporphase

    solubility data of   Morey and Hesselgesser (1951).   Constructed from Geochemists

    Workbench data kindly provided by James Cleverley.

    66   R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

    http://localhost/var/www/apps/conversion/tmp/scratch_7/image%20of%20Fig.%E0%B2%80

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    5/12

    stressed by Sack and Goodell (2002)for theJulcaniand other deposits.

    Fig. 3 shows phase relationships in the Cu–Sb–As–S system; enargite

    (orthorhombic, wurtzite-type, hexagonal close-packed structure) and

    luzonite (tetragonal, sphalerite-type cubic close-packed structure)

    are,  sensu stricto, polytypes of Cu3AsS4  (Posfai and Sundberg, 1998).

    Enargite is the high-temperature form of luzonite above 280–300 °C

    and an immiscibility gap is dened between enargite and the luzonite–

    famatinite solid solution series (Skinner, 1960) where these minerals

    display a wide range of ordered to disordered structures and

    intergrowths at the atomic scale.  Posfai and Buseck (1998) noted that

    the occurrence of enargite is an indicator of its formation well above the

    transitionbetween luzonite andenargite(280 °C)and that luzonite may

    form through solid-state recrystallization of enargite below thetransition temperature. In describing the enargite inclusion samples

    from Lepanto, Mancano (1994) specically noted heterogeneities and

    internal structural defects,3 as well as the dif culties caused by varying

    degrees of IR transparency in establishing the primary or secondary

    nature of inclusions that enabled only 29% of the 97 enargite inclusions

    in 17 samples to be interpreted as potentially primary (and two as

    secondaries) while all quartz inclusions were classied as secondaries.

    We suggest that these data record post-primary phase changes and

    fracturing during the cooling of the enargite assemblages in the

    presence of groundwater.

     2.1.3. Summary of temperatures for alteration and mineralization

    The Sulfate Stage   uid inclusion and stable isotope data reliably

    identify a) a wide range of temperatures (N500 to 100 °C) consistent with

    high thermal gradients and the recrystallization of metastable silica

    polymorphs to quartz during cooling, and b) derivation from partial

    condensation of a magmatic vapor with very limited involvement of 

    groundwater. Since condensation of vapor occurs through heat loss, it is

    implicit that the primary  uid4 had a higher enthalpy and temperature

    andwastherefore, necessarily,also a vapor . Temperature estimatesfor theSulde Stage cannotbe reliably based onuidinclusiondata in enargite or

    other suldes. Micro-analytical data (see   Section 1) from El Indio,

    Summitville and Lepanto provide independent evidence that sulde–

    sulfosalt deposition along with quartz in the earliest part of the Sulde

    Stage were at least 550 °C, consistent with the high temperatures

    indicatedfor thevapor condensateresponsiblefor Sulfate Stage alteration.

     2.2. Constraints on  uid pressure

    Since many of the well-known high-suldation deposits are

    young, pressure estimates for alteration and mineralization can be

    made fromeld data. For example at Chinkuashih, Taiwan, the rugged

    topography and geological setting of the ≈1 Ma Chinkuashih deposit

    suggest that, at most, only one or two hundred meters of rock havebeen removed by erosion. Mining extended to about 850 m below the

    present surface with decreasing enargite and increasing Cu/Au ratio.

    As discussed above (see  Section 2.1.1),   uid inclusion formation is

    likely to have occurred during recrystallization of silica in the

    alteration zones so that interpreted trapping pressures record pore

    uid pressures consistent with the reconstructed paleo-depth at

    which samples were obtained. Jannas et al. (1999) reconstructed the

    original topography at El Indio to suggest a paleo-depth range for

    mineralization of   −500 to  −1100 m (or less if post-mineral fault

    movement occurred).   Hedenquist et al. (1998)   suggested shallow

    depth for the Lepanto mineralization and at Goldeld, Nevada, debris

    ow sediments (Berger, 2007) containing alunite pebbles occur at the

    paleosurface in the eastern part of the district indicating that

    mineralization and alteration developed at ≈200 m depth.

     2.2.1. Summary of pressure constraints on alteration and mineralization

    Since alteration and mineralization at both Chinkuashih and El

    Indio were formed from a vapor phase (see   Section 2.1) to depths

    below the paleosurface of about 1100 m and stable isotope data

    indicate limited, if any, incursion of groundwater, theuid pressure in

    the developing vein system was evidently greater than that due to an

    external groundwater system (≤~110 bars at this depth). At this

    depth, the lithostatic pressure would have been about 290 bars and

    similar pressure estimates apply for other high-suldation vein

    deposits. The pressure constraints provided by geologic and   uid

    Fig. 3.   Phase relations in the system Cu–As–Sb–S redrawn from   Posfai and Buseck

    (1998). The range of compositions of coexisting phases in the enargite–famatinite

    system at Chinkuashih, Goldeld and El Indio is indicated (Huang and Chou, 1975;

    Huang and Tsai, 1984; Mavrogenes et al., 2010; Sung, 2005 ). An estimate of the range

    for Lepanto   ‘enargite–luzonite’   is provided based on a smelter concentrate assay

    (Filippou et al., 2007) and our unpublished FESEM data in order to compare with the

    maximum and minimum ranges of  uid inclusion homogenization data from enargite

    (Mancano, 1994). For comparison the compositional range of high temperature

    volcanic gas condensates from fumaroles with discharge temperatures   N600 °C is

    shown based on the data compiled by Williams-Jones et al. (2002). The average crustal

    antimony and arsenic concentrations of granite and basalt are shown for reference.

    Abbreviations; En, enargite, luz, luzonite, Fam, Famatinite, hcp, hexagonal close packed,

    ccp, cubic close packed, int, interlayered polytype.

    3 An IR image of enargite in Mancano (1994, Fig. 3a), for example, clearly shows a

    relation between an inclusion and a subsolidus structural, not compositional, attribute

    of the crystal.

    4 Fluid-phase terminology. The term “hydrothermal” has a commonly assumed connota-

    tion of referring to a liquid phase. Here we simply use it in reference to the context of the

    phenomenology of H2O-dominated uid phases within an epizonal or sub-volcanic context.The term   “magmatic vapor” has been widely used to refer to the low-density  uid phase

    released from a melt so that the term  “vapor” has a connotation with respect to both phase

    state and provenance. Strictly the term  “vapor” refers to a  gas phase in equilibrium with a

    condensed phase   – solid or liquid   – as for example in the term   “vapor pressure”. Thus a

    “magmatic vapor”  once released from direct interaction with a melt should be termed a

    “magmatic gas” or simply a  “gas” or indeed more precisely a  “gas mixture” because of the

    wide rangeof componentsthatit contains. Sincethis gasmixture is dominated by water it is

    condensibleand therefore on expansionmay intersect thetwo-phaseboundary of thesystem

    resulting in phase separation to fractions of liquid  “condensate” and equilibrium  “vapor” (a

    gas). Whilst the distinctions may appear semantic, it is important to be aware of the precise

    implications of theseterms, particularly with respect to the physicalproperties and behavior

    of   “uid”   phases with respect to   uid properties like density and viscosity in geological

    environments. Here, the term   “vapor”   is used generically where the argument has not

    reached a point of dening physical properties and then ceases to have value in much the

    same way as   Verhoogen (1949)   advocated the avoidance of the non-specic term

    “supercritical”. We have used the term  uid as a non-specic term prior to dening its

    phase state in line with the recommendations of Bodnar (pers.comm., 2010).

    67R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

    http://localhost/var/www/apps/conversion/tmp/scratch_7/image%20of%20Fig.%E0%B3%80

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    6/12

    inclusion data do not therefore sustain the high pressure conditionsrequired by Heinrich et al. (2004) to maintain a liquid phase condition

    for high-suldation mineralization.

    As discussed later (see  Section 5.2), where limited erosion has

    occurred, high-suldation lode mineralization is expressed as low grade

    bulk-tonnage deposits associated with extensive advanced argillic

    alteration, indicating that vapor   ow occurred freely to and through

    the paleo-surface. Ignoring small changes in vapor density due to

    fractional heat loss, thepressure-depth curve fora vapor is much steeper

    than that of liquid water (the boiling point-depth curve or cold

    groundwater hydrostatic gradient) or of rock (lithostatic gradient)

    (Fig. 4). The steep vapor gradient intersects the lithostatic gradient at a

    shallow level so thatpressurerelease after choking by mineraldeposition

    could produce shallow breccias bodies as is observed in high level

    deposits such as Yanacocha, Peru.5 Below this crossover, lithostaticpressuregreatlyexceedsthe vaporpressure inan open fracture. This is an

    unstable situation with respect to the relative  uid pressure within the

    open fracture system which must therefore progressively close by

    wallrock fragmentation due to rock stress as well as mineral deposition.

    Maintenance of fracture permeability is thus dependent on renewed

    fracturing dueto theactive stresseldand thecondition (Cox,2010) that

    PuidNProck+T, where T is the tensile strength of the rock(about 10 bars

    maximum).

    A review of data for alteration and ore mineralization tempera-

    tures (see Section 2.1) indicates that the Sulfate Stage  uid transiting

    to the surface through faults and associated fractures was a

    condensible vapor (i.e., condensate liquid formed as   uid transited

    pressure and temperature gradients in the wallrock) at a temperature

    of  N550 °C with a density of about 0.03 g/cm3.

    3. Magmatic-vapor expansion

    The recognition, through stable isotope data, that the   uid phase

    responsible for bothalteration and ore mineralization in high-suldation

    gold deposits was predominantly of magmatic origin forces

    consideration of the systematics of magmatic-uid expansion from its

    source through a fracture array in the upper few kilometers (b1.5 to

    2 km) of theEarth'scrust to whereit discharges at thesurface. Pressure is

    the criticalcontrol onuidowand the phase state oftheuid, thelatter

    crucially controlling chemical processes such as mineral deposition.

    Critical examination of available   uid inclusion and other data (see

    Section 2) conrms that the   uid responsible for alteration and

    ore mineralization was a low-salinity, high-temperature (N550 °C),

    low-pressure (bb500 bars) vapor phase released from a magmatic

    reservoir comprising multiple sub-volcanic intrusions with alteration

    and ore mineralization occurring within 1500 m of the surface at

    pressures below about 110 bars.

    Sodium chloride (b5 wt.% NaCleq) dominates other non-volatile

    componentsin bothvolcanic gases anduidinclusions in high-suldation

    deposits. By analogy with volcanic gases, CO2 and other insoluble gases

    occur at less than 5 mol% such that their contribution on an ideal gasbasis

    to the total pressure of thevapor phase is lessthan5%. The phase behavior

    of the magmatic uid can appropriately be approximated with respect to

    pressure, enthalpy (heat content), temperature and solute content on the

    basis of data for the simple NaCl–H2O system (Driesner, 2007). Fig. 5a

    shows the extent of the immiscibility gap between liquid and vapor up to

    1000 °C in thissystem and Fig. 5b is a projectionof the two-phase surface

    and isopleths of NaCl concentration for a selection of illustrative

    temperatures. In the high-enthalpy, low-pressure, single-phase region

    and below 5 wt.%.   NaCleq, the properties of the vapor phase can be

    approximated using values for pure water under the same conditions inorder to provide quantication of, for example, the density of the vapor

    phase.

    The melting relations of water-saturated granite (Fig. 5b) with

    respect to pressure and temperature dene the enthalpy range of the

    low-salinity   uid exsolved from a granitic magma. For example, a

    magmatic uid exsolving from melt at ~850 °C in relatively shallow

    volcanic settings hasan enthalpy in excess of 4000 kJ/kg and a density

    of about 0.2 g/cm3 or less. It follows that, for the shallow (b1.5 km)

    volcanic setting of enargite–gold deposits, the low-salinity   uid

    identiedby uid inclusionand stableisotope analyses is of magmatic

    origin and has the properties of a  gas; it expands to  ll all available

    volume.

    It is essentialin considering thebehaviorof hydrothermalsystems to

    track the fate of all the conservative components in the   uid phaseincluding salt content, stable isotopes and heat energy. The thermal

    dynamics of a number of common depressurization paths for magmatic

    vapor are shown in Fig. 5c in order to contrast that for high-suldation

    deposits with others in more permeable environments where

    groundwater interactions dominate ore-forming processes. In settings

    such as geothermal systems (M–G) where high permeability is

    maintained by extensional fracturing, mixing between cold

    groundwater and expanding magmatic vapor occurs across the extent

    of the system and gives rise to the development of magmatic vapor

    plumes as described by Henley and McNabb (1978). In these settings,

    uid enthalpy (and therefore temperature), isotope ratios and solute

    concentration are determined by the progressive change in the ratio of 

    magmatic  uid to groundwater. Intersection of the two-phase region

    may occur on particular enthalpy–

    pressure mixing paths depending on

    Fig. 4. The relationship between pressure and depthin a vaporexpanding to the surface

    through a fracture array. MV represents the  ow through low permeability fractured

    rock from a high pressure magma source, M, to a low pressure sink, V, provided by the

    highly transmissive fracture array. VF is a simplied pressure-depth prole where the

    vapor pressure gradient is greater than hydrostatic (see Section 2.2.1). Hydrothermal

    brecciation occurs above point S as shown. At surface rapid open-system cooling occurs

    due to atmospheric interactions and mixing with shallow groundwater or perched

    layers of condensate providing local regions with hydrostatic gradients that may be

    penetrated by fumarolic discharges. For further discussion see text and  Berger and

    Henley(in press). A usefulmodern example of thecontrastof vapor andliquid pressure

    proles is detailed by   Moore et al. (2008)   for the Karaha-Telaga Bodas volcanic

    geothermal system, Indonesia.

    5 Many of the so-called  hydrothermal breccias in shallow deposits can however be

    shown to have formed during the Sulfate Stage, for example at Tambo, Chile ( Jannas et

    al., 1999), in the presence of aggressively reactive acid condensate indicating

    dissolution and steam explosion as their formation process.

    68   R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

    http://localhost/var/www/apps/conversion/tmp/scratch_7/image%20of%20Fig.%E0%B4%80

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    7/12

    structural-permeability and in these settings porphyry-style

    mineralisation (ϕ) may eventuate (see Section 5.1). In the single-phase

    region near surface, phase separation occurs as the mixed   uid

    intersects the two-phase region and, in favorable structural settings,

    may generate adularia–sericite styles of epithermal precious and base

    metalmineralization (Henleyand Berger, 2000; Henleyand Ellis,1983).

    This mixing scenario differs from path M–FHE  where the isenthalpic

    expansion6 of the magmatic vapor is conned to discrete high-

    permeability fracture arrays through impermeable rocks and heat lossis limited by theconductivity of the wallrock (Toulmin and Clark, 1967).

    It is this setting that corresponds to the formation of high-suldation

    gold deposits and high-enthalpy fumaroles (FHE) in actively degassing

    volcanoes. Where expandingvapor loses a signicant proportionof heat

    by conduction (M–V), a small fraction of vapor condenses as pressure

    drops below M–V –C   (Fig. 5c). Through partitioning of acid gases,

    primarily SO2, to the condensate in wallrock fractures, aggressive

    alteration ensues and gives rise eventually to the characteristic silica–

    alunite alteration of high-suldation deposits (see Section 2.1.1). Path

    S–S′ is illustrative of conductive cooling where a limited  ow of vapor

    loses all of its heat to wallrock; such total condensation retains the

    composition of the primary vapor phase and the liquid condensate may

    deposit sulde minerals in isolated cavity   llings.7 A high   ΔH/ΔP

    trajectory also occurs where vapor condenses directly into a perchedgroundwater or condensate liquid aquifer very close to the surface and

    Fig. 5. (a) Three dimensionalperspective of the 2-phase regionin the systemNaCl–H2O

    based on the data compilation of   Driesner (2007, Supplementary Material). (b)

    Enthalpy–pressure–temperature relations for phases in the system NaCl–H2O calcu-

    lated using SOWAT code kindly provided by Thomas Driesner.). The  gure comprises a

    projection of a selection of isopleths and the two-phase boundary at 400°, 750° and

    1000 °C to illustrate the expansion of the two-phase eld as temperature increases, the

    effect of salt-content on the enthalpy of liquid (L) and gas phases (V) outside the two-

    phase  eld and the relative location of the halite and halite (H)+vapor (V)  elds at

    these temperatures. For discussion purposes, density isochors for pure water are also

    provided. The water-saturated minimum and 2% water content melting curves for

    granite are shown in order to identify the enthalpy–pressure eld for magmatic vapor

    derived from epizonal intrusions. c) Illustrative paths for mixing and heat loss  ow in

    hydrothermal systems; for detail see Section 3.

    Fig. 6. Halite solubility in the halite-vapor  eld of the NaCl–H2O system, redrawn fromthe data of   Sourirajan and Kennedy (1962). An illustrative quasi-isenthalpic

    depressurization and expansion path is shown for a magmatic vapor emphasizing the

    decrease in water vapor pressure and the rapid decrease in halite solubility in the halite

    plus vapor eld both of which lead to rapid sulde and sulfosalt deposition as melt or

    solid phase.

    6 Heat-balance terminology. We here use the term   “isenthalpic”   to specify

    expansion processes where there is no exchange of heat energy with the surroundings.

    The alternative term   “irreversible adiabatic expansion”   distinguishes the non-

    equilibrium aspect of such processes (for example, rapid expansion through a series

    of throttles or dilations in a fracture array) and their rate relative to   “reversible

    adiabatic expansion” processes. The latter refer to the slow constrained expansion of a

    uid where there is an inter-conversion of internal energy and work energy. In such a

    constrained isentropic process, a far more substantial temperature decrease occurs

    relative to that due to irreversible isentropic expansion over the same pressure drop

    (Anderson and Crerar, 1993). Both these processes may control the range of fumarole

    temperatures within the La Fossa volcano on Vulcano, Italy ( Nuccio et al., 1999).7 An example (see Section 2.1.2) is the enargite and pyrite (TH ~317 °C) within late

    drusy cavities in isolated veins cutting the Rosia Poieni porphyry copper deposit

    (Kouzmanov et al., 2010).

    69R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

    http://localhost/var/www/apps/conversion/tmp/scratch_7/image%20of%20Fig.%E0%B6%80http://localhost/var/www/apps/conversion/tmp/scratch_7/image%20of%20Fig.%E0%B5%80

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    8/12

    results in the well-known solfataric environment of active volcanoes

    associated with low-enthalpy fumaroles (FLE).

    Fig. 6 provides the detail of the low-pressure environment where

    quasi-isenthalpic expansion of magmatic vapor enters the halite plus

    vapor   eld in which the vapor-phase solubility of halite decreases

    rapidly with pressure and temperature (or more precisely enthalpy).

    As will be discussed later, it is this rapid decrease in concentration of 

    components such as chloride that, in association with the decrease in

    water pressure during expansion, is likely to be the principal controlon the deposition of metal and metal suldes and sulfosalts as well as

    gold (Mavrogenes et al., 2010).

    4. Chemical processes during alteration and mineralization

    In the magmatic-vapor expansion hypothesis proposed in this

    paper, it is assumed, in commonwith modelsproposedby Hedenquist

    et al. (1998),   Henley and McNabb (1978),   Williams-Jones and

    Heinrich (2005)   and earlier by   Krauskopf (1964), that metals are

    transferred from a magma source as vapor-phase species. The ability

    of a vapor phase to transport signicant quantities of metals does,

    however, remain one of the most controversial areas in the discussion

    of the formation of magmatic-hydrothermal ore deposits despite

     prima facie   evidence such as the transport and deposition of a widerange of metals including gold, iron, and arsenic in high-temperature

    fumarole gases (e.g.,   Africano et al., 2002; Churakov et al., 2000;

    Fulignati and Sbrana, 1998; Larocque et al., 2008; Taran et al., 2000,

    2001; Williams-Jones et al., 2002; Yudovskaya et al., 2006). Further

    support is provided by   uid inclusions in porphyry copper–gold

    systems (e.g.,   Ullrich et al., 1999) and heavy metal sulfosalts and

    suldesin vugs in high-suldation deposits (Mavrogenes et al., 2010).

    Chemical-vapor transport and deposition at low pressures are also

    very well established in the semiconductor and solar technology

    industries for a wide spectrum of metals (cf.  Hastie, 1975; Kodas and

    Hampden-Smith, 1994). However, as stressed by Pokrovski et al. (2008),

    there is a frustrating data vacuum for the high-temperature speciation of 

    metals in a high-temperature gas phase under geological conditions, and

    this reects how dif cult such experimental determinationsare given theproblems of containment, sampling and analysis of highly corrosive,

    low-density uids.

    Pokrovski et al. (2008, p. 357) note that the formation of solvation

    shells around core metal species   “is the essential factor controlling

    vapor-phase transport and vapor–liquid partitioning”   in the vapor

    phase at high temperatures. Species such as Cu(HS)g, nH2O have high

    hydration numbers8 that respond rapidly to decreasing water pressure

    as shown by vapor–liquid partition coef cients: in a sulfur-free system

    for example, base-metal partition coef cients at 500 °C decrease by two

    orders of magnitude between 580 and 480 bars, while partition

    coef cients for gold and copper decrease even more rapidly in systems

    containing reduced sulfur (Pokrovski et al., 2008). The solubility of 

    quartz as SiO2·2H2O is strongly pressure dependent (Fournier and

    Potter, 1982; Morey and Hesselgesser, 1951) as is that of halite asNaCl·2H2O (Hastie, 1975, p. 82).

    Particularly pertinent for high-suldation systems,   Pokrovski

    et al. (2005) have shown that the dominant arsenic species in low-

    density magmatic vapor at 500 °C is As(OH)3  the concentration of 

    which is therefore strongly dependent on PH2O. For illustrative

    purposes the solubility of enargite, as a solid or melt, may then be

    written,

    3CuðHSÞ0

    g þ  AsðOHÞ0

    3g þ H2Sg ¼  Cu3AsS4 þ  3H2Og þ  H2g   ð1aÞ

    so that, with a=activity, the solubility equilibrium constant, K s,gis,

    K s;g  ¼ ½a3

    H2Og  • a H2;g =½a

    3

    CuðHSÞog  • a AsðOHÞo3g  •

     a H2Sg ð1bÞ

    The solubility of enargite in vapor therefore at least a cubic

    dependence on the activity of water or, to a reasonable approximation,

    the pressure of water in the expanding vapor. Its solubility in a

    magmatic vapor, initially at say 1000 bars consequently decreases by a

    factor of 1000 in expanding to 100 bars pressure simply due to the

    changing pressure of the water-dominated vapor phase.Similarly, there

    is an overall solubility decrease of 106 in expandingto a surface pressure

    of 1 bar. Inclusion of solvation water molecules (Eq.  (1c)) shows a

    higher order dependence on the activity or pressure of water; here H2O

    distinguishes the properties of solvation water from free water.

    3CuðHSÞ0

    g ;nH2Og þ  AsðOHÞ0

    3;g þ  H2Sg  ¼  Cu3AsS4 þ  3H2Og þ  H2;g þ  nH2Og

    ð1cÞ

    The release of water molecules from the solvation sphere due to

    vapor expansion and decrease in the dielectric constant has a multiplier

    effecton solubility decrease. Copper, andothermetals in thevaporphase

    may also be present as chloride species so that when vapor expands into

    the halite plus vapor  eld (Fig. 6) the removal of available chloride to

    halite signicantly reduces their solubility; this is potentially most

    signicant for a ‘hard’ metal such as iron in controlling pyritedeposition.

    The solubility of enargite in a vapor is also, through Eq.   (1b),

    dependent upon the rate of change of aH2S/aH2. While wallrock

    reactions are commonly assumed to control the redox state (aH2) of 

    the  uid phase in many ore-forming systems, the chemical processes

    in a vapor expanding within a fracture array through impermeable

    silica–alunite altered wallrocks are dominated by homogenous gas

    reactions. The activity of hydrogen, aH2,g, is controlled by a matrix of 

    reactions but primarily those involving sulfur (Giggenbach, 1987).

    SO2ðgÞ þ  3H2ðgÞ  ¼  H2SðgÞ þ  2H2OðgÞ   ð2aÞ

    K 2a  ¼  aH2Sd  a  2

    H2O =aSO2d  a

      3H2

    ð2bÞ

    SO2(g) dominates the gaseous sulfur species at magmatic tempera-

    turesand, through thenegativevolume changeof Reaction 2a as written,

    is also favored by decreasing pressure. Equilibrium calculations show

    that aH2S/aH2   does not vary by more than a factor of 10 over the

    temperature range 350 to 900 °C or between 10 and 1000 bar pressures

    at 600 °Cand is accordingly farlesssignicant than thevariation of water

    pressure. Giggenbach(1987), Martin et al. (2009) and Montegrossi et al.

    (2008) have drawn attention to the non-equilibrium behavior of the

    H2S–

    SO2 reaction which in turn implies unfavorable kinetics as a redoxcontrol reaction.

    The rapid change in water fugacity and density consequent on

    vapor expansion leads to major changes in the solubility of metals   –

    and silica   –   in the vapor phase. As a consequence, decompression

    results in an immediate very high supersaturation of metals that, in

    turn, triggers a cascade of far-from-equilibrium reactions (Henley and

    Berger, 2000).As shown in Fig. 1, at El Indio, following pyrite and local

    enargite deposition, Cu–Fe–As–S melt condensed from the vapor

    phasewith a range of trace heavy metals. As progressive crystallization

    proceeded the sulfosalt melts dissolved earlier pyrite9 and enargite and8 ) n, the number of solvation sphere water molecules, is a statistical value extracted

    fromexperimental data forthe effective numberof water or other molecules surrounding

    a coremolecular species.The bindingof eachsolvation molecule through,forexample, Van

    der Waals bonding involves a solvation free energy that contributes to the overall

    solubility equilibrium constant for the dissolution reaction in the vapor phase.

    9 The sulfosalt melt+pyrite+vapor reaction occurs via chalcopyritesp  (Fig. 1e) with

    ultimate resorption of the chalcopyrite (Fig. 1e) but leaving islands of pyrite within the

    Fe-tennantite to produce the disease texture described by  Imai (1999) at Lepanto. The

    complex redox reaction process is described elsewhere (Mavrogenes et al., in prep).

    70   R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    9/12

    fractionation of antimony, bismuth, tin, selenium, tellurium, silver (and

    gold) occurred between the melt and vapor phase to produce

    chemically zoned Fe-tennantite and other sulfosalts, tellurides and

    bismuth minerals (Mavrogenes and Henley, in preparation) in an

    analogous manner to that described by Tomkins (2010) for sulfosalts in

    magmatic sulde ore deposits. The simultaneous growth of euhedral

    quartz during melt crystallization and fractionation (Fig. 1c and d) is a

    consequence of high initial vapor phase silica solubility and the limited

    solubilityof water insul

    de–

    andby analogy sulfosalt–

    melts (Wykes andMavrogenes, 2005). The end product of the sulfosalt–sulde crystalliza-

    tion sequence is a vapor phase withrelativelyhigh gold and silvercontent

    which ultimately separates to form silica-rich high grade gold veins.

    Elsewhere vapor deposition may have occurred directly to solid state

    sulfosalt phases with partitioning of heavy metals between solid and

    vapor.

    5. Discussion

    It is well established that during thecrystallization of melts, metals

    partition into magmatic vapor (e.g.,   Candela and Holland, 1984;

    Simon et al., 2008). This process occurs early during crystallization to

    develop an aqueous   uid phase within sub-volcanic intrusions at

    lithostatic pressure. At 4 km depth, for example, the maximum

    pressure is about 1000 bars. Where leakage occurs directly, or more

    likely indirectly via a magmatic-vapor reservoir, into active fracture

    systems open to thesurface (1 to 110 bars, see Section 2.2) the dropin

    uid pressure is more than 890 bars. This disparity in  uid pressures

    drives the expansion and discharge of magmatic vapor and the

    consequent vapor-phase solubility disequilibria that result in the

    sequence of sulfosalt–sulde stages typical of high-suldation

    deposits. Since rapid vapor expansion constrained by active fracture

    arrays within 1500 m or so of the paleosurface involves only minor

    heat losses and temperature decrease the Sulde Stage is initiated by

    high temperature sulde–sulfosalt assemblages as described by

    Mavrogenes et al. (2010). Crystalline enargite and pyrite may form

    directly from the vapor phase at temperatures above about 650 °C. At

    El Indio and elsewhere these phases are subsequently dissolved by

    a Cu–As–S melt that condensed from the expanding vapor phaseand which progressively crystallizes with fractionation of heavy

    metals between melt and vapor giving rise, through the possible

    mixing of groundwater in open fracture arrays, to the later complex

    telluride–bismuth–tin–tungsten enriched vein assemblages with gold

    and silver. The recognition of the sulfosalt melt stage is strongly

    supported by the combination of several lines of evidence of evidence

    as is illustrated in   Fig. 1. Key evidence includes the following: (a)

    euhedral quartzcrystalsare wholly enclosed within cavities and veinlets

    with the sulfosalt (Fig. 1a–c), (b) symmetrical zonation of arsenic,

    antimony and tellurium in Fe-tennantite around quartz crystals, (c)

    Fe-tennantite demonstrably dissolves early pyrite (Fig. 1e),

    (d) preservation of vugs and vug-lling suldes and sulfosalts,

    (e)   uid inclusions in quartz in vugs in Fe-tennantite crystals

    homogenize to a vapor at   N450 °C indicating a minimum trappingtemperature (Mavrogenes et al., 2010).

    5.1. The volcanic context of high-sul dation gold deposits

    It is a matter of common observation that, between eruptive stages,

    magmas beneath composite volcanoes continueto release magmatic-gas

    mixtures that expand via fracture arrays to form high-temperature

    fumaroles and extensive near-surface alteration. Field data for enargite–

    gold lode deposits dene an analogous magmatic-vapor expansion

    context punctuated by eruption and dome extrusion. Following his

    detailed study of Goldeld, Nevada, Ransome (1909, 1910)  suggested

    that the ores there may have been formed by the ascension of hot, acidic

    solutions emanating from a deep magma reservoir. The role of volcanic

    gases in the origin of gold deposits was subsequently debated

    intermittently until substantially brought into focus in a review of 

    volcanic-gas chemistry by Hedenquist (1995). This review showed that

    despitethedif cultiesin samplingvolcanic gases during eruptionor from

    active high-temperature fumaroles between eruptive stages, the

    observed  ux of metals could be argued as suf cient to account for ore

    formation within geologically reasonable time periods (104–105 years)

    depending of course on the ef ciency of   uid focusing and metal

    deposition in the superstructure of the volcanic system.

    Whether or not there is a direct link between the formation of enargite–gold lode deposits and porphyry copper–gold deposits is

    immaterial to our hypothesis for the fracture-controlled expansion of 

    magmatic vapor in the formation of high-suldation gold deposits. The

    one is not   necessarily   tied to the other. At Lepanto, for example, the

    ranges of the K–Ar data for porphyry and high-suldation mineraliza-

    tion (Arribas et al., 1995,   GSA Data Repository Item 9517) do not

    demonstrate contemporaneity of mineralization but simply allow that

    each developed as discrete events of approximate duration 50,000–

    100,000 years during the history of the Mankayan volcanic complex

    over 700,000 years. The relative structural timing of mineralization and

    alteration events is discussed by   Berger and Henley (in press). At

    Famatina, Pudack et al. (2009) carefully detailed the development of a

    magmatic-vapor phaseassociated with earlier high-levelintrusionsthat

    are spatially related to weakly-mineralized but high-grade, enargite–

    famatinite–gold–silver veins associated with alunite. However, the

    chain of evidence between the degassing of the earlier epizonal

    intrusions and the development of the enargite–famatinite veins is

    broken due to extensive erosion of at least 800 m of the host-rock

    sequence prior to the sulfosalt mineralization stage, and the rarity of 

    usable uid inclusions in the sulfosalt stage samples.

    Porphyry copper–gold deposits occur in large but still locally

    constrained areas of extensionalfracturing in contrastto high-suldation

    deposits that are restricted to much more isolated and irregular fracture

    arrays through impermeable rocks. For both, isotope data consistently

    identify the role of magmatic vapor sensu latu, but it is more likely in a

    composite volcanic system that they record a   “magmatic-vapor

    reservoir” containing a mixture of magmatic gases derived from suites

    of discrete melt reservoirs that may also include basalts (e.g.,   Roberge

    et al., 2009) rather than any single or specic reservoir. The dynamics of theuidphase in such a reservoirare determined by crustal stress which,

    in conjunction with pore-uid pressure, leads at some intermediate

    depth to theaccumulation of magmatic vapor from availablesourcesand

    controls higher-level fracture formation and reactivation (Cox et al.,

    2001) assuggested by seismic data at Yellowstone(Husen et al., 2004).In

    turn this suggests that high-suldation and porphyry Cu–Au deposits

    occupy different structural regimes at different times as stress changes

    progress during the history of a large volcanic complex, directly

    determining the available structures within which mixing between

    magmatic vapor and groundwater can occur (Fig. 5c, MG); the ΔP/ΔH

    trajectory determines whether the two-phaseuid regime is intersected

    with the potential for the formation of porphyry copper–gold style

    mineralisation.

    5.2. The near-surface discharge regime

    At and very near the surface in  “degassing” volcanic environments,

    condensation of high-enthalpy volcanic vapor occurs very ef ciently

    through heat transfers due to interaction with the atmosphere, contact

    with highly permeable, epiclastic sediments, or by condensation and

    mixing into groundwater regimes and crater lakes (Brantley et al.,

    1987).10 These processes inevitably lead to the formation of localized

    10 Crater lakes form through the complete condensation of magmatic vapor and

    admixture of meteoric water. Their mineralogy and geochemistry are therefore a

    composite of sulfate and sulde geochemistry with clays, alunite, enargite, pyrite, etc.

    carried in suspension, ejecta and sediment (e.g., Kawah Ijen, Indonesia,  Delmelle and

    Bernard, 1994; Kawah Putih, Indonesia, Sriwana et al., 2000).

    71R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63 –74

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    10/12

    bodies of advanced argillic and associated more extensive zones of 

    advanced argillic and argillic alteration as well as perched layers of 

    acidic   “steam-heated”  water near the surface that mimic topography

    and which are the basic elements of solfatara as observed in active or

    morestrictly quiescentvolcanic environments today.Thesephenomena

    constitute a highly ef cient trap for the remaining metals exiting

    the array of   ‘feeder’   fractures and result in the shallow formation

    of bulk-tonnage, but low-grade, gold deposits. The bulk-tonnage

    Pascua–Lama deposit, for example, has a depth extent of about 500 m

    and is distinguished by the extraordinary dynamic complexity of therelationships between an active structural array and relationships

    between silica–alunite andpyrite–enargite deposition (Chouinard et al.,

    2005). The occurrence of both alunite and sulfosalts in some veins is a

    consequence of rapidly changing pressures and temperatures in the

    near surface part of the system due to complex interplay between

    condensate and magmatic vapor on multiple   ow paths in the

    extensive, active fracture system (Berger and Henley, 2011). In

    more permeable systems, magmatic vapor expanding on isolated

    fracture arrays may become trapped in groundwater  ow systems

    with consequent dilution and dispersion of metals, as is commonly

    observed for arsenic, for example, in groundwater around active

    volcanoes.

    6. Summary 

    High-suldation gold deposits occur in andesite–dacite volcanic

    complexes and relate closely to the discharge of magmatic-vapor gas

    as it expands at depths of only a few hundred meters beneath

    high-temperature fumaroles and solfatara. Their association with

    surcial volcanic rocks however led Lindgren (1913) to classify these

    depositsas epithermal implying formationtemperatures on theorderof 

    180–300 °C. Recent stable isotope,  uid inclusion and micro-analytical

    data however have demonstrated a clearer relationship to much higher

    temperature fumarolic discharges as indeed was suggested by  Cross

    (1891, 1896), and a magmatic uid component of at least 75%mass.

    These deposits are therefore more appropriately termed  “Solfataric

    gold deposits”   in order to distinguish them from epithermal (sensu

    stricto) gold deposits (e.g., McLaughlin, California) whose geological

    context and alteration and mineralization processes relate to the

    shallow   ‘hot-springs’  discharge regime of paleo-geothermal systems

    where, due to mixing, the magmaticuid component is less than about

    30%. The time-integrated elements of Solfataric gold deposits are

    summarized in Fig. 7 in relation to the pressure and temperature within

    an expandingcolumnof magmatic vapor,the spatialcontrol imposed by

    active fracture arrays and the formation of solfataras with extensive

    advanced argillic alteration and low-grade mineralization solfatara at

    the paleosurface. The narrowness of the isotherms localized by owing

    fractures and the high thermal gradients are important to note, alongwith the expansion of the advanced argillic alteration due to heat loss

    and   ‘ponding’  of mixed acid condensate–groundwater aquifers below

    the surface. Recognition of the paleo-environment and controls on

    vapor expansionis of course signicant in exploration forthesedeposits.

    Thissynopsis alsoappliesto solfatarasin activevolcanoesand associated

    acidic geothermal systems (Ruaya et al., 1992) and equivalent sea oor

    volcanic gas discharges, such as at Brothers Volcano in the southern

    Kermadec Arc, New Zealand (de Ronde et al., 2005).

     Acknowledgments

    We thank Richard Arculus, Paul Barton, Phil Bethke, James Cleverley,

    Graham Hughes, Gary Landis, Sue Kieffer, Laurie Madden, John

    Mavrogenes, Agnes Reyes, Cornel de Ronde, and Bob Rye for the helpfulcomments and discussion. Dana Bove, Jeff Hedenquist, Kalin Kouzmanov

    and Francois Robert kindly provided some of the samples illustrated in

    Fig. 1. We thank Steve Kesler and Bob Bodnar for providing constructive

    reviews on the original manuscript and Thomas Driesner for providing

    access to the programs used in construction of the NaCl–H2O phase

    diagrams.

    References

    Africano, F., Bernard, A., 2000. Acid alteration in the fumarolic environment of Usuvolcano, Hokkaido, Japan. Journal of Volcanology and Geothermal Research 97,475–495.

    Africano, F., van Rompaey, G., Bernard, A., Le Guern, F., 2002. Deposition of traceelements from high temperature gases at Satsuma-Iwojima volcano. Earth Planet

    Space 54, 275–

    286.

    Fig. 7.  The relationship between pressure, temperature and depth for a magmatic vapor expanding to the surface through a fracture array, in relation to the scale and geologic

    elements of a typical high-suldation deposit (modied after Bergerand Eimon,1983). At the surface, rapid open-system cooling occurs due to atmosphericinteractions and mixing

    with shallow groundwater or perched layers of condensate to produce a solfataric environment (note that the Solfatara Zone is shown as extending above the zero surface reference

    in order to accommodate topographic variations on hydraulic head). For further discussion see text and  Berger and Henley (in press).

    72   R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

    http://localhost/var/www/apps/conversion/tmp/scratch_7/image%20of%20Fig.%E0%B7%80

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    11/12

    Anderson, G.M., Crerar, D.A., 1993. Thermodynamics in Geochemistry. OxfordUniversity Press, Oxford. 588 pp.

    Arribas Jr., A., 1995. Characteristics of high-suldation epithermal deposits, and theirrelation to magmatic  uid. Mineralogical Association of Canada Short Course 23,419–454.

    Arribas Jr., A., Hedenquist, J.W., Itaya, T., Okada, T., Rogelio, A.C., Garcia Jr., J.S., 1995.Contemporaneous formation of adjacent porphyry and epithermal Cu–Au depositsover 300 ka in northern Luzon, Philippines. Geology 23, 337–340.

    Berger, B.R., 2007. The 3D fault and vein architecture of strike–slip releasing andrestraining bends: evidence from volcanic-centre-related mineral deposits. In:Cunningham,W.D., Mann, P. (Eds.),Tectonics of Strike–Slip Restraining and Releasing

    Bends: Geological Society of London, Special Publication, 290, pp. 447–

    471.Berger, B.R., Eimon, P.I., 1983. Conceptual models of epithermal precious metaldeposits. In: Shanks, W.C. (Ed.), Cameron Volume on Unconventional MineralDeposits. Society of Mining Engineers of AIME, pp. 191–205.

    Berger, B.R., Henley, R.W., 2011. Magmatic-vapor expansion and the formation of high-suldation gold deposits: Structural controls on hydrothermal alteration and oremineralization. Ore Geology Reviews 39, 75–90.

    Bethke, P.M., Rye, R.O., Stoffregen, R.E., Vikre, P.G., 2005. Evolution of the magmatic-hydrothermal acid-sulfate system at Summitville, Colorado: integration of geological, stable-isotope, and   uid-inclusion evidence. Chemical Geology 215,281–315.

    Brantley, S.L., Borgia, A., Rowe, G., Fernandez, J.F., Reynolds, J.R., 1987. Poas volcanocrater lake acts as a condenser for acid metal-rich brine. Nature 330, 470 –472.

    Candela, P.A., Holland, H.D., 1984. The partitioning of copper and molybdenumbetween silicate melts and aqueous  uids. Geochimica et Cosmochimica Acta 48,373–380.

    Chouinard, A., Williams-Jones, A.E., Leonardson, R.W., Hodgson, C.J., Silva, P., Tellez, J.V., Rojas, F., 2005. Geolog y and genesi s of the multistag e high-s uldationepithermal Pascua Au–Ag–Cu deposit, Chile and Argentina. Economic Geology

    100, 463–490.Churakov, S.V., Tkachenko, S.I., Korzhinskii, M.A., Bocharnikov,R.E., Shmuloviich, K.I., 2000.

    Evolution of compositionof high-temperaturefumarolicgases fromKudryavyVolcano,Iterup, Kuril Islands: the thermodynamic modeling. Geochemistry International 38,436–451.

    Claveria, R.J.R., 2001. Mineral paragenesis of the Lepanto copper and gold and theVictoria gold deposits, Mankayan district, Philippines. Resource Geology 51,97–106.

    Cox, S.F., 2010. The application of failure mode diagrams for exploring the roles of  uidpressure and stress states in controlling styles of fracture-controlled permeabilityenhancement in faults and shear zones. Geouids 10, 217–233.

    Cox, S.F., Knackstedt, M.A., Braun, J., 2001. Principles of structural control onpermeability and   uid   ow in hydrothermal systems. Reviews in EconomicGeology 14, 1–24.

    Cross, W., 1891. On alunite and diaspora from the Rosita Hills, Colorado. American Journal of Science 48, 460–475 3rd Series.

    Cross, W., 1896. Geology of Silver Cliff and the Rosita Hills, Colorado. U.S. GeologicalSurvey Seventeenth Annual Report, 2, pp. 263–403.

    De Ronde, C.E.J., Hannington, M.D., Stoffers, P., Wright, I.C., Ditchburn, R.G., Reyes, A.G.,Baker,E.T.,Massoth, G.J., Lupton,J.E.,Walker,S.L., Greene,R.R.,Soong,C.W.R., Ishibashi,

     J., Lebon, G.T., 2005. Evolution of a submarine magmatic-hydrothermal system:Brothers Volcano, Southern Kermadec Arc, New Zealand. Economic Geology 100,1097–1133.

    Deditius, A.P., Utsunomiya, S., Ewing, R.C., Chryssoulis, S.L., Venter, D., Kesler, S.E., 2009.Decoupled geochemical behavior of As and Cu in hydrothermal systems. Geology37, 707–710.

    Delmelle, P., Bernard, A., 1994. Geochemistry, mineralogy, and chemical modeling of the acid crater lake of Kawah Ijen Volcano, Indonesia. Geochimica et CosmochimicaActa 58, 2445–2460.

    Deyell, C.L., Bissig, T., Rye, R.O., 2004. Isotopic evidence for magmatic-dominatedepithermal processes in the El Indio-Pascua Au–Cu–Ag Belt and relationship togeomorphologics etting.Society of EconomicGeologistsSpecial Publication 11, 55–73.

    Deyell, C.L., Leonardson, R., Rye, R.O., Thompson, J.F.H., Bissig, T., Cooke, D.R., 2005.Alunite in the Pascua-Lama high-suldation deposit: Constraints on alteration andore deposition usingstable isotope geochemistry. Economic Geology100, 131–148.

    Driesner, T., 2007. The system H2O–NaCl: II. Correlations for molar volume, enthalpy,andisobaricheat capacity from 0 to 1000 °C, 1–5000 barand 0–1 XNaCl. Geochimica

    et Cosmochimica Acta 71, 4902–

    4919.Filippou, D., St-Germain, P., Grammatikopoulos, T., 2007. Recovery of metal values from

    copper—arsenic minerals and other related resources. Mineral Processing andExtractive Metallurgy Review 28, 247–298.

    Foley, N.K., Bethke, P.M., Rye, R.O., 1989. A reinterpretation of the  δDH2O of inclusionuids in contemporaneous quartz andsphalerite, Creede Mining District, Colorado:a generic problem for shallow orebodies? Economic Geology 84, 1966–1977.

    Fournier, R.O., 1985. The behavior of silica in hydrothermal solutions. In: Berger, B.R.,Bethke, P.M. (Eds.), Geology and Geochemistry of Epithermal Systems. : Reviews inEconomic Geology, 2. Society of Economic Geologists, pp. 45–61.

    Fournier, R.O., Potter II, R.W., 1982. An equation correlating the solubility of quartz inwater from 25° to 900°C at pressures up to 10,000 bars. Geochimica etCosmochimica Acta 46, 1969–1974.

    Fulignati, P., Sbrana, A., 1998. Presence of native gold and tellurium in the active high-suldation hydrothermal system of the La Fossa volcano, Vulcano, Italy. Journal of Volcanology and Geothermal Research 86, 187–198.

    Giggenbach, W.F., 1987. Redox processes governing the chemistry of fumarolic gasdischarges from White Island, New Zealand. Applied Geochemistry 2, 143–161.

    Hastie, J.W., 1975. High Temperature Vapors. New York, Academic Press. 480 pp.

    Heald, P., Foley, N.K., Hayba, D.O., 1987. Comparative anatomy of volcanic-hostedepithermal deposits: acid sulfate and adularia-sericite types. Economic Geology 82,1–26.

    Hedenquist, J.W., 1995. The ascent of magmatic  uid: Discharge versus mineralization.Mineralogical Association of Canada Short Course 23, 263–289.

    Hedenquist, J.W., Matsuhisa, Y., Izawa, E., White, N.C., Giggenbach, W.F., Aoki, M., 1994.Geology, geochemistry and origin of high-suldation Cu–Au mineralization in theNansatsu district, Japan. Economic Geology 89, 1–30.

    Hedenquist, J.W., Arribas, A., Reynolds, T.J., 1998. Evolution of an intrusion-centeredhydrothermal system; Far Southeast-Lepanto porphyry and epithermal Cu–Audeposits, Philippines. Economic Geology 93, 373–404.

    Heinrich, C.A., 2005. The physical and chemical evolution of low-salinity magmaticuids at the porphyry to epithermal transition: a thermodynamic study.Mineralium Deposita 39, 864–889.

    Heinrich, C.A., Driesner, T., Stefansson, A., Seward, T.M., 2004. Magmatic vaporcontraction and the transport of gold from the porphyry environment toepithermal ore deposits. Geology 32, 761–764.

    Hemley, J.J., Hostetler, P.B., Gude, A.J., Mountjoy, W.T., 1969. Some stability relations of alunite. Economic Geology 64, 599–612.

    Hemley, J.J., Montoya, J.W., Marineko, J.W., Luce, R.W., 1980. Equilibria in the Systemand Some General Implications for Alteration/Mineralization Processes. EconomicGeology 75, 210–228.

    Henley, R.W., Berger, B.R., 2000. Self-ordering and complexity in epizonal mineraldeposits. Annual Reviews of Earth and Planetary Sciences 28, 669–719.

    Henley, R.W., Ellis, A.J., 1983. Geothermal systems ancient and modern: a geochemicalreview. Earth Science Reviews 19, 1–50.

    Henley, R.W., McNabb, Alex, 1978. Magmatic vapor plumes and groundwaterinteraction in porphyry copper emplacement. Economic Geology 73, 1–20.

    Huang, C.K., Chou, S.A., 1975. Parageneses of enargite and luzonite–famatinite in theChinkuashih gold–copper deposits, Taiwan. Acta Geologica Taiwanica 18, 26–35.

    Huang, C.K., Tsai, H.L., 1984. Chemical variations of enargite and luzonite–famatinitefrom the Chinkuashih gold–copper deposits, Taiwan. Acta Geologica Taiwanica 22,118–130.

    Husen, S.,Smith, R.B., White,G.P.,2004. Evidence forgas andmagmatic sources beneaththe Yellowstone Volcanic Field from seismic tomographic imaging. Journal of Volcanology and Geothermal Research 131, 397–410.

    Imai, Akita, 1999. Pyrite Disease in Luzonite from the Lepanto Cu–Au Deposit,Mankayan, Philippines: further example of disease texture and its origin. ResourceGeology 49, 163–168.

     Jannas, R.R., 1995. Reduced and oxidizing high-suldation deposits of the El Indiodistrict, Chile. Unpublished Ph.D. dissertation, Harvard University, Cambridge, 421pp.

     Jannas, R.R., Bowers, T.S., Petersen, U ., Beane, R.E., 1999. High-suldation deposit typesin the El Indio district, Chile. In: Skinner, B.J. (Ed.), Geology and Ore Deposits of theCentral Andes: Society of Economic Geologists Special Publication, 7, pp. 219–266.

    Kodas, T., Hampden-Smith, M., 1994. The Chemistry of Metal CVD. WeinheimPublishers, New York. 530 pp.

    Kouzmanov, K., Pettke, T., Heinrich, C.A., 2010. Direct analysis of ore-precipitatinguids: combined IR microscopy and LA-ICP-MS study of  uid inclusions in opaqueore minerals. Economic Geology 105, 351–373.

    Krauskopf, K.B., 1964. The possible role of volatile metal compounds in ore genesis.Economic Geology 59, 341–410.

    Larocque, A.C.L., Stima, J.A., Siebe, C., Greengrass, K., Chapman, R., Mejia, S.R., 2008.Deposition of a high-suldation Au assemblage from a magmatic volatile phase,Volcán Popocatépetl, Mexico. Journal of Volcanology and Geothermal Research170, 51–60.

    Lindgren, W., 1913. Mineral Deposits, 1st edition. McGraw-Hill Book Co, New York. 883pp.

    Losada-Calderon, A., Bloom, M.S.,1990. Geology, paragenesisand uid inclusion studiesof a high sulphidation epithermal/porphyry system, Nevados del Famatina district,La Rioja Province, Argentina. Australasian Institute of Mining and Metallurgy,Pacic Rim Congress 90, Proceedings 2, pp. 457–464.

    Mancano, D.P., 1994. Microthermometry of enargite-hosted  uid inclusions from theLepanto, Philippines, high-suldation Cu–Au deposit using infrared microscopy.Unpublished M.Sc. thesis, New Mexico Institute of Mining and Metallurgy, Socorro,New Mexico, 74pp.

    Mancano, D.P., Campbell, A.R., 1995. Microthermometry of enargite-hosted   uid

    inclusions fromthe Lepanto, Philippines,high-suldation Cu–

    Au deposit.Geochimicaet Cosmochimica Acta 59, 3909–3916.

    Martin, R.S., Roberts, T.J., Mather, T.A., Pyle, D.M., 2009. The implications of H2S and H2kinetic stability in high-T mixtures of magmatic and atmospheric gases for theproduction of oxidized trace species (e.g., BrO and NOx). Chemical Geology 263,143–150.

    Mavrogenes, J., Henley, R.W., in prep, Sulfosalt melts and heavy metal(As–Sb–Bi)fractionation during volcanic gas expansion.

    Mavrogenes, J., Henley, R.W., Reyes, A., Berger, B., 2010. Sulfosalt melts: evidence of high-temperature vapor transport of metals in the formation of high-suldationlode gold deposits. Economic Geology 105, 257–262.

    Meyer,C., Hemley, J.J., 1967.Wall rock alteration.In: Barnes, H.L. (Ed.),Geochemistry of Hydrothermal Ore Deposits. Holt, Rinehart and Winston Inc, pp. 166 –235.

    Montegrossi, G., Tassi, F., Minissale, A.A., Vaselli, O., Buccianti, A., 2008. Naturaluctuations of sulfur species in volcanic fumaroles. Journal of Non-EquilibriumThermodynamics 33, 75–102.

    Moore, J.N., Allis, R.G., Nemcok, M., Powell, T.S., Bruton, C.J., Wannamaker, P.E., Raharjo, I.B.,Norman, D.I.,2008. The evolution of volcano-hosted geothermalsystemsbased on deepwells from Karaha–Telaga Bodas, Indonesia. American Journal of Science 308, 1–48.

    73R.W. Henley, B.R. Berger / Ore Geology Reviews 39 (2011) 63–74

  • 8/20/2019 2011-Epitermal (HS-2;Henley and Berger) 2

    12/12

    Morey, G.W., Hesselgesser, J.M., 1951. The solubility of quartz and some other substancesin superheated steam at high pressures. Economic Geology 46, 821–835.

    Moritz, R., 2006. Fluid salinities obtained by infrared microthermometry of opaqueminerals: implications for ore deposit modeling—a note of caution. Journal of Geochemical Exploration 89, 284–287.

    Moritz, R., Benkhelfa, F., 2009. Fluids in high-suldation epithermal ore deposits:constraints from infrared microthermometry of enargite-hosted  uid inclusions.Extended abstract, 10th Biennial SGA Conference, Townsville, Australia, 16–20August, pp. 533–535.

    Muntean, J.L., Einaudi, M.T., 2001. Porphyry-epithermal transition: Maricunga Belt,northern Chile. Economic Geology 96, 743–773.

    Nuccio, P.M., Paonita, A., Sortino, F., 1999. Geochemical modeling of mixing betweenmagmatic and hydrothermal gases: the case of Vulcano Island, Italy. Earth andPlanetary Science Letters 167, 321–333.

    Okamoto,A., Saishu,H., Hirano,N., Tsuchiya, N., 2010. Mineralogical andtextural variationof silica minerals in hydrothermalow-through experiments: implications for quartzvein formation. Geochimica et Cosmochimica Acta 74, 3692–3706.

    Pokrovski, G.S., Zakirov, I.V., Roux, J., Testemale, D., Hazemann, J.-L., Bychkov, A.Y.,Golikova, G.V., 2005. Experimental study of arsenic speciation in vapor phase to500 °C: implications for As transport and fractionation in low-density crustal uidsand volcanic gases. Geochimica et Cosmochimica Acta 66, 3453–3480.

    Pokrovski, G.B., Borisova, A.Y., Harrichoury, J.-C., 2008. The effect of sulfur on vapor–liquidfractionation of metals in hydrothermal systems. Earthand Planetary ScienceLetters 266, 345–362.

    Posfai, M., Buseck, P.R., 1998. Relationships between microstructures and compositionin enargite and luzonite. American Mineralogist 83, 373–382.

    Posfai, M., Sundberg, M., 1998. Stacking disorder and polytypism in enargite andluzonite. American Mineralogist 83, 365–372.

    Pudack, C., Halter, W.E., Heinrich, C.A., Pettke, T., 2009. Evolution of magmatic vapor togold-rich liquid: the porphyry to epithermal transition at Nevados de Famatina,

    northwest Argentina. Economic Geology 104, 449–477.Ransome, F.L., 1909. The geology and ore deposits of Goldeld, Nevada. U.S. Geological

    Survey Professional Paper, 66. 258 pp.Ransome, F.L., 1910. Geology and ore deposits of the Goldeld district, Nevada.

    Economic Geology 5, 301–311.Roberge, J., Delgado-Granados, H., Wallace, P.J., 2009. Mac magma recharge supplies

    high CO2   and SO2   gas   uxes from Popocatepetl Volcano, Mexico. Geology 37,107–110.

    Ruaya, J.R., Ramos, M.N., Gonantini, R., 1992. Assessment of magmatic components of the  uids at Mt. Pinatubo volcanic-geothermal system, Philippines from chemicaland isotopic data. Report of the Geological Survey of Japan, 279, pp. 141 –151.

    Ruggieri, G., Lattanzi, P., Luxoro, S.S., Dessi, R., Benvenuti, M., Tanelli, G., 1997. Geology,mineralogy, and  uid inclusion data of the Furtei high-suldation gold deposit,Sardinia, Italy. Economic Geology 92, 1–19.

    Rye, R.O., 1993. The evolution of magmatic  uids in the epithermal environment: thestable isotope perspective. Economic Geology 88, 733–752.

    Rye, R.O., 2005. A review of the stable-isotope geochemistry of sulfate minerals inselected igneous environments and related hydrothermal systems. ChemicalGeology 215, 5–36.

    Rye, R.O., Bethke, P.M., Wasserman, M.D., 1992. The stable isotope geochemistry of silica–alunite alteration. Economic Geology 87, 225–262.

    Sack, R.O., Goodell, P.C., 2002. Retrograde reactions involving galena and Ag-sulfosaltsin a zoned ore deposit, Julcani, Peru. Mineralogical Magazine 66, 1043–1062.

    Sakazaki,H., Ohtagaki,T., Chin, W.,1964. Geologyand oredepositsof theChin-qua-shihMine District (2). Mining Geology 14, 274–285.

    Sillitoe, R.H., Hedenquist, J.W., 2003. Linkages between volcanotectonic settings, oreuid compositions and epithermal precious metal deposits. In: Simmons, S.F.,

    Graham, I. (Eds.), Volcanic, Geothermal and Ore-forming Fluids: Rulers andWitnesses of Processes within the Earth: Society of Economic Geologists SpecialPublication, 10, pp. 315–343.

    Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C., 2008. The partitioningbehavior of silver in a vapor–brine–rhyolite melt assemblage. Geochimica etCosmochimica Acta 72, 1638–1659.

    Skinner, B.F., 1960. Assemblage enargite–famatinite, a possible geologic thermometer(abs.). Geological Society of America Bulletin 71, 1975.

    Sourirajan, S., Kennedy, G.C., 1962. The system H2O–NaCl at elevated temperatures andpressures. American Journal of Science 260, 115–141.

    Sriwana, T., van Bergena, M.J., Varekampc, J.C., Sumartib, S., Takanod, B., van Ose, B.J.H.,

    Lengf, M.J., 2000. Geochemistry of the acid Kawah Putih lake, Patuha Volcano,West Java, Indonesia. Journal of Volcanology and Geothermal Research 97, 77–104.Stoffregen, R.E., 1987. Genesis of acid–sulfate alteration and Au–Cu–Ag mineralization

    at Summitville, Colorado. Economic Geology 82, 1575–1591.Stoffregen, R.E., Rye, R.O., Wasserman, M.D., 1994. Experimental studies of alunite: I.

    18O–16O and D–H fractionation factors between alunite and water at 250–450 °C.Geochimica et Cosmochimica Acta 58, 903–916.

    Sung, J., 2005. Mineralogy and geochemistry of gold mineralization in the Goldelddistrict, Nevada. unpublished MSc thesis, University of Utah, Salt Lake City, 155 pp.

    Taran, Y.A., Bernard, A., Gavilanes, J.-C., Africano, F., 2000. Native gold in mineralprecipitants from high temperature volcanic gases of Colima volcano, Mexico.Applied Geochemistry 15, 337–346.

    Taran, Y.A., Bernard, A., Gavilanes, J.C., Lunezheva, E., Cortes, A., Armienta, M.A., 2001.Chemistry and mineralogy of high temperature gas discharges from Colimavolcano, Mexico: implications for magmatic gas–atmosphereinteraction. Journal of Volcanology and Geothermal Research 108, 245–264.

    Tomkins, A.G., 2010. Wetting facilitates late-stage segregation of precious metal-enriched sulfosalt melt in magmatic sulde systems. Geology 38, 951–954.

    Toulmin III, P., Clark Jr., S.P., 1967. Thermal aspects of ore formation. I