Skarn Paper Myers

Download as pdf or txt
Download as pdf or txt
You are on page 1of 46

Skarn Systems and Exploration Concepts

By
Greg Myers Ph.D.
BHP-Billiton, Tintaya, Peru
Summary

Skarn deposits are important sources of iron, copper, gold, lead, zinc, silver, and
tungsten throughout the world and are important exploration targets in Peru and South
America. Skarns form in specific tectonic settings and develop systematic characteristics
of mineral composition and zoning, and igneous rock affinities, which increase the
exploration geologist’s ability to identify favorable systems and the probable metal
associations. The identification of metamorphic and metasomatic features and their areal
extent are the keys to identifying the most favorable targets.
Six skarn deposit types, based on metal content, were described by Einaudi et al.
(1981); these are copper, lead-zinc, tungsten, iron, molybdenum, and tin. The tectonic
setting, igneous rock association, formation conditions, and calc-silicate mineralogy
varies among the skarn types, allowing some distinction between the groups (Einaudi et
al., 1981). For instance, the tungsten skarns of eastern California are thought to represent
formation pressures near 2 KB (Newberry and Einaudi, 1981), zinc-lead skarn such as
Groundhog or Cananea are thought to have formed at relatively shallow depths (0.5Kb)
(Meinert, 1983). These skarn metal groups are associated with different types of igneous
rocks, oxidizing and reducing magmatic settings, and variable protoliths.
Iron skarns are associated with dioritic or mafic igneous rocks under oxidizing
conditions, generally in magnetite or hematite stability. Gold and copper are often
associated with iron skarns but seldom form important gold or copper deposits. Cobalt
and nickel are common trace metals associated with iron skarns. Iron skarns form in
limestone and dolomite. The mineralogy of calcic iron skarns is dominated by andradite.
Alteration assemblages in dolomite include olivine, diopside, actinolite, and serpentine.
Trace metals associated with iron skarns include Co and Ni. Retrograde alterat ion is
common with garnet altering to epidote-quartz- magnetite.
Copper skarns are very important economic deposits and are associated with
calcic to calc-alkalic igneous rocks. Important copper skarn deposits include Tintaya,
Peru, Chino, N.M. USA, and Candeleria, Chile. Copper skarns are associated with
porphyry and non-porphyry intrusive bodies. Byproduct gold, molybdenum, and silver
are common but occur in low concentrations. Copper skarns also form in oxidizing
conditions and usually contain magnetite or hematite. The metal assemblage associated
with copper skarns is generally simple and includes Cu, Au, Mo, and Ag. The
mineralogy of copper skarns is andradite dominant with lesser amounts of light green
diopsidic pyroxene. Retrograde alteration is generally limited to diopside altering to
actinolite.
Gold skarns are important gold and copper producers and include deposits such as
Gratsberg/Ertsberg, Indonesia and Fortitude, NV, USA. Polymetallic gold skarns are
associated with reduced calc-alkalic igneous rocks, which can form porphyry or non-
porphyry bodies. These gold skarns form in more reducing conditions and pyrrhotite is
more common than magnetite and hematite is generally absent. Gold skarns generally
contain important concentrations of copper with significant zones of zinc and lead.
Associated metals include As, Sb, Bi, Te, Se, The mineralogy associated with gold skarns
contains adraditic garnet, but the volume of pyroxene is much greater than iron and
copper skarns and the dark green pyroxene is hedenbergitic. Metals zone from copper,
proximal to the intrusive, to zinc and lead distal to the intrusive.
Lead-zinc skarns are associated with reduced alkalic intrusive rocks and include
deposits such as Santa Eulalia, Mexico, Darwin, CA, USA, and Fresnillo, Mexico. The
mineralogy is pyroxene dominant and the dark green to black pyroxene is hedenbergitic
to johannsenitic or Fe and Mn enriched with a lesser amount of andraditic garnet. Lead-
zinc skarns are polymetallic with a strong bias toward silver and manganese and other
associated metals include As, Sb, Bi, Te, Se, Mo, W, and Au, with minor Cu.
Tungsten skarns are an important source of tungsten worldwide. They form in
deeper settings than the other skarn types and are associated with regional and contact
metamorphic features as well as metasomatic alteration. Garnet and pyroxene
compositions are quite variable depending on the composition of the protolith and are not
indicative of the type of skarn. The major metals associated with Tungsten skarns
include W, Au, Bi, and Cu.
Identification of tectonic setting, igneous rock types, hornfels and skarn
distribution, skarn mineralogy, alteration and mineralization zonation, and trace element
assemblages allow the exploration geologist to define the potential size and probable
metal content of skarn prospects.

Introduction

Skarn deposits occur throughout the world's tectonic belts (Fig. 1) and represent
an important contribution to the world's base and precious metals production. Gabbroic
to granitic rocks associated with skarns range in age from Precambrian to Quaternary.
Skarn deposits are the primary source of tungsten and produce significant amounts of
copper, iron, zinc, silver, lead, and gold. The wide range of igneous rock ages and
compositions, worldwide distribution, and the variety of contained metals make skarns an
important exploration target. Calc-silicate bodies at intrusive contacts have been
recognized and studied for over 150 years, during which time several terminologies have
been used. The term skarn was introduced by Scandinavian miners to refer to the coarse-
grained calc-silicate gangue (amphibole, garnet, pyroxene, etc.), which was associated
with iron deposits hosted by Proterozoic sediments (Geijer and Magnussen, 1952; Burt,
1982). The usage of the term refers to calc-silicate rocks formed by several processes
and from a variety of protoliths. The calc-silicate minerals can replace interbedded
lithologies, such as limestone, dolomite, shale, or marl. Replacement mineralogies and
textures can vary greatly among the different host rocks. Coarse-grained garnet which
replaced limestone may be interbedded with fine-grained diopside which replaced
calcareous siltstone. These alteration/replacement processes range from metamorphic
(isochemical, closed-system) to metasomatic (open-system) and many skarns exhibit a
continuum of these processes, making interpretation and terminology difficult. These
types of occurrences will be referred to as skarns or skarn deposits based on their calc-
silicate mineralogy, without an implied genesis. Impure calcareous rocks or shaley
interbeds within limestone-hosted skarn may have alteration ranging from fine-grained,
biotite hornfels to massive, coarse- grained garnet depend ing on the degree of
metasomatic interaction. Fine-grained, calc-silicate alteration, with limited chemical
changes from the protolith, is often termed hornfels. Skarn is often coarse-grained calc-
silicate minerals resulting from the interaction of abundant fluids adding iron and silica.
Alteration intermediate between hornfels and skarn is commonly referred to as skarnoid,
that is, not strictly metamorphic and not completely metasomatic (Korzkinskii, 1948;
Einaudi et al., 1981; Einaudi and Burt, 1982; Meinert, 1992).

Figure 1.
Distribution of Skarn Mining Districts in relation to the worlds tectonic belts.
(Modified after Theodore et al., 1991)

Hornfels and Skarn Alteration and Zonation

Hornfels alteration is often a good indicator of the potential size of a skarn


system. The mineralogy of the hornfels and the areal extent of alteration indicate the
intensity of dominantly metamorphic reactions surrounding an intrusive complex.
Hornfels formed in calcareous siltstone has distinct mineralogy zoning patterns from
earliest formed biotite hornfels to pyroxene-kspar-albite-quartz to garnet alteration
assemblages, depending on the original composition of the protolith and intensity of
alteration. The relative volume of pyroxe ne to biotite increases with increasing alteration
temperature and the proximity to skarn.
Alteration features common in limestone include bleaching of gray to black
limestone to white. Carbon fronts within the limestone. Recrystallization of the
limestone to marble. The formation of wollastonite around chert nodules and siliceous
beds within limestone. Calcite veining often increases close to skarn contacts.
The lager the volume of the contact metamorphic to metasomatic altered rocks the
greater the chance of encountering significant volumes of skarn alteration. The biotite –
pyroxene hornfels halo surrounding the Fortitude gold skarn deposit covers over 17
square kilometers (Fig. 2).
Skarn alteration shows general zoning trends but because of the co mplex cross-
cutting and overprinting alteration events zoning is observed at both a hand sample scale
as well as a deposit scale. Generally speaking, skarn zones from garnet dominant at the
intrusive-skarn contact to pyroxene dominant at the skarn-limesto ne contact. Quite often
a garnet dominant zone forms between the pyroxene dominanat zone and the limestone-
skarn contact. Garnet color, grain size, and morphology is very erratic and does not
provide a reliable guide to skarn zonation on a deposit wide scale. Purely metasomatic
garnet is andraditic in composition and the color of this garnet can vary from brown to
red to yellow to green, thus garnet color is generally not a guide to composition in a
metasomatic system. Garnet composition is more closely tied to the protolith
composition. Garnet formed in a calcareous siltstone or in endoskarn is more
grossularitic than garnet formed in clean limestone. Pyroxene composition is closely tied
to color. Light green pyroxene is more diopsidic than dark green to brown hedenbergitic
pyroxene. Johannsenitic pyroxene is dark green to brown to black. Retrograde skarn is
commonly formed at the skarn limestone contact or at the skarn-endoskarn contact, but
the distribution of retrograde skarn is most strongly contolled by permeability so
structures and lithologic contacts are strong controls on the distribution of retrograde
skarn.

Figure 2.
Map of the biotite-pyroxene hornfels surrounding the Copper Canyon Stock.
Thermodynamics of hydrothermal skarn systems

Stability conditions for the various calc-silicate minerals, which formed during
hydrothermal alteration, constrain the possible formation conditions. Several variables
define the stability conditions for each mineral. Stability relations vary according to
pressure, temperature, fO 2 , fS 2 , pH, etc.
Pressure variability is very important to the stability temperature range for
andradite as is the concentration of CO2 . Figure 3 shows the effect pressure and CO2
have on the stability of andradite and wollastonite. Andradite stability
increases with increasing temperature, pressure, and CO 2 concentration. The effect of
choosing 0.4 Kb for this evaluation places temperature estimates at a relative minimum.
An estimated CO2 value of 0.1 mo le fraction will also give minimum temperature

Figure 3.
Stability diagram for calc-silicate minerals in relation to temperature and oxidation state.
These conditions ane mineral assemblages are assemblages seen at the Fortitude-Copper
Canyon Deposit in Nevada, USA.
estimates. This approximate concentration is a reasonable estimate as determined from
analyses of fluids contained in fluid inclusions from other skarn systems (Taylor and
Liou, 1978). Wollastonite and garnet both respond to variations in pressure and CO2
molarity, with an increasing stability temperature threshold, as pressure and CO 2 molarity
increase.
The calc-silicate mineral assemblages with respect to temperature and the fluid
oxidation state (fO 2 ) show the buffering trends of the various mineral phases at Copper
Canyon. Temperature estimates based on this plot probably represent a minimum
depending on actual pressure and CO2 molarity. The relative absence of wollastonite,
even as reaction rims surrounding chert nodules in the limestone, places an upper
temperature limit of about 500-575°C on the system under low XCO2 . This temperature
is similar to the fluid inclusion temperatures from garnet and pyroxene. The increasing
concentration of CO 2 or increasing pressure would greatly increase the stability
temperature of wollastonite possibly accounting for its very restricted nature.
The oxidation state of the hydrothermal fluids is constrained by the Hm-Mt buffer
and the C -CO2 buffer. Hematite is absent in the skarn and the carbon in the altered
limestone is oxidized, placing limits on the log fO 2 values at 500°C between -18 to -25.
A more constrained estimate of about -21 at 500°C can be made if the assumption is
made that the system evolved approximately on the andradite-diopside buffer, a
reasonable supposition for the high temperature, Stage 1 alteration.
Decreasing temperature of the hydrothermal system resulted in a change of the
oxidation state of the fluid. The dominance of andradite in the proximal alteration zones
and hedenbergite in the distal zones demonstrates the changing oxidation conditions with
decreasing temperature and distance from the Copper Canyon Stock. A closed system
would evolve along the mineral buffered trend of andradite-diopside with decreasing
temperature and maintaining approximately the same compositions, but the Copper
Canyon system became more reduced with decreasing temperature deviating from a
mineral buffered trend and deviating to more reduced mineralogic associations.
Buffering of the hydrothermal fluid was partially controlled during interaction with the
increasingly reduced outer alteration fringe and reduced host rock. Zones of black
carbonaceous marble, calcareous siltstone, and limestone occur along portions of the
distal skarn- limesto ne contact.
The lower limit of garnet-pyroxene stability is marked by the reaction:

Diopside + magnetite = andradite + quartz + calcite


and
Hedenbergite = andradite + quartz + calcite

This reaction marks the deposition of pyroxene skarn, which lacks garnet and began
between 400 and 450°C with fO 2 between -25 to -28.
The oxidation state of the fluid is also reflected in the composition of the
pyroxene. Hedenbergite contains Fe+2, the abundant ion in reduced settings. The
reaction is approximated by:

diopside = hedenbergite
CaMg+2Si2 O 6 + Fe+2 = CaFe+2Si2 O6 + Mg+2
Diopside is stable under higher oxidation states than hedenbergite at any given
temperature. Hedenbergite stability favors reduced, Fe +2 enriched fluids, while the
concentra tion of Mg+2 , in diopside, is independent of oxidation. The abundance of
hedenbergite indicates deviation, in the oxidation trend during cooling, to more relatively
more reduced conditions.
A water buffered trend maintains the oxidation state of the sys tem at
approximately the same value with falling temperature. Epidote is abundant in a water
buffered, oxidized trend and the retrograde alteration of the garnet to epidote begins at
relatively high temperatures (400-450°C). Epidote dominant retrograde skarn is most
abundant in oxidized, garnet rich, copper skarns (Einaudi et al., 1981).
At any given temperature and pressure, the proportion of iron-rich garnet (Fe +3 ) to
iron-rich pyroxene (Fe+2 ) will partially reflect the oxidation state of the system (Table 1);
more oxidized systems will be garnet-rich and more reduced systems will be
pyroxene-rich (Meinert, 1983).

Table 1.
Mineral assemblages and reactions for different oxidation states
of hydrothermal fluids related to skarn formation.

Oxidation Hydrothermal Fluid Alteration


Conditions Provides Removes Reactions Assemblage
Fe+3 >Fe+2 SiO 2
T>500 C Mg CO 2 Andradite
3CaCO3 +Fe2 O3 +3SiO2 =Ca3 Fe2 Si3 O12 + CO2
Log fO2 –15 to –22
Log fS2 >-4 Fe+3+2 Orthoclase
K
oxidized
Fe+3 =Fe+2 SiO 2 CO 2
3CaCO3 +Fe2 O3 +3SiO2 =Ca3 Fe2 Si3 O12 + CO2
Mg Ad80+Hd20
Fe+3+2 CaCO3 +0.8MgO+0.2Fe3 O4 +2SiO2 =
CaMg0.2 Si2 O6 +CO2
moderately K
oxidized Al CaFe+3 +3AlSi3 O12 =Si2 O+SiO2 +CaCO3 +e-

Fe+3 <Fe+2 SiO 2 CO 2


3CaCO3 +Fe2 O3 +3SiO2 =Ca3 Fe2 Si3 O12 + CO2
T=400-450 C Mg HCO3 Ad60 + Hd40
Log fO2 =-25 to –30
Log fS2 = -10 Fe+3+2 CaCO3 +0.6MgO+0.4Fe3 O4 +2SiO2 =
CaMg0.6 Fe0.4Si2 O6 +CO2
moderately K
reduced Al CaFe+3 +3AlSi3 O12 =Si2 O+SiO2 +CaCO3 +e-

Fe+3 <Fe+2 SiO 2 HCO3


T=350-400 C Mg CaCO3 +0.2MgO+0.8Fe3 O4 +2SiO2 = Hd80
Log fO2 =-25 to –30 CaMg0.2 Fe0.8Si2 O6 +CO2
Log fS2 =< -10 Fe+3+2
reduced
Fe+3 <<Fe +2 SiO 2
T=<350 C Mg 5CO3 + 5CaMgSi 2 O6 = Ca2Mg5 Si8 O22 (OH)2 Actinolite
Log fO2 =-28 to –32 + 2SiO2 + 5CaCO3
Log fS2 =< -10 Fe+3+2 Chlorite
strongly 5CO3 + 5CaFeSi 2 O6 = Ca2Fe5 Si8O22 (OH)2
+ 2SiO2 + 5CaCO3
reduced
Calc-silicate mineral compositions are stated as the proportion of end member
compositions in a solid solution series. For example, the compositional range between
Grossular garnet Ca3 Al2 Si3 O12 and Andradite garnet Ca3 Fe2 Si3O 12 is stated as the
percentage of end member garnet, therefore, a garnet with a composition of 80%
andradite and 20% grossular would be described as Ad 80 or Gr20. The same relationship
and nomenclature applies to the pyroxene solid solution of diopside, hedenbergite, and
johannsenite.

Evolution of Magmatic Systems and Exsolution of Hydrothermal Fluids

Starting with the most basic evaluation of an igneous system, the distinction of
wet and dry magmas needs to be determined. A magma body crystallizes in a specific
sequence based on original composition of the melt. Bowen described the sequence fom
early, high temperature, anhydrous minerals such as olivine and Ca plagioclase to later,
low tempperature quartz (Fig. 4). In order to form primary hornblende or biotite in an
igneous rock a certain concentration of water must exist in the melt. As anhydrous
minerals precipitate the water content of the melt gradually increases. Original
mineralogy of the most mafic igneous rocks in an intrusive complex are a good guide to
the potential of that complex to generate significant volumes of hydrothermal fluids. Dry
magmas do not crystallize mineralogic associations, which are dominated by hydrous
phases. A parent magma which forms early pyroxene gabbro bodies has less potential to
generate hydrothermal fluids capable of forming large skarn systems than a parent
magma which forms early hornblende-biotite gabbro. As the magmatic system evolves
the hydrous phases of the igneous rocks increase until the magma becomes so saturated
with fluids that it exsolves the first hydrothermal phase. The ratio of primary hornblende
to biotite is an important indicator of the relative water content of the melt. This may be
reflected in the concentration of primary biotite which gradually increases during
crystallization of the magma. The size of the biotite is also an indicator of the amount of
contained water in the melt. Large books of biotite, especially elongated along the C -axis
indicate a wet magma. The release of a hydrothermal fluid phase at this point of the
crystallization may revert the magma into hornblende dominant mineralogy, which again
builds water content as non-hydrous phases crystallize. Magmatic systems with the
greatest potential of forming economic skarn deposits may become supersaturated with
respect to water, exsolve a hydrothermal fluid, thus becoming undersaturated in water,
and build to a supersaturated condition several times, releasing several pulses of
hydrothermal fluids.
Olivine Ca Plagioclase

Clinopyroxene

Orthopyroxene

Hornblende

biotite Na Plagioclase

Muscovite

K Feldspar

Quartz

Figure 4.
Crystallization sequence for minerals present in igneous rocks from higher temperature
Olivine and Ca Plagioclase to lower temperature Quartz (modified from Bowen, )

The stepped differentiation and crystallization of a magma body evolves in a


systematic process with different metal and alteration assemblages being concentrated in
different stages of the crystallization process (Fig. 5). Observations of changing igneous
rock mineralogies helps to identify to most favorable stage of the magmatic
differentiation for economic mineralization. Key observations in the stepped magmatic
evolution are the appearance of biotite, the significant decline in the hornblende content
of the igneous rock, the disappearance of hornblende, and the disappearance of biotite. In
a typical porphyry copper system the most important economic stage is at the
disappearance of hornblende (Keith, 1995).
Quartz Feldspar
Porphyry

Stage 4
Porphyry Mo-K
(Ag-Pb-Zn-W-As-Cu-Mn-F)
System associated with fluid exsolution
following biotite disappearance

Biotite Gronodiorite
Porphyry

Stage 3
Porphyry Cu-Mo-K-silicates
(W-Zn-Pb-Ag-Fe-Mg)
System associated with fluid exsolution
following hornblende disappearance

Biotite-Hornblende
Granodiorite

Stage 2
Porphyry Fe-Mg-Na-Ca (Cu)
system associated with fluid exsolution
following major decline in
hornblende abundance

Hornblende
Quartz Diorite

Stage 1
Carbonate-quartz system

Hornblende Gabbro

Figure 5.
Differentiation of a magmatic series and the evolution of stepped hydrothermal fluids in
an oxidized calc-alkalic igneous rock (modified after Keith, 1995).
Skarn Types and Mineralization

Skarn deposits can be divided into seven basic metal types, iron, copper, gold,
zinc, tungsten, which form important economic deposits, molybdenum, and tin skarns,
which are generally small and uneconomic. These skarn metal groups are most common
in the worlds Mesozoic tectonic belts and are associated with different types of intrusive
rocks of variable ages. The division of skarns by metal associations is a guideline, but
not a strict rule, as hybrid groups cross boundaries and are important to consider during
exploration. Skarns can be partially classified by calc-silicate mineralogy (Table 2).
Garnet and pyroxene compositions vary by skarn type. Oxidation state of the
hydrothermal skarn forming fluids is the primary control with lithology and igneous rock
compositions contributing to calc silicate mineralogy. For instance, alkalic igenous rocks
contain more manganese than calc -alkalic igneous rocks and pyroxene formed in
association with reduced alkalic intrusions is enriched in manganese (johansennitic to
bustamitic).

Table 2.
Approximate Skarn Garnet and Pyroxene Compositional Ranges

Skarn Type Garnet Composition Pyroxene Composition


Calcic Iron Skarn Ad 60-100 Di 70-100
Copper Skarns Ad 60-100 Di 70-100
Gold Skarns Ad 60-100 Hd 20-100
Zinc-Lead Silver Skarns Ad 60-100 Hd 20-100 Jo 0-100
Tungsten Skarns Ad 20-85 Di 70-100

Iron Skarns

Iron skarns are generally associated with dioritic or mafic to granodioritic igneous
rocks under oxidizing conditions, generally in magnetite or hematite stability. Iron
skarns are commonly formed in two distinct settings associated with limestone or
dolomite. The calcic iron skarns are most common in island arc settings, whereas
dolomitic iron skarns are more common in Cordilleran orogenic belts. Gold and copper
are often associated with iron skarns but seldom form important gold or copper deposits
(Table 3). Cobalt and nickel are common trace metals associated with iron skarns. The
mineralogy of calcic iron skarns is dominated by andradite. Alteratio n assemblages in
dolomite include olivine, diopside, actinolite, and serpentine. Trace metals associated
with iron skarns include Co and Ni. Retrograde alteration is common with garnet altering
to epidote-quartz-magnetite.
Table 3.
Examples of Iron Skarn Occurrences

Deposit Size (million Fe grade % Cu grade % References


tonnes)
Calcic Iron Skarn
725 45.6 0.04 Sokolov and
Sarbai, Turgai Grigorev (1977)
West Siberia, Russia
Larap, Philippines >20 43 0.12 Frost (1965)
Shinyama, Japan >10 30-35 0.1 Tsusue (1961)
Daiquiri, Cuba >10 Lindgren and Ross
(1916)
Shasta, USA 5 38 Lamey (1946)
Kasaan Peninsula, USA >20 >40 <0.5 Myers (1985)

Dolomitic Iron Skarns


234 35 Sokolov and
Sheregesh , Russia Grigorev (1977)

Eagle, USA >50 45 DuBois and


Brummet (1968)
Fierro, USA >20 45 0.03-2 Hernon and Jones
(1968)
Nabesna, USA >15 >35 <0.5 Newberry et al.
(1996)

The iron skarns on Kasaan Peninsula, Alaska are similar to other island arc hosted
iron skarns of the Pacfic Rim. These skarns are commonly formed in host rocks
consisting of interbedded limestone, calcareous siltstone, and andesitic volcanics,
intruded by dioritic to granodioritic rocks. Metamorphic to metasomatic reactions
commonly produce garnet and garnet pyroxe ne skarns in the calcareous rocks and
endoskarn in the igneous rocks. Epidote and albite are common endoskarn minerals.
Generally, mineralogic zoning from intrusive contact to limestone follows a pattern of
garnet-epidote-albite endoskarn to garnet skarn at the original limestone-intrusive
contact. Garnet-pyroxene assemblages form outward from the garnet zone and often a
second garnet zone forms at the skarn-limestone contact. Retrograde alteration is most
common at the skarn- limestone contact and consists of epidote-quartz- magnetite
replacing garnet. Garnet and pyroxene compositions vary depending on the type of host
rock being replaced, but in general garnet (Ad 70-100 ) and pyroxene (Hd 10-30 ) replacing
limestone. Mineralization consists of massive magnetite which primarily forms near the
skarn-limestone contact with minor veins of chalcopyrite (Myers, 1985).
Copper Skarns

Copper skarns are very important economic deposits and are associated with
oxidized calcic, calc-alkalic, and quartz alkalic igneous rocks. Important copper skarn
deposits include Tintaya, Peru, Santa Rita, N.M. USA, and Candeleria, Chile (Table 4).
Copper skarns are associated with porphyry and non-porphyry intrusive bodies.
Byproduct gold, molybdenum, and silver are common but occur in low concentrations.
Copper skarns also form in oxidizing conditions and usually contain magnetite or
hematite. The metal assemblage associated with copper skarns is generally simple and
includes Cu, Au, Mo, and Ag. The mineralogy of copper skarns is andradite dominant
with lesser amounts of light green diopsidic pyroxene.

Table 4.
Examples of Copper Skarn Deposits

Deposit Size (million Au grade Cu grade % References


tonnes) gm/tonne

Tintaya, Peru 195 0.024 1.4 Myers (2001)


Ruth Nevada >300 0.015 1.1 James, 1976
Christmas, Arizona 100 low 0.7 Perry (1969)
Santa Rita, New Mexico >350 low 0.9 Nielson (1970)
Bingham, USA 1,350 low 0.9 Einaudi (1981)
Morenci, USA >350 low 1.1 Moolick and Dureck
(1966)
Mission, USA 400 low 0.8 Einaudi (1981)
Candaleria, Chile >500 ? >1
Twin Buttes, USA 400 low 0.73 Barter and Kelly
(1981)
Gold Coast, Papua New 30 ? 2.5 Davies et al (1978)
Guinea
Cobriza, Huancavelica, Peru >50 ? ? Cahoon (1970)

Tintaya, Peru

The Tintaya Mine is located in South-eastern Peru (Figure 6). The copper skarn deposits
(Lat. 14° 54’; Long. 71° 20” W) are in the Yauri district, Cusco Department, in the south-
eastern corner of the Andahuaylas-Yauri copper belt of southern Peru. The mine occurs
at 4,100 meters (~13,500 ft), 256 km southwest of Cusco and 255 km northeast of
Arequipa.
Figure 6.
Location of the Tintaya Mine in Southern Peru.

Igneous Rock Assemblages, Magmatic Evolution, and Development of the


Hydrothermal System

The oldest igneous rocks recognized in the Tintaya area are microdiorite sills and
possible feeder dikes or plugs described as fluidal (Figure 7, 8). The microdiorite sills
are up to several hundred meters thick and are often associated with magnetite- garnet
skarns, barren of copper. The sills are fine-grained, hornblende rich diorite and are most
commonly found in the Ferrobamba Formation. The fluidal igneous rocks are
compositionally hornblende diorite with a population of aligned plagioclase phenocrysts
very similar compositionally to the microdiorite. The fluidal rocks are cut by quartz
monzodiorite. Quartz monzodiorite (PM1) is the main skarn forming igneous body and is
characterized by 1-2 vol. Percent quartz “eyes” prominent 3-6 mm wide biotite
“books”a nd subequal hornblende, prominent sphene and only 10-15% matrix. Several
syn and post mineral dikes of quartz monzodiorite compositions form northwest- and
northeast-striking dike swarms cutting Pm1. These dikes are denoted as PM2, PM3,
latite, and porphyritic diorite. Compositionally all of these intrusive rocks are similar and
have only minor variations in mineralogic composition and texture. Geochemically these
rocks are nearly identical and represent a co- magmatic association. The principle
differences in the igneous rocks are the percentage of phenocryst quartz (quartz eyes) and
the ratio of hornblende to biotite. For example, Pm2 typically lacks quartz eyes, contains
significant hornblende (8-10 volume per cent), fewer (3-5 volume per cent) and smaller
biotite phenocrysts, may lack sphene, and has 15 to 25% matrix.
The evolution of the magmatic and hydrothermal system began with the
emplacement and crystallization of the fluidal igneous bodies and associated microdiorite
sills. This period of igneous rock crystallization was mafic rich consisting primarily of
hornblende and plagioclase. During this crystallization the magmatic water content
increased to the point of supersaturation and exsolved, forming the early magnetite skarns
adjacent to the microdiorite sills and minor endoskarn or secondary biotite in the fluidal
rocks. The early magnetite skarns contain only trace amounts of copper. The continuing
evolution of the magmatic system again built the water content within the crystallizing
magma and the ratio of biotite to hornblende increased as crystallization of the quartz
monzodiorite intrusive bodies (PM1) advanced. The high content of magmatic water in
PM1 is noted by the presence of large biotite “books”. This timing of magmatic
differentiation is the

Figure 7.
Stratigraphy of the Tintaya Area.
33500.

34000.

34500.

35000.

35500.

36000.

36500.

37000.
43500. 43500.
33500

34000

34500

35000

35500

36000

36500

37000
43000.
43000 43000.

42500. 42500 42500.

42000.
42000 42000.

41500.
41500 41500.

41000.
41000 41000.

40500.
40500 40500.

40000. 40000.

39500. 39500.

39000. 39000.
33500.

34000.

34500.

35000.

35500.

36000.

36500.

37000.

38500. 38500.
No DATE MADE BY DESCRIPTION

SEDIMENTARIO INTRUSIVOS-DIQUES METASOMATICOS ESTRUCTURAS 1


BHP BILLITON
REVISIONS

MARMOL 523 LUTITA 524 PM - 1 519 ANDESITA 504 SKARN - Gr 508 FALLA (1) 591 FALLA (3) 590
3

4
GEOLOGIA TINTAYA
5 AREA GEOLOGICAL MODELING
DATE DRAWN BY CHECKED APPROVED MAP INDEX NUMBER SCALE DRAWING NUMBER

CALIZA 522 CUARZITA 507 TRAQUITA 619 DIORITA 502 SKARN - Mt 510 FALLA (2) 592 7-16-2001 1:6000.M

Figure 8.
Generalized Geologic Map of the Tintaya Area.
critical point in the evolution of the metasomatic system and marks the first major
exsolution of a skarn forming fluid with significant copper mineralization. During
crystallization of the earliest quartz monzodiorite the magma became supersaturated with
water, which began to separate from the magmatic melt. This exsolved fluid formed the
earliest copper bearing skarn bodies and formed orthoclase veins, quartz orthoclase veins,
and orthoclase flooding in the matrix of PM1. This fluid also became saturated with
elements, which were mobile in high temperature, acidic fluids, namely iron, copper,
gold, silver, and molybdenum. The magmatic system continued to crystallize, but with
the loss of a significant volume of water passed back into hornblende>biotite stability as
noted by the hornblende dominant PM2 dikes, which cut PM1. Magmatic water content
again began to increase with the crystallization of hornblende and silicate minerals and a
second period of subequal biotite- hornblende crystallization is noted in the PM3 dikes.
The exsolution of a magmatic fluid associated with the later PM3 is more irregular and
skarn alteration and mineralization is less extensive. The later dikes reflect this less
voluminous metasomatic event and the relatively dry conditions, after the period of PM3
emplacement, noted by the higher hornblende-biotite ratios in the latite and porphyritic
diorite dikes and the lack of skarn alteration or endoskarn associated with the late dikes.

Alteration of Igneous rocks


Potassic alteration- mineralization in Pm1 includes hairline, irregular,
discontinuous “A-veins”, K feldspar envelopes on barren or sulfide-bearing quartz veins
and K feldspar flooding of the groundmass. There is only local development of
secondary biotite. Cu mineralization related to Pm1 occurs in through going, quartz-
chalcopyrite+/-bornite+/- molybdenite veins and is typically subeconomic, but can locally
exceed 2% total copper in small zones, typically close to skarn contacts. Potassic
alteration in PM-2 includes both hairline, irregular, discontinuous “A-veins” and
pervasive replacement of magmatic amphibole and biotite by secondary biotite. Low
grade (0.5-0.8% total Cu) mineralization occurs in zones of strong secondary biotite.
Alteration in PM-3 typically includes pervasive orthoclase in the groundmass and
generally lacks quartz-orthoclase veining. Sausserization of the feldspar is common with
epidote and chlorite replacing some of the plagioclase phenocrysts.

Endoskarn
Endoskarn alteration in PM1 and PM2 is dominated by garnet and pyroxene, high
temperature mineral phases. Skarn related to PM1and PM2 is also exclusively garnet
pyroxene prograde skarn. Endoskarn alteration varies from very weak, scattered
replacement of plagioclase to garnet or diopside to a texture destructive, total
replacement of the igneous rock.
Endoskarn is generally absent in PM3 and later dikes. Alteratio n in PM3 is
dominantly epidote and chlorite, lower temperature phases. Further evidence that the
hydrothermal event related to PM3 is volumetrically less important is the relative lack of
retrograde skarn where PM-3 dikes cut skarn.
Skarn
Skarn-related alteration and associated Cu-Au-Ag-Mo mineralization occurs along the
contact between Cretaceous limestone of the Ferrobamba Formation and Tertiary quartz
monzodiorite. Skarn occurs over a strike length of at least 2.5 km as a series of steeply
dipping, irregularly shaped bodies up to 100 meters wide and extending down the dip of
the intrusive contact up to at least 500 meters (Refer to Figures 9,10, and 11). Prograde
skarn forms in carbonate rocks and the alteration generally exhibits classic skarn zonation
from proximal garnet to pyroxene to magnetite-sulfide at the marble front. Retrograde
alteration, locally replaces garnet and pyroxene skarn and consists of chlorite-actinolite-
epidote-clay-chalcopyrite veins and chalcopyrite-pyrite-calcite veins with actinolite-
chlorite envelopes.

Hornfels
Hornfels forms in the calcareous siltstone horizons in the Ferrobamba and Mara
Formations. The hornfels consists of biotite with diopside, orthoclase, diopside, and
garnet. Generally the hornfels is unmineralized but with increasing interaction with
metasomatic fluids the hornfels will host mineralization as veins and disseminations of
chalcopyrite.
The limestone also exhibits alteration features, which help lead to metasomatic sources.
The limestone is dark gray in color due to a high carbon content. Bleaching,
recrytallization, and calcite veining occurs with increasing interaction with metasomatic
fluids. Theses reactions are not simply metamorphic reactions as noted by the proximity
of these alteration features to skarn.

Mineralization
Economically important copper mineralization at Tintaya occurs almost solely with
prograde skarn alteration. Copper mineralization occurs as veins, disseminations, and
massive replacement zones. Chalcopyrite is more abundant than bornite. Copper and
gold grades show a strong positive correlation, with gold more strongly associated with
bornite than with chalcopyrite. Primary mineralization is not associated with the late
retrograde alteration and is economically unimportant. Locally copper is remobilised
with the retrograde alteration.
4000 4000

3800 3800

LIMESTONE LATITE Topography

FORM. MARA MONZONITE PIT

SKARN Diorite Porphyry

DD
Figure 9. H

Section 33 looking Northeast in the Chabuca Norte Central Zone.

41400 41600 41800

LIMESTONE LATITE PIT

FORM. MARA PM-2B , PM- Topography


3

MONZONITE SKARN

Diorite Porphyry ENDOSKARN


DDH

Figure 10.
Section 34875 looking West in the Chabuca Este-Oeste Zone.
41800

41300

40800
35100 35600 36100 36600

LIMESTONE Diorite Porphyry SKAR


N

FORM. MARA LATITE ENDOSKAR


N

MONZONITE PM-2B , PM-3

Figure 11.
Composite Bench 3900 through Tajo, to the Eas,t and Chabuca Este-Oeste, to the West.

Structure
Three fault systems are recognized at Tintaya. From oldest to youngest, these include:
1) east-west striking normal faults; 2) conjugate northwest- and northeast-striking faults,
and 3) conjugate north-south striking faults (Fig. 12). The oldest faults dip north and are
most prominent in the eastern portion of Chabuca Este. These faults contain calcite-clay-
pyrite-chalcopyrite, cut prograde skarn, and were important controls on the distribution of
retrograde alteration. The second fault system consists of conjugate northwest-striking,
northeast-dipping sinistral faults and northeast-striking, northwest-dipping dextral faults
formed during a north-south extensional regime. These faults also show normal
displacement. Pm2a dikes were emplaced in these faults. The third and youngest fault
system comprises a conjugate set of north-south striking, west-dipping normal faults and
north-south striking, east-dipping normal faults. These faults offset earlier fault systems,
control the emplacement of the latite and porphyritic diorite dikes, and formed during an
east-west extensional regime.
Figure 12.
1995 Airphoto of the Tintaya Open Pit and the beginning of the Chabuca
Este (ce) and Chabuca Sur (cs) pits. The lines on the photo represent linear feature
related to surface traces of major faults. The dominant sets of EW, NE,
and NW faults are visible.
Gold Skarns

Gold skarns are important gold and copper deposits and include deposits such as
Gratsberg, Indonesia and Fortitude, NV, USA (Table 5) and produce gold as the primary
metal recovered. Polymetallic gold skarns are associated with reduced calc-alkalic,
alkalai calcic, and quartz alkalic igneous rocks, which can form porphyry or non-
porphyry bodies. These gold skarns form in more reducing conditions and pyrrhotite is
more common than magnetite and hematite is generally absent. Gold skarns generally
contain important concentrations of copper with significant zones of zinc and lead.
Associated metals include As, Sb, Bi, Te, Se (Figg. 13). The mineralogy associated with
gold skarns contains adraditic garnet, but the volume of pyroxene is much greater than
iron and copper skarns and the dark green pyroxene is hedenbergitic. Metals zone from
copper, proximal to the intrusive, to zinc and lead distal to the intrusive. The indicative
opapue mineralogy includes pyrrhotite, arsenopyrite, gold, silver, and bismuth tellurides
and selenides, bismuthinite, and native gold, silver, and bismuth.

Gold skarns are generally polymetallic systems and have the characteristics of
copper and zinc- lead skarns. Most are highly enriched in arsenic, tellurium, bismuth,
antimony, selenium, copper, zinc, silver, and gold. Many gold skarn systems are large
and host significant gold mineralization and many are mined for copper as well. Gold
skarn pyroxene compositions are hedenbergite rich, notably different from the diopside
rich copper skarns and the johannsenite rich zinc- lead skarns. Important gold skarn
deposits include Fortitude, Nevada, Hedley, British Columbia, and Gratsberg/Ertsberg,
Indonesia.

Figure 13.
Skarn descrimination diagram based on trace and base metal concentrations.
Table 5.
Examples of Gold Bearing Skarns

Deposit Size (million Au grade Cu grade % References


tonnes) gm/tonne

Copper Canyon Nevada 41.18 4.0 0.4 Theodore et al., 1975


Myers (1994)

Carr Fork Utah 61 0.4 1.8 Tooker, 1989


Mount Hamilton Nevada 7.5 2.0 low Myers et al., 1990
Hedley B.C. 12 6.36 0.1 Ettlinger,
LaLuz Nicaragua 16 4.1 0.4 Theodore et al., 1991
Katanga Peru 2 6.1 3.0 Theodore et al., 1991
Crown Jewel, Washington 9.4 5.5 low Hickey, 1992
Zackly, Alaska 1.2 5.5 2.7 Ford, 198
Concepcion del Oro, Mexico >60 1.5 2 Myers (1997)
Gratsberg -Ertsberg, 523 2.0 1.7
Indonesia
McCoy, Nevada 14.5 1.5 0.1 Brooks (1991)
Cerro Gordo, USA <10 2 low Myers (1991)

Fortitude Nevada

The Eocene Copper Canyon stock, near Battle Mountain Nevada, hosts one of the
larger gold-enriched skarns in the world. The primary ore hosts are the Paleozoic
carbonate rocks of the Antler Peak Sequence (Fig. 14). Intrusive activity was associated
with crus tal thinning related to extension in the Great Basin. The carbonaceous
Pennsylvanian carbonate rocks buffered the magmatic and hydrothermal systems,
resulting in the formation of gold-enriched skarns.
The gold-copper skarns of the Copper Canyon area contain economically
important concentrations of metals. These deposits formed in central Nevada because
metals were concentrated into the hydrothermal phases of the crystallizing magmas. The
sedimentary sequence provided rocks reactive in a hydrothermal setting with buffering
properties, which enhanced the transport and deposition of the base and precious metals.
The tectonic history of Western North America juxtaposes compressional and
extensional events, each with associated magmatism. Subduction related magmatism,
with crustal affinities, may have interacted with Tertiary magmas related to crustal
thinning and a rising mantle. Some of these magmatic systems form zoned magmatic
complexes with associated hydrothermal systems, such as Copper Canyon and McCoy.
Zones of deep-seated crustal weakness and thinning appear to have focused magmatism
along linear trends. These magmatic trends are reflected in the distribution of gold
deposits in the Great Basin. The igneous rocks associated with the gold deposits of the
Great Basin may indicate magma source regions which are related to crustal thickness
and the types of rocks providing partial melts. Many of the Tertiary igneous rocks
contain elevated MgO:SiO 2 , Ce, Sc, Ni, and Cr and depleted Zr and Eu relative to crustal
rocks. These variations are similar to mantle derived magmas and to the Jurassic and
Cretaceous subduction related igneous rocks.

Figure 14.
Stratigraphy of the Fortitude-Copper Canyon area, Nevada.
Figure 15.
Generalized geologic map of the Copper Canyon-Fortitude Area, NV.
The Tertiary intrusive rocks intruded to higher levels (<1Kb) than the bulk of the
Jurassic and Cretaceous plutonic bodies (>1Kb) and interacted with the thick, reduced,
marine sequences of siltstone and limestone (Fig. 15). The interbedded nature of
siltstone and limestone, in the Copper Canyon area, provided relatively restricted
packages of carbonate rocks to interact with the hydrothermal fluids. The relatively thin
layers of limestone allowed the skarn to develop lateral zoning, away from the Copper
Canyon stock (Fig. 16 and 17).

Figure 16.
North-south cross section of the West and Fortitude Area.

More than seven copper and gold deposits occur in the hydrothermal cell around
the Copper Canyon stock. The skarn is zoned from garnet-chalcopyrite skarn near the
stock to pyroxene-pyrrhotite skarn in the distal, gold -enriched Fortitude area, near the
northern limit of skarn formation. alteration halo surrounding the Copper Canyo n stock.
Paragenesis of the skarn began with andraditic garnet (Ad85-100) deposition starting at
over 600°C, based
Hornfels alteration formed in the overlying Pumpernickle Formation and hides the
skarn formed in the underlying Antler Peak Limestone (Fig . 18 and 19). Skarn
assemblages formed in the calcareous sedimentary rocks as interaction with the
dominantly magmatic fluids progressed, leading to a zoned alteration halo surrounding
the Copper Canyon stock. Skarn assemblages formed in the calcareous sedimentary
Figure 17.
East-west cross section through the Fortitude Deposit.

rocks as interaction with the dominantly magmatic fluids progressed, leading to a zoned

on fluid inclusion studies, with minor diopsidic pyroxene (Hd10-30). This Stage 1 skarn
formed along the contact of the Edna Mountain Formation and the Antler Peak
Limestone, extending northward 2 Km, from the Copper Canyon stock. This earliest
skarn was exploited for copper and gold. Stage 2 skarn consists of a diopsidic pyroxene
(Hd20-40) dominant assemblage with lessor andraditic garnet (Ad60-90) and abundant
orthoclase. Stage 2 skarn cuts and overprints Stage 1 skarn. Stage 2 skarn formation
began at temperatures near 500°C. Pyrrhotite, pyrite, marcasite, and chalcopyrite are the
primary opaque minerals and telluride mineral deposition began with this skarn forming
period. Gold is widespread but concentrations are lower in Stage 2 skarn. Stage 3 skarn
is focused in the Fortitude area north of a zone of northeast trending faults. Stage 3 skarn
consists of hedenbergitic pyroxene (Hd>60), orthoclase and minor garnet (Ad40-70).
This stage of skarn formation is restricted to the Fortitude area north of the cross-fault
feeder zone. Retrograde skarn is minor in the Antler Peak Limestone, forming irregular
pods consisting of actinolite, chlorite, and epidote. The largest gold deposits are related
to this period of skarn formation and the associated metals include Cu, Bi, Te, As, Sb, Zn,
Figure 18.
Alteration distribution on the no rth-south A-A’ cross-section.

and Pb. Dominantly magmatic fluids (d18 O=5-10) reacted with the reduced carbonate
rocks, resulting in the formation of increasingly more reduced skarn mineralogies.
The evolution of the moderately reduced metal-bearing magma and the interaction
of the developing magmatic hydrothermal cell (d18 O = 4-8 per mil) with the surrounding
reduced sedimentary rocks (d 18 O = 20 per mil) and developing skarn assemblages (d 18 O =
7-12 per mil), created conditions favorable for metal transport, concentration, and
deposition. Gold deposition began at moderately high temperatures (>450°C) as metals
precipitated from chloride complexes during reactions under the reducing and
neutralizing conditions developing at the skarn-limestone reaction front. The
hydrothermal fluids became increasingly reduced with falling temperature, time, and
distance from the stock, increasing the solubility of bisulfide complexes while depositing
gold from chloride complexes. Dilution of the hydrothermal fluid, increasing oxidation,
or falling temperature lead to the deposition of gold from the bisulfide complexes during
Stage 3 skarn. Base metals are also strongly zoned with copper being most prevalent
with higher temperature alteration assemblages near the stock while zinc and lead are
dominant in the lower temperature Stage 3 skarn assemblage
Trace elements associated with this gold-enriched skarn include As, Bi, Te, and
Sb. Arsenic and Te were deposited with Stage 2 and Stage 3 skarn, while Bi and Sb are
more closely associated with late Stage 2 and Stage 3 skarn.
The Fortitude cross-fault zone influenced the location of late Stage 2 and Stage 3
skarn. This apparent feeder zone focused metal-laden hydrothermal fluids, which cooled
or mixed with oxidized fluids and deposited precious and base metals. This association
of distal feeder- like fault zones and high- grade gold mineralization also occurs at Hedley,
B. C. (Ettlinger et al., 1992).

Figure 19.
Alteration on the east-west B-B’ cross section in the Fortitude area.
Figure 20.
Distribution of copper, silver, and gold in the Fortitude area.
Figure 21.
Distribution of copper, silver, and gold in the Fortitude deposit.
Figure 22.
Schematic diagram of the paragenesis of the skarn system in the Fortitude area.
Zinc-Lead-Silver Skarns
Zinc-lead-silver skarns are associated with slightly oxidized to reduced alkalai-
calcic intrusive rocks and include deposits such as Santa Eulalia, Mexico, Darwin, CA,
USA, and Fresnillo, Mexico. The mineralogy of the more reduced skarns is pyroxene
dominant and the dark green to black pyroxene is hedenbergitic to johannsenitic or Fe
and Mn enriched with a lesser amount of andraditic garnet. The more oxidized Zn skarns
aare garnet dominant. Lead- zinc skarns are polymetallic with a strong bias toward Ag
and Mn and other associated metals include Cu, As, Sb, Bi, Te, Se, Mo, W, and Au, with
minor Cu. This group of skarns is quite variable in metal content with the Zn-Cu rich
skarns, such as Antamina, Peru, being more oxidized while the Pb-Ag rich are more
reduced.

Table 6.
Examples of Zinc -Lead Skarn Ocurrences
Deposit Size Ag grade Zn grade % Pb grade % References
(million gm/tonne
tonnes)
Antamina, Peru 559 13.7 1.03 Teck 2000 Annual
(1.23% Cu) Report
Darwin, USA >15 >20 >2% >2% Myers (1991)
Hanover, USA >5 150 14 0.3 Meinert (1980)
Santa Eulalia, Mexico >35 150 11 10 Hewitt (1968)
Fresnillo, Mexico >15 800 >2 >2 Koch (1956)
Myers (1993)
Santa Barbara, Mexico >150 250 10 15 Allen and Fahey
(1957)
Myers (1993)
Groundhog, USA 3 60 12 4 Meinert (1980)
Naica, Mexico 25 450 10 13 Erwood et al. (1979)
Myers (1993)
Uchucchacua, Peru small? 500 3 3 Alpers (1980)

Tungsten Skarns
Tungsten skarns are an important source of tungsten worldwide. They form in
deeper settings than the other skarn types and are associated with regional and contact
metamorphic features as well as metasomatic alteration. The igneous rock associations
are widely ranging from calcic to quartz alkalic and are both oxidized and reduced. The
igneous rocks are generally peraluminous whereas the other skarn groups are associated
with metaluminous igneous rocks (Keith, 1994). Peraluminous versus metaluminous
igneous rocks are define using A/CNK ratios (Al2 O 3 /CaO+Na2O+K2 O) where
peraluminous igneous rocks have a ratio greater than 1. Garnet and pyroxene
compositions are quite variable depending on the composition of the protolith and are not
indicative of the type of skarn. The major metals associated with Tungsten skarns
include W, Au, Bi, and Cu.

Table 7.
Examples of Tungsten Skarn Occurrences
Deposit Size (million WO3 grade % References
tonnes)
MacMillan Pass, Canada 65 0.95 Dick (1976)
Canada Tungsten, Canada >6 1.6 Archibald et al.
(1978)
Pine Creek, USA >6 0.5 Newberry (1980)
Sangdong, SE Korea >10 1.0 John, 1978
King Island, Australia 14 0.8 Kwak (1978)

Conclusions

Skarn deposits are important sources of iron, copper, gold, zinc, lead, silver, and
tungsten. The characteristics of each skarn metal type are identified primarily on calc-
silicate mineralogy and base and trace metal associations. Exploration begins with the
identification of favorable tectonic terranes. Andean type convergent margins generate
favorable igneous rocks. The second feature to identify are intrusive centers which
intrude “wet” sediments, which added to the water content of the rising and
differentiating magma. Features to look for in this evaluation are the ages of intrusives
and any associated mineralization. If a district contains multiple Cretaceous aged
intrusive bodies and related mineralization the later Tertiary igneous rocks may not have
been able to assimilate sufficient water to produce large ore bodies. The igneous rocks
again can help determine the potential to produce a large hydrothermal event by
evaluating the original igneous mineralogy (pyroxene or hornblende). The differentiation
of the igneous complex enables the determination of the number of hydrothermal events.
The cyclical crystallization of hornblende>biotite and biotite >hornblende may indicate
the exsolution of multiple hydrothermal events. The density of veins in the igneous rocks
is also an indicator of how wet the system was. A good tool to use for vein studies is a
contour map based on the number of veins per meter. In strongly altered systems these
vein density contour maps may require several maps for multiple vein types.
The identification of a favorable intrusive complex now requires that reactive
rocks are in contact with the hydrothermal fluids. Calcareous rocks are the most
favorable because they neutralize the acidic, metal bearing fluids causing the
precipitation of the metals. These chemical traps are as important as physical traps and
both are needed to help with the formation the best deposits. The final steps to skarn
exploration are the careful and detailed mapping of igneous alteration and alteration in
the sediments, whether they are calcareous siltstones with hornfels alteration or
carbonates with calc-silicate alteration. Geochemical sampling is very important to
determine if metal zoning is an important factor. Leakage of metals in the hydrothermal
fluids along fault zones or lithologic contacts is often the only way to identify underlying
skarn mineraliztion.

REFERENCES

Ahmad, S.N. and Rose, A.W., 1980, Fluid inclusions in porphyry and skarn ore at Santa
Rita, New Mexico: Econ. Geol., v. 75, p. 229-250.

Allen, V.T., and Fahey, J.J., 1957, Some Pyroxenes associated with pyrometasomatic
zinc deposits in Mexico and New Mexico: Geol. Soc. America Bull.v. 68, p. 881-
896.

Alpers, C.N., 1980, Mineralogy, paragenesis, and zoning of the Luz vein, Uchucchacua,
Peru: B.A. Thesis, Harvard Univ., 138 p.

Archibald, D.A., Clark, A.H., Ferrar, E., and Zaw, U.K., 1978, Potassium-Argon ages of
intrusion and scheelite mineralization, Cantung Tungsten, Northwest Territories:
Canadian Journal Earth Sci., v. 15, p.1205-1207.

Atkinson, Jr., W. W., Kaczmarowski, J. H., and Erickson, Jr., A. J., 1982, Geology of a
Skarn-Breccia Orebody at the Victoria Mine, Elko County, Nevada: Econ. Geol.,
vol. 77, no. 4, p. 899-918.

Axen, G. J., Taylor, W., and Bartley, J. M., 1993, Space-time patterns and tectonic
controls of Tertiary extension and magmatism in the Great Basin of the western
United States:Geological Society of America Bull., v. 105, no. 1, p.56-76.

Bakken, B. M. and Einaudi, M. T., 1986, Spatial and temporal relations between wall
rock alteration and gold mineralization, Main pit, Carlin Gold Mine, Nevada,
U.S.A.,: in Macdonald, A. J., ed., Proceedings of Gold '86, an International
Symposium on the Geology of Gold: Toronto, Canada, 1986, p. 388-403.

Bakken, B. M., 1991, Gold Mineralization, Wall-Rock Alteration, and the


Geochemical Evolution of the Hydrothermal System in the Main Ore Body,
Carlin Mine, Nevada: in Raines, G.L., Lisle, R.E., Schafer, R.W., and Wilkinson,
W.H. eds., Geology and Ore Deposits of the Great Basin Symposium
Proceedings, Geological Society of Nevada, p. 233-234.
Baker, E. D., 1991, Geology and Ore Deposits of the Bootstrap Subdistrict, Elko County,
Nevada: in Raines, G.L., Lisle, R.E., Schafer, R.W., and Wilkinson, W.H. eds.,
Geology and Ore Deposits of the Great Basin Symposium Proceedings,
Geological Society of Nevada, p. 619-624.

Barter, C.F., and Kelly, J.L., 1981, Geology of the Twin Buttes mineral deposit, Pima
County, Arizona in Titely, S.R., ed., Advances in the geology of porphyry copper
deposits, southwestern North America: Tucson, Univ. Ariz. Press p.407-432.

Bartlett M. W., Enders, M. S., and Hruska, D. C., 1991, Geology of the Hollister Gold
Deposit, Ivanhoe District, Elko County, Nevada: in Raines, G.L., Lisle, R.E.,
Schafer, R.W., and Wilkinson, W.H. eds., Geology and Ore Deposits of the Great
Basin Symposium Proceedings, Geological Society of Nevada, p. 957-978.

Barton, P.B., and Skinner, B.J., 1979, Sulfide mineral stabilities: in H.L. Barnes ed.,
Geochemistry of Hydrothermal Ore Deposits, J. Wiley, p. 278-403.

Blake, D.W., Wotruba, P.R., and Theodore, T.G., 1984, Zonation in the skarn
environment at the Tomboy-Minnie gold deposits, Lander County, Nevada:
Arizona Geol. Soc. Digest, v. 15, p. 67-72.

Brooks, J.W., Meinert, L.D., Kuyper, B.A., and Lane, M.L., 1991, Petrology and
Geochemistry of the McCoy Gold Skarn, Lander County, NV.: in Raines, G.L.,
Lise, R.E., Schafer, R.W., and Wilkinson, W.H. eds., Geology and Ore Deposits
of the Great Basin Symposium Proceedings, Geological Society of Nevada, p.
419-442

Brown, I.J. and Nesbitt, B.E., 1987, Gold -copper-bismuth mineralization in hedenbergitic
skarn, Tombstone Mountains, Yukon: Can. Jour. of Earth Sc., vol. 24, p.
2362-2372.

Burnham, C. W. and Jahns, R. H., 1962, A method for determining the solubility of water
in silicate melts: Amer. Jour. of Sci., v. 260, p. 721-745.

Burnham, C. W. and Davis, N. F., 1974, The role of H2Oin silicate melts:II.
Thermodynamics and phase relations in the system NaAlSi3O8-H2O to 10
kilobars, 700 to 1100 degrees Celcius: Am. Jour. of Sci., v. 274, p. 902-940.

Burt, D. M., 1982, Skarn Deposits - Historical Bibliography through 1970: Economic
Geology, v. 77, no. 4, p. 755-763.

Buseck, P. R., 1967, Contact metasomatism and ore deposition: Tem Piute, Nevada:
Econ. Geol., v. 77, p. 1818-1836.
Cahoon, B.G., 1970, Geologia de la Mina Cobriza, in Geologia de los yacimientos
minerales operados por la Cerro de Pasco Corporation: La Oroya, Peru, Cerro de
Pasco Corp. p. 41-61.

Carmichael, I. S. E., Turner, F. J., and Verhoogen, J., 1974, Igneous Petrology: McGraw-
Hill, 739 p.

Davies, H.L., Howell, W.J.S., Fardon, R.S.H., Carter, R.J., and Brumstead, E.D., 1978,
History of the Ok Tedi porphyry copper prospect, Papua-New Guinea: Econ.
Geol. v. 73, p.287-292.

Deer, W. A., Howie, R. A., and Zussman, J., 1982, An Introduction to the Rock
Forming Minerals:Longman Group Limited, Esex, England, 528 p.

Dennis, M.D., Myers, G.L., Wilkinson, W., and Wendt, C.J., 1989, Precious metal
mineralization at Mount Hamilton, White Pine County, N V: AIME preprint 89-
180, Mining Engineering, v. 41, no. 10, p. 1029-1031.

Dick, L.A., 1976, Metamorphism and metasomatism at the MacMillan Pass tungsten
deposit, Yukon and District of McKenzie, Canada: M.S. Thesis, Queens Univ.,
Kingston, 226p.

Dickinson, W. R., 1981, Plate tectonic evolution of the southern Cordillera, in Dickinson,
W. R. and Payne, W. D., eds., Relation of tectonics to ore deposits in the southern
Cordillera: Tucson, Arizona Geological Society Digest, v. 14, p. 113-135.

Dilles, J. H. and Wright, J. E., 1988, The chronology of early Mesozoic arc magmatism
in the Yerington district, Nevada, and its regional implications: Geol. Soc. of
America Bull., v. 100, p. 644-652.

Dilles, J. H. and Einaudi, M. T., 1992, Wall- rock alteration and hydrothermal flow paths
about the Ann-Mason porphyry copper deposit, Nevada-A 6km vertical
reconstruction: Econ. Geol., v. 87, no. 8, p. 1963-2001.

Dubois, R.L. and Brummet, R.W., 1968, Geology of the Eagle Mountain mine area, in
J.D. Ridge, ed., Ore Deposits of the United States, 1933-1967: New York, Am.
Inst. Mining Metall.Petroleum Engineers, p. 1592-1606.

Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits: Econ. Geol.,
75th Anniv. Vol., p. 317-391.

Einaudi, M. T., and Burt, D. M., 1982, Introduction - Terminology, Classification, and
Composition of Skarn Deposits: Economic Geology, v. 77, no. 4, p. 745-754.

Emmons, D. L. and Coyle, R. D., 1988, Echo Bay details exploration activities at its
Cove gold deposit in Nevada: Mining Engineering, v. 40, no. 8, p. 791-795.
Erwood, R.J., Kesler, S.E., and Cloke, P.L., 1979, Compositionally distinct saline
hydrothermal solutions, Naica Mine, Chihuahua, Mexico: Econ. Geol., v. 74, p.
95-108.

Ettlinger, A. D., and Ray, G. E., 1988, Gold-enriched skarn deposits of British Columbia,
in Geological fieldwork, 1987: British Columbia Ministry of Energy, Mines, and
Petroleum Resources Paper 1988-1, p. 263-279.

Ettlinger, A. D., and Meinert, L.D., 1989, Mineralogy and hydrothermal fluid
characteristics of the Nickel Plate Deposit, Hedley, British Columbia:, Geol.
Soc.Amer., Abstracts with Programs, v. 21, p. 76.

Ettlinger, A. D., and Ray, G. E., 1989, Precious metal enriched skarns in British
Columbia: An overview and geological study: British Columbia Ministry of
Energy, Mines, and Petroleum Resources Paper 1989-3, 128 p.

Ettlinger, A. D., Meinert, L. D., and Ray, G. E., 1992, Gold Skarn Mineralization and
Fluid Evolution in the Nickel Plate Deposit, British Columbia: Econ. Geol., v.87,
no. 6, p. 1541-1565.

Fergussen, H. G., 1924, Geology and ore deposits of the Manhattan district, Nevada:
U.S.G.S. Bull., 723, 163 p.

Ford, M., 1985, Geology of the Zachly Skarn Deposit, Alaska: unpub. M.S. thesis, Univ.
of Alaska, Fairbanks,

Frost, J.E., 1965, Controls of ore deposition for the Larap mineral depsoits, Camarines
Norte, Philippines: unpub Ph.D. dissertation, Stanford, Univ. 173p.

Geijer, P., and Magnussen, N. H., 1952, The iron-ores of Sweden: International
Geological Congress, 19th Algiers 1952, v. 2, p. 477-499.

Gott, G. B., and McCarthy, Jr., J. H., 1966, Distribution of gold, silver, tellurium, and
mercury in the Ely Mining District, White Pine County, Nevada: U.S. Geological
Survey, Circular 535, 15 p.

Hamilton, D. L., Burnham, C. W., and Osborn, E. F., 1964, The solubility of water and
effects of oxygen fugacity and water content on crystallization in mafic magmas:
Jour. Petrol., v. 5, p. 21-39.

Hamilton, W. B., 1987, Crustal extension in the Basin and Range province, southwestern
United States, in Coward, M. P., Dewey, J. F., and Hancock, P. L., eds.,
Continental extensional tectonics: Geological Society of London Special
Publication 28, p. 155-176.
Hernon, R.M. and Jones, W.R., 1968, Ore deposits of the Central Mining District, New
Mexico, in J.D. Ridge, ed., Ore Deposits of the United States, 1933-1967: New
York, Am. Inst. Mining Metall.Petroleum Engineers, p. 1592-1606.

Hewitt, W.P., 1968, Geology and mineralization of the main mineral zone of the Santa
Eulalia district, Chihuahua, Mexico: Soc. Mining Engineers AIME Trans., v.241,
p. 228-260.

Hickey, 1992, Geology of the Crown Jewel Gold Skarn, Washington: unpub. M.S. thesis,
Washington State Univ., Pullman.

Ishihara, S., 1977, The magnetite-series and ilmenite-series granitic rocks: Mining
Geology, v. 27, p. 293-305.

Ishihara, S., 1981, The granitoid series and mineralization: Econ. Geol., 75th Anniversary
Volume, p. 458-484.

Jacobs, D.C. and Parry, W.T., 1979, Geochemistry of biotite in the Santa Rita porphyry
copper deposit, New Mexico: Econ. Geol., v. 74, p. 860-887.

James, L.P., 1976, Zoned alteration in limestone at porphyry copper deposits, Ely,
Nevada: Econ. Geol., v. 71, p. 488-512.

John, Y.W., 1978, Sangdong mine, Korea, in Imai, H., ed., Geological studies of the
mineral deposits of Japan and east Asia: Tokyo, Univ. Tokyo Press, p. 196-200.

Jones, W.R., Hernon, R.M., and Moore, S.J., 1967, General geology of Santa Rita
Quadrangle, Grant County, New Mexico: U. S. Geol. Surv. Prof. Paper. 555, 144
p.

Jones, S. K., 1984, Geology and mineralization in the zone of contact metamorphism
associated with the Seligman Stock; White Pine County, NV: unpub. M. S. thesis,
Univ of Nevada, Reno, NV, 93 p.

Keith, S. B., 1984, Magma series and mineral deposits: MagmaChem private report, 111
p.

Keith, 1994, Pluton Vectoring for Porphyry Metal Deposits, MagmaChem private report,
112 p.

Keith, 1995, Explanation of MagmaChem Magma -Metal Series Models, MagmaChem


private report, 100 p.

Knopf, A., 1942, Ore deposition in the pyrometasomatic deposits, in Newhouse, W.H.,
ed., Ore deposits as related to structural features: Princeton, Princeton Univ.
Press, p. 63-72.
Koch, G.S., 1956, The Frisco Mine, Chihuahua, Mexico: Econ. Geol., v. 51, p. 1-40.

Korzkinskii, D. S., 1948, Petrology of the Tur'insk skarn deposit of copper: Akad. Nauk
SSSr, Inst. Geol. Nauk Trudy, vyp. 68, Ser. Rudnykh Mestorozhdenii, no. 10, 147
p.

Kuyper, B. A., 1987, Geology of the McCoy gold deposit, Lander County, Nevada:
Geological Society of Nevada Program with Abstracts, Bulk Mineable Precious
Metal Deposits of the Western United States, April 6-8, 1987, A Symposium, p.
40.

Kwak, T.A.P., 1978, The conditions of formation of the King Island scheelite contact
skarn, King Island, Tasmania, Australia: Am. Jour. Sci., v. 278, p. 969-999.

Lamey, C.A., 1946, Shasta and California iron ore deposits, Shasta County, California:
California Div. Mines Bull. 129, p. 27-38.

LaPointe, D. D., Tingley, J. V., and Jones, R. B., 1991, Mineral resources of Elko
County, NV:Nevada Bur. of Mines and Geology, Bull. 106, 236 p.

Leveille, R. A., Newberry, R. J., and Bull, K. F., 1988, An oxidation state-alkalinity
diagram for discriminating some gold- favorable plutons: An empirical and
phenomenological approach: Geological Society of America Abstracts with
Programs, v. 20, no. 7, p. A142.

Lindgren, Waldemar, 1933, Mineral deposits: New York, McGraw-Hill, 930 p.

Lindgren, W. and Ross, C.P., 1916, The iron deposits of Daiquiri, Cuba: Am. Inst.
Mining Engineers Trans., v. 53, p.40-66.

Lipman, P. W., 1992, Magmatism in the Cordilleran United States; Progress and
problems: in Burchfiel, B. C., Lipman, P. W., and Zoback, M. L., eds., The
Cordilleran Orogen: Conterminous U. S.: Boulder, Colorado, Geological Society
of America, The Geology of North America, v. G-3., p. 481-514.

Little, T. A., 1987, Stratigraphy and structure of metamorphosed upper Paleozoic rocks
near Mountain City, NV: Geol. Soc. of America Bull., v. 98, p. 1-17.

MacMillan, J. R., 1972, Late Paleozoic and Mesozoic tectonic events in west-central
Nevada: Ph. D. thesis, Northwestern Univ., Evanston Ill.

Madrid, R. J., 1987, Stratigraphy of the Roberts Mountains Allochthon in north-central


Nevada: Unpub. Ph. D. dissertation, Stanford Univ., Stanford, Calif., 346 p.
Madrid R. J. and Roberts, R. J., 1991, Origin of gold belts in north-central Nevada: in
Madrid, R. J., Roberts, R. J., and Mathewson, D., eds., Stratigraphy and structure
of the Battle Mountain gold belt and their relationship to gold deposits:
Geological Society of Nevad 1990 Great Basin Symposium, Field Trip Guide,
145 p.

McKee, E. H. and Silberman, M. L., 1970, Geochronology of Tertiary igneous rocks in


central Nevada: Geol. Soc. of America Bull., v. 81, no. 8, p. 2317-2328.

McKee, E. H., Noble, D. C., and Silberman, M. L., 1970, Middle Miocene hiatus in
volcanic activity in the Great Basin area of the western United States: Earth and
Planetary Sci. Letters, v. 8, no. 2, p. 93-96.

Meinert, L.D., 1983, Variability of skarn deposits - Guides to exploration: in Boardman,


S.J., ed., Revolution in the Earth Sciences, Kendall- Hunt Publishing Co., p.
301-316.

Meinert, L.D., 1987, Skarn zonation and fluid evolution in the Groundhog Mine, Central
Mining District, New Mexico, Econ. Geol., v. 82, p. 523-545.

Meinert, L. D., 1988a, Gold in skarn deposits-A preliminary overview, in Zachrisson, E.,
ed., Proceedings of the Symposium of the 7th Quadrennial International
Association of Geochemistry of Ore Deposits: Stuttgart, E. Schweizerbart'sche, p.
363-374.

Meinert, L.D., 1988b, Gold and silver in skarn deposits, in Goode, A. D. T. Smythe, E.
L., Birch, W. D., and Bosma, L. I., compilers, Bicentennial Gold 88, extended
abstracts, poster programmev. 2: Geological Society of Australia, p. 614-616.

Meinert, L. D., 1989, Gold skarn deposits-Geology and exploration criteria, in Keays,
Reid, Ramsey, Ross, and Groves, David, eds., The geology of gold deposits: The
perspective in 1988: New Haven, Conn., Economic Geology Publishing Co.,
Economic Geology Monograph 6, p. 537-552.

Meinert, L.D., Brooks, J., and Myers, G.L., 1991, Whole rock geochemistry and contrasts
among skarn types: in Skarn deposits of Nevada, geology, mineralogy, and
petrology of Au, Cu, W, and Zn skarns: Field Trip Guidebook, 1990 Geological
Society of Nevada, Great Basin Symposium, Reno, NV

Meinert, L. D. 1991, Skarns and skarn deposits: Geoscience Canada, v. 19, no. 4, p. 145

Moolick, R.T. and Dureck, J.J., 1966, The Morenci Distric, in Titely, S.R., and Hicks,
C.L., eds., Geology of the porphyry copper deposits, southwestern North
America: Tucson, Univ. Tucson Press. p.221-232.
Moore, W.J., 1978, Chemical characteristics of hydrothermal alteration at Bingham,
Utah: Econ. Geol., v. 73, p. 1260-1269.

Muller, S. W., Ferguson, H. G., and Roberts, R. J., 1951, Geology of the Mount Tobin
quadrangle, Nevada: U.S.G.S. Geol. Quad. Map GQ -7.

Mutschler, F. E., Griffen, M. E., Stevens, D. S., and Shannon, S. S., Jr., 1985, Precious
metal deposits related to alkaline rocks in the North American Cordillera: An
interpretive review: Transactions of the Geological Society of South Africa, v. 88,
p. 355-377.

Myers, G.L., 1984, Geology of the Fe-Cu-Au Skarns of Kasaan Peninsula, Southeast
Alaska: Geol. Soc. Amer., Abstracts with Programs, v. 16, no. 5, p.324

Myers, G.L., 1985, Gold Distribution in the Fe-Cu-Au Skarns of Kasaan Peninsula,
Southeast Alaska: Geol. Soc. Amer., Abstracts with Programs, v. 17, no. 6, p.397

Myers, G.L., 1985, Geology and Geochemistry of the Fe-Cu-Au Skarns of Kasaan
Peninsula, Southeast Alaska: unpub MS thesis, Univ. of Alaska, Fairbanks, AK
155p.

Myers, G.L. and Meinert, L.D., 1988, Zonation of the Copper Canyon-Fortitude Gold
Skarn System: Geol. Soc. Amer. Abstracts with programs, v. 20, no. 7, p.A93.

Myers, G. L., 1989, Alteration zonation of the Fortitude gold skarn deposit, Lander
County, Nevada: Mining Engineering, v. , p. 360-368.

Myers, G.L. and Meinert, L.D., 1990, Fluid inclusion and stable isotope systematics of
the Fortitude gold skarn, NV: 1990 IAGOD Symposium, Ottawa, Canada

Myers, G.L., Dennis, M. D., Wilkinson, W. H., and Wendt, C. W., 1991, Precious metal
distribution in the Mount Hamilton poly-metallic skarn: in Raines, G.L., Lisle,
R.E., Schafer, R.W., and Wilkinson, W.H. eds., Geology and Ore Deposits of the
Great Basin Symposium Proceedings, Geological Society of Nevada, p.

Myers, G. L., and Meinert, L. D., 1991, Alteration mineralogy and gold distribution in
the Fortitude gold skarn: in Raines, G.L., Lisle , R.E., Schafer, R.W., and
Wilkinson, W.H. eds., Geology and Ore Deposits of the Great Basin Symposium
Proceedings, Geological Society of Nevada, p.

Myers, G.L., 1991 Geology of the Cerro Gordo Gold Skarn, Inyo County, California;
Phelps Dodge Private Report, 35 p.

Myers, G.L., 1991 Geology of the Darwin District, Inyo County, California; Phelps
Dodge Private Report, 42 p.
Myers, G.L., 1993 Geology of the Zinc-Lead-Silver Skarn, Replacement Bodies, and
Vein Systems, Central Mexico; Phelps Dodge Private Report, 68 p.

Myers, G.L., 1994, Geology of the Copper Canyon-Fortitude Skarn System, Battle
Mountain, Nevada; Unpub. Ph.D. dissertation, Washington State University, 337
p.

Myers, G.L., 1997, Geology of the Concepcion del Oro District, Zacatecas, Mexico, Sand
River Resources Private Report, 75 p.

Newberry, R. J., 1980, The geology and chemistry of skarn formation and tungsten
deposition in the central Sierra Nevada, California; Unpub. Ph.D. thesis, Stanford
Univ., 325 p.

Newberry, R. J. and Einaudi, M. T., 1981, Tectonic and geochemical setting of


tungsten skarn deposits in the Cordillera: Symposium on Tectonics and Ore
Deposits, Tucson, Ariz., 1981 Proceedings.

Newberry, R. J., 1986, Compilation of data on Alaskan skarns: Alaska Division of


Geological and Geophysical Surveys PDF 86-21, 835p.

Newberry, .R.J., Allegro, G.L., S.E. Cutler, J.H. Hagen-Leveille, D.D.Adams, L.C.
Nicholson, T.B. Weglarz, A.S. Bakke, K.H. Clautice, G.A. Coulter, M.J. Ford, G.L.
Myers, and D.J. Szumigala, 1996, Skarn Deposits of Alaska, in, Mineral Deposits of
Alaska, Economic Geology Monograph No. 9., Edited by R.J. Goldfarb and L.D.
Miller,

Nielson, R.L., 1970, Mineralization and alteration in calcareous rocks near the Santa Rita
stock, New Mexico: New Mexico Geol. Soc. Guidebook, 21st Field Conf., p. 133-
139.

Oldow, J. S., 1984, Spatial variability in the structure of the Roberts Mountains
allochthon, western Nevada: Geological Society of America Bull., v. 95, p. 174-
185.

Oldow J. S., Bally, A. W., Ave La llemant, H. G., and Leeman, W. P., 1989, Phanerozoic
evolution of the North American Cordillera; United States and Canada, in Bally,
A. W. and Palmer, A. R., eds., The Geology of North America; An overview:
Boulder, Colorado, Geological Society of America, The Geology of North
America, v. A, p. 139-232.

Orris, G. J., Bliss, J. D., Hammarstrom, J. M., and Theodore, T. G., 1987, Descriptions
and grades of tonnages for gold-bearing skarns: U. S. Geol. Survey Open-File
Report 87-273, 50 p.
Perry, D.V., 1969, Skarn genesis at the Christmas Mine, Gila County, Arizona: Econ.
Geol., v. 64, p. 255-270.

Poole, F. G., Stewart, J. H., Palmer, A. R., Sandberg, C. A., Madrid, R. J., Ross, Jr., R. J.,
Hintze, L. F., Miller, M. M., and Wrucke, C. T., 1992, Latest Precambrian to
latest Devonian time; Development of a continental margin: in Burchfiel, B. C.,
Lipman, P. W., and Zoback, M. L., eds., The Cordilleran Orogen: Conterminous
U. S.: Boulder, Colorado, Geological Society of America, The Geology of North
America, v. G- 3., p. 9-56

Potter, 1977, Pressure corrections for fluid-inclusion homogenization temperatures based


on the volumetric properties of the system NaCl- H2 O: U.S. Geol. Surv. Jour. Res.
5, p. 603-607.

Putney, T. R., 1985, Geology , geochemistry, and alteration of the Seligman and Monte
Cristo Stocks, White Pine Mining District, White Pine County, Nevada: unpub.
M. S. thesis, Univ. of Nevada, Reno, NV, 152 p.

Ray, G. E., Dawson, G. L., and Simpson, R., 1987, The Geology and Controls of Skarn
Mineralization in the Hedley Gold Camp, Southern British Columbia: in
Geological Fieldwork 1986, B. C. Ministry of Energy, Mines, and Petroleum
Resources, Paper 1987-1, p. 65-79.

Ray, G. E., Dawson, G. L., and Simpson, R., 1988, Geology, Geochemistry, and
Metallogenic Zoning of the Hedley Gold-skarn camp: in Geological Fieldwork
1987, B. C. Ministry of Energy, Mines, and Petroleum Resources, Paper 1988-1,
p. 59-80.

Reynolds, S. J. and Keith, S. B., 1983, Geochemistry and mineral potential of


peraluminous granitoids: Fieldnotes, Ariz. Bur. of Geology and Mineral
Technology, v. 12, no. 4, p. 4-6.

Roberts, R. J., 1949, Geology of the Antler Peak quadrangle, Nevada: unpublished
U.S.G.S. open file report.

Roberts, R. J., 1951, Geology of the Antler Peak quadrangle, Humbolt and Lander
Counties, Nevada: U.S.G.S. Prof. Paper 459-A, p. A1-A93.

Roberts, R.J., 1964, Stratigraphy and structure of the Antler Peak Quadrangle, Humbolt
and Lander Counties, Nevada: U. S. Geol. Surv. Prof. Paper 459A, 93p.

Roberts R. J., Montgomery, K. M., and Lehner, R. E., 1967, Geology and mineral
resources of Eureka County, Nevada: Nevada Bur. of Mines Bull. 64.
Schwab, K. J., Keith, S. B., and Burt, D. M., 1989, Influences of magma oxidation state
on biotite and amphibole compositions and possible relations to metallogeny:
Geol. Soc. of America, Abstracts with Programs, v. 21, p. A118.

Scott, S.D. and Barnes, H.L., 1971, Sphalerite geothermometry and geobarometry: Econ.
Geol., v.66, p. 653-669.

Seedorf, E., 1991, Magmatism, Extension, and Ore Deposits of Eocene to Holocene Age
in the Great Basin-Mutual Effects and Preliminary Proposed Genetic
Relationships: in Raines, G.L., Lisle, R.E., Schafer, R.W., and Wilkinson, W.H.
eds., Geology and Ore Deposits of the Great Basin Symposium Proceedings,
Geological Society of Nevada, p. 133-178.

Seward, T.M., 1984, The transport and deposition of gold in hydrothermal systems: in
R.P. Foster ed., Gold '88: The geology, geochemistry, and genesis of gold
deposits: p. 165-181.

Sokolo v, G.A. and Grigorev, V.M., 1977, Deposits of iron, in Smirnov, V.I., ed., Ore
deposits of the USSR: London, Pittman, v. 1, p. 7-113.

Stager, H. K., 1977, Geology and mineral deposits of Lander County, Nevada, Part II
Mineral deposits: Nevada Bureau of Mines and Geology, Bulletin 88, p. 66-106.

Stewart J. H. and McKee, E. H., 1977, Geology and mineral deposits of Lander County,
Nevada, 1, Geology, Nev. Bur. of Mines Bull., 88, p. 1-59.

Stewart, J. H., 1980, Geology of Nevada-A discussion to accompany the Geologic Map
of Nevada: Nevada Bur. of Mines and Geology Spec. Pub., 4,

Taylor, B. E. and O'Neil, J. R., 1977, Stable isotope studies of metasomatic Ca-Fe-Al-Si
skarns and associated metamorphic and igneous rocks, Osgood Mountains,
Nevada: Contrib. to Minl. and Petrol., v. 63, p. 1-49.

Theodore, T.G., Silberman, M.L., and Blake, D.W., 1973, Geochemistry and K-Ar ages
of plutonic rocks in the Battle Mountain mining district, Lander County, Nevada:
U.S. Geol. Surv. Prof. Paper 798-A, 24 p.

Theodore, T. G., and Blake, D. W., 1975, Geology and geochemistry of the Copper
Canyon porphyry copper deposit and surroundig area, Lander County, Nevada: U.
S. Geol., Survey Prof. Paper 798-B, 86 p.

Theodore, T.G., and Blake, D.W., 1978, Geology and geochemistry of the West orebody
and associated skarns, Copper Canyon porphyry copper deposits, Lander County,
Nevada (with a section on electron microprobe analyses of andradite and diopside
by N.G. Banks): U.S. Geol. Surv. Prof. Paper 798-C, 85 p.
Theodore, T. G., Orris, G. J., Hammarstrom, J. M., and Bliss, J. D., 1991, Gold -Bearing
Skarns: U. S. Geol. Survey Bulletin 1930, 61 p.

Theodore, T.G., Blake, D. W., Loucks, T. A., and Johnson, C. A., 1992, Geology of the
Buckingham Stockwork Molybdenum Deposit and Surrounding Area, Lander
County, Nevada: U. S. Geol. Survey Prof. Paper 798-D, 307 p.

Thorman, C. H., Ketner, K. B., Brooks, W. E., Snee, L. W., and Zimmerman, R. A.,
1991, Late Mesozoic-Cenozoic Tectonics in Northeastern Nevada: in Raines,
G.L., Lisle, R.E., Schafer, R.W., and Wilkinson, W.H. eds., Geology and Ore
Deposits of the Great Basin Symposium Proceedings, Geological Society of
Nevada, p. 25-46.

Tooker, E.W., 1989, Gold in the Bingham district, Utah, in Shawe, D.R., Ashley, R.P.,
and Carter, L.M.H., eds., Geology and resources of gold in the United States: U.S.
Geological Survey Bulletin 1857-E, p. E1-E27.

Torrey, C. E., Karajalainen, H., Joyce, P. J., Erceg, M., and Stevens, M., 1986, Geology
and mineralization of the Red Dome (Munga) gold skarn deposit, north
Queensland, Australia, in Macdonald, A. J., ed., Proceedings of Gold '86, an
International Symposium on the Geology of Gold: Toronto, Geological
Association of Canada, 1986, p. 3-22.

Tsusue, A., 1961, Contact metamorphic iron and copper ore deposits of the Kamaishi
mining district, northeastern Japan: Jour. Fac. Sci. Univ. Tokyo, sec. 2, v. 13, pt.2,
p. 133-179.

Turner, R. J. W., 1982, The geology of the east-central Tobin Range, Nevada: M. S.
thesis, Stanford Univ., Stanford, CA., 113 p.

Westra, G., 1982, Alteration and Mineralization in the Ruth Porphyry Copper Depsoit
near Ely, Nevada: Econ. Geol., v. 77, no. 4, p. 950-970.

Wotruba, P.R., Benson, R.G., and Schmidt, K.W., 1986, Battle Mountain describes the
geology of its Fortitude gold -silver deposit at Copper Canyon: Mining Eng., v. 38,
p. 495-499.

Zharikov, V. A., 1970, Skarns: International Geology Review, v. 12, p. 541-559, 619-
647, 760-775.

You might also like