Sei sulla pagina 1di 62

EXPLORATION MODELS FOR SKARN DEPOSITS

Lawrence D. Meinert
Department of Geology
Washington State University

ABSTRACT
Most large skarn deposits are zoned in both space and time relative to
associated intrusions, and this zonation can be used to guide exploration.
Zonation occurs on scales from km to m to m, and reflects infiltrative fluid flow,
wallrock reaction, temperature variations, and fluid mixing. In most skarn
systems, there is a general zonation pattern of proximal garnet, distal
pyroxene, and vesuvianite (or a pyroxenoid such as wollastonite, bustamite, or
rhodonite) at the marble front. In addition, individual skarn minerals may
display systematic color or compositional variations within the larger zonation
pattern. Such patterns are reviewed for several well-studied examples of Cu,
W, Au, and Zn-Pb-Ag skarns. In addition, many deposits have endoskarn or
other alteration of the associated intrusion and recrystallization or other subtle
changes in the surrounding wallrocks.

Cu skarns, such as Mines Gasp in Quebec and Big Gossan in Irian Jaya,
have high garnet:pyroxene ratios, and are zoned from intrusion to garnet to
pyroxene to massive-sulfide replacement and vein deposits. Garnets in Cu
skarn are iron-rich and change from dark red-brown near the intrusive contact
to paler brown, green or yellow in distal locations. Pyroxenes in Cu skarns are
pale and diopsidic near the intrusion, and become darker and more iron- and
manganese-rich away from the intrusion. W skarns, such as Salau and
Costabonne in France and Pine Creek and Garnet Dike in California, have
intermediate garnet:pyroxene ratios, are more extensive vertically and along
strike than perpendicular to the intrusive contact, and have zonation patterns
commonly complicated by overprinting of metamorphic lithologies. In W skarns,

garnet is often subcalcic and pyroxene iron-rich, reflecting particularly reducing


wallrocks or great depth of formation. Many high-grade Au skarns, such as
Hedley in British Columbia and Fortitude in Nevada, have low garnet:pyroxene
ratios and are associated with both reduced plutons and reduced wallrocks.
Gold rich zones occur in iron-rich, pyroxene-dominant, distal skarn. Zn-Pb
skarns, such as the Yeonhwa-Ulchin district in Korea and Groundhog in New
Mexico, have low garnet:pyroxene ratios and tend to form distal to associated
intrusions. They also are zoned from proximal garnet to distal pyroxene and
pyroxenoid (bustamite-rhodonite) with significant zones of massive-sulfide
within and beyond skarn. Manganese enrichment of most mineral phases,
particularly pyroxene, is characteristic of distal zones.

Fundamental controls on skarn zonation include temperature, depth of


formation, composition and oxidation state of associated plutons and wallrocks,
and tectonic setting. W skarns tend to form at relatively great depth, 5-20 km,
with extensive high-temperature metamorphic and metasomatic mineral
assemblages. In contrast, most other skarn types are relatively shallow, <10 km
and mostly <5 km, with limited, lower temperature metamorphic aureoles.
Differences in oxidation state correlate well with different skarn zonation
patterns, particularly garnet:pyroxene ratios and compositions, and can be
used in both classification of and exploration for skarn deposits. Zonation
models, especially where quantified, can be used predictively in exploration for
both known and blind targets.

INTRODUCTION
Most large skarn deposits are associated with relatively shallow magmatic
systems emplaced within or near carbonate rocks (see reviews by Einaudi et
al.,

1981;

Meinert,

1992;

or

the

skarn

Internet

site

at:

http://www.wsu.edu:8080/~meinert/skarnHP.html). In a simple sense, these


deposits form by the transfer of heat and fluid from a cooling magma to the
surrounding rocks. This leads, inescapably, to hydrothermal alteration

(skarnification) and, more importantly, to zoned systems in both space and


time, similar to that documented for other hydrothermal ore deposits such as
porphyry Cu deposits (e.g., Gustafson and Hunt, 1975). Zonation in skarn
deposits occurs on scales from km to m to m, and reflects fluid flow, wallrock
reaction, temperature change and fluid evolution. The most spectacular
examples of skarn zonation usually occur at the skarn-marble contact where
transitions between monomineralic bands can be knife sharp (e.g., Nakano,
1978; Ochiai, 1987). Other small-scale examples occur in zoned veins (Bussell
et al., 1990) and individual mineral crystals (e.g., epidote - Nakano et al.,
1989b; garnet - Jamtveit, 1991; monticellite - Katchan, 1984; pyroxene Nakano, 1989; Nakano et al., 1989c).

Investigations of individual zoned crystals have focused upon elemental and


isotopic variations as records of fluid evolution and skarn development. For
example, in most skarn systems, two general periods of garnet growth can be
identified, with a sharp boundary between early usually poikiolitic cores and
later more homogeneous rims (Vlasova et al., 1985). Garnet cores generally
reflect the major element composition of the protolith and record progressive
reaction with infiltrating hydrothermal fluids (Jamtveit et al., 1993). In many
skarn systems, garnet cores contain high levels of Ca, Al, Ti, and Mn, and are
relatively poor in Fe and LREEs. In contrast, garnet rims tend to be depleted in
Ca, Al, Ti, Zr, Y, and Mn, and enriched in Fe, LREEs, and various ore elements
such as W, Mo, As, Cu, and Zn (Nakano et al., 1989a; Jamtveit and Hervig,
1994). These patterns can be complicated by unusual protoliths or oxidation
states; for example, early garnets can be enriched in Ni, Cr, and V from a
basalt protolith (e.g., Meinert, 1984), and late garnets in strongly reducing
environments can be enriched in aluminous sub-calcic components such as
spessartine and almandine (e.g., Newberry, 1983). Although, visually striking
and scientifically interesting, these small-scale variations are less useful in
exploration than deposit or district-scale zonation involving mineralogy and
whole rock geochemistry.

In most skarns there is a general zonation pattern of proximal garnet, distal


pyroxene and vesuvianite (or a pyroxenoid such as wollastonite, bustamite, or
rhodonite) at the contact between skarn and marble. Variations on this general
pattern reflect differences in magma chemistry, wallrock composition, depth of
formation, and oxidation state. These variations can be quite complex, in some
cases complimentary and in others, compensatory. For example, very reduced
wallrocks may lessen or completely mask the effects of a relatively oxidized
magma, producing much lower garnet:pyroxene ratios or Fe+3/Fe+2 ratios than
would normally be the case for a given skarn type. For this reason, many
researchers

use

variations

of

oxidation

state

or

some

other

geologic/geochemical variable to classify skarns within a particular deposit


class (e.g. W skarns - Newberry, 1982; Cu skarns - Einaudi et al., 1981; Zn-Pb
skarns - Einaudi et al., 1981; Megaw et al., 1988; Au skarns - Brooks, 1994;
Ray and Webster, 1997; Sn skarns - Kwak, 1987). These variations were
discussed in a general way by Meinert (1992), and the purpose of the present
paper is to illustrate common zonation patterns for some of the major skarn
classes and to explore some of the underlying causes of such variation.

DEPOSIT-SCALE ZONATION
Copper Skarns
Mines Gasp Cu-Mo
Perhaps the simplest deposit scale zonation patterns are associated with Cu
skarns, particularly those associated with shallow-level porphyritic intrusions
(Einaudi, 1982). A typical example is Mines Gasp in Quebec, that has
mineralization in porphyry, skarn, and distal mantos (Figure 1). At Mines
Gasp, the Devonian host rocks have been intruded by the quartz monzonite
Copper Mountain stock, that has a K-Ar date of 358 18 Ma (Allcock, 1982;
Gower and Walker, 1993). The dominant alteration within the Copper Mountain
stock is secondary K-feldspar and biotite with relatively minor late quartzsericite-pyrite. The margins of the Copper Mountain stock are irregular, with
numerous small dikes and sills intruded along bedding planes. A crackle

breccia up to 800 m in diameter is centered on the stock and extends for up to


200 m into the hornfelsed clastic rocks exposed on the surface. The crackle
breccia is mineralized with quartz-pyrrhotite-chalcopyrite molybdenite veinlets
that make up the porphyry copper style mineralization, constituting 275 million
tons averaging 0.4% copper and 0.03% molybdenum, that was mined from the
Copper Mountain open pit (Allcock, 1982).

Skarn occurs at depth in the more calcareous beds and extends asymmetrically
almost 2 km updip from the intrusive contact (Figure 1). Skarn mineralization is
mined underground and is higher grade than the disseminated porphyry ore,
with production/reserves of 62 million tons averaging 1.35% copper and 0.03%
molybdenum, with minor lead, zinc, silver, and gold (Williams-Jones, 1986).
Like most copper skarns (e.g. Atkinson and Einaudi, 1978), Mines Gasp is
zoned within favorable carbonate units from high garnet:pyroxene ratios near
the stock to lower garnet:pyroxene ratios towards the marble front (Figure 1).
The overall garnet:pyroxene ratio is estimated at 2:1. Contours of
garnet:pyroxene ratio define paleo fluid flow channels along more permeable
structures and stratigraphic horizons, and point directly to outlying massivesulfide zones. In addition to the zonation of garnet:pyroxene ratios, there also
is a progressive iron enrichment in pyroxene with distance from the Copper
Mountain stock (Figure 2).

At the contact of skarn and marble, there are local zones of massive-sulfide
including pyrite, pyrrhotite, chalcopyrite, sphalerite, and galena. These
massive-sulfide zones are high-grade (the E zone contains 8.9 million tons
averaging 3.16% copper and 14.2 g/t silver); relatively flat-lying; and occur both
at the skarn contact and as isolated mantos in marble (Wares and Bernard,
1993). A marble line marks the relatively sharp transition from recrystallized
white marble to fine-grained, dark gray limestone about 10-200 meters beyond
the limit of skarn (Figure 1). Beyond the marble line, the thermal aureole of the
Copper Mountain stock is evidenced in the Devonian sedimentary rocks as a

very subtle change in illite crystallinity, detectable for distances of several


kilometers beyond visible alteration (Williams-Jones, 1986). Thus, the Mines
Gasp system is zoned from a central porphyry copper zone with disseminated
low-grade mineralization, through a 2 km intermediate skarn zone with higher
grade mineralization, to distal, small mantos of very high-grade massive-sulfide
surrounded by bleached marble (Wares and Bernard, 1993).

Fluid inclusion and stable isotope investigations at Mines Gasp show a


complex mixing pattern for mineralizing fluids involving magmatic and meteoric
end-members (Shelton, 1983). The proximal porphyry mineralization was
dominated by magmatic fluids (DH2O = -42 to -61) at relatively high
temperatures. Fluid inclusions in skarn samples also are high temperature (TH
= 334-506C) and high salinity (15-56 eq. wt. % NaCl). Later mineralizing fluids
were lower temperature (~250C) and define a mixing trend (DH2O = -41 to -78)
with meteoric water. Post-mineralization fluids were low temperature (138192C), low salinity (4-15 eq. wt. % NaCl), and dominantly meteoric (DH2O = 105).

Big Gossan Cu-Au


The Big Gossan Cu-Au skarn deposit is the highest grade copper deposit in the
world-class Ertsberg district, Irian Jaya. Current reserves are 37.4 Mt, grading
2.69% Cu, 1.02 g/t Au, and 16 g/t Ag (Meinert et al., 1997). Mineralization is
associated with a series of 3-4 Ma granodioritic dikes that have intruded close
to the near-vertical faulted contact between the Shale Member of the
Cretaceous Ekmai Formation and the stratigraphically overlying Paleocene
Waripi

and

Eocene

Faumai

formations

(figures

3-4).

Most

mineralization/alteration occurs in the purer carbonate rocks of the Waripi


Formation, although biotite and calc-silicate hornfels alteration also occur in the
clastic footwall rocks adjacent to mineralization. The hornfels color varies with
mineralogy and is dark gray-brown with biotite, light gray-brown with biotiteorthoclase-plagioclase, and near the Waripi Formation contact and along

fractures, gray-green with epidote and diopsidic pyroxene. Near contacts with
massive mineralized skarn in the overlying Waripi Formation, the Ekmai biotite
hornfels is cut by veins of red-brown garnet with envelopes of pyroxenefeldspar hydrothermal anhydrite. This dark red-brown color of garnet is
restricted to proximal skarn zones near the Ekmai-Waripi contact.

The calcic magnesian skarn assemblage in the Waripi Formation that hosts
the bulk of the Big Gossan orebody is characterized by relatively coarsegrained prograde garnet and pyroxene, with an average ratio of about 1:2.
Individual

garnet

and

pyroxene

crystals

typically

are

0.1-1

cm

in

length/diameter and range up to 10 cm. Garnet exhibits a wide range in color


from dark red-brown near fluid conduits to intermediate browns and greens in
more distal locations. Pyroxene ranges in color from almost white to dark
green. This change in color corresponds directly with iron content and is zoned
in both space and time. The pale, proximal, and early pyroxene is nearly pure
diopside, and the very dark green, distal, and late pyroxene ranges up to 75%
hedenbergite. Complicating this pyroxene compositional trend is the effect of
hostrock composition; skarn formed from pure dolomite or anhydrite has a more
diopsidic composition than would normally be the case for a particular
spatial/paragenetic position.

Perhaps, the strongest influence on pyroxene composition is elevation (relative


position) within the hydrothermal system. As shown in Table 1, pyroxene from
all parts of the skarn system becomes more iron-rich, and to a lesser extent
more manganese-rich, higher in the system. The average pyroxene
composition for the highest third of the skarn system is Di58Hd38Jo4, whereas
the average for the lowest third is Di86Hd13Jo1. In general, pyroxene becomes
more iron- and manganese-rich towards the western and eastern margins of
the system, suggesting a hydrothermal center in the middle of the system,
coincident with the largest igneous dike mass. Average pyroxene compositions

for the western, middle, and eastern thirds of the skarn system are
Di67Hd30Jo3.5, Di81Hd17Jo2, and Di71Hd26 Jo3.

Unlike pyroxene, garnet is zoned on the scale of individual crystals as well as


on the scale of the deposit. Convoluted, concentric, planar and sector zoning
are present in different garnet crystals. In general under crossed nicols, more
isotropic zones are more andraditic, whereas more birefringent zones are more
grossularitic. There is no consistent compositional zonation from core to rim.
Most garnet rims are more andraditic, but some rims are up to 40% more
grossularitic than the garnet core. Different compositional bands within an
individual crystal can display as wide a compositional range as is seen at the
deposit scale. The average composition of all analyzed garnets is
Ad84.7Gr13.5Sp1.5Py0.3. Garnet higher in the system and on the western and
eastern margins is very slightly enriched in iron relative to deeper more central
locations, but this zonation is not nearly as pronounced as that previously
described for pyroxene.

Fluids associated with prograde skarn are high-temperature, low-CO2 (<0.05


mole %) NaCl-KCl brines. Homogenization temperatures for fluid inclusions in
pyroxene range from 320-485C and average 410C. Most fluid inclusions in
pyroxene contain multiple daughter minerals including halite, sylvite,
chalcopyrite, hematite, and anhydrite. Total salinity ranges from 38-65 wt. %
NaCl+KCl, and mean salinities are 22 wt. % KCl, 35 wt. % NaCl, and 57 wt. %
NaCl+KCl. Clear evidence for boiling in some fluid inclusions indicates a
pressure of 20 MPa that corresponds to a depth of 2 km under hydrostatic
conditions (Meinert et al., 1997).

Chalcopyrite is the dominant ore mineral with bornite comprising less than 1
volume % of total Cu sulfides. Other metallic minerals include electrum
(Au>Ag), pyrite, sphalerite, galena, and pyrrhotite. Pyrite is the most abundant
of these minerals, locally reaching 20 volume percent. Sphalerite, galena, and

pyrrhotite combined are <1 volume percent, except at the margins of the skarn
where pyrrhotite forms a subeconomic "cap" on the chalcopyrite mineralization.
This sulfide zone is analogous to massive-sulfide bodies in other skarns, as
previously discussed for Mines Gasp and the mantos and chimneys of
some zinc skarns (e.g., Megaw et al., 1988). At Big Gossan, this "cap" occurs
at high elevations (Figure 4) for almost the entire length of the deposit.

Beyond skarn, but within the marble aureole surrounding the Big Gossan
deposit, there are numerous planar to wavy veinlets, usually less than 1 mm
thick, that appear to represent fluid conduits. In places these look like stylolites
except their orientation is systematic (usually perpendicular) relative to the
skarn front. The dark centerline of these veins is marked by a concentration of
carbon, sulfides (pyrite, sphalerite, and galena), chlorite, serpentine, and/or
clay. Marble occurs for 10s to 100s of meters beyond the Big Gossan skarn in
the Waripi Formation hanging wall. This marbleization decreases in grain size
with distance from the skarn and/or intrusive breccia.

Mineralization is zoned within the Big Gossan system (Table 2). Copper, Au,
Ag, Pb, Zn, As, Mo, and Co assay data from 192 drill holes totaling 13,215
assay intervals were composited every 3 m down the hole, and categorized by
rock type, elevation and mine location along strike. Cu, Au, Ag, Pb, Zn, As, and
Co grades in skarn all increase towards the top of the system whereas Mo
grades increase with depth. Similarly, Cu, Au, Ag, Pb, Zn, As, and Co grades
increase (for a given elevation) towards the western and, in most cases,
towards the eastern margin of the system. Only Mo is inverse to this trend,
defining a central core zone that is interpreted to represent the main locus of
fluid flow overlying the source pluton. Pb and Zn show the strongest enrichment
towards the distal margins of the system, and even higher concentrations of
these elements (up to several percent) occur as veins and disseminations in
marble beyond the limit of skarn. However, nowhere do Pb and Zn reach
economic proportions within the Big Gossan system.

Tungsten Skarns
French Pyrenees W-Mo
Salau and Costabonne are the two largest of many W-Mo-bearing skarns in the
French Pyrenees. Salau has production/reserves of 1.2 million tons averaging
1.6% WO3 and Costabonne is smaller with average grades of 0.3-0.4% WO3
(Fonteilles et al., 1989). These skarns are associated with 270-290 Ma
granodioritic stocks and batholiths that have intruded Paleozoic carbonate and
clastic sedimentary rocks (Figure 5). The igneous rocks have minor
plagioclase-amphibole endoskarn on a scale of meters near some contacts with
mineralized skarn, but otherwise are relatively unaltered (Soler, 1977;
Raimbault and Kaelin, 1987). Particularly lacking are the mineralized stockwork
quartz veinlets with K-feldspar and biotite envelopes, that occur in intrusions
associated with many Cu skarns. This is consistent with the estimated pressure
of 2 kb, corresponding to a depth of at least 8 km, depending on the range of
lithostatic to hydrostatic load (Van Marcke de Lummen, 1983).

At Salau, the host rocks consist of Silurian black carbonaceous shales overlain
by Devonian platform carbonates of the Salau Formation (Derr and Krylatov,
1976). These strata have been folded during the Hercynian orogeny and are
now subparallel to the generally steep intrusive contact (Figure 6). Skarn forms
a relatively thin band (10-50 m) at the intrusive contact and its mineralogy
depends on the composition of the protolith (Fonteilles et al., 1989). In marble,
skarn is surrounded by a narrow bleached zone that is marked by an increase
in grain size, destruction of graphite, and the formation of wollastonite along
the skarn-marble contact (Figure 7). The most distal skarn is pyroxenedominant and the pyroxene is almost pure hedenbergite (Figure 8). Proximal
skarn is garnet-dominant and the garnet is low in ferric iron (Figure 8). The
abundance of pyroxene relative to garnet (Fonteilles et al., 1989, estimate an
overall garnet:pyroxene ratio of 1:10 to 1:20) and the low Fe+3/Fe+2 ratios of
both garnet and pyroxene are indicative of relatively reduced conditions,
consistent with the presence of graphite in the protolith. At Salau, as with most

10

W skarns, there is a late subcalcic garnet that is not associated with scheelite
mineralization and which cuts the earlier garnet and pyroxene skarn (Figure 8).
Thus, there is both a spatial and temporal zonation to the skarn system. In
some locations, skarn is in direct contact with granodiorite and, in other
locations, there is a zone of endoskarn developed in the granodiorite.
Endoskarn is characterized by epidote-amphibole-titanite apatite.

Less calcareous rocks at Salau are converted to biotite hornfels within 900 m of
the La Fourque stock and to calc-silicate hornfels within 50-100 m of the
intrusive contact (Fonteilles et al., 1989). The garnet and pyroxene in such
hornfels are distinctly less iron-rich than skarn formed from marble and, in
many cases, approximate the composition of the protolith (Zahm, 1987b). In
areas of intense hydrothermal alteration, this early calc-silicate hornfels is
overprinted by garnet and pyroxene (with scheelite) that are closer in
composition to skarn formed directly from marble (Fonteilles et al., 1989).
Compositions of all analyzed garnet and pyroxene at Salau form a continuum
between hornfels and pure marble end-members (Figure 8). Thus, the
composition of the minerals in a particular skarn sample directly reflects the
composition of the host rock and the intensity of alteration, with skarn formed
from marble being the end-member composition produced by this particular
metasomatic system.

Low grade scheelite mineralization occurs with the early garnet-pyroxene skarn
with typical grades of 0.1-0.5% WO3 (Fonteilles et al., 1989). Higher grades are
associated with retrograde alteration of garnet-pyroxene to epidote-amphibolequartzcalcite. The highest grades (up to 40% WO3) are coincident with both
retrograde alteration and precipitation of sulfide minerals including abundant
pyrrhotite and minor chalcopyrite, sphalerite, arsenopyrite, and molybdenite. In
general, the highest grades are associated with skarn formed from marble
rather than from clastic/argillaceous rocks and, in most cases, tungsten grades
increase towards the marble contact.

11

Costabonne is a very similar skarn system to Salau except that most of the
carbonate host rock is dolomitic rather than calcic limestone (Van Marcke de
Lummen and Verkaeren, 1986; Guy, 1988). The skarn zonation pattern is
similar to that at Salau except that there is a border zone with magnesian skarn
minerals (mostly forsterite, talc, serpentine, and brucite) instead of wollastonite
at the marble front (Guy et al., 1988). In addition, the outer pyroxene skarn is
zoned with more diopsidic pyroxene towards the marble front (Figure 9). This is
another example of wallrock compositional control. At Salau, the graphitic
carbonate rocks caused very iron-rich pyroxene to form in the most distal skarn
zone whereas at Costabonne, the dolomitic carbonate rocks stabilized
diopsidic pyroxene instead.

Skarns at Salau and Costabonne are estimated to form at temperatures of 470560C, based upon mineral equilibria, and 455-570C, based upon pressurecorrected fluid inclusion measurements (Fonteilles et al., 1989). Salinities in
skarn silicates range from 23-30 wt. % NaCl eq. with halite daughter minerals
present in many inclusions. Stable isotopes indicate a dominantly magmatic
fluid during initial skarn formation and a mixture of magmatic and other,
possibly meteoric, fluids during retrograde alteration (Guy et al., 1988; Van
Marcke de Lummen, 1988).

Sierra Nevada W-Mo


The Sierra Nevada region of California contains hundreds of W-bearing skarns
that range from isolated occurrences to substantial mines (Newberry, 1982).
The largest, the Pine Creek district, produced more than 6.6 million tons of ore
prior to 1966, and a possibly similar amount since, at an average grade of 0.5%
WO3, 0.13% Mo, and 0.15% Cu (Gray et al., 1968). The Pine Creek skarn
occurs in a septum of Paleozoic sedimentary rocks surrounded by parts of the
multiphase Sierra Nevada batholith. Skarn is associated with the 94 Ma
Morgan Creek quartz monzonite, an equigranular pluton with pegmatitic,
alaskitic, and quartz diorite phases (Newberry, 1982). The quartz monzonite

12

contains anastomosing zones of granoblastic quartz and intergranular


myrmekite that Newberry (1982) interprets as due to the gradual, nonexplosive
release of magmatic fluids, that in a shallower porphyry Cu environment might
result in stockwork quartz veining, brecciation, and mineralization. This is
consistent with an estimated maximum pressure of 2 kb at the time of intrusion
(Brown et al., 1985).

The

host

rocks

at

Pine

Creek

include

30-60

of

fine-grained

clastic/argillaceous rocks overlain by 300 m of calcic carbonate and 1000 m of


medium-grained clastic/argillaceous rocks (Newberry, 1982). These rocks have
been metamorphosed to biotite hornfels/schist, diopsidic marble, and
micaceous quartzite, respectively. These strata were deformed during intrusion
and are subparallel to the intrusive contact (Figure 10). Skarn forms a narrow
(1-15 m), relatively continuous zone along the intrusive contact. The deposit is
large, because this thin skarn zone is continuous for approximately 1.5 km
vertically and 2+ km laterally. The thickest and highest grade skarn formed
from marble, although biotite hornfels also is overprinted by garnet-pyroxene
skarn. Skarn is systematically zoned relative to the intrusive and marble
contacts (Figure 11). Marble close (cm) to skarn is bleached and recrystallized.
The outermost skarn zone consists of abundant vesuvianite and lesser
pyroxene, wollastonite, and scheelite. The main skarn zone consists of redbrown garnet and green pyroxene in about a 2:1 ratio; this zone constitutes
approximately 80% of the total skarn volume and contains ore grade scheelite.
Skarn at the intrusive contact consists of dark orange-red garnet, green
pyroxene, and quartz, but this inner skarn zone is barren of scheelite.

The garnet and pyroxene in these different zones are visually and
compositionally distinct (Figure 12). Main/outer zone pyroxene is more iron-rich
(up to Hd80) and less manganese-rich than inner zone pyroxene. Similarly,
inner zone garnet is iron-rich and manganese-poor relative to main zone
garnet. Importantly, the inner zone garnets are subcalcic; the iron is ferrous

13

and is mostly present as almandine component, whereas, the iron in main zone
garnets is ferric and is mostly present as andradite component (Newberry,
1983). A more detailed study of garnet and pyroxene zonation by Newberry
(1991) suggests a compositional plateau for the main skarn zone. Profiles from
the intrusive contact to marble do not suggest a statistically significant
compositional gradient for either garnet or pyroxene within the main skarn
zone, even though the relatively thin inner and outer skarn zones are quite
distinct. Furthermore, the subcalcic garnets in the inner skarn zone apparently
show a systematic decrease in subcalcic (almandine + spessartine) component
towards the marble front.

A similar mineralogy and zonation pattern exists at the Garnet Dike mine in the
western Sierra Nevada (Newberry, 1980). The main difference between Pine
Creek and Garnet Dike skarns is that the latter is estimated to have formed at a
pressure of 3-4 kb, and thus is a substantially deeper exposure. This is
reflected in a larger subcalcic garnet zone, more almandine-rich garnets, and
more hedenbergitic pyroxene. Thus, for most known W-Mo skarns, there is a
similar

zonation

pattern

of

proximal/late

subcalcic

garnet-quartz,

an

intermediate (and large) zone of garnet-pyroxene skarn, and a thin skarn front
of vesuvianite and/or wollastonite surrounded by <1 m of bleached marble. In
all the W-Mo skarns, the thickness of skarn perpendicular to the intrusive
contact is small (10s of meters) relative to the vertical and lateral dimensions
that are measured in 100s to 1000s of meters. Most of the associated plutons
are equigranular, lack stockwork veining/mineralization, and formed at depths
corresponding to pressures of >2 kb (200 MPa).

Gold Skarns
The highest grade (5-15 g/t Au) gold skarn deposits (e.g., Hedley district,
British Columbia, Ray and Dawson, 1994; Fortitude, Nevada, Myers and
Meinert, 1991; Myers, 1994) are relatively reduced, are mined solely for their
gold content, and lack economic concentrations of other metals. Other gold

14

skarns (e.g. McCoy, Nevada, Brooks et al., 1991; Brooks, 1994) are more
oxidized, have lower gold grades (1-5 g/t Au), and contain subeconomic
amounts of other metals such as Cu, Pb, and Zn. Several other skarn types,
particularly Cu skarns, contain enough gold (0.01 to >1 g/t Au) to be produced
as a byproduct of base metal mining (Meinert, 1989). Most high-grade gold
skarns are associated with reduced (ilmenite-bearing, Fe2O3/( Fe2O3+FeO)
<0.75) diorite-granodiorite plutons and dike/sill complexes. Such skarns are
dominated by iron-rich pyroxene (typically >Hd50), but proximal zones can
contain abundant intermediate grandite garnet. Other common minerals include
K-feldspar, scapolite, vesuvianite, apatite, and amphibole. Distal/early zones
contain biotite+K-feldspar hornfels, that can extend for 100s of meters beyond
massive skarn. Due to the clastic-rich, carbonaceous nature of the sedimentary
rocks in these deposits, most skarn is relatively fine-grained.

Hedley District Au
The Nickel Plate mine in the Hedley district, British Columbia, is the largest and
highest grade gold skarn in Canada. Discontinuous production from 1904 until
the mine closed in 1995 was 13.4 million tons averaging 5.3 g/t Au, 1.3 g/t Ag,
and 0.02% Cu (Ray et al., 1996). Of this, more than 3 million tons of ore was
mined underground at an even higher grade, averaging 14 g/t Au. Skarn
formed in dominantly clastic rocks of the upper Triassic Nicola Group, that is
part of the allochthonous Quesnel Terrane of the Intermontane Belt. Skarn is
spatially and genetically associated with the dioritic Hedley intrusions, that
comprise the Toronto Stock and a series of dikes and sills. Attempts to date,
these intrusions have been inconclusive with a suggested age range of 194219 Ma (Ray and Dawson, 1994). The Toronto Stock is a very reduced
ilmenite-bearing intrusion with an average Fe2O3/( Fe2O3+FeO) value of 0.15,
the lowest of any gold skarn (Ray et al., 1995) and the lowest of any major
skarn class (Meinert, 1995).

15

As first recognized by Billingsley and Hume (1941), skarn is zoned in both


space and time relative to the Toronto Stock and associated dikes and sills
(Figure 13). The earliest and most distal alteration is a fine-grained biotite
hornfels that affects both clastic rocks and some of the early sills (Ray et al.,
1988). With time and proximity to massive skarn, biotite occurs with K-feldspar
and pyroxene, and is slightly coarser grained (Ettlinger, 1990). This forms an
aureole around the massive garnet-pyroxene skarn that is zoned from garnet >
pyroxene near the Toronto Stock to pyroxene-dominant (garnet:pyroxene <0.1)
skarn in distal ore zones (Ettlinger et al. 1992). Garnet is intermediate grandite
in composition, whereas pyroxene is relatively iron-rich (Figure 14). The most
iron-rich garnet (Ad73-82) occurs in distal ore zones (Ettlinger et al., 1992),
whereas pyroxene composition changes systematically away from the Toronto
Stock and the larger dikes, becoming more iron-rich and slightly more
manganese-rich (Figure 15). The sulfide minerals associated with garnet and
pyroxene skarn are dominantly arsenopyrite and pyrrhotite. Other sulfides, in
decreasing order of abundance, are chalcopyrite, pyrite, sphalerite, hedleyite
(Bi2+XTe1-X), native bismuth, gold, galena, maldonite (Au2Bi), and loellingite
(Ettlinger, 1990). This latter group of minerals is mostly associated with lower
temperature alteration including amphibole, ferroan wollastonite, scapolite, and
prehnite. The scapolite and some amphiboles are unusually chlorine-rich and
this feature has been suggested as an exploration guide to gold-rich systems
(Pan et al., 1994).

Garnet-pyroxene skarn at the Nickel Plate mine is estimated to have formed at


a depth of 5 km and at an average temperature of 460-480C, although fluid
inclusions in some garnet and pyroxene samples homogenized at temperatures
above 600C (Ettlinger, 1990). Salinities of garnet and pyroxene fluid
inclusions average 18.3 and 9.7 wt. percent NaCl equivalent, respectively, with
sparse halite daughter minerals in garnet, pyroxene, and quartz yielding a
maximum of 37.9 wt. percent NaCl equivalent.

16

Fortitude Deposit, Battle Mountain District, Au


The Fortitude deposit occurs in the Battle Mountain District of north-central
Nevada. The Battle Mountain District contains several different skarn types
ranging from a typical copper skarn with low gold grades, called the West Ore
Body, to the gold-rich, copper-poor skarns such as the Upper and Lower
Fortitude ore bodies (Figure 16). The West ore body is adjacent to the 38-38.5
Ma (Theodore et al., 1978, K-Ar on primary biotite) Copper Canyon
granodiorite porphyry. The skarns nearest the intrusive body are dominated by
garnet with minor pyroxene and are typically high in copper with low
concentrations of gold. The Fortitude deposit, like most high-grade Au skarns,
has an unusual reduced skarn mineralogy and trace-element signature that
distinguishes it from most other skarn types. The reduced skarn mineralogy
reflects the very reduced nature of the associated Copper Canyon granodiorite
(Fe2O3/(Fe2O3+FeO) <0.5) that is quite distinct from typical oxidized porphyry
copper deposit plutons. The more distal skarns contain more pyroxene than
garnet and contain the highest concentrations of gold in the district. An
extensive biotite+diopsidic pyroxene+K-feldspar alteration halo surrounds the
main skarn zone and extends up to 3 km from the Copper Canyon stock
(Theodore and Blake, 1975).

Like much of Nevada, detailed geologic relations in the Battle Mountain District
are complicated by numerous thrust faults (De Witt, Golconda, and Roberts
Mountain) that have juxtaposed a wide variety of rock types, all of which have
been affected by hydrothermal alteration (Blake et al., 1984). Host rocks range
from lower Cambrian through Permian, and consist of sandstone, arkose,
shale, chert, argillite, greenstone, limestone, and quartzite (Blake et al., 1984).
Most skarn occurs in the basal mid-Pennsylvanian Battle Formation, the
Pennsylvanian to Permian Antler Peak Limestone, and the Permian Edna
Mountain Formation. The main mineralized unit is the Antler Peak Limestone,
that consists of well bedded dark gray limestone and minor chert nodules
(Theodore and Blake, 1978).

17

Skarn mineralogy has been investigated for most of the deposits in the Battle
Mountain district. For example, the West Ore Body occurs mostly in the Antler
Peak Limestone and is a typical copper skarn with a prograde skarn
assemblage of grandite garnet (Ad39-99) + diopsidic pyroxene (Hd20-50) and
a retrograde assemblage of actinolite + epidote + K-feldspar (Theodore and
Blake, 1978). Pyrrhotite, pyrite, chalcopyrite, and marcasite are the main
sulfide minerals.

The Fortitude deposit also occurs in the Antler Peak Sequence, primarily in the
Antler Peak Limestone and the Battle Formation. The stratigraphic continuity
between the West and Fortitude skarns is illustrated in the cross section of
Figure 17. The Fortitude ore body contains a prograde skarn assemblage of
andraditic garnet + hedenbergitic pyroxene (Hd20-95, Jo<10 ), with only minor
retrograde alteration to epidote + actinolite + chlorite (Myers, 1994). As
illustrated in Figures 16 and 17, pyroxene shows a general increase in iron
content moving toward the marble front (Hd30 proximal to the Copper Canyon
stock and Hd>90 at the distal skarn fringe). This trend is mimicked by a Mn
enrichment moving towards the marble front, with the pyroxene nearest the
intrusion rarely exceeding 3% Jo (except for crosscutting veinlets and crystal
rims), whereas the pyroxene near the marble front is generally between 2-8%
Jo (Myers, 1994).

As illustrated in Figure 17, the distribution of most metals parallels the skarn
zonation in the Antler Peak Limestone. Copper is highest in garnet-rich skarn
near the intrusive contact, whereas gold is concentrated in pyroxene-dominant
skarn, particularly where the pyroxene is iron-rich (>Hd50). Silver has elevated
concentrations proximal to the stock and at the distal skarn front, beyond the
main Fortitude gold zone. The skarn system contains several sulfide species
including chalcopyrite, pyrite, pyrrhotite, arsenopyrite, marcasite, sphalerite,
and galena, that occur roughly in the order listed from intrusion to marble.
Arsenopyrite is locally massive and native Bi is commonly visible in hand

18

specimens. Native gold occurs at grain boundaries between skarn and sulfide
minerals indicating a possible reaction relationship (Wotruba et al., 1988). In
general, gold is associated with native bismuth, hedleyite, pearcite, and
stannite. Trace elements are also zoned within the system with anomalous Co,
Mo, Cr, and Ni in proximal zones and anomalous As, Bi, Cd, Mn, Pb, Zn, Sb,
and Hg in distal zones (Figure 17).

Fluid inclusion work shows that the skarn formed at relatively high
temperatures (300->550C) that parallel fluid inclusion homogenization
temperatures measured in the adjacent Virgin dike apophysis of the Copper
Canyon granodiorite (Myers, 1994). The distribution of measured fluid inclusion
temperatures parallels the skarn zonation of Figure 17. Garnet closest to the
main stock (drill hole #500) ranges from 360-590C. More distal garnet and
pyroxene (drill holes 2723 and 1997) range from 380-440C and 320-430C,
respectively and the most distal (and iron-rich) pyroxene (drill hole 1994)
ranges from 350-400C (Myers, 1994). In addition, high salinity conditions have
been documented, with multiple daughter minerals in fluid inclusions identified
by SEM and STEM analysis. Limited fluid inclusion measurements indicate
pyroxene skarn had salinities of 25-44 wt. percent NaCl equivalent. Based
upon limited evidence for boiling, Myers (1994) estimated a formation pressure
of 0.4 kb for the Fortitude system, in close agreement with the stratigraphic
estimate of 1.5 km and a pressure of 375 bars by Theodore and Blake (1975).

Similar zonation occurs in 18O and 13C values that indicate progressive
reaction of a magmatic fluid with isotopically heavy carbonate wallrocks as
summarized by Zimmerman et al. (1992) and Myers (1994). Skarn garnets are
progressively enriched in 18O outward from the Copper Canyon stock with
garnet 18O values of 6.9 per mil in the proximal skarn and values as high as
8.2 per mil in distal skarn. Pyroxene (18O 8.6 to 10.3 per mil), amphibole
(18O 8.6 to 9.2 per mil), and quartz (18O 11.4 to 13.2 per mil) are less
systematic, but in each case the highest 18O values are most distal to the

19

granodiorite stock. Skarn formation can be modeled as resulting from the


progressive reaction of magmatic fluids with isotopically heavier carbonate
wallrocks (18O = 24.0 per mil). The variation in 13C values in calcite also
can be explained by progressive reaction of magmatic fluids with carbonate
wallrocks. 18O and 13C values decrease from unaltered limestone (18O
24.0 per mil, 13C 2.4 per mil) to blocks of residual limestone in skarn (18O
15.4 to 19.3, 13C -4.5 to 1.7 per mil) to calcite intergrown with skarn minerals
(18O 11.8 to 13.1 per mil, 13C -10.3 to -1.7 per mil). The absence of mineral
phases with 18O less than magmatic values suggests that meteoric fluids
(18O <-10) probably did not play a significant role in the formation of this
deposit (Zimmerman et al., 1992). This is consistent with the relatively small
amount of retrograde alteration observed in the Fortitude skarn (Myers, 1994).

Zinc-Lead Skarns
Most zinc skarns occur in continental settings associated with either subduction
or rifting. They are mined for ores of zinc, lead, silver, and sometimes copper,
although zinc is usually dominant. They are also high-grade (10-20% Zn+ Pb,
30-300 g/t Ag) and sulfide-rich. Many are transitional to massive-sulfide veins,
mantos, and chimneys, that lack significant calc-silicate alteration (e.g., Megaw
et al., 1988). Besides their Zn-Pb-Ag metal content, zinc skarns can be
distinguished from other skarn types by their distinctive manganese- and ironrich mineralogy, by their occurrence along structural and lithologic contacts,
and by the absence of significant metamorphic aureoles centered on the skarn.
Almost all skarn minerals in these deposits can be enriched in manganese
including garnet, pyroxene, olivine, ilvaite, pyroxenoid, amphibole, chlorite, and
serpentine. In some deposits, the pyroxene:garnet ratio and the manganese
content of pyroxene increase systematically along the fluid flow path.
Yeonhwa-Ulchin District Zn-Pb-Ag
Numerous Zn-Pb-Ag skarns occur in the Kyongsang tectonic belt of eastern
Korea (Yun, 1979). The three largest of these occur in the Yeonhwa-Ulchin
district and provide an interesting contrast in terms of metal ratios, skarn

20

mineralogy, and associated igneous rocks, as summarized by Yun and Einaudi


(1982). Collectively, the Zn skarns of the Yeonhwa-Ulchin district have
produced more than 13.5 million tons of ore. The host rocks for these skarns
are interbedded Cambrian shallow marine shales (Myobong Slate) and platform
carbonates (Pungchon Limestone) with the bulk of ore grade mineralization in
the Pungchon Limestone (Yun, 1978). These rocks were folded and intruded by
granodiorite to quartz monzonite porphyries of Late Cretaceous to early
Tertiary age along the Yeonhwa-Ulchin axis. The intrusive rocks appear to
young from northwest to southeast and limited K-Ar dating suggests that
Yeonhwa II is about 72.62.2 Ma and Ulchin is 49.32.0 Ma (Yun and
Silberman, 1979).

The Yeonhwa I mine is the largest producer of zinc and lead in Korea with 7.6
million tons averaging 6% Zn and 2% Pb (Yun and Einaudi, 1982). Ore bodies
occur as tabular lenses along the contact of the Pungchon Limestone and
Myobong Slate as well as vertical pipes at intersections of north-northeast and
north-northwest faults that cut both the lower limestone and upper dolomitic
limestone facies of the Pungchon Limestone (Figure 18a). Some pipes also
penetrate the overlying Hwajeol calcareous shales and form anastomosing,
sulfide-rich veins close to the surface. The only igneous rocks in the mine area
are minor quartz porphyry dikes. Although these are thought to connect at
depth to the mineralizing source, the skarn and mineralization are not zoned
relative to dike contacts. Instead, skarn is zoned relative to stratigraphic
contacts and fractures.

At the deepest mine levels, skarn occurs as a tabular body zoned relative to
the Pungchon/Myobong contact (Figure 18a). This skarn displays a deep,
central (proximal to the fluid conduit) garnet>>bustamite zone with pyrrhotite,
sphalerite, galena, and chalcopyrite and a shallow, distal pyroxene-sulfide zone
with pyrite, sphalerite, galena, and rhodocrosite (Yun, 1979). Overall, pyroxene
is much more abundant than garnet and ranges up to 23 mole % johannsenite.

21

At shallower levels, garnet is absent and pyroxene occurs with massive-sulfide


lenses, pipes, and veins. The overall ratio of sulfide to silicate skarn is
estimated at 10:1. Mineralization is zoned within the Yeonhwa I mine in terms
of grade and Pb/Zn ratio, both increasing higher in the system from 5 to 20 %
combined Pb+Zn and from 0.06 to 1.3, respectively (Yun and Einaudi, 1982).

The Yeonhwa II mine is the second largest producer of zinc and lead in Korea
with 4.5 million tons averaging 4.1% Zn, 0.2% Pb, and 0.1% Cu (Yun and
Einaudi, 1982). Although no large stocks are present at Yeonhwa II, there are
several sizable sills (up to 300 m thick) and, unlike Yeonhwa I, skarn is present
along igneous contacts. Tabular contact and fracture-controlled skarns have a
central garnet>>rhodonite-quartz zone with envelopes of 80-90% pyroxene
skarn. The garnet:pyroxene ratio decreases upward in the mine although,
unlike Yeonhwa I, garnet-free zones are rare (Yun, 1979). Overall, pyroxene is
estimated to be slightly more abundant than garnet and the maximum iron and
manganese content is Hd74 and Jo21, respectively. Garnet is both iron and
manganese rich (up to Sp24) and generally is later than pyroxene, a feature that
Yun and Einaudi (1982) interpreted to represent contemporaneous zonal
growth with garnet overprinting pyroxene as the system expanded. Like
Yeonhwa I, skarn contains zones of massive-sulfide (sphalerite>pyrrhotitegalena-chalcopyrite), but there does not appear to be a systematic vertical
increase in sulfide content or metal ratio. This may be related to the proximity
of skarn to igneous contacts and the lack of vertically oriented pipes as at
Yeonhwa I.

Compared to Yeonhwa I and II, Ulchin is smaller (1.4 million tons), has slightly
higher grades of zinc (5.9%), lead (2.3%), and copper (0.3%) (Yun and
Einaudi, 1982). Ulchin skarn occurs in roof pendants surrounded by Ulchin
granite and appears to be related to a series of rhyodacite dikes (Figure 18b).
The large granite body appears to be much older than the rhyodacite dikes and
skarn but has had an important influence in metamorphosing the sedimentary

22

roof pendants. Thus, there are barren metamorphic calc-silicate rocks that are
similar to those previously described for W skarns; such features are absent at
most Zn skarns and specifically absent at Yeonhwa I and II.

Skarn occurs at igneous contacts and along east-northeast-striking faults that


have been intruded by rhyodacite dikes (Yun and Einaudi, 1982). Garnet is
concentrated close to igneous contacts and epidote endoskarn occurs within
the intrusions. Outer skarn zones contain more pyroxene, but overall, garnet
and pyroxene are present in subequal proportions. Both garnet and pyroxene
are

iron-rich,

up

to

Ad96

and

Hd91,

respectively.

Ore

(sphalerite~pyrrhotite>>galena-chalcopyrite) is localized in pyroxene zones


and commonly forms vertical pipe-like bodies. Pyroxene becomes more
manganese-rich towards marble (Jo5-20) and rhodonite occurs at the marble
front, completing a general zonation pattern of dike-epidote-garnet-pyroxenerhodonite-marble. Like Yeonhwa I, mineralization is zoned within the Ulchin
mine in terms of grade and Pb/Zn ratio, both increasing higher in the system
from 5 to 20 % combined Pb+Zn and from 0.01 to 1.1, respectively (Yun and
Einaudi, 1982).

Groundhog Zn-Pb-Ag
The Groundhog mine, New Mexico, is the largest Zn skarn in the United States,
with total production of approximately 3 million tons and average skarn grades
of 13.5 % Zn, 2.8% Pb, 0.6% Cu, and 46 g/t Ag (Meinert, 1987). The deposit is
systematically zoned relative to a nearly vertical dike swarm of Tertiary
granodiorite porphyry dikes that range in thickness from 1-50 meters, extend
for more than 3 km in a northeast-southwest direction, and intrude almost
horizontal strata of Mississippian Lake Valley Limestone. Based upon metal
ratios, skarn mineralogy, and fluid inclusion temperatures, the hot, proximal
part of the system lies to the northeast, and the cooler, distal part of the
system, to the southwest. Zinc and lead grades both increase and copper
decreases towards the distal southwest part of the system (Table 3). Silver

23

does not vary as dramatically and has a maxima in the central part of the
system. Similarly, metal ratios change systematically with high Zn/Pb, Cu/Zn,
and Cu/Pb in the proximal part of the system (Table 3).

Skarn has formed at the contact of the granodiorite porphyry dikes and
receptive carbonate units. At all skarn contacts, the granodiorite porphyry dikes
are converted to epidote endoskarn, with pistachio green epidote close to the
sedimentary contact and pink, manganese-rich epidote further from the contact.
Exoskarn also is zoned relative to the dike/limestone contact, both along strike
and

perpendicular

to

the

contact.

Proximal

(northeast)

zones

have

garnetchalcopyritemagnetite near the igneous contact, pyroxene>garnet


further away, and pyroxene plus pyroxenoid (bustamite-rhodonite) at the
marble front (Figure 19). Sphalerite>galena-pyrite occurs in all pyroxene zones
and can be quite massive at the marble front. A narrow (<1 m) bleached
selvage mantles the skarn and is marked by recrystallization of carbonate and
removal of finely disseminated carbon. Sphalerite, galena, and pyrite can occur
as disseminations, veins, or massive replacements within this bleached zone.
Some of the massive-sulfide replacements are connected to skarn by tiny dark,
sometimes wavy, veinlets similar to stylolites. In other cases, even with
complete three dimensional exposure due to mining, there is no visible
connection between these sulfide mantos and skarn.

Towards the distal end of the system, garnet-chalcopyrite is absent at the dike
contact and zones of pyroxenoid and massive-sulfide near the marble front are
much larger (Figure 19). Although garnet-chalcopyrite is not present in distal
parts of the system and the copper grade is lower than in proximal locations,
copper is still present due to chalcopyrite inclusions in sphalerite. Although
sphalerite is still the dominant sulfide in this part of the skarn system, galena is
much more abundant and iron oxide, if present, is hematite rather than
magnetite. The massive-sulfide replacement bodies in carbonate rock beyond
skarn can be very high grade and form a significant part of the ore reserves in

24

distal parts of the mine. They also account for the higher overall grades in
distal (21.3%) versus proximal (14.8%) parts of the system (Table 3).

In addition to metals and skarn mineralogy, the Groundhog system is zoned


with respect to the composition of individual mineral phases. This is most
evident in pyroxene, although garnet, ilvaite, amphibole, chlorite, pyroxenoid,
carbonate, and sphalerite also show solid-solution compositional variations.
For pyroxene, the main compositional change is an enrichment in manganese
along the fluid flow path (Figure 20). For a given location within the mine,
pyroxene becomes more johannsenitic away from the dike contact towards the
marble front (Figure 20). Similarly, for a given location in a dike to marble
transect, pyroxene becomes more johannsenitic towards the distal southwest
part of the skarn system (Figure 20). Although the Groundhog strata are
essentially horizontal compared to the vertical zonations observed in the
Yeonhwa-Ulchin district, there is a detectable manganese enrichment in
pyroxene with elevation. A larger variation is caused by the composition of the
carbonate protolith, particularly carbon. In general, dark carbonaceous
limestones are less reactive than white, carbon-poor limestones, presumably
due to the inhibitory effect of CO2 on calc-silicate forming reactions. Thus,
pyroxene formed from carbonaceous limestone is less reacted and has a
lower manganese content than it would otherwise for a given position along
strike and perpendicular to the dike contact (Figure 20). These differences in
pyroxene composition are far greater than the negligible iron and manganese
differences among carbonate protoliths.

One cause of the described variations in skarn mineralogy and metal content of
the Groundhog system is the temperature and composition of the hydrothermal
fluids (Meinert, 1987). Fluid inclusion homogenization temperatures in the
proximal part of the system and close to the dike contact exceed 400C. In the
central part of the skarn system, they are less than 350C, and in the distal
southwest part of the system, they are less than 325C. In addition, for a given

25

location along strike, fluid inclusion homogenization temperatures decline from


the

dike

contact

towards

marble.

Thus,

the

highest

fluid

inclusion

homogenization temperatures were measured at the dike contact in the


proximal northeast part of the system and the lowest fluid inclusion
homogenization temperatures were measured near the marble front in the
distal southwest part of the skarn system. A similar but less dramatic decline
was measured for fluid inclusion salinities (from >26 to <3 wt. % NaCl).

DISCUSSION
As has been appreciated from the earliest studies (e.g., Trnebohm, 1875;
Goldschmidt, 1911), most skarns are zoned. This paper has summarized
examples of skarn deposits from four major classes (Cu, W, Au, and Zn) that
illustrate some of the variations in mineralogy, metal content, geochemistry,
and geologic setting that are possible. Even though each skarn type is
somewhat different and, in some sense, each individual deposit is unique,
there are some common themes that warrant further discussion. Since the
definition of skarn is based upon mineralogy and because mineralogy is the
basis for mapping, identifying, and studying skarns in the field, it is appropriate
to focus on zonation of skarn mineralogy. In terms of calc-silicate mineralogy,
all of the skarn deposits discussed in this paper contain garnet and pyroxene.
Furthermore, all of the deposits show at least some zonation in the distribution
of these two minerals, their relative abundance, their appearance, and in many
cases, their compositions. Thus, one could ask, Is there a general pattern that
applies to all skarns?

General Patterns and Causes of Variation


As suggested earlier, most skarns seem to have a proximal zone of garnet,
distal zones of pyroxene, and distinctive minerals such as vesuvianite or a
pyroxenoid such as wollastonite, bustamite, or rhodonite near the marble
contact. There are several variations on this theme controlled by depth of
formation and wallrock composition and permeability as illustrated in Figure 21.

26

As a specific example, W skarns generally form at significant depth and are


associated with large plutons. This combination of high pressure and
temperature leads to widespread and high grade metamorphism such that W
skarns typically overprint metamorphic lithologies. Thus, there may be barren
metamorphic calc-silicate rocks interspersed with mineralized skarn. Also, at
high temperatures and pressures, permeability may be greatly reduced such
that skarn forms as relatively narrow zones along plutonic contacts (Figure 21).
In contrast, shallower and cooler skarn systems, such as some of the Cu and
Zn skarns previously discussed, appear to form from fluids that have infiltrated
for long distances relative to the intrusive contact, resulting in a more
pronounced mineralogical zonation. This may be why many W skarns show
less dramatic and less systematic mineralogical variations than most Cu and
Zn skarns. Thus, the scale and direction of mineralogical zonation appears to
be systematically different among different skarn classes. Conversely, some of
the deposits within particular skarn classes, for example Pine Creek and Salau
for W skarns and Mines Gasp and Big Gossan for Cu skarns, share many
features even though they formed at different times and on different continents.

Additional causes of skarn variation include compositional variations of plutons


and protoliths. Several studies have identified systematic associations of skarn
deposits with particular pluton compositions (e.g., Ray et al., 1995; Meinert,
1995), and some of the deposits discussed in this paper have mineralogical
variations controlled by protolith composition (e.g., dolomitic host rocks at
Costabonne and Big Gossan and carbonaceous host rocks at Salau and
Groundhog) (Figure 21). For a given skarn type, reduced plutons or wallrocks
may cause lower garnet:pyroxene ratios coupled with relatively iron-poor
garnet (dominantly ferric iron) and iron-rich pyroxene (dominantly ferrous iron);
oxidizing conditions may cause the reverse to happen (Figure 21). Thus, in a
general way, the overall garnet:pyroxene ratio and mineral compositions give
important clues about several fundamental geological variables. This concept
of oxidation state also has been used in classifying some skarn deposits (e.g.,

27

Newberry, 1982, 1983, 1991, for W skarns, and Brooks, 1994; Ray and
Webster, 1997, for Au skarns). Additionally, it has been suggested that the
major- and trace-element composition of a particular mineral phase, such as
pyroxene, can be used to classify skarn deposits (e.g., Nakano et al., 1994).

If many skarns are zoned from proximal garnet to distal pyroxene, then the
overall garnet:pyroxene ratio should serve as a crude indicator of the general
oxidation state of a system, including the effects of pluton, wallrocks, and depth
of formation (Figure 21). Figure 22 shows the effect of oxidation state of
plutons and wallrocks on a variety of skarn parameters. For the different skarn
types discussed in this paper, oxidation state can be manifested in different
ways. For example, the inner zone of a typical W skarn would contain
subcalcic garnet, whereas it would contain andraditic garnet in a copper skarn.
The outer zone of a typical high grade Au skarn would contain very iron-rich
pyroxene, whereas in a copper skarn it would be more diopsidic. Each skarn
type can be thought of as having a typical range of mineralogical
characteristics and that range can shift based upon external factors such as
the composition and oxidation state of the wall rocks (Figure 21). Thus, in
evaluating a specific skarn deposit, it is necessary to interpret the mapped
mineralogy in terms of the larger geologic context.

Skarn deposits can be thought of as variations upon a theme and the central
theme is that most skarns are zoned. This zonation occurs on scales from m
to km. For exploration purposes, zonation on a deposit scale is the most useful.
Zonation models exist for many different skarn types and although every ore
deposit in a sense is unique, the common threads among different skarns of a
given type provide powerful predictive tools for both exploration and
understanding. Figure 23 illustrates a possible exploration application of the
concept of mineralogical zonation within a skarn deposit. This cross section is
generalized, but perhaps corresponds most closely to a copper skarn. The
deep drill holes have penetrated skarn, but not the causative pluton. Using the

28

zonation

patterns

for

typical

Cu

skarns,

it

can

be

predicted

that

garnet:pyroxene ratios will increase towards the pluton, that the appearance
(color and texture) of garnet and pyroxene will change, and that distal pyroxene
zones will be more hedenbergitic and johannsenitic than proximal zones. These
predictable patterns can be used to target future exploration with considerable
confidence. Furthermore, for skarn models that attempt to quantify spatial
changes in mineralogy, composition, and geochemistry, such as that presented
for the Groundhog system, it may be possible to provide semi-quantitative
estimates of distances between different skarn zones.

Regional Variations
Since most skarns result from the interaction of magmatic systems and crustal
rocks, it is not surprising that there should be regional variations that reflect
underlying tectonic and petrogenetic processes. Early attempts at quantifying
such regional zonations focused on the economic metals in skarns, either
geographically (Shimazaki, 1975) or by skarn type (Zharikov, 1970; Smirnov,
1976). More recently, Nakano et al . (1990) showed that Sr isotopes are
relatively constant (<0.001

87

Sr/86Sr)for individual skarn deposits but vary

systematically as a function of skarn type (economic metals) and crustal


composition. In terms of

87

Sr/86Sr ratios, Cu skarns are very similar to their

associated plutons, whereas Zn-Pb skarns, that generally form in a distal


environment relative to causative igneous rocks, show very little similarity to
their associated plutons. Another way of assessing regional variations is to
examine all the different skarn types in a particular tectonic or geologic area.
Recent examples of such an approach by Newberry et al. (1997) for Alaska and
Ray and Webster (1997) for British Columbia demonstrate not only the
abundance of skarns in different parts of the world but also some patterns of
association and non-association that we are just beginning to understand.
ACKNOWLEDGMENTS
This paper is slightly modified from previous papers on skarn deposits,
zonation, and exploration techniques. For details on any of the deposits

29

discussed in this paper, the reader is referred to the more detailed studies cited
within.

REFERENCES
ALLCOCK, J.B., 1982. Skarn and porphyry copper mineralization at Mines
Gasp, Murdochville, Quebec. Economic Geology, 77, P. 971-999.
ATKINSON, W.W., Jr., and EINAUDI, M.T., 1978. Skarn formation and
mineralization in the Contact Aureole at Carr Fork, Bingham, Utah. Economic
Geology, 73, p. 1326-1365.
BILLINGSLEY, P., and HUME, C.B., 1941. The ore deposits of Nickel Plate
Mountain, Hedley, B.C. Canadian Institute of Mining and Metalllurgy, Bulletin,
44, p. 524-590.
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 Geological Society Digest, 15, p. 67-72.
BROOKS, J.W., 1994. Petrology and geochemistry of the McCoy gold skarn,
Lander County, Nevada. Unpublished Ph.D. thesis, Washington State
University, Pullman, Washington, 607 p.
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
Geology and Ore Deposits of the Great Basin, Edited by G.L. Raines, R.E.
Lisle, R.W. Schafer and W.H Wilkinson, Geological Society Nevada, Reno, 1,
p. 419-442.
BROWN, P.E., BOWMAN, J.R., and KELLY, W.C., 1985. Petrologic and stable
isotope constraints on the source and evolution of skarn-forming fluids at Pine
Creek, California. Economic Geology, 80, p. 72-95.
BUSSELL, M.A., ALPERS, C.N., PETERSEN, U., SHEPHERD, T.J.,
BERMUDEZ, C., BAXTER, A.N., 1990. The Ag-Mn-Pb-Zn vein, replacement,
and skarn deposits of Uchucchacua, Peru; studies of structure, mineralogy,
metal zoning, Sr isotopes, and fluid inclusions. Economic Geology, 85, p. 13481383.
DERR, C., and KRYLATOV, S., 1976. Comparaison entre la srie de Salau
(Arige) et kautres sries du Dvonien de la partie centrale des Pyrnes. Un
caractre original de la partie infrieure du Dvonien. Prsence de
phosphates. Acad. Sci. (Paris) Comptes Rendus, 282, ser.D, p. 2051-2054.
EINAUDI, M.T., 1982. Descriptions of skarn associated with porphyry copper
plutons, southwestern North America, In Advances in Geology of the Porphyry

30

Copper Deposits, Southwestern North America, Edited by S.R. Titley,


University of Arizona Press, Tucson, p. 139-184.
EINAUDI, M.T., MEINERT, L.D., and NEWBERRY, R.J., 1981. Skarn deposits.
Economic Geology, 75th AnniVol., p. 317-391.
ETTLINGER, A.D., 1990. A geological analysis of gold skarns and precious
metal enriched iron and copper skarns in British Columbia, Canada;
Unpublished Ph.D. Thesis, Washington State University, 246 pages.
ETTLINGER, A. D., MEINERT, L.D., and RAY, G.E., 1992. Gold skarn
mineralization and fluid evolution in the Nickel Plate Deposit, Hedley, District,
British Columbia. Economic Geology, 87, p. 1541-1565.
FONTEILLES, M., SOLER, P., DEMANGE, M., DERRE, C., KRIERSCHELLEN. A. D., VERKAEREN, J., GUY, B., and ZAHM, A., 1989. The
scheelite skarn deposit of Salau (Ariege, French Pyrenees). Economic
Geology, 84, p. 1172-1209.
GOLDSCHMIDT, V.M., 1911. Die kontakmetamorphose im Kristianiagebiet.
Oslo Vidensk. Skr., I., Mat.-Natur K1., no 1, 483 p.
GOWER, S.J., and WALKER, J.A. 1993. Skarn-type, base-metal deposits in
northern N.B. Geology and skarn occurrences, in Guidebook to the porphyry
copper and copper skarn mineralization in northern New Brunswick and Gasp,
Quebec, edited by S.R. McCutcheon & G.A.Woods, Trip #1 of Bathurst '93. 3rd
Annual Field Conference, Geological Society of CIM, p. 5-21.
GRAY, R.F., HOFFMAN, V.J., BAGAN, R.J., and MCKINLEY, H.L., 1968.
Bishop tungsten district, California, in Ridge, J.D. (ed.), Ore deposits of the
United States, 1933-1967 (Graton-Sales vol.), New York, Am. Inst. Mining
Metall. Petroleum Engineers, p. 1531-1554.
GUSTAFSON, L.B., and HUNT, J.P., 1975. The porphyry copper deposit at El
Salvador, Chile. Economic Geology, 70, 5, p. 856-912.
GUY, B., 1988. Contribution ltude des skarns de Costabonne Pyrnes
Orientales, (France) et la thorie de la zonation mtasomatique.
Unpublished Ph.D. thesis, University of Paris, Paris, France, 645 p.
GUY, B., SHEPPARD, S.M.F., FOUILLAC, A.M., LE GUYADER, R.,
TOULHOAT, P., and FONTEILLES, M., 1988. Geochemical and isotope
(H,C,O,S) studies of barren and tungsten-bearing skarns of the French
Pyrenees. Soc. Geology Appl. Mineral Deposits Spec. Pub. 6, Berlin-New York,
Springer-Verlag, p. 53-75.
JAMTVEIT, B., 1991. Oscillatory zonation patterns in hydrothermal grossular
andradite garnet, nonlinear dynamics in regions of immiscibility. American
Mineralogist, 76, p. 1319-1327.

31

JAMTVEIT, B., WOGELIUS R. A, and FRASER D. G., 1993. Zonation patterns


of skarn garnets, records of hydrothermal system evolution. Geology, 21, p.
113-116.
JAMTVEIT, B. and HERVIG, R. L., 1994. Constraints on transport and kinetics
in hydrothermal systems from zoned garnet. Science, 263, p. 505-508.
KATCHAN, G., 1984. Zoned monticellite-glaucochroite from the Gunung Bijih
(Ertsberg) skarn, Irian Jaya, Indonesia. 27th International Geol. Congress
Abstracts, 5, p. 70.
KWAK, T.A. P., 1987. W-Sn skarn deposits and related metamorphic skarns
and granitoids. Developments in Economic Geology, No. 24, Elsevier,
Amsterdam, 451 p.
MEGAW, P. K. M., RUIZ, J., and TITLEY, S. R., 1988. High-temperature,
carbonate-hosted Ag-Pb-Zn(Cu) deposits of northern Mexico. Economic
Geology, 83, p. 1856-1885.
MEINERT, L.D., 1984. Mineralogy and petrology of iron skarns in western
British Columbia, Economic Geology, 79, p. 869-882.
MEINERT, L.D., 1987. Skarn zonation and fluid evolution in the Groundhog
Mine, Central Mining District, New Mexico, Economic Geology, 82, p. 523-545.
MEINERT, L.D., 1989. Gold skarn deposits - Geology and exploration criteria.
In The Geology of Gold Deposits: The Perspective in 1988, Edited by R. Keays,
R. Ramsay and D. Groves, Economic Geology Monograph #6, p. 537-552.
MEINERT, L.D., 1992. Skarns and skarn deposits. Geoscience Canada, 19, p.
145-162.
MEINERT, L.D., 1995. Compositional variation of igneous rocks associated
with skarn deposits - Chemical evidence for a genetic connection between
petrogenesis and mineralization. In Magmas, fluids, and ore deposits, Edited by
J.F.H. Thompson, Mineralogical Association of Canada Short Course Series,
23, p. 401-418.
MEINERT, L.D., HEFTON, K.K., MAYES, D., and TASIRAN, I., 1997. Geology,
zonation, and fluid evolution of the Big Gossan Cu-Au skarn deposit, Ertsberg
district, Irian Jaya. Economic Geology, 92, p. 509-526.
MYERS, G. L., 1994. Geology of the Copper Canyon-Fortitude skarn system,
Battle Mountain, Nevada. Unpublished Ph.D. thesis, Washington State
University, Pullman, Washington, 356 p.
MYERS, G. L., and MEINERT, L.D., 1991. Alteration, mineralization, and gold
distribution in the Fortitude gold skarn. In Geology and Ore Deposits of the
Great Basin, Edited by G.L. Raines, R.E. Lisle, R.W. Schafer and W.H
Wilkinson, Geological Society Nevada, Reno, 1, p. 407-418.

32

NAKANO, T., 1978. The zoned skarn developed in diorite porphyry in the
Shinyama area, Kamaishi mine, Japan. Mining Geology, 28, p. 99-109.
NAKANO, T., 1989. Fluctuation model for compositional heterogeneity in skarn
clinopyroxenes. Geochemical Journal, 23, p. 91-99.
NAKANO, T., SHIMAZAKI, H., and SHIMIZU, M., 1990. Strontium isotope
systematics and metallogenesis of skarn deposits in Japan. Economic Geology,
85, p. 794-815.
NAKANO, T., TAKAHARA, H., and NORIMASA, N., 1989a. Intracrystalline
distribution of major elements in zoned garnet from skarn in the Chichibu Mine,
central Japan. Canadian Mineralogist, 27, p. 499-507.
NAKANO, T., TAKAHARA, H., and FUJII, T., 1989b. Development of
compositionally-zoned epidote from the Yaguki skarn deposit, northeastern
Japan. Mining Geology, 39, p. 1-8.
NAKANO, T., YOSHINO, T., and SHIMAZAKI, H., 1989c. Local evolution and
fluctuation of skarn clinopyroxene composition from the Horado mine, central
Japan. Annual Report of the Institute of Geoscience, University of Tsukuba, 15,
p. 102-106.
NAKANO, T., YOSHINO, T., SHIMAZAKI, H., and SHIMIZU, M., 1994.
Pyroxene composition as an indicator in the classification of skarn deposits.
Economic Geology, 89, p. 1567-1580.
NEWBERRY, R.J., 1980. The geology and chemistry of skarn formation and
tungsten deposition in the central Sierra Nevada, California. Unpublished Ph.D.
thesis, Stanford University, Stanford, California, 325 p.
NEWBERRY, R.J., 1982. Tungsten-bearing skarns of the Sierra Nevada. I. The
Pine Creek Mine, California. Economic Geology, 77, p. 823-844.
NEWBERRY, R.J., 1983. The formation of subcalcic garnet in scheelitebearing skarns. Canadian Mineralogist, 21, p. 529-544.
NEWBERRY, R., 1991. Scheelite-bearing skarns in the Sierra Nevada Region,
California. Contrasts in zoning and mineral compositions and tests of infiltration
metasomatism theory, In Skarns - Their Genesis and Metallogeny.
Theophrastus Publications S.A., Athens, p. 343-384.
NEWBERRY, R.J., ALLEGRO, G. L., CUTLER, S.E., HAGEN-LEVEILLE, J.H.,
ADAMS, D.D., NICHOLSON, L.C., WEGLARZ, T.B., BAKKE, A.A., CLAUTICE,
K.H., COULTER, G.A., FORD, M.J., MYERS, G.L., and SZUMIGALA, D.J.,
1997. Skarn deposits of Alaska. In Ore Deposits of Alaska, Edited by R.J.
Goldfarb, Economic Geology Monograph 9, (in press).

33

OCHIAI, K., 1987. A reaction model relating skarn zones and ore formation at
the Nippo copper ore deposit, Kamaishi Mine, northeastern Japan. Economic
Geology, 82, p. 1001-1018.
PAN, Y., FLEET, M.E., and RAY, G.E., 1994. Scapolites in two Canadian gold
deposits. Nickel Plate, British Columbia and Hemlo, Ontario. Canadian
Mineralogist, 32, p. 825-837.
RAIMBAULT, L. and KAELIN, J.L., 1987. Gologie, ptrographie et gochimie
de la granodiorite de la Fourque et de ses facis hydrothermaliss associs
aux skarns scheelite de Salau (Pyrenes, France). Bull. Minralogie, 110, p.
633-644.
RAY, G.E., and DAWSON, G.L., 1994. The geology and mineral deposits of the
Hedley Gold Skarn District, Southern British Columbia; B.C. Ministry of Energy,
Mines and Petroleum Resources, Bulletin 87, 156 p.
RAY, G.E., DAWSON, G.L., and SIMPSON, R., 1988. Geology, geochemistry
and metallogenic zoning in the Hedley Gold-Skarn Camp. British Columbia
Ministry of Energy Mines and Petroleum Resources, Geological Fieldwork,
1987, Paper 1988-1, p. 59-80.
RAY, G.E., DAWSON, G.L., and WEBSTER, I.C.L., 1996. The stratigraphy of
the Nicola Group in the Hedley district, British Columbia and the chemistry of
its intrusions and Au skarns. Canadian Journal of Earth Science, 33, p. 11051126.
RAY, G.E., and WEBSTER, I.C.L., 1997. Skarns in British Columbia, B.C.
Ministry of Employment and Investment, Bulletin 101, 260 p.
RAY, G.E., WEBSTER, I.C.L., and ETTLINGER, A.D., 1995. The distribution of
skarns in British Columbia and the chemistry and ages of their related plutonic
rocks, Economic Geology, Volume 90, pages 920-937.
SHELTON, K.L., 1983. Composition and origin of ore-forming fluids in a
carbonate-hosted porphyry copper and skarn deposit. A fluid inclusion and
stable isotope study of Mines Gasp, Quebec. Economic Geology, 78, p. 387421.
SHIMAZAKI, H., 1975. The ratios of Cu/Zn-Pb of pyrometasomatic deposits in
Japan and their genetic implications. Economic Geology, 70, p. 717-724.
SMIRNOV, V.I., 1976. Skarn deposits, In Geology of Mineral Deposits. MIR
Publications, Moscow, p. 156-188.
SOLER, P., 1977. Ptrographie, thermochimie et mtallognie du gisement de
scheelite de Salau (Pyrnes Arigeoises - France). Unpublished Ph.D. thesis,
Ecole des Mines de Paris, Paris, France, 220 p.

34

THEODORE, T.G., and BLAKE, D.W., 1975. Geology and geochemistry of the
Copper Canyon porphyry copper deposit and surrounding area, Lander
County, Nevada. United States Geological Survey Professional 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). United States Geological
Survey Professional Paper 798-C, 85 p.
THEODORE, T.G., SILBERMAN, M.L, and BLAKE, D.W., 1978. Geochemistry
and K-Ar ages of plutonic rocks in the Battle Mountain mining district, Lander
County, Nevada. United States Geological Survey Professional Paper 798-A,
24p.
TRNEBOHM, A.E., 1875. Geognostisk beskrifning ofver Persbergets
Grufvefalt. Sveriges Geologiska Underskning, P.A. Norstedt and Sons,
Stockholm, 21 p.
VAN MARCKE DE LUMMEN, G., 1983. Ptrologie et gochimie des skarnodes
du site tungstifre de Costabonne (Pyrnes Orientales). Unpublished Ph.D.
thesis, University Catholique Louvain-la-Neuve, Belgium, 293 p.
VAN MARCKE DE LUMMEN, G., 1988. Oxygen and hydrogen isotope
evidence for influx of magmatic water in the formation of W-, Mo- and Snbearing skarns in pelitic rocks at Costabonne, France, and Land's End,
England. In Proceedings of the 7th Quadrennial IAGOD Symposium, Edited by
E. Zachrisson, E. Schweizerbart'sche Verlagsbuchhandlung, Stuttgart, p. 355362.
VAN MARCKE DE LUMMEN, G. and VERKAEREN, J., 1986. Physicochemical
study of skarn formation in pelitic rock, Costabonne peak area, eastern
Pyrenees, France. Contributions to Mineralogy and Petrology, 93, p. 77-88.
VLASOVA, D.K., PODLESSKIY, K.V., KUDRYA, P.F., BORONIKHIN, V.A., and
MURAVITSKAYA, G.N., 1985. Zoning in garnets from skarn deposits.
International Geology Review, 27, p 465-482.
WARES, R. and BERNARD, P., 1993. Skarn-type, base-metal deposits in
Gasp P.Q.: The copper deposits at Mines Gasp, Quebec. In Guidebook to
the porphyry copper and copper skarn mineralization in northern New
Brunswick and Gasp, Quebec, Edited by S.R. McCutcheon and G.A.Woods,
Trip #1 of Bathurst '93. 3rd Annual Field Conference, Geological Society of
CIM, p. 54-62.
WILLIAMS-JONES, A.E., 1986. Low-temperature metamorphism of the rocks
surrounding les Mines Gasp, Quebec. Implications for mineral exploration.
Economic Geology, 81, p. 466-470.

35

WOTRUBA, P.R., BENSON, R.G., and SCHMIDT, K.W., 1988. Geology of the
Fortitude gold-silver skarn deposit, Copper Canyon, Lander County, Nevada. In
Bulk mineable precious metal deposits of the western United States, Edited by
R.W. Schafer, J.J. Cooper and P.G. Vikre, Geological Society of Nevada,
Reno, p. 159-172.
YUN, S., 1978. Petrography, chemical composition, and depositional
environments of the Cambro-Ordovician sedimentary sequence in the Yeonhwa
I mine area, southeastern Taebaegsan region, Korea. Geological Society
Korea Journal, 14, p. 145-174.
YUN, S., 1979. Geology and skarn ore mineralization of the Yeonhwa-Ulchin
zinc-lead mining district, southeastern Taebaegsan region, Korea. Unpublished
Ph.D. thesis, Stanford University, Stanford, California, 306 p.
YUN, S., and EINAUDI, M.T., 1982. Zinc-lead skarns of the Yeonhwa-Ulchin
district, South Korea. Economic Geology, 77, p. 1013-1032.
YUN, S., and SILBERMAN, M.L, 1979. K-Ar geochronology of igneous rocks in
the Yeonhwa-Ulchin zinc-lead district and southern margin of Taebaegsan
basin, Korea. Geological Society Korea Journal, 15, p. 89-99.
ZAHM, A., 1987a. Ptrologie, minralogie et gochimie des cornennes
calciques et des skarns minraliss dans le gisement de scheelite de Salau
(Arige, France). Unpublished Ph.D. thesis, University of Paris, Paris, France,
330 p.
ZAHM, A., 1987b. The compositional evolution of calc silicates from the Salau
skarn deposit (Arige, Pyrnes). Bull. Minralogie, 110, p. 623-632.
ZHARIKOV, V.A., 1970. Skarns. International Geology Review, 12, p. 541-559,
p. 619-647, and p. 760-775.
ZIMMERMAN, B. S., MYERS, G.L., MEINERT, L.D., and LARSON, P.B, 1992.
Stable isotopic evidence for magmatic fluid dominance in the Fortitude Gold
Skarn Deposit, Lander County, Nevada. Geological Society of America,
Abstracts with Programs, 24, p. 144-145.

36

TABLE AND FIGURE CAPTIONS


Table 1: Summary of Big Gossan pyroxene compositions
West
mole %
Hd
Di
Jo
# analyses
mole %
Hd
Di
Jo
# analyses
mole %
Hd
Di
Jo
# analyses
mole %
Hd
Di
Jo
# analyses
Location

Middle

East

Average/
Total

37,0%
58,7%
4,3%
13

Higher than 2760 meters


38,9%
37,2%
58,1%
58,7%
3,0%
4,1%
12
15

37,7%
58,5%
3,8%
40

28,7%
68,0%
3,3%
14

Between 2495 and 2760 meters


13,3%
18,2%
84,8%
79,5%
1,9%
2,3%
16
15

19,7%
77,8%
2,5%
45

10,2%
88,6%
1,2%
4

29,8%
66,7%
3,5%
31
<19100E

Lower than 2495 meters


15,9%
13,0%
82,1%
85,8%
2,0%
1,3%
4
3
Average/total
16,9%
81,0%
2,1%
32

26,4%
70,6%
3,0%
33

19100-19400E

>19400E

13,0%
85,5%
1,5%
11

24,3%
72,8%
2,8%
96

Summary of Big Gossan (Ertsberg district, Irian Jaya) pyroxene compositions


classified by elevation and mine location (from Meinert et al., 1997).

37

Table 2: Big Gossan metal zonation


West
<19100E

Cu (%)
Au (ppm)
Ag (ppm)
Pb (ppm)
Zn (ppm)
As (ppm)
Mo (ppm)
Co (ppm)
kilotons

2,95%
1,26
25,0
1.025
1.567
252
7,1
114
7.602

Cu (%)
Au (ppm)
Ag (ppm)
Pb (ppm)
Zn (ppm)
As (ppm)
Mo (ppm)
Co (ppm)
kilotons

3,01%
1,18
21,6
242
1.028
81
10,2
134
14.447

Cu (%)
Au (ppm)
Ag (ppm)
Pb (ppm)
Zn (ppm)
As (ppm)
Mo (ppm)
Co (ppm)
kilotons

1,71%
0,59
9,3
26
83
119
19,5
43
4.414

Cu (%)
Au (ppm)
Ag (ppm)
Pb (ppm)
Zn (ppm)
As (ppm)
Mo (ppm)
Co (ppm)
kilotons

2,78%
1,11
20,5
431
1.025
136
10,9
113
26.463

Middle
19100-19400E

East
>19400E

Higher than 2760 meters


2,07%
2,68%
0,71
0,96
11,3
15,9
34
54
426
662
74
50
4,1
4,6
50
79
10.132
7.955
Between 2495 and 2760 meters
2,22%
1,33%
0,53
0,37
8,0
7,0
29
71
215
131
50
35
11,6
3,5
30
29
14.398
7.596
Lower than 2495 meters
1,09%
0,53%
0,26
0,22
5,9
3,7
34
34
94
105
50
67
14,4
16,3
19
15
3.485
950
Average/total
2,03%
0,56
8,9
32
276
58
9,2
36
28.015

1,93%
0,64
11,1
61
386
44
4,8
52
16.501

Average/
Total

2,52%
0,95
16,8
333
837
119
5,1
78
25.689

2,35%
0,75
13,2
122
520
59
9,4
71
36.441

1,34%
0,42
7,4
30
89
86
17,1
31
8.849

2,28%
0,78
13,8
61
196
26
2,7
23
70.979

Summary of Big Gossan (Ertsberg district, Irian Jaya) Cu,Au, Ag, Pb, Zn, As, Mo, and
Co grades classified by elevation and mine location (from Meinert et al., 1997).

38

Table 3: Groundhog metal zonation

Zone
Location

Proximal Intermediate
Distal
Northeast
Central
Southwest

Total/
Average

Tons

486.403

1.867.106

84.408

2.437.917

Grade
Zn %
Pb %
Cu %
Ag g/t

12,60
1,60
0,64
31

13,66
2,94
0,61
51

14,17
6,69
0,47
28

13,5
2,8
0,61
46

Zn+Pb+Cu

14,8%

17,2%

21,3%

16,9%

7,9
19,7
2,5
4018
510
204

4,6
22,4
4,8
2668
574
119

2,1
30,1
14,2
4975
2349
165

4,8
22,0
4,6
0,29
0,06
76

Ratios
Zn/Pb
Zn/Cu
Pb/Cu
Zn/Ag
Pb/Ag
Cu/Ag

Metal grades and ratios from the Groundhog Zn skarn (after Meinert, 1987).

39

Schematic geologic map and cross section through the Mines Gasp Copper
Mountain skarn system, Quebec, Canada (modified from Allcock, 1982). Unpublished
Mines Gasp mapping, and garnet:pyroxene observations by L. Meinert (1992)).

40

Composition of Mines Gasp pyroxene from proximal (Copper Mtn.) to distal


(E-zones) skarn zones (Unpublished data, Meinert, 1992).

41

Plan view of Big Gossan 2930 m elevation geology and underground drill stations,
Irian Jaya, Indonesia. Vertical projection of skarn zones is approximate due to
steep dip of units. More accurate skarn geometry is illustrated in two cross sections
(Figure 4) indicated by dark lines on plan map (modified from Meinert et al., 1997).

42

Cross section through Big Gossan drill stations BGU 26 and BGU 14 illustrating
typical geology and distal skarn geometry (Modified from Meinert et al., 1997).
Lines of section are shown on Figure 3.
43

Geologic map of the Salau area (France) showing skarn deposits on the contact of
the L Fourque stock (modified from Fonteilles et al., 1989). Location of Figure 6 cross
section is shown on the southern margin of the stock.

44

Vertical north-south cross section of the Vronique orebodies of the Salau


(France) W skarn (modified from Fonteilles et al., 1989).

45

Zonation of skarn formed from marble in the Salau (France) W skarn (modified
from Soler, 1977).

46

Composition of garnet and pyroxene from the Salau (France) W skarn (data from
Soler, 1977; Zahm, 1987a).

47

Zonation of skarn from granite outward to dolomitic marble, Costabonne (France)


(modified from Guy et al., 1988).

48

Generalized cross section through the upper north end of the Pine Creek
mine, California, showing the distribution of skarn relative to intrusive contacts
(modified from Newberry, 1982).

49

Skarn zonation in underground drift at Pine Creek, California (modified from


Newberry, 1982).

50

Composition of garnet and pyroxene from the Pine Creek and Garnet Dike
skarns (modified from data in Newberry, 1980). M = metamorphic occurrences.
All other symbols are metasomatic occurrences. Inner, main, and outer skarn
zones are shown in Figure 11 and Newberry (1980).

51

Plan view of skarn zonation relative to the Toronto Stock in the Hedley district, British
Columbia, Canada (after Ray et al., 1996; Ettlinger, 1992).

52

Composition of garnet and pyroxene from the Nickel Plate mine, British Columbia
(modified from Ettlinger et al., 1992).

53

Cross section from the Toronto Stock through distal skarn zones of the Nickel Plate
mine, British Columbia, showing change in garnet-pyroxene abundance and
compositional variations in pyroxene (modified from Ettlinger et al., 1992).

54

Plan view of skarn in the West and Fortitude deposits, Nevada, illustrating
mineralogical zonation and inferred direction of fluid flow (after Myers, 1994).
North-South cross section through West and Fortitude deposits is shown in
Figure 17.

55

North-South cross section through Copper Canyon stock and West and
Fortitude deposits, Nevada, illustrating mineralogical and metal zonation (after
Myers and Meinert, 1991).

56

Cross sections through A) the Yeonhwa I skarn system, Korea, showing the
distribution of ores as both tabular bodies along bedding and near-vertical pipes
across bedding, and B) the Ulchin skarn system showing the distribution of
skarn at igneous contacts (after Yun and Einaudi, 1982).

57

Idealized zonation pattern for the Groundhog skarn system, New Mexico,
illustrating variations in mineralogy and composition, fluid inclusion
characteristics, and metal ratios (after Meinert, 1987). Cp = chalcopyrite, gar =
garnet, gl = galena, hm = hematite, jo = johansennite, mt = magnetite, and pyx =

58

Ternary plot of Groundhog, New Mexico, pyroxene composition as a function of


protolith composition, and location along strike and perpendicular to the dike
contact as discussed in the text (after Meinert, 1987).

59

Illustration of the general effect on skarn size of variations in a) depth of


formation, and b) geometry, c) porosity/permeability, and d) composition of the
replaced strata. Abbreviations: A = amphibole, G = garnet, O = olivine, P =
pyroxene, S = serpentine, W = wollastonite.

60

Oxidation state of skarn deposits in terms of plutonic and host rock characteristics
(modified from Newberry, 1991). The oxidation state of plutonic rocks is measured
by whole rock oxide Fe2O3/( Fe2O3+FeO). Alternative indicators of plutonic oxidation
state include oxide mineralogy (e.g., ilmenite, magnetite, hematite) and iron content
of mafic minerals such as pyroxene, amphibole, and biotite. The oxidation state of
host rocks is measured by the abundance of carbon (e.g., graphite, carbon,
hydrocarbon), sulfides (pyrrhotite, pyrite), and oxides (ilmenite, magnetite, hematite).

61

Cartoon illustrating application of zonation models in skarn exploration in a typical


situation where a blind skarn target is being explored by surface drilling. Skarn
intercepts are evaluated in terms of skarn mineralogy and mineral compositions.
Comparison of these data with models based upon specific deposits as discussed
in the text allows determination of relative location within the skarn and
approximate vectors to specific skarn or ore zones.

62

Potrebbero piacerti anche