Anda di halaman 1dari 36

Economic Geology

Vol. 86, 1991, pp. 318-353


Geochemistry of Mineralizing Fluids in the Bralorne-Pioneer Mesothermal
Gold Vein Deposit, British Columbia, Canada
C. H. B. LEITCH*, C. I. GODWIN, T. H. BROWN,
Department of Geological Sciences, University of British Columbia, Vancouver, British Columbia, Canada V6T 2B4
AND B. E. TAYLOR
Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario, Canada K1A OE8
Abstract
Gold quartz veins at the Bralorne mesothermal vein deposit have extensive quartz-ankeritic
carbonate-muscovite hydrothermal alteration envelopes that grade outward to chlorite-calite-
albite. Minor pyrite with traces of sphalerite, galena, and tetrahedrite are found in the veins
with native gold; more abundant pyrite and arsenopyrite, with lesser pyrrhotite and chalcopyrite
occur in altered wall rock adjacent to the veins. Envelopes are characterized by addition of
KgO, COg, S, As, and Au, with corresponding depletion of NagO, FeO total, and MgO; SiOg
and CaO are locally depleted and reconcentrated. Fluid inclusions suggest that deposition of
gold was from dilute aqueous solutions (<5 wt % NaC1 equiv) that had a significant carbonic
component (5-15 mole % COg + CH4), at 350C and 1.75 kbars. Progressive dilution of this
fluid by cooler meteoric water led to fluids at 250C and 0.5 to I kbar with 1 wt percent or
less NaC1 equivalent and no detectable carbonic component. Sulfur isotope ratios of sulfides
associated with gold mineralization range from -7 to +9 per mil, clustering about a magmatic
signature of 0 per mil. The ore fluid had a so value of 13 _ I and 3C of-11 _ 2 per mil,
based on measurements in coexisting vein quartz, carbonate, and muscovite. Temperature of
the mineralizing fluids appears to have increased with depth at a normal geothermal gradient
of approximately 30 C/km.
Gold, sulfides, and the associated alteration assemblages are modeled, using observed min-
eralogy and thermodynamic data for chloride complexes, as having been deposited from a
slightly acid solution (pH 4.5) with an Na/K ratio of at least 8:1 and a high content of dissolved
COg (log fugacity = 2.5). Conditions were strongly reducing, as suggested by the log fugacity
of CH4 (0.5), with fo2 about 10 -30 bars andfs2 about 10 -7 bars. Precipitation of gold in the
immediately adjacent, highly quartz-sericite-ankerite-pyrite-altered wall rock, was due to re-
action with (carbonation of) the wall rock that caused the pH of the ore fluid to rise, destabilizing
aurous chloride complexes. Such a shift is in the direction of slightly increasing or constant
stability of gold thiosulfide ions, and at a relatively constant fo2. The most significant result
of this modeling is that in detail, the actual precipitation of gold is by reduction of aurous ions
by electrons donated through concurrent oxidation of S -g in HgS to S- in pyrite (FeSg). This
explains the empirically observed correlation between pyrite and gold in this type of deposit,
and in the model between consumption of Fe +g and production of pyrite on the one hand,
with precipitation of gold on the other.
Introduction
THE Bralorne-Pioneer mesothermal gold vein pro-
duced 130 metric tons of gold over a 70-year span,
the only deposit in the Canadian Cordillera to ap-
proach the output of major deposits in the Precam-
brian shield (Bertoni, 1983; Phillips, 1986). In this
study, petrographic and chemical analyses of hydro-
thermal alteration, fluid inclusion and stable isotope
studies, and thermodynamic modeling, have led to
estimates of the pressure-temperature-composition
(P-T-X) of the mineralizing fluids. Stable isotope
studies confirm that significant interaction of wall
* Present address: Geological Survey of Canada, 100 West
Pender Street, Vancouver, British Columbia V6B 1R8.
rocks with the ore fluid took place and shed light on
the source of the fluids.
Geologic Setting of the Bralorne Deposit
The Bralorne gold vein is in the Bridge River min-
ing camp, 180 km north of Vancouver (Fig. 1). The
tectonic setting, just east of the Coast Plutonic Com-
plex, is described by Rusmore et al. (1988); detailed
geology and previous work on the deposit are in
Leitch (1990). In the Bridge River camp, two main
lithologic assemblages can be distinguished: one
dominantly oceanic and the other dominantly island
arc. The former is represented by the Permian(?) to
Jurassic Bridge River Complex that comprises basalts
and thick accumulations of ribbon chert with minor
0361-0128/91/1185/318-3653.00 318
BRALORNE Au DEPOSIT, B. C. 319
E'/dox"'d-/Xtx'/ \\ I 5100
Geologic Contact
h--hornblende k = K -Ar BR
b=biotite r =Rb-Sr J
in o LAKE 15
z=zircon u =U-Pb 1
ngress
0 5 10
EOCENE L E G E N D k44
u hk2 BR BR v';
Rexmount Porphyry n,o. r 63 9 ' vvv
Bendor Plutons Lake
coas,.,u,o.c co..,ex c.c co
hk284zu272* * * + + + -
MethowTerrane '( ''-t,n--.+:+fi7'+ + +*+*$
./
EARLY PERMIAN-EAR1. Y CRETACEOUS - -2 + + + +
Cadwallader ,errane
PERMIAN-EARLY JURASSIC .... YUKON BR
BRIDGE RIVER TERRANE , -- .... ] XX % '+l /'Anderson
:. +'+ + + hk+'/ / Lake
Bridge River Assembledge ' + :5 + BR Prince % + + t'x x / +hk63. + '"
( George
ULTRAMAFICROCKS [ -
Cadwallade. Fault Zone '+ usa hk73 :t -- 5034'
FIG. 1. Geology, mineral deposits, and isotopic dates of the Bridge River district, southwestern
British Columbia. Geology is after Woodsworth (1977) and Church (1987). Mineral deposits correspond
to the following numbered open triangle symbols: 1 = Bralorne, 2 = Ida May, 3 = Pioneer, 4 = P. E.
Gold, 5 = BRX (Arizona), 6 = BRX (California), 7 = Veritas, 8 = Waterloo, 9 = Lucky Gem, 10 = Lucky
Strike, 11 = Minto, 12 = Golden Sidewalk, 13 = Congress, 14 = Peerless, 15 = Summit, 16 = Kelvin,
17 = Olympic, 18 = Matson, 19 = Greyrock, 20 = Piebiter (Chopper).
limestone and coarse elastic rocks. Alpine-type ultra-
marie rocks in lensold to very elongate bodies form
part of the stratified assemblage (Schiarizza et al.,
1989). The ultramarie rocks may mark the sites of
major crustal shortening that were later foci for trans-
current movements. The island-arc assemblage is
represented by the Cadwallader Group of Permian(?)
to Triassic age (Cairnes, 1937; Rusmore, 1987). It is
a basaltic andesitc pile with minor felsic volcanics (Pi-
oneer Formation) overlain by volcaniclastics with mi-
nor limestone (Hurley Formation).
The Bridge River and Cadwallader terranes con-
taining these two assemblages form small lozengelike
fault-bounded slices sutured between the Insular su-
perterrane on the west and the Intermontane super-
terrane on the east (Monger, 1984). Potter (1986)
and Rusmore (1987) proposed that a collapsed back-
arc basin (the Bridge River terrane) was thrust onto
the continent margin together with its offshore vol-
canic arc (the contemporaneous Cadwallader terrane).
Others propose that the two terranes were unrelated
(Wheeler and McFeely, 1987). If so, they represent
true suspect terranes that could have been transported
for large distances from their original places of de-
position. In any case, the terranes have been displaced
northward for up to 150 km from correlative rocks
of the Hozameen terrane by Tertiary movement along
the Fraser strike-slip fault (Haugerud, 1985; Ray,
1986). The Hozameen terrane hosts the Carolin gold
deposit, which, although of a more dispersed or
stockwork nature and hosted in younger (Methow)
rocks, has similarities to Bralorne in alteration style
and mesothermal character.
The Bralorne block containing the Bralorne and
Pioneer deposits, shown on the 8 level underground
in Figure 2, is bounded by the northwest-trending
Fergusson and Cadwallader strike-slip faults, which
are marked along their length by sinuous serpentin-
ized ultramarie bodies. Latest movement may have
been right lateral (Stevenson, 1958), the same as the
Fraser fault system. Bralorne diorite makes up the
bulk of the elongate stock hosting the vein gold de-
posits; its northeastern contact with Pioneer and Hur-
ley Formations has been intruded by a tabular body
320 LEITCH, GODWIN, BROWN, AND TAYLOR
/
/
/ \
*'o
/
BRALORNE Au DEPOSIT, B. C. 321
of soda granite, almost 3 km long by up to 500 m
wide. These intrusions taper southeast near the Pi-
oneer mine; farther southeast at the P. E. Gold prop-
erty they are dikelike (Nordine, 1983). Thus the ma-
jor ore host is diorite at Bralorne but volcanics at Pi-
oneer. The diorite-soda granite complex is spatially
associated with gold mineralization but predates it by
at least 150 m.y. (Leitch et al., 1991). Mineralization
was related instead to a swarm of early Late Creta-
ceous (90 Ma) "albitite" dikes that parallel the veins
and postdate the metamorphic peak at ca. 100 Ma.
These dikes are cut by, and altered along, veins. They
are gradational to usually unaltered and therefore
largely postmineral green hornblende porphyry dikes
(86 Ma).
Workings at the Bralorne mine extend over a ver-
tical range of 2 km from the 0 level at surface at 1,400-
m elevation, where the Cosmopolitan vein crops out
in the old King mine area (Fig. 2), down to the 44
level at 600 m below sea level. Much of the produc-
tion of the upper levels of the mine came from the
51 vein; the 51B FW vein still contains some 150,000
metric tons at 15 g/metric ton gold above the 8 level
(Bellamy and Arnold, 1985). Farther to the southeast,
the largest vein in the mine, the 77, accounted for
most of the production at depth. These major gold-
bearing veins strike about 110 and dip north at 70 .
Slickensides plunge 45 east and steps indicate the
last movement was reverse. Major ore shoots in the
veins occupy less than 20 percent of the vein and
plunge steeply west, roughly perpendicular to the
slickensides. The veins contain thin dark ribbons of
fine-grained sulfide in massive milky quartz with mi-
nor calcite. The "fault valve" hypothesis of Sibson et
al. (1988) offers an explanation for fluid migration
and is consistent with the main features of the ribbon-
banded yet coarsely crystalline quartz veins at Bra-
lorne. A more detailed structural analysis is in Leitch
(1990).
Quartz-carbonate-sericite-chlorite envelopes, up to
several meters wide, are extensive around the veins.
They contain disseminated pyrite, pyrrhotite, and
lesser chalcopyrite; arsenopyrite is confined to vein
margins. Minor sphalerite and especially galena cor-
relate with portions of the veins that are richer in
gold. Traces of tetrahedrite and stibnite have been
observed but tellurides have not. Gold is principally
found as thin smeared flakes of native metal in black
sulfidic septae of the strongly ribboned shear veins.
Gold is rarely found by itself in the quartz, usually as
rich pockets in the relatively rare extensional veins
(oriented at 060), such as the 27 vein in the Pioneer
mine.
Analytical Methods
Mineral analysis utilized X-ray diffraction (XRD),
scanning electron microscope-energy dispersive sys-
tem (SEM-EDS), and CAMECA wavelength dispersive
electron microprobe (WDEM). Operating conditions
and correction procedures for the SEM-EDS and
WDEM studies are in Leitch (1989). Routine analysis
of secondary standards by WDEM were within _+5
percent of the accepted values. Compositions of both
the precursor (feldspar and hornblende) and the main
alteration minerals were studied. The variation about
a median value of major element oxides for all the
minerals studied was less than 10 percent in an in-
dividual grain or within a sample. Analyses for whole-
rock chemical compositions were by X-ray fluores-
cence spectrometry (XRF), mainly on pressed powder
TABLE 1. Summary of Modal Mineralogy for Primary Rocks and Alteration Envelopes around the Bralorne Veins
Unaltered (model vol %)
Diorite Soda granite
Fresh Fresh
Altered (modal vol %)1
Outer Central Inner
cl-ep cb-asqns qz-ms-cb
Di SG DI SG DI SG
Qz 10 37 5 43 12 40 25 50
Ab 55 52 30 42 30 35 5 5
Hb 33 102 5 a
Ms 10 3 13 7 30 25
Ca 15 3 23 13 15 10
Ak 2 5 2 20 7
C1 20 7 12
Bi 2 a
Ep 10
Ox 2 <1 2 1 1 1 tr tr
Sx tr i i i 2 2 5 3
Abbreviations: ab = albite, ak = ankerite, bi = biotite, ca = calcite, cb = carbonate, el = chlorite, DI = diorite, ep = epidote, hb
= hornblende, ms = serieite (muscovite), ox = oxides (ilmenite, rutile, leucoxene, sphene), qz = quartz, SG = soda granite, sx
= sulfides (pyrite, arsenopyrite, pyrrhotite, ehalcopyrite), tr = trace
i Averaged from visual thin section estimates in Leitch and Godwin (1988, table 2-4-,5)
2 Identification not certain
a Not common
322 LEITCH, GODWIN, BROWN, AND TAYLOR
TABLE 2. Electron Microprobe Analyses
Carbonates
Sample no. C094A 84-49/ 84-49/ C116-1
812' 795'
Type Calcite Calcite Calcite Calcite
Location
Spot (n) 2-10(2) 9 2-1(4)
MgCOa 0.51 2.19 1.4 1.82
CaCOa 97.10 95.90 97.1 95.22
FeCOa 2.38 3.01 1.5 1.57
Total 99.99 101.1 100.0 98.61
Fe/Mg 0.77 0.50 0.43 0.39
Sample no.
SB84-49/795'
Ankerite
Cores
7(3) 3(6)
30.57 31.96
62.92 62.32
7.79 6.59
101.3 100.9
0.16 0.15
Siderite
Rims
2(4)
47.39
15.26
38.27
100.9
0.38
19-51 FW I
Type Calcites Ankerites
Location Cores Rim Margins
Spot 4 5B 6A 1 6B 3 6 5
MgCOa 3.3 1.6 2.1 34.11 36.94 34.81 37.99 31.12
CaCO3 95.2 96.6 96.8 51.79 51.48 54.75 51.23 49.29
FeCO3 1.5 1.8 1.1 16.23 11.35 11.57 11.82 15.64
Total 99.9 99.8 100.0 102.1 99.77 101.1 101.0 96.06
Fe/Mg 0.25 0.46 0.27 0.26 0.18 0.19 0.18 0.27
8(3)
44.44
18.18
34.60
97.22
0.35
Sample no. C093A
Type
Site (n) 5(1) 6(7)
SiO2 67.79 67.21
A12Oa 19.51 19.87
CaO 0.69 0.72
NaO 11.33 11.11
KO 0.12 0.12
Total 99.44 99.03
Ab 96 96
An 3 3
Or 1 1
Rims
5A
33.5
5O.5
16.0
100.0
0.26
Feldspars
C094A C182-2
Albite
1-2(6) (6)
67.77 68.58
19.82 19.85
0.52 0.26
11.67 11.15
0.03 0.05
99.81 99.89
97 99
3 1
0 0
19-51 FW1
68.76
19.72
0.25
11.84
0.09
100.7
98.5
1.0
0.5
Sample no.
Type
Site (n)
SiO
AlcOa
TiO2
MgO
FeO
KO
HO
Total
Fe/Mg
C182-2
(3)
48.26
37.19
0.30
1.10
0.90
7.81
4.67
100.2
0.31
Micas
84-49/795' C094A
Muscovite
6(2) 5(2) 2-8
48.48 49.46 54.30
34.18 34.47 35.30
0.25 0.42 0.05
1.60 2.63 0.19
1.06 2.28 0.73
8.44 5.47 4.65
4.54 4.72 4.77
98.55 99.45 99.99
0.26 0.33 0.68
19-51FW1 C116-1 84-49/795'
Sericite Biotite
7, 8(2) 4A, 2(2) 4 4
49.27 47.17 47.60 39.72
33.00 37.01 32.33 18.55
0.23 0.08 0.51 1.00
2.16 0.82 2.62 16.90
0.88 0.77 1.89 10.41
9.54 8.40 8.29 7.39
4.56 4.56 4.47 4.15
99.64 98.81 97.72 98.13
0.19 0.34 0.28 0.26
Cations
Si 6.69 6.40 6.48 6.82 6.48 6.21 6.39 5.74
Ti 0.02 0.02 0.04 0.00 0.02 0.01 0.05 0.10
A1 5.14 5.32 5.35 5.24 5.10 5.75 5.11 3.16
Mg 0.18 0.31 0.53 0.04 0.42 0.16 0.53 3.64
Fe 0.08 0.11 0.25 0.08 0.09 0.08 0.21 1.26
K 1.24 1.42 0.91 0.75 1.60 1.41 1.42 1.36
Total 13.3 13.6 13.6 12.9 13.7 13.6 13.7 15.3
BRALORNE Au DEPOSIT, B.C. 323
TABLE 2. (Cont.)
Chlorites Hornblende and others
Sample C182-2 C116-1 84-49/795 C093A C094A C093A
Type RIP P'TH RIP C'PH RIP APR HBL PREH EP
(,0 (7) (2) (5) (5) (5) (6)
SiO2 25.70 19.74 26.13 17.47 27.44 24.83 52.74 44.32 41.23
TiO2 0.03 0.09 0.27 ND 0.10
AlcOa 22.94 18.49 23.94 15.17 19.63 19.89 4.07 24.66 30.21
MgO 18.38 13.5 ! 19.03 28.57 20.40 8.78 17.01 0.06 0.08
FeO 20.05 33.82 18.96 27.85 18.53 32.74 12.27 0.27 3.88
CaO 0.07 0.06 11.48 27.17 23.12
NaO 0.09 0.14 0.71 ND 0.79
KO 0.15 0.07
HO 11.88 10.33 11.91 10.94 11.56 10.73
Total 98.95 95.89 100.3 100.0 97.85 96.97 98.62 96.49 99.42
Fe/Mg 0.38 0.59 0.36 0.35 0.33 0.67
Cations
Si 5.19 4.58 5.25 3.83 5.68 5.55
A1 5.51 5.06 5.71 3.92 4.78 5.24
Mg 5.87 4.67 5.75 9.34 6.37 2.93
Fe 3.35 6.56 3.20 5.11 3.19 6.12
Total 19.9 20.9 19.9 22.2 20.0 19.8
Operating conditions and standards used are in given Leitch (1989); chlorite classifications from Hey (1954)
Sample locations are given in text
Rock types: C093A, SB84-49/795, 812, 19-51FW1, Cl16-1 = diorite; C094A, C182-2 = soda granite
(n) -- number of separate sites analyzed; where no number is given, only one analysis was obtained; carbonate compositions are
plotted in Figure 8; Ab, An, Or = mole fractions of albite, anorthite, orthoclase, respectively; not reliable for calcites due to low Fe
and Mg contents; blanks indicate elements not analyzed for; ND = not detected; Fe/Mg -- total Fe (as FeO) divided by (total Fe as
FeO, plus Mg); H20 contents assigned by difference; chlorite compositions are plotted in Figure 10
Abbreviations: APR -- aphrosiderite; C'PH -- corundophillite; EP -- epidote; HBL = hornblende; PREH = prehnite; P'TH
= psuedothuringite; RIP = ripidolite
pellets but with checks on fused discs (precisions var-
ied from 5% for major elements to 20% for certain
minor elements; details of procedure, error analysis,
and interlaboratory comparisons are in Leitch, 1989).
Complete whole-rock chemical analyses and norms
of altered and unaltered host rocks are in Leitch and
Godwin (1986, 1987).
Petrography and Chemical Composition
of Altered Wall Rocks
A striking feature of the Bralorne-Pioneer deposit
is the intense wall-rock alteration that accompanied
vein formation and gold deposition. Vein envelopes
vary from less than 0.1 up to 10 m wide, and in places
coalesce to 50 m thickness. The dominant alteration
is carbonitization, although silicification and sericit-
ization commonly occur adjacent to the veins. Biotite
alteration, rare in the Bralorne mine, is more common
at the Pioneer and at the P. E. Gold property, 2 km
southeast of the Pioneer. Mineralogical and textural
features were observed in about 300 thin and polished
sections in ten detailed traverses across altered wall
rock at various levels from surface to 2-km depth.
Visually estimated modal compositions of the altered
rocks are in Table 1.
Mineralogical zoning
Unaltered wall rocks display sub- or lower green-
schist facies metamorphic mineral assemblages, which
set broad P-T limits of 2 to 3 kbars and 325 to 375C
(Leitch, 1989). Altered wall-rock envelopes adjacent
to the veins including occasional biotite indicate sim-
ilar or slightly higher temperatures. The three prin-
cipal rock types hosting veins are Bralorne diorite,
soda granite, and Pioneer greenstone. These three
rock types vary in their response to alteration, but in
general there is a consistent zoning of alteration min-
erals, with an associated progressive destruction of
texture, over about 5 m from unaltered wall rock to
the vein. The sequence is usually from (1) a green
chlorite-epidote outer zone that faithfully preserves
the original rock textures, to (2) a buff carbonate-al-
bite + sericite intermediate zone, divided into outer
texture-preserved and inner texture-destroyed sub-
zones, to (3) a foliated, cream-colored quartz-sericite
(_+ fuchsite)-carbonate innermost zone that locally
becomes a "paper schist." The composition of the
carbonate changes from calcite to ankerite as the vein
is approached. The zonal arrangement is similar to
that described by Robert and Brown (1986b) at the
324 LEITCH, GODWIN, BROWN, AND TAYLOR
Sigma mine in Quebec, by Kishida and Kerrich (1987)
at the Kerr-Addison mine in Ontario, and by Landfeldt
(1987) and Albino (1987) for the Mother Lode district
in California.
The chlorite-epidote outer zone is gradational out-
ward with surrounding rocks of greenschist facies
metamorphism. It can also be confused with an early,
widespread, barren quartz-epidote-carbonate-
prehnite stockwork alteration in the diorite and soda
granite. Chlorite occurs as fine (0.05 mm or less), dark
green, scaly aggregates usually replacing primary
hornblende in the diorite and soda granite and com-
monly intimately intergrown with lesser amounts of
calcite and sericite. Chlorite in the diorite lacks pleo-
chroism or anomalous interference colors; in contrast,
chlorite in the soda granite is coarser (to 0.5 mm),
bright green and pleochroic and displays strong pur-
ple to blue anomalous interference colors. These op-
tical characteristics suggest that the chlorite in the
diorite is a prochlorite with higher Mg and A1, and
less Si than the variety found in the soda granite that
may be more Fe-rich pennine (Kerr, 1959). This is
supported by the mineral chemistry (Table 9.) and re-
fleets the more marie (Mg-rich) character of the diorite
compared to the soda granite. Epidote usually appears
to be early and is replaced by chlorite, carbonate,
sericite, and quartz. It most commonly occurs as fine
(0.01-0.1 ram), anhedral, semitranslucent grains. The
common brown color in plane-polarized light appears
to be due to an alteration mineral of unknown identity;
the epidote is darkest where it is most altered. Massive
epidote (clinozoisite) locally forms a creamy, dense
(specific gravity about 3.1), hard rock that looks like
a separate intrusive phase but is actually an alteration
product. These areas appear to be the product of un-
usually intense, early barren quartz-epidote-carbon-
ate-prehnite stockworking that escaped later altera-
tion.
The carbonate-albite-sericite intermediate zone is
divided into an outer texture-preserved and an inner
texture-destroyed subzone. Carbonates (up to 65%)
are usually the most abundant alteration minerals in
the intermediate zone in altered diorite but are less
common in altered soda granite (Table 1). Albite may
make up to 45 percent; chlorite remnants from the
outer zone may not be completely replaced in the
intermediate zone, forming up to 15 percent of the
rock. The outer subzone, marked by the onset of car-
bonate alteration as a vein is approached, is distin-
guished in the diorite as a buff calcite replacement of
hornblende crystals (Fig. 3a) that preserves the orig-
inal rock texture but makes the rock look like a dif-
ferent phase. This transition from the outer chlorite-
epidote zone to the onset of carbonate alteration
roughly corresponds to the "cryptic-visible" bound-
ary of Robert and Brown (1986b). The buff or creamy
inner subzone is marked by progressive destruction
of original igneous texture toward the vein. This sub-
FIG. 3. (a). Thiu sectiou view of outer subzone of central zone
aheration (ab = albite, ca = calcite, eli = ohiorite, ms = muscovite),
3 m in the hanging wall, 79 vein, 41 level. Width of field of view
is I cm; crossed polars. (b). Thin section view of inner subzone
of central zone alteration, 0.2 m into footwall, 79 vein, 32 level.
Large porphyroblastie grains of calcite with brigliter ankerite rims
are set iu a matrix of sericite and calcite, plus minor chlorite that
is dark. Width of field of view is 1.2 cm; crossed polars.
zone commonly weathers a rusty orange-brown due
to increasing amounts of ankeritic carbonate (which
does not react to cold dilute HC1) as the vein is ap-
proached. The range of ankerite-calcite ratios is il-
lustrated in Figure 4a to d (iron and magnesium con-
tents are confirmed by microprobe analyses in Table
2). Ankeritic carbonate forms anhedral interlocking
grains ranging from 0.1 to 0.5 mm in diameter. Com-
posite grains with clear calcite cores and ankerite rims
are common (Fig. 3b). The lack of a strong X-ray peak
for ankerite (Fig. 4a and b) in many samples may be
due to the volumetrically minor ankeritic rim to the
composite grains. Calcite is the only carbonate in late
reopenings at the centers of major veins, in crosscut-
ting veinlets, and in the early barren quartz-carbon-
ate-epidote-prehnite stockwork. It may be of explo-
'ration significance that only the main-stage (ore-re-
lated) carbonate is ankeritic. Coarse secondary albite
BRALORNE Au DEPOSIT, B. C. 325
5[3 84-49 / 794.5'
, I , , I , , , I
31 30 29
5B 8i-49/795'
31 30 29
b
19-51 FW1
31 3 2
C
C080
,,,1.
3 30 28
d
FIG. 4. X-ray diffraction scans of carbonate minerals from the
intensely altered inner zone. (a). Sample SB-84-49 at 794.5 ft.
(b). The same hole at 795 ft. Both show a small peak for ankerite
compared to calcite. (c) Sample 19-51 FW1. It shows a larger
ankerite peak. (d) Sample C080. It shows that ankerite is m. ore
abundant than calcite. Full-scale deflection is 4 X 102 counts in
(a) md (b) but l03 counts in (e) and (d).
long. It completely replaces original plagioclase and
forms up to 30 percent of the rock. Bright green
fuchsite is locally present in altered diorite and forms
up to 5 percent of the rock, usually rimming cores of
chromite. Green mica, variably identified as Cr-bear-
ing, is also known in Archean mesothermal gold vein
deposits such as the Casa Berardi (Pattison et al.,
1986) and Lac Shortt (Morasse et al., 1986), both in
Quebec. Carbonate includes siderite, ankerite, and
minor calcite. The modal estimates of the carbonate
minerals in Table 1 are tentative because of the dif-
ficulty in distinguishing them in thin section; even
SEM studies failed to identify them confidently, and
microprobe analyses were required.
Quartz-albite silica-flooded zones form in the cen-
tral portions of the soda granite and in albitite dikes.
This alteration is fracture controlled and grades in
is next in abundance after carbonate and forms grains
up to 0.5 mm across, in veins, vein envelopes, or with
replacement textures that include "chessboard,"
"patchwork," and "irregular" (untwinned) albite
(Battey, 1951; Leitch, 1981). Albite alteration is more
prevalent in the soda granite, reflecting the differ-
ences in original composition between the more marie
diorite and the more felsic soda granite.
The quartz-sericite-carbonate innermost zone
consists of intense replacement by secondary quartz,
sericite (_ fuchsite), ankeritic carbonate, and pyrite
adjacent to the vein. This zone, up to 50 cm wide, is
schistose (Fig. 5a and b) or occasionally brecciated.
Fine (0.1 mm) quartz forms up to 60 percent of the
rock, although some of this, especially in the soda
granite, is merely isochemically recrystallized.
Chemical analyses corrected for volume changes
confirm this observation, although the inclusion of
minor quartz veins in the samples cause fluctuations
in SiO2 content (cf. Sketchley and Sinclair, 1987a).
Sericite (muscovite: analyses in Table 2) occurs as pale
green to yellowish masses of flakes less than 0.05 mm
.
'L ' '"P' ":*', ' -.. :'' .. :*% ..
C. 5. (a). Section o[core around the 5] vein on the
level, sowin the progressive destruction o[ texture in both te
soda ra)ite (top) an iorite (bottom) [roma coarse-raine,
msive rote to a [oliateS, ki!y uz-sedcite-earbonate-altered
rock in the inner zone. The position of the vein is at the gap at
the center of the box. (b). Thin section view of foliated, intensely
quz-seeite-carbonate-altered inner zone rock, from immediate
footwall of the 51 vein, 8 level, near the Empire sh. Field of
view is 1 em wide; crossed polars.
326 LEITCH, GODWIN, BROWN, AND TAYLOR
intensity from a crackling or sparse stockwork of
quartz-sericite-carbonate-chlorite-epidote-albite mi-
croveinlets, to a breccia composed of a network of
the same minerals enclosing fragments of soda granite,
to a rock composed almost entirely of quartz and al-
bite. Accompanying changes in feldspars proceed
from chessboard albite through veinlet albite to
stockworked soda granite (Fig. 6) and finally to an
intensely silicified white rock composed of an unusual
symplectic intergrowth of quartz and albite that lo-
cally destroys the original texture (Fig. 7). These in-
tergrowths are up to 3 mm across, with a radiating,
crudely hexagonal pattern, probably controlled by
quartz crystallography because the quartz and albite
crystallized simultaneously. Since the alteration is
later than the albitite dikes, it appears to be hydro-
thermal. However, there are no well-defined quartz
veins or gold values associated with these silica-
flooded zones.
Biotite envelopes occur at Bralorne only near the
surface in green hornblende porphyry dikes near the
51 vein on the 4 level and in altered diorite below
1,700 m depth around the 77 vein. It is more common
to the southeast on the Pioneer (Joubin, 1948) and
P. E. Gold (Nordine, 1983) properties where the Pi-
oneer greenstones, prevalent on these two properties,
alter more readily to biotite. The occurrence of biotite
may be similar to that at the Sigma mine in Quebec
(Robert and Brown, 1986) where biotite is prominent
as an alteration mineral in the lower levels of the mine.
Consequently, the Pioneer and P. E. Gold exposures
might represent deeper levels of the vein system than
at Bralorne.
Black calcite replacement is a distinctive alteration
present only locally in the soda granite, crosscutting
the main-stage alteration. It begins along hairline
fractures and may replace the whole rock. The black
FIG. 6. "Crackled" soda granite, cut by a stockwork of fine-
grained quartz, sericite, calcite, epidote, pyrite, and minor ohio-
rite, from drill hole 8E-1174 at 880 ft. Crossed polars, field of
view 1 era.
FIG. 7. Primary texture of soda granite replaced by symplectic
growths of quartz and albite in a radiating, psuedohexagonal pat-
tern (DDH SB-14-80/93 fl.). Field of view is 1 era; crossed polars.
and nearly opaque character is caused by myriads of
extremely fine (1-2 jam) opaque inclusions that may
be amorphous carbon. Tests by X-ray diffraction,
SEM-EDS, and electron microprobe failed to confi-
dently identify this material.
Garnet-quartz-calcite-pyrite and quartz-tourmaline
(schorl) alteration was found in specimens collected
from Pioneer volcanics in the Pioneer mine by F. R.
Joubin, which are now in the University of British
Columbia Economic Geology collection. Neither al-
teration facies is common, although tourmaline is
abundant in late fractures cutting the extensional 27
vein (Joubin, 1948). The presence of borosilicate may
be significant; although apparently rare at Bralorne,
it is common in mesothermal systems in the Canadian
Shield (e.g., Sigma: Robert and Brown, 1986a).
Mineral chemistry
Chemical zonation in carbonates is from calcite in
the outer and intermediate zones to ankerite in the
inner zone (Table 2 and Fig. 8). In the least altered
rocks (sample C094A), the carbonate is almost pure
calcite (0.97 mole fraction CaCOa). Similar calcite is
found in the intermediate zone (sample SB-84-49, 812
ft depth), 3 m away from the major 5lB vein. SEM-
EDS studies revealed that inner zone carbonates are
composite grains with calcite cores rimmed by an-
kerite, or less commonly, ankerite cores with siderite
rims (Fig. 9a and b). The fine grain size (equal to the
beam diameter of the probe) causes contamination of
some ankerite analyses by calcite and of siderite anal-
yses by ankerite (Fig. 8), resulting in the apparently
metastable plotted points (cf. Anovitz and Essene,
1987). Adjacent to the vein (sample SB-84-49, 795
ft depth), much of the carbonate is ankerite, with the
cores of 0.03-mm-diameter grains (spots 3 and 7, Fig.
9a and b; analyses in Table 2) being about
Cco.Dolo.aSido. (Cc, Dol, and Sid indicate mole frae-
BRALORNE Au DEPOSIT, B. C. 327
Calc le
SB 84-49/795 812'
19-51FW1, Cll 6-1
Ca( O 3 CALCITE
Ca[c te
DOLOMITE
cores
SB 84-49 / 795'
Ankente
r., m .s_._.../
ANKERITE
SI DERI TE _g] S,der,te r,ms
SB 84-49/795 . -
MC
..
FeCO 3
FIG. 8. Triangular diagram showing fields of carbonate compositions from altered wall rocks (atomic
proportions). Analyses are given in Table 2. Stability fields at 400C (stippled) for calcite and siderite
(calcite group) and for ankerite (dolomite group) are from Anovitz and Essene (1987).
tion CaCO3, MgCO3, and FeCO3, respectively). The
rims of these grains are only 5 m thick but are dis-
tinguished in thin section by higher relief than the
cores. These rims (spots 8 and 2, Table 2) are Mg and
Fe rich (Cc0.0s_0.3.sDo10.sSid0. Ls_0.45), and plot as sid-
erites (Fig. 8) with higher Fe/(Fe + Mg) of around
0.4 than the 0.1 to 0.2 ratios typical of the ankeritic
core. However, even adjacent to this major vein a
few grains of calcite are present (Table 2, spot 9; see
also sample Cl16-1 adjacent to the 79 vein on the
41 level). Analyses from the inner zone around the
51 vein on the 19 level (sample 19-51 FW1, Table
2) are similar to those just discussed. Calcite probably
formed initially but was progressively replaced with
iron-magnesium carbonate by advancing alteration;
calcite formed simultaneously farther from the vein.
This explains the bulk zoning and zoning in the car-
bonate grains from Fe-Mg-rich rims to Ca-rich cores.
The Ca, Fe, and Mg in the carbonates was probably
derived from primary hornblende. However, a portion
of the Ca, Fe, and Mg became available for deposition
elsewhere, since total Ca, Fe, and in particular Mg,
show depletion near the veins (see below). Zoning of
calcite farther from, and ankerite closer to, mesother-
mal gold veins has been described at other deposits.
It is noted at the Yellowknife district (Boyle, 1961);
Lac Shortt, Quebec (Morasse et al., 1986); Casa Ber-
ardi, Quebec (Pattison et al., 1986); the McDermott
deposit at Kirkland Lake, Ontario (Workman, 1986);
Cameron Lake, Ontario (Melling et al., 1986); and
the Red Lake district, Ontario (MacGeehan et al.,
1989.).
Feldspars are ubiquitous, almost pure albite, as
confirmed by X-ray, SEM, and microprobe (Table 2);
no K feldspar was found. The average albite in least
altered rock (samples C093A and C094A) is Ab6_,
with a negligible orthoclase molecule; this may be
greenschist facies metamorphic albite. There is a
slight increase in the albite molecule, to Ab,_00, in
the texturally distinct alteration albite in and around
veins (samples 19-51 FW1 and C182-2, respectively,
in Table 2). The very low anorthite component of
both the metamorphic and the hydrothermal plagio-
clase implies a formation temperature of less than
400C (Winkler, 1971). This agrees with the esti-
mates of ore fluid temperatures of 350C from fluid
inclusion and sulfur isotope studies (below). The nar-
row range of compositional variability observed sug-
gests thorough homogenization of the feldspars by
metamorphic and hydrothermal fluids.
Micas in all the vein envelope zones (inner to outer)
are primarily sericite (muscovite). Sericites analyzed
were from various rock types and various parts of the
deposit (sample SB-84-49, 795-ft depth in green
hornblende porphyry, C 182-2 in crackled soda gran-
ite, 19-51 FW1 and Cl16-1 in diorite: Table 2). They
all have similar compositions, with low Fe, Mg, and
Ti contents. Only one analysis of biotite is reported
328 LEITCH, GODWIN, BROWN, AND TAYLOR
FIC. 9. (a) and (b). SEM-EDS backscattered pictures of com-
posite carbonate grains in sample SB-84-49 at 795 ft, found in
the inner zone alteration of diorite in the hanging wall of the 5lB
FW vein near the 5 level. The grains show ankerite cores (gray;
spots 3 and 7) and siderite rims (brighter due to higher atomic
number of Fe compared to calcium and magnesium; spots 2 and
8). Dark spots are holes; analyses are given in Table 2.
(Table 2), from an unusual biotitic envelope in a green
hornblende porphyry dike on the 4 level near the
5lB vein.
Chlorite, distinguished optically, is pennine or
prochlorite; the two types are also chemically distinct.
Pennine, present only in the least altered soda granite
(C094A), is a high Fe chlorite with Fe/(Fe + Mg)
about 0.7 and Si (cations) about 5.5 that approximates
aphrosiderite, a variety ofripidolite (Table 2 and Fig.
10). A few analyses plot as pseudothuringite, with Si
cations about 4.7 and Fe/(Fe + Mg) about 0.6. Pro-
chlorite, more abundant than pennine, occurs both
in unaltered diorite (sample C093A: Si cations about
5.6-6.0) and strongly altered diorite (C 116-1) and in
soda granite (C182-2: Si cations about 5.0 to 5.4). It
is richer in Mg with Fe/(Fe + Mg) about 0.35, i.e.,
between ripidolite and pycnochlorite (Fig. 10); one
analysis plots as corundophillite. The general ripi-
dolite-pseudothuringite compositions are similar to
those described by Robert and Brown (1986b) for
chlorite at the Sigma gold quartz vein deposit in
Quebec.
The original, possibly metamorphic, prochlorite in
the diorite is thus more magnesian than that in the
soda granite. This magnesian chlorite is also stable in
the presence of hydrothermal fluids, as it is present
in altered diorite and soda granite. It implies a high
Mg/Fe ratio in the hydrothermal fluids, which is sup-
ported by the results of computer modeling (below).
Analyses of hornblende (Table 2), the precursor min-
eral to the chlorite in the diorite, also show Mg/Fe
ratios greater than unity. The similarity of Mg/Fe ra-
tios in the precursor hornblende and alteration chlo-
rite suggests that the chlorite replaced some horn-
blende grains on a one-for-one basis, although in other
grains, some Fe, Mg, and Ca cations were trapped as
carbonate or released to the fluid.
Chemical changes
Losses and gains of elements during alteration,
corrected for volume changes (Leitch and Day, 1990),
were determined by Gresens' (1967) technique from
whole-rock analyses (Leitch and Godwin, 1988) along
ten detailed traverses across envelopes in the footwall
and hanging wall of major veins. Traverses were made
in both soda granite and diorite host rocks, from sur-
face to the 44 level, an interval of 2,000 m. This cor-
0'8
0'67
6-O
FIG. 10. Classification of the observed chlorite compositions,
according to the scheme of Hey (1954). A -- aphrosiderite, filled
circles -- least altered host rocks, crosses = altered rocks.
BRALORNE Au DEPOSIT, B. C. 329
responds to studies elsewhere (Kerrich and Watson,
1984; Robert and Brown, 1986a; Sketchley and Sin-
clair, 1987).
Two examples of the chemical changes adjacent to
the major veins are in Figure 11a and b. There are
no significant differences in gains and losses between
the footwall and the hanging wall. Other traverses
(Leitch, 1989) show the same general patterns. The
main changes within the alteration envelopes on ap-
proaching the vein, with mineralogical explanations,
are (1) loss of Na20 and gain of K20, due to destruc-
tion of albite in the inner zone and sericite replace-
ment; (2) loss of MgO and FeO3, due to destruction
of amphibole and replacement by sericite and car-
bonate; (3) net loss of SiO in the inner and inter-
VOLUME
FACTOR
VOLUME
FAC TO R
Sampl e.
160t AS ..... /No.
/ \",, coo3/2
1.20.__ .......... ................. _'__ '1
O'801 .""- /3
AI203 TiO 2 Zr' Avg.
O/ ..........
AI203 Ti 02 Zr Avg,
Fe203
2 2
-2
-
K2 o
0:5 b
Dstance toVe,n (m)
K20
i i i i i
2 4 6 8 10 1
Distance to vein (m)
FIG. 11. Oresens' plots of weight percent loss or gain of oxides,
plotted versus distance from the vein, from program ORESPLOT
(Leitch and Day, 1990). Volume factor is based on alumina and
titania contents in sample C093, the least altered diorite host,
from drill hole UB-81o17 at 350 to 400 ft. (a). Hanging wall of
the 51 vein, 8 level, near the Empire shaft. (b). Footwall of the
51 vein, 15 level, near the Crown shaft.
FIG. 12. Volume factor plot (ratio of final volume to initial
volume for immobile components) for sample series C003 (a) and
C032 (b). Avg = average of volume factors for AI2Oa, TiO2, Zr:
losses and gains are plotted in Figure 11.
mediate zones of the vein envelope--this is variable,
due to quartz veinlets being included in some samples
as the major vein is approached; and (4) increase in
loss on ignition (L.O.I.) and CaO due to development
of sericite and carbonate in all zones. Increases are
not always linear, depending on composition of the
original host rock and the amount of included quartz
vein. CaO commonly increases in the outer envelope
as calcite becomes stable and decreases in the inter-
mediate and inner envelopes as ankerite becomes sta-
ble. The parallel increase of L.O.I. and the calculated
volume factor implies that the changes in volume es-
timated from immobile elements are real, and there-
fore, the calculated losses and gains in other elements
reflect real processes.
The similarity of detailed traverses down to the 44
level indicates that there is no major change in en-
velope alteration chemistry from the surface to a
2,000-m depth; thin section examination supports this
conclusion. Differences between the way diorite and
soda granite react to alteration are slight. In the al-
tered soda granite there is a smaller loss of FeO3 and
MgO, since there are only minor marlcs to destroy,
and a greater loss of NaO, since there is more albite
to destroy. Changes in CaO are less noticeable in the
soda granite, again since most of the CaO is in mafic
minerals in both soda granite and diorite, and the
former is less mafic.
Diagrams of losses and gains (Fig. 11) demonstrate
the relative immobility of A1203 and TiO, the com-
ponents used to estimate volume changes (Gresens,
1967). An example of the volume changes or "volume
factor" involved in alteration, for the altered series
330 LEITCH, GODWIN, BROWN, AND TAYLOR
of Figure 11a, is given in Figure 12a. In this case
there is an initial volume decrease in the intermediate
zone sample (C003/3), possibly due to destruction of
albite and hornblende. This is followed by a volume
increase in the inner zone samples immediately ad-
jacent to the vein (C003/1,2), caused by strong car-
bonate alteration as shown by correlated increases in
L.O.I. and CaO (Leitch and Day, 1990). Figure 12b
also shows the typical agreement in using A1203 or
TiO2 in estimation of volume change and the consis-
tent overestimation of volume factor using Zr. In such
cases, Zr is rejected from the computation of the av-
erage volume factor used in the profiles.
Minor elements also show patterns. There is a
marked increase in S, As, and Sb as the veins are ap-
proached (Leitch and Godwin, 1988). This is asso-
ciated with the addition of pyrite, arsenopyrite, and
tetrahedrite (like COz, HzO, and KzO) in immediate
vein envelopes, by hydrothermal fluids rather than
mere remobilization (like CaO, MgO, and FezO3: cf.
Ludden et al., 1984). Chalcophile elements, such as
Cu and Zn, show a slight decrease toward the vein or
are constant. Lithophile elements follow the patterns
of certain major elements due to substitution: Ba and
Rb follow K20, whereas Sr mimics CaO. Yttrium, like
Zr, shows mainly immobile behavior, falling as the
volume factor increases and vice versa. Cr and Ni,
dependent on the amount of mafic inclusions in the
diorite, show correlated behavior. Co and V do not
show discernible patterns. Elements present at or near
the detection limit (Pb, Nb, Mo, Ag, and W) are er-
ratic.
These patterns are broadly similar to those detailed
by Boyle (1979) in his study of the chemical changes
attending alteration of the diorite at Bralorne (cor-
rected only for density differences, not for volume
changes). The changes are also similar to those for
the Yellowknife deposits (Boyle, 1961), which show
a marked increase in KzO/NazO and a consistent de-
crease in SiO2/COz and SiOz/L.O.I. near the veins.
Similar chemical profiles are reported for mesother-
mal gold vein deposits at Macassa in Quebec (Kerrich
and Watson, 1984), Kerr-Addison in Ontario (Kishida
and Kerrich, 1987), Wawa in Ontario (Studemeister
and Kilias, 1987), Cameron Lake in Ontario (Melling
et al., 1986), and the Mother Lode in California (Al-
bino, 1987). Most vein-type Archcan gold deposits in
the Canadian Shield have similar patterns of alteration
(Kerrich, 1983).
Fluid Inclusion Studies
Samples for fluid inclusion studies were collected
from quartz veins at surface and over a 2-km depth.
Sites were the same as those for oxygen isotope sam-
pies. Vein quartz forms euhedral crystals up to i cm
long (Fig. 13) that are outlined by concentric growth
../laye ring
sulf des i' calci t e
quartz k../clear quartz
J (..7-- lcrn
cloudy quart z-- (
s,lfides-- '/ I 0.5 C rn I
(b)
FIG. 13. Sketch of textures in quartz from the main, ribboned
shear veins in the Bralorne deposit; vein walls are oriented east-
west in both diagrams. (a). Growth zoning in coarse euhedral
quartz crystals that grew perpendicular to vein walls and ribbons
(Woodchuck vein, 3 level adit portal). (b). Growth zoning in large
euhedral quartz crystals (79 vein, 41 level, Queen mine area,
hole Q-144 at 201 ft. Note also the small clear euhedral crystals
enclosed in the larger crystals.
zones of minute (1-2 gm diameter) primary fluid in-
clusions (Fig. 14). None of these inclusions were large
enough for microthermometry. Instead, larger iso-
lated (assumed primary) and fracture-controlled
(pseudosecondary) inclusions were studied (Fig. 15).
Fluid inclusions were measured from the two main
paragenetically distinct types of quartz found in the
veins: gray main-stage quartz adjacent to sulfides, and
clear quartz, both earlier and later. The main mass of
milky quartz provided useable fluid inclusions only
in local areas where it was clearer with fewer but
larger inclusions or along fracture planes where larger
inclusions were present. Fluid inclusions were also
measured in late-stage calcite, mainly as ribbons in
the quartz veins, less commonly in apparently con-
temporaneous calcite grains included within the
quartz veins. No particles of gold were seen directly
associated with any fluid inclusions.
Preliminary fluid inclusion data for the Bralorne
deposit were obtained with Chaixmeca equipment
(Leitch and Godwin, 1987, 1988). Further studies,
carried out with a FLUID Inc.-adapted U.S.G.S. type
gas-flow stage calibrated to better than _0.4C from
-56.6 to +660.4C, characterize the compositions
of the fluids in the inclusions in greater detail. Mole
fractions of carbonic fluid are derived from visually
BRALORNE Au DEPOSIT, B.C. 331
FIG. 14. Growth zones outlined by primary fluid inclusions
in quartz from the 79 vein on the 32 level. Note hexagonally
arranged 1-tam primary inclusions and swarms of 3- to 5-tam see-
ondary fluid inclusions on microfractures, that give the quartz its
milky color. Width of field of view (a) is 2.5 ram, 1.25 mm in
enlargement of lower right-hand portion (b).
estimated vol percent at 25C (Fig. 16). Although
there are uncertainties inherent in this method, pre-
cise calculation by the method of Parry (1986) does
not change the results significantly for the more com-
mon inclusions wih modest carbonic contents. Details
of calibration and error analysis are in Leitch (1989).
Classification and description of fluid inclusions
Inclusions were classed as primary, pseudosecond-
ary, and secondary by the criteria in Roedder (1984;
of. Hollister, 1981). Results for the major types of
inclusions suitable for measurements are summarized
in Table 3 and Figure 16. The visual division into
primary and pseudosecondary is supported by the
general bimodal distribution of Th and T data in
Table 4 (summarized from histograms in Leitch and
Godwin, 1988).
Primary (type 1) inclusions display spheroid or
negative crystal shapes in quartz (Fig. 15), or are
rhomboid in calcite. They range from 3 to 60 /m
across and occur as isolated inclusions or in discrete
clusters not in linear arrays. Many contain three visible
phases (Fig. 17): an outer aqueous liquid, an inner
carbonic liquid, and an innermost carbonic vapor
bubble. Some appear to be simple two-phase inclu-
sions since small amounts of carbonic fluid rimming
the vapor bubble are hard to see (but are confirmed
by CO or clathrate melting). The presence of another
component in the type 1 primary inclusions, probably
mostly methane (CH4) but possibly including trace
amounts of hydrogen sulfide (HoS), is indicated by
the depression of the melting point of CO9.
Type la inclusions in quartz, marked "P" in Figure
16, form the majority of the three-phase inclusions
FIG. 15. Isolated, and therefore assumed primary, type la
COu-bearing inclusions of typical 5- to 10-tam size, in quartz from
breccia portion of the 51 vein on the 8 level. Trails of pseudo-
secondary and secondary inclusions (types 2 and 3) are 3- to 5-
tam size (width of field of view is 200 tam). Higher vapor/liquid
ratios distinguish the type I inclusions, which may be either neg-
ative crystal shape as in (a) or rounded as in (b). Irregular inclusions
in (b) are type 2 (lower vapor/liquid ratio).
332 LEITCH, GODWIN, BROWN, AND TAYLOR
20
15
-70 -65 -60 -55
Tmco 2
I.Id .
o ,o , o , o
Thco 2
3O
20
10
0 -1 -2 -3 -& - -
Tm ic
30
20
0
, [- ,
o' 5 10 15
Trnc lat hrat
-4O -35 -30 -25 -20
Teut ectc
0 10 20 30 40 50 60 70 60 90 10C
V/V+ L (vol %)
FIG. 16. Summary of data for carbonic fluid inclusions for the Bralorne deposit. P, PS, S correspond
to primary, pseudosecondary, and secondary inclusions, respectively; black areas to methane-rich (type
lb) inclusions. All inclusions are in quartz except for hachured areas which are in calcite. Other ab-
breviations are given in Table 3.
and usually homogenize to liquid. They contain mod-
est amounts of carbonic fluid (0.1-0.4 mole fraction
of the total contents of the inclusions, with a mode of
0.15) and consist of an aqueous phase, carbon dioxide,
and minor methane (Xco2 = 0.10, Xcu4 = 0.05). The
volume percent of carbonic fluid represents the H20-
CO2 ratio at the time of trapping. Variations in car-
bonic fluid could be due to any one of, or a combi-
nation of, necking, trapping of immiscible fluids, or
fluid mixing.
Type lb inclusions in quartz, shown in black in
Figure 16, are rare, with significantly higher carbonic
fluid contents than type 1 a; they usually homogenize
to vapor. They are vapor rich and contain a minor
aqueous phase plus abundant methane and carbon
dioxide (0.3-0.9 mole fraction carbonic component
of the total contents, with a mode of 0.5, consisting
of Xcu4 = 0.25, Xco2 = 0.25). They occur both in
clear quartz that may be fragments included in the
main-stage milky quartz, or in main-stage quartz a
few millimeters from, but not mixed with, the more
common type la inclusions. The high CH4 contents
of type lb inclusions are evidenced by strong depres-
sion of COe melting points to below -60C (Fig. 16).
The COe homogenization temperatures of type lb
inclusions are lower than those of type la, but type
lb densities are lower (Table 3) because of high CH4
contents.
Primary inclusions in calcite (hachured in Fig. 16)
have homogenization temperatures and fluid com-
positions similar to pseudosecondary (type 2) inclu-
sions in quartz. They generally have a minor COe
content and occur as isolated inclusions or as clusters
not related to fractures. Because most calcite is para-
genetically late, it seems reasonable that the primary
inclusions in calcite and the pseudosecondary inclu-
sions in quartz both represent the same phase of later,
lower temperature, more dilute fluids.
Pseudosecondary (type 2) inclusions are the most
common in the Bralorne vein quartz, arranged along
myriads of tiny fractures that crisscross the quartz
grains with a brushlike or wispy texture. This texture,
together with the presence of the primary three-phase
inclusions, indicates a mesothermal environment
(T. J. Reynolds, pers. commun., 1987). Many of these
fractures do not cross grain boundaries; thus, the fluid
inclusions in them represent fluids trapped in frac-
tures during growth of the quartz host. Type 2 inclu-
sions in quartz, marked "PS" in Figure 16, homog-
enize to the liquid phase and contain no methane and
only minor carbon dioxide (mode of less than 0.1 mole
fraction of the total contents) that may not be visible
as a separate phase but is inferred from clathrate
melting temperatures. They range from 3 to 20
across, averaging about 5/zm, and have rounded to
irregular shapes (Fig. 18) or flattened lensold shapes
which may display the highly irregular outlines typical
of necking. Even if these inclusions have necked, they
still have homogenization temperatures similar to
others in the same fracture.
Secondary (type 3) inclusions in both quartz and
calcite are localized along through-going fractures,
BRALORNE Au DEPOSIT, B.C. 333
TABLE 3. Summary of Fluid Inclusion Characteristics
for the Bralorne Deposit, Southwest British Columbia
(all temperatures in C)
Type la primary: three-phase, CO2-bearing, CH4-poor
M.F.: H20 = 0.8, CO = 0.1, CH4 = 0.05, NaC1 = 0.03
Th: range = 235 to 425, average of modes 1 = 280
Thcol: range = 15 to 27, mode = 20.5
Tmo: range -- 8.2 to 11.0, mode = 9.6
Tm, (.9): range = -2 to -7, average= -3.5
Tmc%: range -- -57.3 to -59.8, mode = -58.8
V/(V + L): range = 10-45 vol %, mode = 20 vol %
r carbonic: 0.7 g/cm a
r bulk: 0.95 g/cm a
Vbar: 23 moles/cm a
Type lb primary: Three-phase, CO-CH4-rich
M.F.I: HO = 0.45, CO2 = 0.25, CH4 = 0.25,
NaCI = 0.03
Td: range = 230 to 330, mode = 270
Thco,: range = 3.3 to 9.0, mode = 7.5
Tmc: range = 11.1 to 13.0, mode -- 12.4
Tml (?): not observable
Tmco,: range = -61.3 to -66.5, mode = -63(?)
V/(V + L): range = 30-85 vol %, mode = 50 vol %
r carbonic: 0.6 g/cm a
r bulk: 0.8 g/era a
Vbar: 31 moles/cm
Type 2 psuedosecondary: Two-phase, HaO-liquid-rich
M.F.l: H20 -- 0.95, COa = 0.05, NaC1 < 0.01
Th: range = 160 to 260, average of modes 1 = 200
Thco,: range = 20 to 31, mode = 27.5
Tmo: range = 7.0 to 10.5, mode = 9.5
Trot (.9): range = -0.5 to -2.5, average = -1.8
Tracon: range = -56.6.to -57.5, mode = -56.8
V/(V + L): range = 5-15 vol %, mode = 10 vol %
r carbonic: 0.7 g/era a
r bulk: 0.97 g/cm a
Vbar: 20 moles/cm a
Type 3 Secondary: Two-phase, HO-liquid-rich
M.F.I: HO = 0.99, NaCI < 0.01
Th: range = 120 to 180, average = 150
Tm,: range = -0.1 to - 1.0, average = -0.5
V/(V + L): 2-7 vol %, mode = 5 vol %
r bulk: 1.00 g/cm a
Vbar: 18 moles/era a
1 Average Of Th histogram means in Table 4; mean value is 280,
standard deviation +__ 36, 276 measurements
2 Mode is from histograms in Figure 16
Abbreviations: M.F. = approximate mole fraction, Th = final
homogenization temperature, Td = decrepitation temperature,
Thco, = homogenization temperature of CO, Tmcl = final melting
poiut of clathrate, Tmt = filial melting point of ice, Tincol = final
melting point ofCO, V/(V + L) -- ratio of vapor bubble to liquid,
r = density, Vbar = molar volume
some oriented with their long axes perpendicular to
their controlling fracture (cf. Roedder, 1979). These
inclusions trapped fluid that was circulating after the
quartz had been deposited and fractured. They are
simple two-phase aqueous inclusions with small (5 vol
%) vapor bubbles, no detectable carbonic content,
and they homogenize to the liquid phase. They are
generally small (less than 10 tm), with lensoidal, an-
gular, or negative crystal shapes.
Homogenization temperatures
Homogenization temperatures (Th) for fluid inclu-
sions in quartz and calcite range from 120 to 440C.
Most inclusions homogenize to the liquid phase; only
a few to the vapor phase. This is unlikely to be due
to necking, because the Th for liquid and vapor is sim-
ilar. Inclusions homogenizing to vapor (mainly type
lb vapor- and CH4-rieh inclusions) commonly de-
erepirated before homogenization. For type la and
type 2 inclusions the few deerepitation temperatures
(Ta) measured were usually close to the Th values of
other inclusions in the same duster. Only a minimum
estimate of 230 to 330 (mode 270C) can be made
for the trapping temperature of inclusions that de-
erepirated.
Two main populations of Th are indicated in Table
4. The higher temperature group, with a total range
of 240 to 420C (avg 280C), corresponds to the
type 1 primary inclusions. The lower temperature
peak, at 180 to 220C (avg 200C), corresponds to
the type 2 pseudosecondary inclusions. Occasionally,
a weak third peak at lower temperatures (130 -
160C, avg 150C) corresponds to type 3 secondary
inclusions. A similar pattern of primary homogeni-
zations at 290 to 400C, with early secondary ho-
mogenizations at 250C and later secondary homog-
enizations at 190C, was observed at the Kolar deposit
in India, which is similar to Bralorne in strike (8 km)
and depth (3.2 kin) extent (Santosh, 1986). A similar
bimodal distribution in Th was found at the Big Hurrah
deposit in Alaska (Read and Meinert, 1986) and for
homogenization and ice-melting temperatures in the
mesothermal Au deposits of the Jungwon area of Ko-
rea (Shelton et al., 1988).
With one notable exception (sample 8-51B), in
which clear quartz (late vug crystals or early frag-
ments) has anomalously high homogenization tem-
peratures, there is a tendency for T h in primary (type
1) inclusions to increase with depth, from about 235
to around 330C. There is, however, overlap in the
temperatures and the zonation would not be con-
vincing without the corroboration of deerepitation
studies (Sugiyama, 1986; Table 4). The deerepitation
method suffers from difficulties relating data to spe-
cific inclusion types and to significant or variable
overshoot of Th depending on size of the inclusions
and the grains studied. However, the advantages are
that data may be rapidly gained from many inclusions,
too small to measure optically. Optically studied see-
tions from surface to the 2,000-m depth show general
agreement between Ta and Th (Leiteh, 1989). The
gradient suggested by the deerepitation data is about
334 LEITCH, GODWIN, BROWN, AND TAYLOR
TABLE 4.
Summary of Fluid Inclusion Data, Bralorne Gold Quartz Vein System, Southwestern British Columbia
Level (vein)
Depth (m)
Tillice (C) Th (C)
Pseudo- Psudo-
Primary secondary Secondary Primary secondary
Sample no. Type 1 Type 2 Type 3 Type 1 Type 2
Secondary
Type 3
0 (Lorne) C1013
(Cosmopolitan) C1002
0
3 (Woodchuck) C1024
(5lB) C1027
130
8 (5lB FW)
(51B Bxa)
(51B)
270
15 (51)
580
16 (51) 16-5 ivV(D
630
26 (85)
1,100
32 (79)
1,400
-2.o _ 1.5
-1.,5 + 1.0 (27)
-2.2 + 0.7 (11)
8-51B(V"vV) -2.0 + 1.1 (8)
8-51B -7 + 1..5 (11) -2.5 + 1.7 (39)
cos1-1/2
290+ 15(10) 230+30(5)
280 +_ 20 (7)
325-335
235 + 25 (21) 195 q- 25 (9)
250 + 30 (16)
345-380 175
310 i 60 (46) 225+ 15 (5)
270 q- 30 (26)
?200 (4)
350-430
140 q- 20 (16)
15-51(C) -2.5 q- 1.3 (9) 260 + 20 (21)
360-400
-2.3 q- 1.5 (20) -0.5 q- 0.2 (6) ?,280 (2)
-1.5 + 0.6 (6) e 220 + 25
Cl18-11 -1.6 q- 0.6 (11) 300 q- 40
365-385
(9)
(20)
220
190 q- 20 (4)
190
190 q- 20 (18) 160 q- 20 (5)
lS0
225 q- 25 (6)
230
Cl17-7 -4 + 1.0 (14) -1.5 q- 0.5 (8) 320 q- 60 (15) 190 _ 30 (3,2) 7140 (1)
-2.0 q- 1.o (9) e 205 _+ 35 (16) c
340-385 180
Cl17-5 -3.5 q- 1.5 (9) -1.5 q- 0.7 (18) 270 q- 40 (18) 180 q- 30 (26)
41 (79) Cl16-14 -3.5 q- 1.7 (15) -0.5 q- 0.2 (8) ?265 q- 15 (9) 215 q- 20 (5) 7150 (1)
1,8oo 57o-as5 21o
44 (77) C128-20 -5.0 (3) -1.5 q- 0.8 (11) 280 q- 50 (,25) 190 q- 40 (10)
2,000 C128-19 -5.0 q- 1.1 (9) -1.5 q- 1.3 (29) 330 q- 60 (21) 200 q- 40 (45)
410-425 200
T .... (melting point of ice) figures can only be regarded as crude estimates (see text for details); Tl = homogenization temperature
Data are reported as mean q- one standard deviation, followed by number of determinations in brackets; ? indicates uncertain data
(too few to calculate standard deviation); italic figures beneath Th are decrepitation data (Sugiyama, 1986), with Ta (start of decrepitation)
followed by a dash and Tp (peak of decrepitation)
Measured in calcite; all others from quartz
30C/km, which corresponds to a reasonable geo-
thermal gradient (cf. Goldfarb et al., 1988). The lower
temperature type 2 and 3 inclusions show a signifi-
cantly lower gradient of 0 to 10C/km.
Salinities (ice-melting temperatures)
Salinities of type 1 and 2 inclusions at Bralorne are
difficult to estimate because of the presence of vari-
able amounts of both carbon dioxide and methane (cf.
Collins, 1979). Thus the T,, e data for type 1 and 2
inclusions (5 and 3 wt % NaC1 equiv, respectively;
Table 4) overestimates the salinity of the aqueous
solutions using the equation of Potter et al. (1978).
If only carbon dioxide is present, the salinity can be
correctly estimated from clathrate-melting tempera-
tures (Bozzo et al., 1973). A clathrate-melting tem-
perature of 9.5C (mode from P, PS peak in Fig. 16),
suggests 1 wt percent NaC1 equiv for the type 1 and
2 inclusions in quartz and calcite. However, in type
lb inclusions, ice melting is not seen, and clathrate
melting is above 10C. This reflects methane and in-
dicates that the method of Bozzo et al. (1973) is not
applicable. Correlation between mole fraction CH4
and Tinice in type la and lb inclusions suggests roughly
equivalent salinities (cf. Linnen, 1985). Therefore the
salinity of the type 1 inclusions is between 1 and 5
(i.e., 3 _ 2) wt percent NaC1 equiv. Type 3 inclusions
have ice-melting temperatures just below that of pure
water (-0.5C), implying 0.8 wt percent NaC1 equiv
salinities. Vaguely detectable eutectic melting in in-
clusions in quartz (Fig. 16) shows modes at -20.5C
in type 2, suggesting only NaC1, and -23C in type
1, suggesting the presence of minor KC1 (Roedder,
1984). This conclusion is supported by the computer
BRALORNE Au DEPOSIT, B. (2. 335
;-.
.
,
.
FIG. 17. Large (30-50 #m) three-phase type la inclusions in
quartz from the 51B FW vein on the 8 level. The inclusions contain
an easily visible innermost carbonic (COg + CH4) vapor bubble,
surrounded by a carbonic liquid phase and an outer aqueous phase
(at 25C). In most of the inclusions, vapor/liquid ratios are about
20 to 50 vol percent; molar carbonic fractions are about 0.15 to
0.25. Width of field of view: 375 #m in (a), 200 #m in (b).
type' 2. Bulk densities, equivalent mole fraction CO2,
and molar volumes (V,) for the inclusions were es-
timated from Brown and Lamb (1986), assuming a
density of 1.02 g/cm a for a 3 wt percent NaC1 equiv
solution (Potter and Brown, 1977). Equivalent CO2
contents for inclusions containing both methane and
carbon dioxide were computed by Swanenberg's
(1979) method.
Pressure estimates
Entrapment pressures for primary and pseudosec-
ondary fluids, estimated by several methods, are
compared in Table 5. Independent estimates of trap-
ping temperatures (Tt) from sulfur isotope fraction-
ations (350C for primary inclusions and 250C for
pseudosecondary inclusions, see below) give pressure
estimates of 1 and 0.5 kbars, respectively, based on
the difference between Th and Tt of 70 and 50C
for primary and pseudosecondary inclusions, and the
modeling of the ore fluid, below, which suggests an
Na/K ratio of about 8:1.
Densities
Data in Table 3 indicate that the carbonic (non-
aqueous) component of type lb methane-rich inclu-
sions has a lower density (mode 0.60, range 0.57-
0.64) than in type la carbon dioxide-rich inclusions
(mode 0.70, range 0.67-0.75) or type 2 inclusions
(mode 0.70, range 0.63-0.75). These densities are
based on the temperatures of homogenization of the
carbonic portion of the inclusion (Thco2) and data from
Swanenberg (1979) and Hollister (1981). The Thco2
is difficult to measure reliably for type lb inclusions
because of the overlap with the clathrate-melting
temperatures, but it is well defined for type l a and
FI;. 18. Typical pseudosecondary (type 2) inclusions, without
visible carbonic fluid (i.e., containing only two phases, aqueous
liquid and vapor, with vapor/liquid ratios about 5 to 10 vol percent)
in quartz from the 51 vein on the 16 level. The characteristic
planar distribution along healed microfractures is shown in (a),
in which the field of view is 400 #m. An enlargement of the lower
left portion (field of view = 160 #m) is shown in (b).
336 LEITCH, GODWIN, BROWN, AND TAYLOR
TABLE 5. Pressure Estimates for the Bralorne Deposit
Gathered from the Different Approaches Used
in the References Listed
Primary luciusions Pseudosecondary
inclusions
Type la Type lb Type 2
kbars (MPa) kbars (MPa) kbars (MPa) Reference
1.o (100) o.s (so) 1
>1.0 (>100) >1.0(>100) 2
1.5 (50) 0.7,5 (7,5) 3
1.75(175) 1.5(150) 1.O (lO0) 4
Figures given are rounded off; uncertainty is 0.25 kbars
References (see text for details of the method used in each
case): 1 = Leitch and Godwin (1988), 2 = Leroy (1979), 3 = Hol-
lister and Burruss (1975), 4 = Bowers and Helgeson (1983b)
assumption of pure HsO (Roedder, 1984). However,
such estimates do not consider the presence of car-
bonic fluid. Estimates of the bulk density of these in-
clusion fluids define isochores on a pressure-temper-
ature (P-T) plot, which can be projected to the esti-
mated entrapment temperatures to provide trapping
pressures of 1.5 and 0.75 kbars, respectively (Hollis-
ter and Burruss, 1976).
The P-T conditions of entrapment of the primary
and pseudosecondary fluids can be evaluated more
rigorously using solubility relations for HsO and COs
in salt solution (Bowers and Helgeson, 1983a). Fluids
in type l a inclusions, with mole fractions of Xcoa
= 0.1 and XCH4 = 0.05, salinities of about 3 wt percent
NaC1 equiv, and Th less than or equal to 300C, would
have been supercritical at pressures above 1.5 kbars.
Projection of the 0.95 g/cm 3 isochore to the solvus
for these inclusion compositions yields a pressure of
entrapment of 1.75 kbars at the estimated trapping
temperature of 350C (Bowers and Helgeson,
1983b). Similarly, type 2 fluids would have been su-
percritical above 0.75 kbars; projection of the 0.97
g/cm 3 isochore to the solvus for these inclusions at
the estimated trapping temperature of 250C con-
strains pressure of entrapment to 1.0 kbar, lower than
we suggest for most vein filling. For type lb fluids
(mean carbonic mole fraction of 0.5, bulk density of
0.8 g/cm and salinity of 3 _ 2 wt % NaC1 equiv)
decrepitation temperatures of 230 to 300C give
minimum estimates of trapping pressures of 1.5 kbars.
The presence of occasional vapor- and CH4-rich
type lb inclusions with the far more abundant vapor-
and CH4-poor type 1 a inclusions raises the possibility
of trapping at subcritical temperatures. However,
type lb inclusions are so rare in comparison to type
la that most trapping was probably at or above the
solvus. Also, continued fault movement at lower tem-
peratures could have produced heterogeneous pop-
ulations of inclusions when earlier inclusions were
broken.
Most type lb (vapor-rich), and a few type la, in-
clusions decrepitate before homogenizing at temper-
atures ranging from 230 to 330C. Deerepitation
over such a small range of temperature is consistent
with rapid pressure increases in carbonic fluids at
temperatures above 250C (Malinin, 1974). Since in-
ternal pressures of up to 1.2 kbars are required to
decrepitate 12- to 13-#m-size inclusions in quartz,
and up to 2.7 kbars for smaller inclusions (Leroy,
1979), minimum trapping pressures implied are in
agreement with the pressure estimates above.
Evolution of mineralizing fluids
Paragenetie relations are obvious between type 2
and 3 inclusions, but less certain between type 1 and
type 2. However, based on the temperature drop from
type 1 to type 2, a progression in fluid composition
at Bralorne from type 1 to type 3 is postulated as
follows. Type 1 primary (?early) fluids were high
temperature (350C) and carbon dioxide and methane
rich (Xco + XCH4 = 0.15), with low salinities (ap-
proximately 3 wt % NaCl equiv). Type 2 pseudosec-
ondary fluids were lower temperature (250C), with
less carbon dioxide (Xco = 0.05), no detectable
methane, and had lower salinities (less than 1 wt %
NaG1 equiv). Later type 3 secondary fluids were cool-
est (180C), with no detectable carbon dioxide
(<0.03 mole fraction), but the same low salinity as
the pseudosecondary fluids. The fluid evolution can
be explained by mixing of hotter, more saline, carbon
dioxide- and methane-rich fluid with cold, more dilute
meteoric water as the hydrothermal system waned.
Estimated pressures of entrapment are lower for
type 2 pseudosecondary inclusions than for type 1
primary inclusions, and the geothermal gradient was
probably lower (< 10C/km) at the time of entrapment
of the pseudosecondary fluids than it was for the pri-
mary fluids (30C/km). These two features suggest
that the deposit had been partly unroofed and the
rock mass had cooled by the time fluids in the pseu-
dosecondary and secondary inclusions were trapped.
Pressure estimates from primary and pseudosecond-
ary inclusions (Table 5) indicate that the early (main)
mineralization took place at depths of 4 to 7 km from
fluids of type 1. Later deposition took place at 2- to
4-km-depth from fluids of type 2. These pressures
and depths are comparable to those estimated by
Coveney (1981) for the Alleghany mesothermal gold
lodes in California, or the 1.3 to 2.9 kbars (5-12 km)
estimated by Smith et al. (1984, but el. Brown and
Lamb, 1986) for a similar mesothermal deposit at
Timmins, Ontario.
Stable Isotope Studies
Stable isotope analyses of sulfides, carbonates, and
silicates constrain the origin of ore fluids at Bralorne
and permit estimates of the temperature of mineral-
BRALORNE Au DEPOSIT, B. C. 3 37
ization. Lateral and vertical zonation within the de-
posit and extensive fluid-wall rock interactions are
indicated. Minerals observed to be in contact or close
proximity to each other prior to crushing were sep-
arated by handpicking. For most samples, isotopic
equilibrium is indicated by a normal enrichment or-
der; samples not displaying this order are "reversed."
Inconsistent equilibrium relations among some sam-
ples may reflect deposition over a range of temper-
atures and/or chemical conditions.
Sulfur isotopes
Sulfur isotopes were analyzed in galena, sphalerite,
pyrite, pyrrhotite, chalcopyrite, and tetrahedrite from
samples at different levels of the mine and from nearby
deposits in the Bralorne area (Pioneer, P. E. Gold,
and B.R.X.). The results in Table 6 and Figure 19
show that i34S values range from -7 to q-11 per mil.
They cluster around -5 and q-5 per rail and have an
overall average of q-2. There is no horizontal or ver-
tical zonation for any given sulfide. The narrow range
of b34S values for Bralorne and other deposits in the
Bridge River camp, averaging '2 per mil, is character-
istic of some vein deposits (Rye and Ohmoto, 1974)
and implies an igneous source. The absence of sulfate
minerals at Bralorne is characteristic of Mesozoic me-
sothermal vein deposits (Taylor, 1987) and suggests
that H2S was the dominant sulfur species in the ore
fluid (Ohmoto and Rye, 1979). Computer modeling,
below, indicates that H2S was the dominant sulfur
species in the ore fluid.
The b34S value of pyrite, from -5 to 4-11 per mil,
is similar to that in deposits of the Mother Lode dis-
trict, California (Weir and Kerrick, 1987; Taylor,
1987), interpreted by Taylor to be consistent with
derivation of sulfur from shales, greenstones, or felsic
plutons, or a combination of all three. In the Bridge
River camp, Bralorne and other deposits are spatially
and temporally associated with Late Cretaceous felsic
intrusions, and Pb isotope evidence suggests that
TABLE 6. Sulfur Isotope Data for Several Deposits of the Bridge River Camp, Southwestern British Columbia
Data are plotted in Figure 19.
Measured ba4S for minerals
Deposit and sample no. gn sl tt py cp po As.g, T (C)
Bralorne host rocks
C082A
C082B
C092B
C093A
C094A
C095A
Bralorne-Pioneer
E73.004.048
Bralorne Mines
Suite 60
Bralorne (41 level, 79 vein)
Cl16-14
Bralorne (surface, Ida May)
E73.004.047
Pioneer (14 level)
E3519
Pioneer (5 level)
Joubin
P.E. Gold
P-85-03.450.5m
2.35 4.13
-6.48 -3.35
[-6.321
5.95 4.68
2.36
-5.30 -2.63
1.94 1.00
2.15
4.53
10.7
8.7
3.5
8.5
4.0
3.4
1.78 370
3.03 260
8.99 6.42 4.17 (R)
2.67 295
I1.141 (R)
-0.08 1.78 2.85 2.02 1.86 360
[1.78]
-4.22 -4.92 -5.76 3.28 240
BRX Arizona
Unlocated -7.50
Blackdome
Uniocated 0.64
Sulfur isotope ratios are reported as rSa4S, in per mil, relative to the Canyon Diablo troilite (CDT) standard, determined on SOs gas
extracted by high-temperature combustion of the sulfide with cupric oxide; analyses were carried out at the Stable Isotope Laboratory,
University of Calgary, under the supervision of Roy Krouse, and at the University of Ottawa, under the supervision of Bruce Taylor;
no systematic differences are apparent in the data sets from the two laboratories; precision of analysis was evaluated by duplicate
analyses, which are given as bracketed figures; difference in duplicate rSa4S values is less than 0.1 per rail; equilibration temperatures
were calculated from formulas of Ohmoto and Rye (1979)
Abbreviations: cp = chalcopyrite; gn = galena, po = pyrrhotite, py = pyrite, sl = sphalerite, tt = tetrahedrite, R = reversed
338 LEITCH, GODWIN, BROWN, AND TAYLOR
po
, py
x $1
gn
the main ore event. Sphalerite and galena are prob-
ably in equilibrium; experimental curves for tetra-
hedrite are not available. Sphalerite and galena pairs
give temperatures of formation ranging from 240 to
370C. These temperatures cannot be compared di-
rectly with fluid inclusion homogenization data since
the only sample for which both sulfur isotope and
fluid inclusion data are available (C 116-14) showed
reversed b34S values.
-10 -5 0 *5 -1 0
/34S,O/0o
FIG. 19. Sulfur isotope data for various sulfides from the Bridge
River camp. Mineral abbreviations: cp = chalcopyrite, gn = galena,
po = pyrrhotite, py = pyrite, sl = sphalerite, tt = tetrahedrite.
Solid circles are for pyrite from least altered host rocks; open
circles are for vein pyrite (see Table 6 for individual data).
large-scale hydrothermal systems were operative
(Leitch et al., 1989). It therefore seems likely that
sulfur could have been derived from a felsic igneous
source as well as from sedimentary and greenstone
sources in the intruded rocks.
Calculated temperatures of equilibration between
various sulfide pairs are given in Table 6. Tempera-
tures calculated using data from minerals other than
galena and sphalerite are erratic and unrealistically
high. Pyrite is stable over a wide temperature range
and typically does not equilibrate as well as galena
and sphalerite (cf. Godwin et al., 1986). Pyrrhotite,
chalcopyrite, and pyrite are more widespread in the
deposit and probably precipitated at several times
over the history of ore formation. Tetrahedrite,
sphalerite, and galena, which are closely associated
with the gold, only formed contemporaneously during
Carbon and oxygen isotopes of carbonates
Carbon and oxygen isotope measurements in vein
calcites and coexisting quartz are given in Table 7.
Values vary over only small ranges, t3C = -10.7
_ 1.5 and (180 ---- +15.2 _+ 1.5 per mil. The one ex-
ception is the tlsO of 7.2 for sample C1035, a clear,
vug-filling calcite that is paragenetically late and of
restricted occurrence (near the 77 vein on the 20
level). Despite the relative insensitivity of the quartz-
calcite geothermometer (20C per 0.2%0 variation in
delta quartz-calcite), the 2 to 3 per mil quartz-calcite
fractionation indicates calcite formation temperatures
of 145 to 290C (Table 7), similar to the 150 to
250 C range for secondary and pseudosecondary fluid
inclusions. The single high temperature of 345C
(sample 8-51Bsp) is anomalous, but fluid inclusions
in clear quartz in this sample also yield anomalously
high Th data. The calcites are enriched (Fig. 20) in
180 compared to the mean values (13 ___ 2%0) of car-
bonates from mesothermal gold vein deposits in the
Precambrian shield in Ontario (Timrains, Hollinger)
and in Western Australia (Kalgoorlie) but are depleted
in lSO relative to carbonates in the Mother Lode veins
of California (17 ___ 5%0: Taylor, 1987).
Calcite analyzed from Bralorne has a range oftlC
values from -12.1 to -9.5 per mil. This is more neg-
ative in t13C (Fig. 20, from Taylor, 1987) than vein
carbonates from other similar deposits such as the
TABLE 7.
Carbon and Oxygen Isotope Ratios in Vein Calcite, and Oxygen Isotope Ratios in Coexisting Vein Quartz,
in the Bralorne Deposit, Southwestern British Columbia
Data are plotted in Figure 20.
Sample no. Level Vein ( 13Cca]cite I ( 18Ocalcite2 ( 18Oquartz3 Aquart .... lcite 4 T(C)
8-51B(FW) 8 5lB -11.3 14.6 18.0 3.4 145
8-51B(SP) 8 51B -9.9 15.0 16.6 1.6 345
C081-2 8 5lB -12.1 15.0 18.4 3.4 145
15-51(c) 15 51 -9.6 17.0 18.9 1.9 290
16-51(E) 16 51 -11.0 15.6 18.5 2.9 180
C1035 20 775 -9.5 7.2 -- -- low
Cl17-7 32 79 -11.3 13.8 -- -- --
In per mil relative to PDB; CO2 liberated from calcite by reaction with 100 percent phosphoric acid at 25C (McCrea, 1950)
and collected until reaction was essentially complete; reproducibility of b3C values is ___0.2 per mil
2 In per mil relative to SMOW
a Dash indicates there was no quartz present in the vein
4 Per mil quartz-calcite fractionation
s Taken from a small carbonate veinlet near the major 77 vein
BRALORNE Au DEPOSIT, B. C. 3 3 9
o
.4
-8
MARBL//
""__joVEIN & ALTERATION
H - CARBONATE
0 T o* o
- 1 I I I I I I I I I I I
'8 -12 -16 *20 +24 +28 32
FIG. 20. Plot of $]3C vs. (]SO (Taylor, 1987) for carbonates
in the Mother Lode area of California, showing averages and fields
for Bralorne (B), Timmins (T), Hollinger (H), and Kalgoorlie (K).
Mother Lode in California (-5 _+ 4%o) or the Timrains
and Hollinger deposits in Ontario and the Kalgoorlie
deposit in Western Australia (-3 _+ 1%0). The Bralorne
carbonate 3C values are most similar to those mea-
sured at Panasquiera, Portugal, and Pasto Bueno, Peru
(high-temperature Sn-W deposits), which range from
-12 to -7 per mil (Rye and Ohmoto, 1974). At Bra-
lorne (see below), the dominant carbon species in the
ore fluids was H2603(aq), and the temperature of car-
bonate deposition was higher than 200C, so Coaltit e
13
= CHco3 -- SCuld (Rye and Ohmoto, 1974). Thus
the increase in sC to -9.5 per mil in sample C1035,
which is a late-stage calcite, suggests that the sCf
increased in the later stages of mineralization.
The sC of the hydrothermal fluids at Bralorne is
too negative for carbon to have been derived from
purely magmatic sources (-3 to -7.5: Taylor and
Gerlach, 1986). If magmatic and/or sedimentary car-
bonate (3C = -2 to +2) carbon were the dominant
source(s), then these more negative values are pos-
sibly due to oxidized organic carbon present in sed-
iments through which the fluids have passed (Ohmoto
and Kerrick, 1977), or to fractionation between CH4
and CO2 (Taylor, 1987). Similar sC values (-10
_+ 0.05%0) and ]so values (+16.4 _+ 0.15%0) have
been observed at the Agnico-Eagle deposit in Quebec
but only for siderite in premineral chemical carbonate
sediments; ankerite associated with gold has the more
typical values noted above of 3C = -3 _+ 0.12 and
So = +11.4 _+ 0.29 per rail (Kerrich, 1987). As
Ohmoto and Rye (1979) point out, the source of car-
bon for carbonates in hydrothermal vein deposits with
sC values between -5 and -10 per mil is difficult
to define.
Oxygen isotopes of silicates
One hundred oxygen isotope analyses (Table 8) of
mineral separates and whole-rock samples from 15
different vein systems in the deposit were obtained,
including ten duplicate analyses of quartz and one of
sericite. Vein quartz was sampled at the surface over
6 km, from the Cosmopolitan vein at the northwest
end to the P. E. Gold vein at the southeast end of the
vein system (Fig. 21a). Variation with depth was in-
vestigated in samples collected to 2 km deep in the
Bralorne mine (Fig. 2lb).
Lateral and vertical isotopic zoning in quartz veins
at Bralorne is not well defined. Quartz veins are ho-
mogeneous across their width (see the 51 vein on the
16 level, samples 1-7, in Table 8). Analyses of vein
quartz from five surface localities (Fig. 21a) suggest
a possible district-scale variation in so from 14.9
in the southeast to 19.4 per mil in the northwest
(analyses from the Pioneer mine and in the 77 vein
at the Bralorne mine are almost within analytical error
of each other). Mineralogy (more biotite to the south-
east; Joubin, 1948; Nordine, 1983) and structural data
(a deeper level of exposure to the southeast; Joubin,
1948) suggest that the geographic increase in so
of quartz reflects a variation in vein temperature
rather than so of the hydrothermal fluid. Similarly,
a decrease in so of quartz with depth (Fig. 21b)
suggests increasing temperatures with depth as in-
dicated by fluid inclusions.
Geothermometry: Quartz-sericite,-chlorite,-albite,
and -hornblende mineral pairs were analyzed from
veins and wall rocks to estimate the temperature of
mineralization based on oxygen isotope fractionations
(Table 8). For quartz-sericite, calculated tempera-
tures for milky quartz from the main veins (Clayton
et al., 1972; O'Neil and Taylor, 1969; Table 8) vary
from 200 to 240C; paragenetically distinct, clear
euhedral quartz (with anomalously high Th) and as-
sociated mica yielded a higher temperature of 305C.
The temperature of 530C from the 79 vein on the
41 level is too high due to contamination of the sericite
by fine quartz. Temperature estimates for quartz-ser-
icite in altered wall rocks range from 370 to 400C.
Using the equation of Wenner and Taylor (1971),
most of the quartz-chlorite pairs analyzed do not pro-
vide valid temperatures. Only one sample, from al-
tered wall rock around the 51 vein on the 8 level
(C002FW4), gave a geologically reasonable temper-
ature of 325C. The indicated isotopic disequilibrium
between quartz and chlorite in the veins may stem
from crystallization at different times and/or reequil-
ibration of the chlorite with lower temperature fluids.
Using the quartz-albite pair (Matthews et al., 1983)
a weakly altered diorite sample, 10 m from the 51
vein on the 15 level (C033-9), gives a temperature
of 290C, apparently reset toward lower tempera-
340 LEITCH, GODWIN, BROWN, AND TAYLOR
TABLE 8. Oxygen Isotope Compositions of Minerals and Rocks in the Braiorne-Pioneer Mesothermai Gold Vein System
Sample Level Vein Host rock bX8OX Tt 18Ofiuid
Vein quartz (all milky quartz except clear quartz, marked*)
C1002 0 Cosmopolitan Diorite 19.4
C 1027 3 51 (surface) Diorite 19.1 320 12.9
C048 4 77 (surface) Diorite 17.3
C082 4 Pioneer HW Soda gr 18.0
C049 710 PE gold Pioneer 14.9 360 9.9
8-51B(FW) 8 5lB, in FW Diorite 18.0 385 13.6
8-51Bsp 8 5lB split Diorite 16.6
8-51Bsp* 8 5lB split Diorite 17.4 380 12.8
C081-2 8 5 lB main Diorite 18.4 350 13.1
Cl11-28 8 5lB FW Soda gr 18.4 375 13.7
15-51(c) 15 51, center Diorite 18.9 340 13.3
16-51(E) 16 51, in FW Diorite 18.1 ?350 11.7
16-51 C 16 Composite Diorite 18.4
16-51(1) 16 1 cm from HW Diorite 19.4
(2) 16 3 cm from HW Diorite 19.0
(3) 16 6 cm from HW Diorite 18.3
(4) 16 10 cm from HW Diorite 18.9 ?350 12.5
(5) 16 13 cm from HW Diorite 18.6
(6) 16 17 cm from HW Diorite 18.9
(7) 16 20 cm from HW Diorite 18.6
C118-11 26 85 vein Soda gr 14.8
C118-11' 26 85 vein Soda gr 15.5 365 10.6
C116-3/16 41 79 vein Diorite 17.3 400 13.2
Cl16-14 41 79 vein Diorite 17.1 400 13.3
C128-20 44 77 vein Diorite 17.1 410 13.0
Quartz-mineral
fractionation
Sample Level Vein Mineral jXSOmi .... I (%0) Tf
Wt
Vein sericite and chlorite
Pioneer 0 HW Main Sericite 13.6
CO 49 ? 10 PE Gold Chlorite 12.0
8-51Bsp 8 51B Sericite 13.9
C111-28 8 5lB FW Sericite 14.1
C117-7 32 79 Sericite 10.7
C 116-3/16 41 77 Sericite 16.0
C128-20 44 79 Sericite 12.1
Rock J 18Omineral
Sample type qz ab hb
mica
4.4 400 (230)
2.8 1,260
3.5 465 (305)
4.3 405 (240)
(no qz)
1.3 845 (530)
S.O 360 (200)
Quartz-
mineral fractionation
ab hb mica
360
380
375
400
410
T
(c)
Wall-rock minerals: Least altered rocks
C092A Albitite 10.6 14.0
C071 Albitite 9.5
C4141 Albitite 12.8
C093A Diorite 14.3 13.8
C094B Soda gr 11.5 12.3
C033-9 Diorite 16.1
C033-5/6 Diorite 16.2
C033-1/2 Diorite 17.5
C002FW4 Diorite 15.3
C002FW1 Diorite 16.4
C080 Albitite
11.7 c
6.2
9.9 c
Wall-rock minerals: Altered rocks
14.5 8.3
14.7 TM
7.8
13.7 TM
13.8 TM
(R)
(R)
0.5
1.6
8.1
8.3
(R)
1.6 e
2.8 TM
2.7
2.7 TM
730
340
1,040
290
350
530
325
560
BRALORNE At DEPOSIT, B. C. 3 41
TABLE 8. (Cont.)
Sample Host rock 61sO D (m) a Sample Host rock 61SO 1 D (m)
Whole rocks: Least altered rocks
C083A Hb porphyry 10.0 > 10 C094B Soda granite 12.7 > 10
C092A Albitite 13.5 > 10 C095A Pioneer 6.5 > 10
C093A Diorite 10.6 >10 C4141 Albitite 13.4 >10
Whole rocks: Altered rocks
C1027HW3 Diorite 11.1 2.3 C116-23 Diorite 11.7 10.0
HW2 Diorite 12.9 0.3 -22 Diorite 13.0 5.0
HW1 Diorite 13.7 0.1 -21 Diorite 13.2 3.0
C002FW4 Diorite 13.9 3.5 -20 Diorite 13.0 2.0
FW3 Diorite 14.2 1.5 -19 Diorite 13.4 1.0
FW2 Diorite 15.2 0.4 - 18 Diorite 12.8 0.3
FW1 Diorite 16.1 0.1 C128 Diorite 11.0 6.0
C033-9 Diorite 11.8 10.0 -5 Diorite 11.9 4.0
-10 Diorite 13.0 10.0 -3 Diorite 11.6 2.0
-8 Diorite 14.1 8.0 - 1 Diorite 12.1 0.1
-7 Diorite 13.4 5.0 C111-29 Soda gr 13.7 0.5
-5/6 Diorite 14.9 3.5 -30 Soda gr 13.8 1.0
-3/4 Diorite 16.1 1.5 -31 Soda gr 13.9 3.5
-1/2 Diorite 16.4 0.3
Data are reported in the usual 6 notation relative to SMOW (standard mean ocean water); oxygen was extracted with CF3 at 575C
(Borthwick and Harmon, 1982), isotopic ratios were determined on CO2 produced by reaction with a hot carbon rod at 750C
(Clayton and Mayeda, 1963); mass spectrometric analyses are normalized to both SMOW and SLAP (standard light Antarctic precip-
itation); mean variation of 6 from duplicate analyses is 0.09 _+ 0.04 per mil; our 6so value for NBS-28 is 9.6 +__ 0.1
Tt = trapping temperature estimated from fluid inclusions (italic figure is from sulfur isotopes); for quartz-mineral temperatures
(Tf), figures for quartz-muscovite are from Bottinga and Javoy (1975); parenthetical figures are from Clayton et al. (1972) and O'Neill
and Taylor (1969); all temperatures are in C
Abbreviations: FW = footwall; HW = hanging wall; (R) = reversal of bSo values; ab = albite, ch = chlorite, hb = hornblende, qz
= quartz; c = chlorite, m = muscovite; Soda gr = soda granite, Pioneer = Pioneer greenstones, Albitite = albitite dike, Hb porphyry
= hornblende porphyry dike
Average of duplicates is used where available
2 Distance from vein in meters
tures by the hydrothermal alteration. Least altered
wall rock (C093A) far from veining gives a quartz-
albite temperature of 730C (although the albite here
is not likely of magmatic origin). For quartz-horn-
blende (Bottinga and Javoy, 1975), least altered Bra-
lorne diorite (C093A) far from the main veins gives
a similar temperature (340C) to that from quartz-
hornblende in sample C033-9, 10 m from the 51 vein
on the 15 level, which gives 350C. Both these figures
are extrapolated slightly outside the range of the
equation used. However, as for the quartz-albite pairs
above, they imply a strong interaction of the wall rock
with the altering fluids, (i.e., a high water/rock ratio).
The interaction is apparently more marked for the
hydrous minerals, chlorite and hornblende, than for
albite.
Implications for water/rock ratios: Samples of
whole rocks and separate minerals from five traverses
across altered wall rocks (the 0, 8, 15, 41, and 44
levels, at 0, 350, 650, 1,800, and 2,000 m depth)
were analyzed for their oxygen isotope ratios to test
for zoning. Alteration varies from quartz-sericite-an-
kerite up to 1 m from the vein, through albite-calcite
from 1 to 5 m, to chlorite-epidote 5 m or more from
the vein. Both whole-rock and mineral oxygen isotope
ratios increase as the veins are approached (Table 8).
For instance, in the 10-m series in altered diorite near
the 51 vein on the 15 level (C033-1 to 10), there is
a progression from 5sO = 10.6 per mil in least altered
rock (estimated by C093A), to 16.4 per mil imme-
diately adjacent to the vein; the minor reversal be-
tween 7 and 8 is due to proximity to a subsidiary frac-
ture off the main vein. The 5sO varies in a similar
way among samples from the other series in diorite,
although gradients are not as pronounced at deeper
mine levels (the Cl16 and C128 series from the 41
and 44 levels) which are in much more mafic wall
rock than the typical diorite, containing less than 5
percent quartz. In soda granite, whole-rock 5sO val-
ues are about 1 per mil higher within 3.5 m of the
vein. There is thus a clear pattern of reequilibration
of the whole-rock 5sO values due to hydrothermal
alteration for up to 10 m from veins.
Quartz in the veins and altered envelopes has a
high 5sO value (up to 19.4%0) compared to quartz
in the least altered rock (14.3%0), and its 5SO in-
creases systematically toward the veins. The other
major rock-forming minerals in the diorite also show
342 LEITCH, GODWIN, BROWN, AND TAYLOR
(14.9),
, co49
, ,,,P. E.
(19.4) ,%_ ,_ /GOL D
T '-- -' 'd107 z + C082 /
/ - CADWALLADER VAULT (Z PIoer HW a
LEVEL DEPTH (m) COSMOPOLITAN' '51'VEIN '77'VEIN'51BFW' 'HW' PE,GOLD
O 13 19.4 3 19.1 9'
15 65o 5
26 70
32 450
41
44 2oo0 1 ? b
FIG. 2I. (a), Geographic zoning of oxygen isotope ratios of vein quartz from the Bralorne-Pioneer
vein system. The "+" symbols indicate extent of Bralorne intrusions; sample numbers are C1002, etc.;
80 values are given in parentheses; vein designation is contained in single quotes, with vein strike
and steep northerly dip indicated. (b). Vertical zoning in oxygen isotope ratios in the Bralorne-Pioneer
vein system is shown by variation in {518Oquartz values with reference to the mine levels and schematic
position in the plane of the vein system. The host rock is diorite except for those samples marked *
(51BFW and Pioneer HW veins) in soda granite, and the P, E. Gold vein marked * in Pioneer greenstone.
small increases in h80 values toward the veins until
they become unstable (plagioclase from 13.8-14.5%0,
and hornblende from 6.2-8.3%0). These effects ex-
plain the shifts in whole-rock i]sO patterns toward
higher values as the vein is approached in hydro-
thermally altered wall-rock envelopes. In the more
mafic host rocks at deeper levels, the lack of alteration
quartz may explain the smaller shifts observed. The
marked shifts in h]80 values for both mineral separates
and whole-rock samples imply high water/rock ratios
during the alteration, a feature common to other ear-
bonate alteration zones around mesothermal gold
quartz veins (e.g., Kerrich, 1983; Taylor, 1987).
Oxygen isotope composition of the ore fluids (Table
8) were calculated from isotopic analyses of vein
quartz, trapping temperatures from fluid inclusion
studies, and the quartz-water fraedonation equation
of Clayton et al. (1972). Values appear to have been
reasonably constant at about 12.6 + 1.3 per rail from
the bottom to top of the vein system except for two
samples, from the P. E. Gold vein and the 51 vein on
the 26 level, which have lower values.
The small-scale variation in h]sO values of vein
quartz (decreasing h values with depth, i.e., increasing
temperature) is similar to the regional lateral variation
in observed h]sO values of quartz described by Ma-
heux et al. (1987) and Nesbitt et al. (1987). Deposits
in the camp are zoned from the high-temperature
Bralorne-type Au quartz veins (18Oquartz = 17.5
+ 1.0%0), to intermediate temperature Sb-Ag-Au
veins (21.0 + 1.0%0), to low-temperature Hg deposits
(29.0 __+ 2.0%0). Deposition was from the same or sim-
ilar deeply circulating, highly evolved (]SO-shifted)
fluids of around 11.5 __+ 2 per mil. This value, cor-
rected to the estimated trapping temperature of
350C, is within analytical uncertainty of the mean
value obtained for Bralorne in this study. The strong
enrichment in i]sO values of quartz, combined with
D/H studies of fluid inclusions, led Nesbitt et al.
(1986, 1987) to propose that the ore-forming fluids
were composed of deeply circulated meteoric water
(cf. Magaritz and Taylor, 1976, 1986).
The calculated h]80 fluid of 12.6 per mil at Bra-
lorne is close to values for the Coquihalla deposits
near Hope, B. C., and the Mother Lode deposits in
California (Taylor, 1987) and is in the middle of the
range for metamorphic waters. However, the concept
of "metamorphic water" must be clarified, because
its original isotopic definition (e.g., Taylor, 1974) is
not meant to imply water of dehydration (see Taylor,
1987). Inasmuch as most water/rock reactions involve
either seawater or meteoric water, the evolved me-
teoric water hypothesis seems reasonable from the
evidence to date but might be further constrained by
data from hydrogen isotope studies (in progress). In
similar deposits in the Juneau area of Alaska, iD values
of-20 to -30 per mil indicate a deep-seated meta-
morphic fluid source (Goldfarb et al., 1989). In the
Mother Lode of California, sericite and mariposite
associated with alteration and gold mineralization
formed from waters with somewhat higher iD than
waters in some fluid inclusions in adjacent quartz
BRALORNE Au DEPOSIT, B. C. 343
veins, indicating mixing of waters with different evo-
lutionary histories (Taylor, 1987). Lead isotope evi-
dence (Leitch et al., 1989) suggests that a magmatic
source was involved in the formation of galena.
Although meteoric water may end up with a meta-
morphic buffered CO2 value (Nesbitt et al., 1987),
the high CO2 contents typically found in the ore fluids
are usually interpreted as derived by metamorphic
dehydration (e.g., Kerrich, 1983; Andrews et al.,
1986). Such an origin is espoused by many workers
on gold vein deposits in the Precambrian shield of
Ontario and Quebec (e.g., Colvine et al., 1984).
Hodgson and MacGeehan (1982) favor a variation of
metamorphic origin for the fluids by expulsion of pore
water (after Lydon, 1977). At the Mother Lode de-
posit in California, Landfeldt (1987) proposed that
the ore fluids may represent a CO-rich fluid phase
separation (Trommsdorffand Skippen, 1986) derived
from a reservoir generated at 20- to 30-km depth
during rebound of an isotherm depressed by subduc-
tion. The high CO contents are also consistent with
fluids produced by degassing of mantle magmas (e.g.,
Cameron, 1989a, b).
A more complex, mixed origin for H-O-C-S in the
mineralizing fluids must be considered for fluids cir-
culating along structural and tectonic boundaries such
as the Bralorne fault zone. For example, emplacement
of mafic magmas along major tectonic zones provides
a source of CO isotopically similar to that found in
the Mother Lode (Taylor, 1979, 1987). Thus the ore
fluids at Bralorne could also have been derived from
mafic magmas. Sulfur isotope values in the sulfides,
with a range from -7 to +9 per mil, clustered around
zero, would support such an interpretation. Carbon
and oxygen isotope ratios in carbonate minerals are
most similar to those in high-temperature (Sn-W) vein
deposits, but the observed range in 3C values could
be produced from many sources. Contamination of
magmatic carbon by oxidized organic carbon or by
dissolved carbonate carbon in sediments could pro-
duce the observed range of 3C (-12 to -7%0). It
seems likely that deeply circulating meteoric waters
were mixed with a more uniform crust-equilibrated
reservoir of waters being driven off by metamorphism
(cf. Kerrich, 1987; Taylor, 1987). Early Late Creta-
ceous intrusions, closely associated with mineraliza-
tion, provided heat to drive the flow of the ore fluids;
they may also have provided a magmatic component.
Discussion: Deposition, Transport,
and Source of Gold
Characteristics of the ore-forming fluids
The results presented above suggest that the min-
eralizing fluids at Bralorne evolved from initially high-
temperature, CO-rich to low-temperature, H20-
dominant ones (cf. the deposits in the Red Lake camp
of Ontario; Andrews et al., 1986). It is not certain
that the gold was deposited with the earlier carbonic
fluids (as found at many mesothermal gold vein de-
posits, e.g., Robert and Kelly, 1987, or Walsh et al.,
1988) or with the later HO-rich fluid (cf. Andrews
et al., 1986). However, since the later fluids are like
those found in inclusions in calcite, and no gold is
associated with calcite, it seems most likely that gold
deposition accompanied the earlier fluids. Thus the
mineralizing fluids at Bralorne are probably similar
to other mesothermal gold vein deposits in the Ca-
nadian Shield, such as the Hollinger-McIntyre, where
the composition of the primary fluid was CO rich
and low salinity: 93 mole percent HO, 6 equiv mole
percent CO and 1 equiv mole percent NaC1, and
deposition was at 280 +__ 50C (Smith et al., 1984;
Wood et al., 1986). Similar fluids have also been de-
scribed at Doyon, Quebec (Guha et al., 1982) and
Pamour, Ontario (Walsh et al., 1988), as well as for
the Kalgoorlie-Kambalda area, Western Australia
(Groves et al., 1984). The salinity of the Bralorne ore
fluid, at about 3 wt percent NaC1 (approximately 0.5
m) is the same as predicted by the experiments of
Wood (1987) to be best capable of mobilizing gold
(with a high Au/Ag ratio, as observed at Bralorne)
from source rocks while leaving base metals behind,
a process suggested by Kerrich (1983) as important
in forming mesothermal gold deposits as opposed to
base metal deposits.
Further details of the ore fluid were modeled using
the PATH computer program (Helgeson et al., 1970;
Perkins, 1980). Thermodynamic data for aqueous
species are from Helgeson (1969); data for solid
phases are from Helgeson et al. (1978). Data for gold
chloride complexes were included (Wood, 1987, has
shown experimentally that gold solubilities in the 5
to 500 ppb range are attained in solutions of 0.5 m
NaC1). Other workers have postulated that gold,
present as reduced sulfur complexes, was deposited
in response to fluid oxidation and the lowering of total
sulfur in solution, both as a result of pyritization (e.g.,
Phillips and Groves, 1983, 1984; Romberger, 1986).
Carbonyl complexes may also be responsible for so-
lution and transport of Au in a reduced species, in
accordance with the low redox state (Fe+2/Fetotal
= 0.9) and the ubiquitous gold-carbonate association
(Fyfe and Kerrich, 1984). Carbonyl, cyanide, and/or
thiocyanide complexes were also proposed as gold-
transporting agents by Hutchinson and Burlington
(1984). At the time of the study, data for carbonyl,
cyanide, and thiosulfide complexes were not available.
Including thiosulfide, however, would only increase
the gold in solution and total gold deposited (Seward,
1973, 1984); the mechanism of precipitation would
not be altered.
The PATH program models reactions and changes
in the ore fluid by progressive, stepwise titration that
344 LEITCH, GODWIN, BROWN, AND TAYLOR
TABLE 9.
Reaction between Modeled Fluid and Wall Rock at the Bralorne Deposit, Predicted by the PATH Program
Only species at concentrations greater than 10 to 6 m were considered.
0.6NaAISiaOs (albite) + 0.06Ca2Mg4SisO22(OH)2 (tremolite) + 0.34FeSiOa (ferrosilite) + 0.19793K + + 0.22535Ca +2
+ 0.54649H2S + 0.70719H2COa + 0.007058Mg +2 + 0.003401SO 2 + 0.0003107H + + 0.00006433Al(OH) +
+ 0.002067KSO + 0.001771CASO4 + 0.039548MGSO4 + 0.000108550AuC1 + 0.031207HSO q- 0.0021105HC1
+ 0.000000015Au + + 0.0005135Mg(OH) + ::> 2.01805SiOg (quartz) + 0.199993KAISiaO0(OH)g (muscovite)
+ 0.347118CaMg(COa) (dolomite) + 0.312016FeS (pyrite) + 0.000108565Au (native) + 0.60002Na + + 0.0023276C1-
+ 0.012928HCO + 0.016685Fe(OH) + + 0.011300Fe +2 + 0.001975H4SIO4 + 0.0000111OH- + 0.0000839Al(OH)
+ 0.0004263HS- q- 1.1116H20
All species are aqueous except where noted and are expressed in moles/liter (m)
represents water-rock reactions. The program writes
detailed reactions (Table 9), considering as many as
25 reactants and as many products. It first calculates
the fluid that would be in equilibrium with the ob-
served alteration assemblage. Such a model for the
Bralorne deposit must account for the abundant car-
bonate in a typical alteration assemblage that also in-
cludes quartz, muscovite, albite, chlorite, pyrite, and
native gold (thermodynamic data for arsenic com-
pounds are not available, so the implications of the
arsenopyrite commonly seen in the observed assem-
blage cannot be assessed). The ore fluid in equilibrium
with this typical alteration assemblage is characterized
in Table 10. Except for the fugacities of CO2 (10 +2")
and CH4 (]0+'5), the characteristics of these fluids
are similar to those modeled by Helgeson and Garrels
(1968) for deposition of gold, pyrite, and quartz, al-
though their study considered precipitation in re-
sponse to a temperature drop rather than reaction
with the wall rock. However, the gold content of the
ore fluid predicted by the model (1 to 2 x 10 - m,
or 0.1 to 0.2 ppb) is much lower than previously es-
timated (e.g., 20 ppb in both Helgeson and Garrels,
1968, and Kerrich, 1983; 1.5 to 10.5 ppb in Brown,
1986, and 1 to 10 ppb calculated by Shenberger,
1985). The value estimated in this study is much
closer to the value of 0.04 ppb measured at Broad-
lands, New Zealand (Seward, 1984), or the average
from mineralized areas of 0.1 ppb that was measured
by McHugh (1988). Such low predicted gold contents
emphasize the importance of gold-depositing mech-
anisms (rather than transporting mechanisms) in order
TABLE 10. Characteristics of Ore Fluid at the Bralorne
Deposit, Predicted by the PATH Program
Temperature = 350C, pressure = 1.75 kbars, pH = 4.5
(slightly acid at 350C and 1.75 kbars)
Na/K = 8:1, [Na] = 0.4 m, [K] = 0.0.5 m, [CI l = 0.5 m, (total
salinity = 0.5 m, approximately 3 wt %)2
/coz = 102'S,/ca, = 10S,/oa = 10-a,Ja = 10 -7
lee+21 -- X X 10 -7 m, [ng+l -- 0.003 m, [Ca+21 -- 0.01 m, lS-l
: 10 - m, lAu +] = 10 - m (0.1 ppb, as AuCI)
Estimates from fluid inclusion studies
Agrees with estimate from fluid inclusion studies
to explain adequately the formation of large gold de-
posits.
Next, the model allows a fluid of the derived com-
position to react with a fresh rock of the observed
composition. The least altered Bralorne diorite (Table
1) was modeled as 60 percent albite, 6 percent trem-
olite, and 34 percent ferrosilite. The latter two min-
erals approximate the Fe/Mg ratios in probe analyses
of the 40 percent hornblende typically found in the
fresh diorite (thermodynamic data are not available
for hornblende, a compositionally complex mineral).
The pressure and temperature constraints applied
were constant at 300C and 1 kbar, selected to be as
close as possible to the fluid inclusion information
without extending beyond the experimental data in
PATH.
Deposition of gold
Theoretical predictions: When the modeled fluid
is allowed to react with the "fresh" (least altered)
rock, the alteration mineralogy predicted by the
PATH program closely matches the sequence ob-
served outward from the veins at Bralorne (modal
mineral data in Table 1). Close to the vein, the model
predicts dissolution of the original albite and tremolite
q- ferrosilite and precipitation of major amounts of
quartz, dolomite, and muscovite, minor amounts of
pyrite, and minute amounts of native gold. Gold pre-
cipitation is most favored in the early steps of the
process. About two-thirds of the gold is deposited im-
mediately adjacent to the vein as the COa-bearing
fluids react with the wall rock. Gold precipitation is
predicted to decrease sharply as soon as carbon (in
trace amounts) becomes stable. Farther out from the
vein (i.e., in later steps of the modeled process), chlo-
rite becomes stable (approximated thermodynami-
cally by talc, for which data are available). Farther
out still, albite becomes stable where muscovite is no
longer stable. This would be expected because close
to the vein, hydrolysis of albite to muscovite consumes
K from the fluid and releases Na into solution; thus
the Na/K in the fluid would evolve to higher values
as it passed outward (cf. Kerrich, 1983). Concur-
rently, as the fugacity of sulfur in the fluid drops, the
stable sulfide becomes pyrrhotite rather than pyrite
BRALORNE Au DEPOSIT, B. C. 345
and the precipitation of Au is no longer favored (cf.
Phillips et al., 1983). Where magnetite becomes sta-
ble (farthest from the vein), Au is dissolved. The model
thus predicts a strong correlation of gold with pyrite
and an absence of significant gold from pyrrhotite- or
magnetite~bearing assemblages. This is corroborated
by evidence from other mesothermal gold vein de-
posits where magnetite is found as a precursor iron
mineral supplying Fe to form pyrite and enhance gold
deposition (Phillips et al., 1983; Hall and Rigg, 1986).
The calculated solubilities for gold, as both chloride
complexes such as AuCI and thiosulfide complexes
such as Au(HS), are illustrated in Figure 22. Under
the predicted reducing, slightly acid conditions of log
fo = -30 and pH = 4.5 (neutral at 300C is 5.2), in
the field of stability of pyrite, carbonate, and sericite
(as observed in the wall-rock assemblage) and in the
H2S stability field (as confirmed by PATH), the dom-
inant Au species appears to be the chloride complex.
The observed alteration assemblage, where K feldspar
is not stable, requires this pH (contrast this with the
-28
-30
-32
-36
-38
-4(
FIG. 22. Log fo-pH diagram to illustrate conditions of min-
eralization at Bralorne, at 300C and 1 kbar. Stability fields for
carbonate (ca), graphite (gr), methane (CH4), muscovite (muse),
K-feldspar (kspar), kaolinitc (kaolin), pyrite (py), pyrrhotite (po),
hematite (hm), and magnetite (mt) are shown. The solution com-
position used is total sulfur = 0.5 m, ionic strength = 0.5 m, [Ca +]
-- 0.01 ra, [K +] = 0.05 ra, from PATH modeling. The ore fluid at
Bralorne, modeled at log fo = -30, pH = 4.5, is shown by the
cross in the field ofpyrite, muscovite, carbonate, and HS. As CO
is removed from the solution by carbonation of the wall rock, the
solution becomes less acid, moving in the direction of the arrow
and causing precipitation of gold (down the chloride complex sol-
ubility gradient). The abundances of Au as the chloride complex
and the thiosulfide complexes are from Seward (1984).
very high pH conditions, where thiosulfide complexes
are more stable, predicted by Walsh et al., 1988).
Reaction of this solution with the wall rock would
remove CO2 due to the formation of carbonate, caus-
ing the pH to increase and a corresponding shift to-
ward lower AuCI concentrations, i.e., precipitation
of gold (arrow in Fig. 22). Since this is in the direction
of slightly increasing or constant stability of gold thio-
sulfide ions, it is not supportive of Au deposition by
destabilization of thiosulfide complexes. Note that it
is also at relatively constant fo (cf. Goldfarb et al.,
1989), so that sulfide complexing of Au would not be
expected to vary during this process.
A correlation between pyrite and gold has been
empirically observed at mesothermal gold mines
throughout the world; this model offers a theoretical
explanation. The actual precipitation of the gold in
the model involves reduction of the aurous gold (Au +)
in the AuCI complex, to native gold (Au), by do-
nation of an electron. Electrons appear to be donated
by a concurrent oxidation reaction, of S -2 in H2S, to
sulfur in pyrite, FeS, which may be thought of as S-.
This is suggested by the strong correlation in the
model between Fe + consumption and pyrite pro-
duction on the one hand and gold produced on the
other. Consideration of charge balance requirements
for the equation written by the program (Table 9)
shows that in this reaction, the small surplus of S -2
(in HS) oxidized to S- (in pyrite) corresponds to the
small amount of gold being precipitated. As soon as
carbon becomes stable (C +4 + 4e- = CO), the com-
petition for available electrons appears to reduce
sharply the possibility of precipitating gold. A similar
conclusion has been reached for the mesothermal gold
lode deposits of Western Australia (cf. the reactions
for reduction of gold in thiosulfide complexes: Phillips
et al., 1983). According to the experiments of Bancroft
and Jean (1982), the gold ions may be adsorbed on
the surface of sulfide grains such as pyrite and then
reduced, similar to reduction of gold on activated
carbon.
Comparison of predicted and observed assem-
blages: The predicted outward transition from quartz-
sericite-ankerite-pyrite to chlorite-albite-calcite is as
observed. Amorphous carbon is seen only rarely in a
late-stage carbonate alteration that is black from in-
cluded finely divided carbon. Assay data show that
gold concentrations drop off sharply immediately
outside the veins. The bulk of the gold is in the thin
black ribbons in the quartz veins, which are highly
altered wall-rock slivers composed of finely divided
pyrite (carbon has not been identified here), arse-
nopyrite, sericite, and minor ankerite. Higher gold
contents correlate with galena, and to a lesser extent,
sphalerite (Dolmage, 1934; Joubin, 1948; Poole,
1955). Gold grains up to several hundred microns
across are intergrown with galena (this study) and
346 LEITCH, GODWIN, BROWN, AND TAYLOR
sphalerite (Poole, 1955) in some Bralorne-Pioneer
specimens, whereas gold inclusions in pyrite and ar-
senopyrite tend to be less than 15 #m across. Sphal-
erite is also an excellent indicator of gold at Pamour,
Ontario (Walsh et al., 1988), where it contains inclu-
sions of gold. At Bralorne, pyrrhotite is present in
some altered wall rocks, but there is insufficient data
to confirm the predicted inverse relation between
gold and pyrrhotite. Magnetite is not found in any
altered rocks at Bralorne.
The best gold ore shoots were in diorite, peripheral
to soda granite bodies (D. H. James and J.P. Weeks,
1961, unpub data; Campbell, 1975). This has been
explained by suggesting either a genetic relationship
of ore to the soda granite (disproved by isotopic dat-
ing: Leitch et al., 1991) or by brittleness of the soda
granite causing an inability to sustain large openings.
However, the PATH model suggests an alternative
explanation: the more abundant iron in the diorite
and in the greenstone, the favored host rocks, could
increase the precipitation of pyrite, and hence, of
gold. A similar correlation between iron-rich host
rocks, the best ore, and the most abundant sulfides
occurs in the mesothermal gold veins of the North
Western mining camp in Zimbabwe (Fuchter and
Hodgson, 1986). At the Hunt mine, Kambalda, West-
ern Australia (Phillips and Groves, 1984), fluid-wall
rock reaction is also interpreted to have caused the
deposition of gold and associated ankerite, calcite,
biotite, chlorite, and pyrite. Similarly, Au has been
found to be concentrated where the shear crosses a
competent, chemically favorable Fe-rich host rock at
the Lac Shortt deposit in Quebec (Morasse et al.,
1986). There is no net increase in iron during such
an alteration process (Phillips et al., 1983; Hall and
Rigg, 1986; Melling et al., 1986).
Transport of gold
A model for ore deposition at Bralorne must ac-
count for two outstanding features of the quartz veins.
These are the presence of the ribbons that contain
the bulk of the gold, and the syntaxial, coarsely crys-
talline milky quartz that grew perpendicular to the
ribbons or walls of the veins. The latter observation
is supported by the strong induced piezoelectric re-
sponse of the quartz, which indicates alignment of
the c-axes of many grains (M. M. Gomshei, pers. com-
mun., 1988). The dark ribbons in the quartz imply a
cyclic process, with repeated fracturing and em-
placement of quartz. The coarse quartz crystals imply
growth in open space (under high, but fluid, pressures:
cf. Wood et al., 1986, and Nesbitt, 1988). This con-
flicts with the general lack of open space in the me-
sothermal environment, under pressures of 1 to 2
kbars (4-8 km depth, see above). Hydrostatic pres-
sures are found as deep as 6 to 7 km in Witwatersrand
drill holes and to 12 km in the Kola Peninsula hole
(B. E. Nesbitt, pers. commun., 1990), but these are
both in noncompressional cratonic settings, unlike
Bralorne.
The characteristics of repeated fracturing and de-
position of quartz, in a high fluid pressure regime,
may be explained by a "fault-valve" model (Sibson
et al., 1988). Fluids transporting ore are derived from
a geopressured reservoir, possibly formed by meta-
morphic devolatilization at amphibolite grade,
ponded below the ductile-brittle transition zone. The
build-up of pressure occurs because of the inappro-
priate orientation for slip, in a horizontally compres-
sive stress field, along the steeply dipping faults host-
ing the Bralorne veins. In order to permit slip, and
therefore rupture and fluid flow, the pore pressure
must be very high. Such extremely high fluid pres-
sures thus provide the environment for coarse crystal
growth to occur at depths of over 4 km. When rupture
occurs, fluid will flow into the open space of the fault
where the sudden drop in pressure would promote
deposition of quartz (Walther and Helgeson, 1977),
sulfides (Helgeson and Lichtner, 1987), and gold.
Fluid flow would tend to die away slowly in such a
system over a period of a few months (Sibson, 1981),
possibly allowing time for coarse crystal growth.
As mineral deposition occurred, and the fracture
sealed itself, pressure would rise again, leading to a
repeat of the same process. This cyclic process ex-
plains the ribboning in the Bralorne veins. Each ribbon
of quartz would have associated with it a black layer
of intensely altered host rock consisting of pyrite, ar-
senopyrite, sericite, and occasional ankerite, with
gold, that formed the vein margin before rupture took
place again. The PATH model predicts that the bulk
of the gold is precipitated in the immediate envelope
to the vein, so that incorporating even a sliver of wall
rock into the vein would include most of the gold in
the vein.
Source of fluids and gold
Lead isotope ratios suggest large-scale mixing of
mantle and upper crustal components, implying
widespread circulation of fluids that were then fo-
cused along major fault zones to form the deposits.
It seems most likely that these fluids were of meta-
morphic derivation (cf. Kerrich, 1983; Groves et al.,
1987), with possible contributions of highly evolved
(80-shifted) meteoric (Nesbitt et al., 1986; Nesbitt,
1988) and magmatic waters. It is not possible to
choose between evolved meteoric and metamorphic
waters due to a lack of hydrogen isotope data for Bra-
lorne. A similar conclusion, that the available data are
not sufficient to define a source for the mineralizing
fluids of Archean lode gold deposits with conviction,
even with hydrogen data, was reached by Roberts
(1987). At the Mother Lode in California, stable iso-
tope data merely confirm that the fluids have inter-
BRALORNE Au DEPOSIT, B. C. 3 4 7
acted extensively with rocks (Bohlke and Kistler,
1986; Taylor, 1987), and the ultimate source or
sources of the Mother Lode fluids is not clear (Weir
and Kerrick, 1987). Objections to the evolved me-
teoric water model relate mainly to the lack of reli-
ability of hydrogen isotope analyses of fluids, which
are often extracted from fluid inclusions (Pickthorn
et al., 1987). For a few deposits, such as the Jungwon
in Korea, the isotopic data are not compatible with
ore formation from a metamorphic fluid, and there-
fore, favor an evolved meteoric fluid (Shelton et al.,
1988).
Wall rocks adjacent to veins at Bralorne are
strongly enriched in lsO, toward isotopic equilibrium
with the veins. This signifies that a fluid-dominated
vein system was operative at Bralorne (el. Kerrieh,
1983). A corollary is that lateral diffusion of silica,
gold, etc. (el. Boyle, 1979) from wall rock into veins
(i.e., a rock-dominated system) is not compatible with
the isotopic relations. These results support the con-
clusion that the fluids involved in transporting the
gold at Bralorne were derived from deep metamor-
phic waters or deeply circulating, evolved meteoric
waters, in a fashion similar to that proposed by Fyfe
and Henley (1973) or Nesbitt (1988).
A further constraint on the origin of the ore fluids
at Bralorne is required to explain the extensive ear-
bonate alteration. This abundant CO2, and the K re-
quired for muscovite replacement of albite adjacent
to veins, could be supplied by fluids released during
greensehist-amphibolite metamorphism, where the
relative proportion of CO2/H20 is on the order of 0.2
to 0.5, and K/Na is about 1 (Kerrieh and Fyfe, 1981).
A reduced, Na- and Au-bearing solution was found
to be produced during the dehydration of amphibo-
lites in the metamorphism of spilitized marie volcanic
rocks at the Opapimiskan Lake gold deposit in Ontario
(Hall and Rigg, 1986); such studies at other deposits
(Colvine et al., 1988) support the derivation of the
ore fluids from metamorphic dehydration. In the deep
crust "magmatie" and "metamorphic" processes are
so intertwined that distinction between the two is
probably irrelevant (Cameron, 1989e). At Bralorne,
the timing of ore deposition, synchronous with ret-
rograde conditions after the metamorphic peak as
suggested by isotopic dating and pressures of min-
eralization lower than metamorphic pressures, sup-
ports metamorphic dehydration (el. Phillips and
Groves, 1984; Thompson, 1986; Nesbitt and Mueh-
lenbachs, 1988).
Ultramarie rocks have been cited as the source of
gold in vein deposits (e.g., Pyke, 1975, 1976, in the
Porcupine camp, Ontario) because of their presumed
elevated Au abundances. However, analysis of Au by
Tilling et al. (1973) suggested that ultramarie rocks
have abundances of 0.5 to 2 ppb Au, in the same range
as other igneous rocks. More recent work suggests
that carbonation provides a mechanism for release of
Au from magnetite and secondary sulfides where it
may have been concentrated earlier during serpen-
tinization of ultramafic rocks. Thus the gold may be
reconcentrated in carbonate-altered ultramafic
equivalents, or listwanites (3-5 ppb: Buisson and
Leblanc, 1987). Serpentinized ultramafic rocks are
prominent along the faults of the Bralorne zone, ad-
jacent to the gold deposits, although their carbonate-
altered equivalents are not. Thus it is not likely that
the nearby or subjacent ultramafic rocks of the Bra-
lorne area are the source for the gold. Instead,
sweeping large volumes of fluid (as suggested by the
isotopic studies) through large bodies of rock with
slightly elevated gold contents could provide a mech-
anism for extraction of sufficient gold (cf. Hutchinson
and Burlington, 1984). High Mg basalts of the Pioneer
Formation in the Bralorne area (Leitch, 1989) could
be such a source (cf. Keays, 1984). The low-salinity
(0.5 m) solutions found at Bralorne would be capable
of nearly complete extraction of gold from the source
rock, yet are close enough to saturation to allow for
efficient precipitation at the site of deposition (Wood,
1987).
The total production of gold from the Bralorne de-
posit was 130 metric tons, or 1.3 X 108 g. If this mass
of gold was derived by 50 percent efficient leaching
(cf. Buisson and Leblanc, 1987) from rocks containing
1 ppb Au (Tilling et al., 1973), then the source volume
would be on the order of 100 km 3 of rock (1 km 3 of
rock contains 2.8 X 106 g of gold at 1 ppb). If this
gold were leached and transported during metamor-
phic outgassing within the greenschist facies with 2
to 3 percent water liberated from the volcanics and
sediments, then the solvent volume would be on the
order of 200 km a, and the concentration of gold in
solution would be 1.3 X 108 g Au divided by 200
X 10 l' g of rock, or 0.65 ppb (cf. Kerrich, 1983). If
the leaching were less than 50 percent efficient, this
value would be less (at 10% efficiency, it would be
0.1 ppb). This is similar to the 0.1 to 0.2 ppb predicted
by the PATH model, or the 0.04 to 0.1 ppb range
measured in natural waters associated with mineral-
ization presently forming (Seward, 1984; Krupp and
Seward, 1987; McHugh, 1988). Calculations by
Shenberger (1985) indicate that a concentration of 1
to 10 ppb Au in solution is sufficient to produce eco-
nomic Au grades, consistent with observations by
Brown (1986) and Krupp and Seward (1987) in New
Zealand, where economic grades of mineralization are
being formed from fluids with 1.5 to 10.5 ppb Au in
solution. The length of time necessary to form a de-
posit of economic size (50-100 metric tons, or 2-3
Moz) may be only several thousand years (Krupp and
Seward, 1987).
It is possible that the high heat flow in evidence at
Bralorne today, and presumably present at the time
348 LEITCH, GODWIN, BROWN, AND TAYLOR
of mineralization, was both indicative of the flow of
large volumes of hydrothermal solutions, and favor-
able for efficient gold mobilization from source rocks
at depth. The importance of a high geothermal gra-
dient to generation of auriferous metamorphic fluids
has been noted by Groves et al. (1987). The well-
developed fracturing within the major Bralorne fault
zone provided excellent channelways, or plumbing,
to focus solutions from large volumes of rock and is
probably the main reason that these deposits are found
in this zone (cf. Groves et al., 1987; Nesbitt, 1988).
Conclusions
The Bralorne deposit is located within a regional
brittle fault zone. Such major crustal structures are
also associated with many of the large mining camps
of the Superior province in Canada (Hodgson et al.,
1982) and the Yilgarn block in Australia (Phillips,
1986). Rocks in the mine area show sub- to lower
greenschist facies metamorphism, at 325 to 375C
and 2 to 3 kbars H20 pressures. As at other mesother-
mal gold quartz vein deposits, alteration envelopes
grade outward from intensely foliated quartz-ankeritic
carbonate-sericite (_ fuchsite) to less sheared calcite-
chlorite-albite to unsheared epidote-calcite. Chemical
studies of the altered rocks show that there has been
addition of K20, CO=, S, As, and Au; Na=O, FeO, and
MgO have been depleted; SiO2 and CaO are locally
depleted and reconcentrated. These results reflect
replacement of albite by muscovite, and hornblende
by chlorite and ankerite. Gold is found mainly as thin
smeared flakes in the black sulfidic septae of the
strongly ribboned shear veins. Concentrations of py-
rite, pyrrhotite, and much lesser chalcopyrite drop
off sharply within a few meters of the veins; arseno-
pyrite is found immediately adjacent to the veins.
Small amounts of sphalerite, and especially galena,
correlate with gold-rich portions of the veins.
A fault-valve model best explains the ribboned, yet
euhedral, coarsely crystalline milky quartz veins.
Cyclic build-up of pressure in a reservoir of geopres-
sured fluids below the brittle-ductile transition caused
overpressuring, invoking failure by reactivation of the
previously formed steeply dipping faults, unfavorably
oriented for failure at a high angle to the maximum
compressive stress in a transpressive regime. Failure
provoked strong discharge of the ponded fluids and
the pressure release promoted deposition of quartz
and sulfides, with zoned quartz crystals deposited in
space held open by the high pore pressures. Sealing
of the fault by this mineral deposition allowed pres-
sure to build and the cycle to repeat. Each of the
ribbons of sulfide, with minor gold, may represent a
sliver of highly replaced wall rock that was included
in the vein when the next episode of fracturing, di-
lation, and mineral deposition occurred.
A vertical zonation of gradually increasing tem-
perature with depth at about 30C/km is apparent
from studies of fluid inclusions by both optical and
decrepitation methods and is reflected by the system-
atic variation in the o;xygen isotope ratios measured
in vein quartz. However, there is no appreciable
change in ore grade, sulfide or gangue mineral assem-
blages, or rock alteration, from surface to an almost
2-km depth. Such a lack of zoning implies ore de-
position from a large, widely circulating hydrothermal
system that essentially cooled only slowly upward at
a geothermal gradient. The Bralorne fault zone was
instrumental in providing abundant ground prepa-
ration, or plumbing, for the veins, and in focusing the
discharge of large volumes of fluid containing minute
quantities of gold.
Oxygen isotope ratios of vein quartz and associated
micas show that there has been significant wall-rock
reaction with the mineralizing fluids, suggesting high
water-rock ratios near the veins. Calculated oxygen
isotope compositions of mineralizing fluids are ap-
proximately constant at 13 _ 1.5 per rail over the 2-
km vertical extent of the deposit. Measures of vein
carbonates suggest that the carbon isotope ratio
ranged from -12 per mil in the early ore fluids to
-9.5 per mil in the latest stages. These values are too
negative to be derived purely from magmatic sources
but are consistent with the circulation of fluids
through sediments containing organic matter. Sulfur
isotope ratios of sulfides associated with the gold min-
eralization range from -7 to +9 per mil, clustering
about 0 per mil. It is possible to derive these values
from a felsic igneous source as well as from sedimen-
tary and greenstone sources in the intruded rocks.
The isotopic evidence, coupled with the need for CO
to cause the widespread carbonate alteration, suggests
that the source for the ore fluid probably was meta-
morphic devolatilization at greenschist or amphibolite
facies. Contributions from deeply circulated, highly
evolved meteoric and magmatic waters are possible.
Although the source(s) of the ore fluid remains un-
resolved, its characteristics can be reasonably well
defined. Based on fluid inclusion studies, primary ore
deposition appears to have been from fluids of low
salinity (3 wt % NaC1 equiv) with a significant CO
content (mole fraction of 0.10, rarely 0.25), minor
CH4 content (mole fraction of 0.05, rarely 0.25), and
densities of 0.95 g/cm 3, at temperatures of 350C
and pressures of up to 1.75 kbars (7 km depth). Later
pseudosecondary fluids were even more dilute (1 wt
% NaC1 equiv), with much lower CO2 (0.05 mole
fraction), no CH4, and densities of 0.97 g/cm 3, at
temperatures below 250C and pressures of 1.0 kbar.
Secondary fluids contained less than 1 wt percent
NaC1 equiv, with nondetectable CO and densities of
1.0 g/cm 3, at temperatures of 180C. The progression
of fluids may be explained by dilution of a hot, car-
BRALORNE Au DEPOSIT, B.C. 349
bonic fluid with increasing amounts of cool, less saline
meteoric water.
Computer modeling of the fluid responsible for the
observed alteration assemblages and gold deposition,
using chloride complexes, is supported by the ob-
served CO2 and CH4 in the fluid inclusions. The ore
fluid was slightly acid (pH = 4.5) at the 350C and
1.75 kbars suggested by fluid inclusion studies and
was a dilute solution of about 3 wt percent NaC1 equiv
(Na/K = 8:1), containing significant CO2 (log fugacity
-- 2.5) and minor CH4. It was reducing (fo about
10 -a bars, andJ about 10 -7 bars) and contained a
calculated 0.2 X 10-" m AuC12 (0.2 ppb Au). This is
significantly less than most previously published es-
timates of 20 ppb but is about the same as average
present-day mineralizing fluids. The main deposition
of gold was accompanied by deposition of quartz,
muscovite, ankerite, pyrite, and minor graphite.
Chlorite, pyrrhotite, and albite-stable areas are not
compatible with gold deposition. Gold precipitation
appears to have been by reduction ofAu + by donation
of an electron from a concurrent oxidation of sulfur
in HS to form pyrite. Deposition occurred mainly in
response to reaction with, and pyritization of, wall
rock. The results of this study suggest that further
thermodynamic modeling, using thiosulfide com-
plexes and considering temperature variations, is
warranted.
Acknowledgments
Thanks are extended to the Corona Corporation
for access to underground workings and field support.
Critical reviews of the manuscript by two Economic
Geology reviewers are appreciated. Research at the
University of British Columbia has been supported
by an I. W. Killam predoctoral fellowship to the senior
author and grants to Godwin from the Natural Sci-
ences and Engineering Research Council of Canada,
the British Columbia Ministry of Energy, Mines and
Petroleum Resources (BCMEMPR), and the Canada-
British Columbia Mineral Development Agreement.
Fluid inclusion studies were completed using equip-
ment at the BCMEMPR in Victoria. CO2 from oxygen
isotope extractions, performed at the Geological Sur-
vey of Canada (Ottawa) with the assistance of W. C.
Cornell, was analyzed in a mass spectrometer at the
Ottawa Center for Geoscience Studies/Geological
Survey of Canada Stable Isotope Laboratory by G. St.
Jean.
May 15, December 6, 1990
REFERENCES
Albino, G. V., 1987, Contrasting styles of gold mineralization in
the Sierra Nevada metamorphic belt, California: Unpub. manu-
script, Univ. Western Ontario, Dept. Geology, 26 p.
Andrews, A. J., Jugon, H., Durocher, M., Corfu, F., and Lavigne,
M. J., 1986, The anatomy of a gold-bearing greenstone belt:
Red Lake, northwestern Ontario, Canada, in Macdonald, A. J.,
ed., Gold '86: Willowdale, Ontario, Konsult Internat., p. 3-26.
Anovitz, L. M., and Essene, E. J., 1987, Phase equilibria in the
system CaCO-MgCO3-FeCO: Jour. Petrology, v. 28, p. 389-
414.
Bancroft, G. M., and Jean, G., 1982, Gold deposition at low tem-
perature on sulphide minerals: Nature, v. 298, p. 730-731.
Battey, M. H., 1951, Alkali metasomatism and the petrology of
some keratophyres: Geol. Mag., v. 92, p. 104-126.
Bellamy, J. R., and Arnold, R., 1985, Bralorne mine project: Report
on the 1984 exploration work: Vancouver, British Columbia,
E&B Explorations Inc., March 1985, unpub. rept., 25 p.
Bertoni, C. H., 1983, Gold production in the Superior province
of the Canadian Shield: Canadian Inst. Mining Metallurgy Bull.,
v. 76, no. 857, p. 62-69.
Bohlke, J. K., and Kistler, R. W., 1986, Rb-Sr, K-At, and stable
isotope evidence for the ages and sources of fluid components
of gold-bearing quartz veins in the northern Sierra Nevada
foothills metamorphic belt, California: ECON. GEOL., v. 81, p.
296-322.
Borthwick, J., and Harmon, R. S., 1982, A note regarding CIF3
as an alternative to BrF5 for oxygen isotope analysis: Geochim.
et Cosmochim. Acta, v. 46, p. 1665-1668.
Bottinga, Y., and Javoy, M., 1975, Oxygen isotope partitioning
among the minerals in igneous and metamorphic rocks: Rev.
Geophysics Space Physics, v. 13, p. 401-418.
Bowers, T. S., and Helgeson, H. C., 1983a, Calculation of the
thermodynamic and geochemical consequences of nonideal
mixing in the system HO-CO-NaCI on phase relations in geo-
logic systems: Equation of state for HO-CO-NaCI fluids at
high pressures and temperatures: Geochim. et Cosmochim.
Acta, v. 47, p. 1244-1275.
1983b, Calculation of thermodynamic and geochemical
cousequenees of nonideal mixing in the system HO-CO-NaCI
on phase relations in geologic systems: Metamorphic equilibria
at high pressures and temperatures: Am. Mineralogist, v. 68,
p. 1059-1075.
Boyle, B. W., 1961, The geology, geochemistry and origin of the
gold deposits of the Yellowknife district: Canada Geol. Survey
Mem. 310, 135 p.
1979, The geochemistry of gold and its deposits: Canada
Geol. Survey Bull. 280, p. 228-232.
Bozzo, A. T., Chen, J. B., and Barduhn, A. J., 1973, The properties
of the hydrates of chlorine and carbon dioxide, in Delyanni, A.,
and Delyanni, E., eds., Internat. Symposium on Fresh Water
from the Sea, 4th, v. 3, p. 437-451.
Brown, K. L., 1986, Gold deposition from geothermal discharges
in New Zealand: ECON. GEOL., v. 81, p. 979-983.
Brown, P. E., and Lamb, W. M., 1986, Mixing of HO-CO in
fluid inclusions; geobarometry and Arebean gold deposits:
Geochim. et Cosmochim. Acta, v. 50, p. 847-852.
Buisson, G., and Leblanc, M., 1987, Gold in mantle peridotites
from upper Proterozoic ophiolites in Arabia, Mall, and Morocco:
ECON. GEOL., v. 82, p. 2091-2097.
Cairnes, C. E. (1937): Geology and mineral deposits of the Bridge
River mining camp, B.C.: Canada Geol. Survey Mem. 213, 140
P.
Campbell, D. D., 1975, Bridge River gold camp: Vancouver,
B.C., Dolmage Campbell and Assoc., Ltd., unpub. rept., 7 p.
Cameron, E. M., 1989a, Derivation of gold by oxidative meta-
morphism of a deep ductile shear zone: Part 1. Conceptual
model: Jour. Geochem. Explor., v. 31, p. 135-144.
-- 1989b, Derivation of gold by oxidative metamorphism of a
deep ductile shear zone: Part 2. Evidence from the Ramble
belt, south Norway: Jour. Geochem. Explor., v. 31, p. 145-
149.
-- 1989c, Review of"Archean Lode Gold Deposits in Ontario,"
by A. C. Colvine, J. A. Fyon, K. B. Heather, S. Marmount,
P.M. Smith and D. G. Troop, Ontario Geological Survey Mis-
350 LEITCH, GODWIN, BROWN, AND TAYLOR
cellaneous Paper 139, 1988: Jour. Geochem. Explor., v. 31, p.
329-334.
Church, B. N., 1987, Geology and mineralization of the Bridge
River mining camp (92J/15, 920/2, 92J/10): British Columbia
Ministry Energy, Mines and Petroleum Resources, Geol. Field-
work, 1986, Paper 1987-1, p. 23-29.
Clayton, R. N., and Mayeda, T. K., 1963, The use of bromine
pentafiuoride in the extraction of oxygen from oxides and sili-
cates for isotopic analyses: Geochim. et Cosmochim. Acta, v.
27, p. 43-52.
Clayton, R. N., O'Neil, J. R., and Mayeda, T. K., 1972, Oxygen
isotope exchange between quartz and water: Jour. Geophys.
Research, v. 77, p. 3057-3067.
Collins, P. L. F., 1979, Gas hydrates in COg-bearing fluid inclu-
sions and the use of freezing data for estimation of salinity:
ECON. GEOL., v. 74, p. 1435-1444.
Colvine, A. C., Andrews, A. J., Cherry, M. E., Durocher, M. E.,
Fyon, A. J., Lavigne, M. J. Jr., Macdonald, A. J., Marmount, S.,
Poulsen, K. H., Springer, J. S., and Troop, D. G., 1984, An
integrated model for the origin of Archean lode gold deposits:
Ontario Geol. Survey Open-File Rept. 5524, 103 p.
Colvine, A. C., Fyon, J. A., Heather, K. B., Marmount, S., Smith,
P.M., and Troop, D. G., 1988, Archean lode gold deposits in
Ontario: Ontario Geol. Survey Misc. Paper 139, 136 p.
Coveney, R. M., Jr., 1981, Gold quartz veins and auriferous granite
at the Oriental mine, Alleghany district, California: ECON.
GEOL., v, 76, p. 2176-2199.
Dolmage, V., 1934, The Cariboo and Bridge River goldfields,
British Columbia: Canadian Inst. Mining Metall. Trans., v. 37,
p. 405-430.
Fuchter, W. A. H., and Hodgson, C. J., 1986, Gold deposits of
the North Western mining camp, Gwanda greenstone belt,
Zimbabwe, in Macdonald, A. J., ed., Gold '86: Willowdale, On-
tario, Konsult Internat., p. 255-269.
Fyfe, W. S., and Henley, R. W., 1973, Some thoughts on chemical
transport processes with particular reference to gold: Minerals
Sci. Eng., v. 5, p. 295-303.
Fyfe, W. S., and Kerrich, R., 1984, Gold: Natural concentration
processes, in Foster, R. P., ed., Gold '82: Rotterdam, A. A.
Balkema Pub., p. 99-127.
Godwin, C. I., Watson, P. H., and Shen, K., 1986, Genesis of the
Lass vein system, Beaverdell silver camp, south-central British
Columbia: Canadian Jour. Earth Sci., v. 23, p. 1615-1626.
Goldfarb, R. J., Leach, D. L., Pickthorn, W. J., and Paterson,
C. J., 1988, Origin of lode-gold deposits of the Juneau gold
belt, southeastern Alaska: Geology, v. 16, p. 440-443.
Goldfarb, R. J., Leach, D. L., Rose, S.C., and Landis, G. P., 1989,
Fluid inclusion geochemistry of gold-bearing quartz veins of
the Juneau gold belt, southeastern Alaska: Implications for ore
genesis: ECON. GEOL. MON. 6, p. 363-375.
Gresens, R, L., 1967, Composition-volume relationships of meta-
somatism: Chem. Geology, v. 2, p. 47-55.
Groves, D. I., Phillips, G. N., Ho, S. E., Henderson, C. A., Clark,
M. E., and Woad, G. M., 1984, Controls on distribution of Ar-
chean hydrothermal gold deposits in Western Australia, in Fos-
ter, R. P., ed., Gold '82: Rotterdam, A. A. Balkema Pub., p.
689-711.
Groves, D, I., Phillips, G. N., Ho, S. E., Houston, S. M., and Stand-
ing, C. A., 1987, Craton-scale distribution of Archeau green-
stone gold deposits: Predictive capacity of the metamorphic
model: ECON. GEOL., v. 82, p. 2045-2058.
Guha, J., Gauthier, A., Vallee, M., Descarreaux, J., and Lange-
brard, F., 1982, Gold mineralization patterns at the Doyon mine
(Silverstack, Bousquet, Quebec): Canadian Inst. Mining Met-
allurgy Spec. Vol. 24, p. 50-57.
Hall, R. S., and Rigg, D. M., 1986, Geology of the West Anticline
zone, Musselwhite prospect, Opapimiskan Lake, Ontario, Can-
ada, in Macdonald, A. J., ed., Gold '86: Willowdale, Ontario,
Konsult Internat., p. 124-136.
Haugerud, R., 1985, Geology of the Hozameen Group and Ross
Lake shear zone, Maselpanic area, northern Ctscades, south-
western British Columbia: Unpub. Ph.D. thesis, Seattle, Univ.
Washington, 263 p.
Helgeson, H. C., 1969, Thermodynamics of hydrothermal systems
at elevated temperatures and pressures: Am. Jour. Sci., v. 167,
p. 729-804.
Helgeson, H. C., and Garrels, R. M., 1968, Hydrothermal transport
and deposition of gold: ECON. GEOL., v. 63, p. 622-635.
Helgeson, H. C., and Lichtner, P. C., 1987, Fluid flow and mineral
reactions at high temperatures and pressures: Geol. Soc. London
Jour., v. 144, p. 313-326.
Helgeson, H. C., Brown, T. H., Nigrini, A., and Jones, T. A., 1970,
Calculation of mass transfer in geochemical processes involving
aqueous solutions: Geochim. et Cosmochim. Acta, v. 34, p. 569-
592.
Helgeson, H. C., Delany, J. M., Nesbitt, H. W., and Bird, D. K.,
1978, Summary and critique of the thermodynamic properties
of rock forming minerals: Am. Jour. Sci., v. 278-A, 229 p.
Hodgson, C. J., and MacGeehan, P. J., 1982, A review of the
geologic characteristics of "gold only" deposits in the Superior
province of the Canadian Shield: Canadian Inst. Mining Met-
allurgy Spec. Vol. 24, p. 211-229.
Hodgson, C. J., Chapman, R. S. G., and MacGeehan, P. J., 1982,
Application of exploration criteria for gold deposits in the Su-
perior province of the Canadian Shield to gold exploration in
the Cordillera, in Levinson, A. A., ed., Precious metals in the
northern cordillera: Calgary, Alberta, Assoc. Explor. Geo-
chemists, p. 174-206.
Hollister, L. S., 1981, Information intrinsically available from fluid
inclusions: Mineralog. Assoc. Canada Short Course Handbook,
v. 6, p. 1-12.
Hollister, L. S., and Burruss, R. C., 1976, Phase equilibria in fluid
inclusions from the Khtada Lake metamorphic complex: Geo-
chim. et Cosmochim. Acta, v. 40, p. 163-175.
Hutchinson, R. W., and Burlington, J. L., 1984, Some broad char-
acteristics of greenstone belt gold lodes, in Macdonald, A. J.,
ed., Gold '86: Willowdale, Ontario, Konsult Internat., p. 339-
371.
Joubin, F. R., 1948, Structural geology of the Bralorne and Pioneer
mines, Bridge River district, British Columbia: Western Miner,
July 1948, p. 39-50.
Keays, R. R., 1984, Archeau gold deposits and their source rocks:
The upper mantle comection, in Macdonald, A. J., ed., Gold
'86: Willowdale, Outario, Konsult Internat., p. 17-51.
Kerr, P. F., 1959, Optical mineralogy, 3rd ed.: New York,
McGraw-Hill, 442 p.
Kerrich, R., 1983, Geochemistry of gold deposits in the Abitibi
greenstone belt: Canadian Inst. Mining Metallurgy Spec. Vol.
27, 75 p.
-- 1987, The stable isotope geochemistry of Au-Ag vein de-
posits in metamorphic rocks: Mineralog. Assoc. Canada Short
Course Handbook, v. 13, p. 287-336.
Kerrich, R., and Fyfe, W. S., 1981, The gold-carbonate tssociation:
Source of COg, and COg fixation reactions in Archeau lode de-
posits: Chem. Geology, v. 33, p. 265-294.
Kerrich, R., and Watson, G. P., 1984, The Mactssa mine Archeau
lode gold deposit, Kirkland Lake, Ontario: Geology, patterns
of alteration, and hydrothermal regimes: ECON. GEOL., v. 79,
p. 1104-1130.
Kishida, A., and Kerrich, R., 1987, Hydrothermal alteration zoning
and gold concentration at the Kerr-Addison Archeau lode gold
deposit, Kirkland Lake, Ontario: ECON. GEOL., v. 82, p. 649-
690.
Krupp, R. E., and Seward, T. M., 1987, The Rotokawa geothermal
system, New Zealand: An active epithermal gold-depositing
environment: ECON. GEOL., v. 82, p. 1109-1129.
Landfeldt, L. A., 1987, The geology of the Mother Lode gold
belt, Sierra Nevada Foothills metamorphic belt, California, in
BRALORNE An DEPOSIT, B.C. 351
Kisvarsanyi, G., and Grant, S. R., eds., North American con-
ference on tectonic control of ore deposits and the vertical and
horizontal extent of ore systems: Proceedings volume: Rolla,
Univ. Missouri-Rolla Press, p. 47-56.
Leitch, C. H. B., 1981, Secondary alkali feldspars in porphyry
systems: Canadian Inst. Mining Metallurgy Bull., v. 74, no. 831,
p. 83-88.
-- 1989, Geology, wallrock alteration, and characteristics of
the ore fluid at the Bralorne mesothermal gold quartz vein de-
posit, southwestern B.C.: Unpub. Ph.D. thesis, Vancouver,
Univ. British Columbia, 483 p.
-- 1990, Bralorne: A mesothermal, shield-type vein gold de-
posit of Cretaceous age in southwestern British Columbia: Ca-
nadian Inst. Mining Metallurgy Bull., v. 83, no. 941, p. 53-80.
Leitch, C. H. B., and Day, S. J., 1990, NEWGRES: A TurboPascal
program to solve a modified version of Gresens' hydrothermal
alteration equation: Computers Geosci., v. 16, p. 925-932.
Leitch, C. H. B., and Godwin, C. I,, 1986, Geology of the Bralorne-
Pioneer gold camp (92J/15): British Columbia Ministry Energy,
Mines Petroleum Resources, Geol. Fieldwork 1985, Paper
1986-1, p. 311-317.
-- 1987, The Bralorne gold vein deposit: An update (92J/15):
British Columbia Ministry Energy, Miues Petroleum Resources,
Geol. Fieldwork 1986, Paper 1987-1, p. 35-38.
-- 1988, Isotopic ages, wallrock chemistry and fluid inclusion
data from the Bralorne gold vein deposit (92J/15w): British
Columbia Ministry Energy, Mines Petroleum Resources, Geol.
Fieldwork 1987, Paper 1988-1, p. 301-324.
Leitch, C. H. B., Dawson, K. M., and Godwin, C. I., 1989, Early
Late Cretaceous-early Tertiary gold mineralization: A galena
lead isotope study of the Bridge River mining camp, south-
western British Columbia, Canada: ECON. GEOL., v. 84, p.
2226--2236.
Leitch, C. H. B., van der Heyden, P., Godwin, C. I., Armstrong,
R. L., and Harakal, J., 1991, Geochronometry of the Bridge
River camp, southwestern British Columbia: Canadian Jour.
Earth Sci., v. 28, no. 2, p. 261-274.
Leroy, J., 1979, Contribution t l'etalinnage de la pression interne
des inclusions fluides lor de leur decrepitation: Bull. Mineral-
ogle, v. 102, p. 584-593.
Linnen, R., 1985, Contact metamorphism, wallrock alteration,
and mineralization at the Trout Lake stockwork molybdenum
deposit, southeastern British Columbia: Unpub. M.Sc. thesis,
McGill Univ., 220 p.
Ludden, J. N., Daigneault, R., Robert, F., and Taylor, R. E., 1984,
Trace element mobility in alteration zones associated with Ar-
chcan Au lode deposits: ECON. GEOL., v. 79, p. 1131-1141.
Lydon, J. W., 1977, The significance of metal ratios of hydro-
thermal ore deposits: Unpub. Ph.D. thesis, Kingston, Queen's
Univ, 353 p.
MaeGeehan, P. J., Sanders, T., and Hodgsou, C. J., 1982, Meter-
wide veins and a kilometer-wide anomaly: Wall-rock alteration
at the Campbell Red Lake aud Dickenson gold mines, Red Lake
district, Ontario: Canadian Inst. Mining Metallurgy Bull., v. 75,
no. 841, p. 90-102.
Magaritz, M., and Taylor, H. P., Jr., 1976, O/60 and D/H studies
along a 500 km traverse across the Coast Range batholith and
its country rocks, central British Columbia: Canadian Jour. Earth
Sci., v. 13, p. 1514-1536.
1986, SO/60 and D/H studies of plutonie granitic and
metamorphic rocks across the Cordilleran batholiths of southern
British Columbia: Jour. Geophys. Research, v. 91, p. 2193-
2217.
Maheux, P. J., Muehlenbachs, K., aud Nesbitt, B. E., 1987, Evi-
dence of highly evolved ore fluids responsible for sulfide as-
sociated gold mineralization iu the Bridge River district, B.C.
labs.l: Geol. Assoc. Canada Program with Abstracts, v. 12, p.
70.
Malinin, S. D., 1974, Thermodynamics of the H20-CO2 system:
Geochem. Internat., v. 11, p. 1060-1085.
Matthews, A., Goldsmith, J. R., and Clayton, R. M., 1983, Oxygen
isotope fractionations involving pyroxenes: The calibration of
mineral-pair geothermometers: Geochim. et Cosmochim. Acta,
v. 47, p. 631-644.
McCrea, J. M., 1950, The isotopic chemistry of carbonates and a
palcotemperature scale: Jour. Chem. Physics, v. 18, p. 849-
853.
McHugh, J. B., 1988, Concentration of gold in natural waters:
Jour. Geochem. Explor., v. 30, p. 85-94.
Melling, D. R., Watkinson, D. H., Poulsen, K. H., Chorlton,
L. B., and Hunter, A.D., 1986, The Cameron Lake gold deposit,
northwestern Ontario, Canada: Geological setting, structure,
and alteration, in Macdonald, A. J., ed., Gold '86: Willowdale,
Ontario, Konsult Internat., p. 149-169.
Monger, J. W. H., 1984, Cordilleran tectonics: A Canadian per-
spective: Soc. Geol. France Bull., ser. 7, v. 26, no. 2, p. 255-
278.
Morasse, S., Hodgson, C. J., Guha, J., and Coulombe, A., 1986,
Preliminary report on the geology of the Lac Shortt gold deposit,
Demaraisville area, Quebec, Canada, in Macdonald, A. J., ed.,
Gold '86: Willowdale, Ontario, Konsult Internat., p. 191-196.
Nesbitt, B. E., 1988, Gold deposit continuum: A genetic model
for lode Au mineralization in the continental crust: Geology,
v. 16, p. 1044-1048.
Nesbitt, B. E., and Muehlenbachs, K., 1988, Genetic implications
of the association of mesothermal gold deposits with major
strike-slip fault systems, in Kisvarsanyi, G., and Grant, S. R,,
eds., North American conference on tectonic control of ore
deposits and the vertical and horizontal extent of ore systems:
Proceedings volume: Rolla, Univ. Missouri-Rolla Press, p. 57-
66.
Nesbitt, B. E., Murowchick, J, B. and Muehlenbachs, K., 1986,
Dual origins of lode gold deposits in the Canadian Cordillera:
Geology, v. 14, p. 506-509,
-- 1987, Genesis ofAu, Sb, and Hg deposits in accreted ter-
ranes of the Canadian Cordillera labs.l: Geol. Assoc. Canada
Program with Abstracts, v. 12, p. 70.
Nordine, G., 1983, Geological report on the Pacific Eastern prop-
erty, Lillooet Mining Division: Vancouver, British Columbia,
Amir Mines, Ltd., priv. rept., 48 p.
Ohmoto, H., and Rye, R. O., 1979, Isotopes of sulfur and carbon,
in Barnes, H. L., ed., Geochemistry of hydrothermal ore de-
posits: New York, Wiley Intersci., p. 509-567.
Ohmoto, H., and Kerrick, D. M., 1977, Devolatilization equilibria
in graphitic systems: Am. Jour. Sci., v. 277, p. 1013-1044.
O'Neil, J. R., and Taylor, H. P., Jr., 1969, Oxygen isotope equi-
librium between muscovite and water: Jour. Geophys. Research,
v. 74, p, 6012-6022.
Parry, W. T., 1986, Estimation of Xco, P, and fluid inclusion
volume from fluid inclusion temperature measurements in the
system NaCI-CO2-H20: ECON. GEOL., v. 81, p. 1009-1013.
Pattison, E. F., Sauerbrei, J. A. Hannila, J. J., and Church, J. F.,
1986, Gold mineralization in the Casa-Berardi area, Quebec,
Canada, in Macdonald, A. J., ed., Gold '86: Willowdale, Ontario,
Konsult Internat., p. 170-183.
Perkins, E. H., 1980, A re-investigation of the theoretical basis
for the calculation of isothermal-isobaric mass transfer in geo-
chemical systems involving an aqueous phase: Unpub. M.Sc.
thesis, Vancouver, Univ. British Columbia, 176 p.
Phillips, G. N., 1986, Geology and alteration in the Golden Mile,
Kalgoorlie: ECON. GEOL., v. 81, p. 779-808.
Phillips, G. N., and Groves, D. I., 1983, The nature of Archcan
gold-bearing fluids as deduced from gold deposits of Western
Australia: Geol. Soc. Australia Jour., v. 30, p. 25-39.
-- 1984, Fluid access and fiuid-wallrock interaction in the
genesis of the Arcbean gold-quartz vein deposit at Hunt mine,
352 LEITCH, GODWIN, BROWN, AND TAYLOR
Kambalda, Western Australia, in Foster, R. P., ed., Gold '82:
Rotterdam, A. A. Balkema Pub., p. 389-416.
Phillips, G. N., Groves, D. I., and Martyn, J. E., 1983, An epi-
genetic origin for Archean banded iron-formation-hosted gold
deposits: E(ON. GEOL., v. 79, p. 162-171.
Pickthorn, W. L., Goldfarb, R. J., and Leach, D. L., 1987, Com-
ment and reply on "Dual origins of lode gold deposits in the
Canadian Cordillera": Geology, v. 15, p. 471-472.
Poole, A. W., 1955, The geology and analysis of vein and fault
structure of the Bralorne mine: Canadian Inst. Mining Metal-
lurgy Bull., November 1955, p. 733-737 (Trans., v. 58, p. 433-
437).
Potter, C. J., 1986, Origin, accretion, and postaccretionary evo-
lution of the Bridge River terrane, southwest British Columbia:
Tectonics, v. 5, p. 1027-1041.
Potter, R. W., and Brown, D. L., 1977, The volumetric properties
of aqueous sodium chloride solutions from 0 to 500C at pres-
sures up to 2000 bars based on a regression of available data
in the literature: U.S. Geol. Survey Bull. 1421-C, 36 p.
Potter, R. W., Clynne, M. A., and Brown, D. L., 1978, Freezing
point depression of aqueous sodium chloride solutions: ECON.
GEOL., v. 73, p. 284-285.
Ray, G. E., 1986, The Hozameen fault system and related Co-
quihalla serpentine belt of southwestern British Columbia: Ca-
nadian Jour. Earth Sci., v. 23, p. 1022-1041.
Read, J. J., and Meinert, L. D., 1986, Gold-bearing quartz vein
mineralization at the Big Hurrah mine, Seward Peninsula,
Alaska: ECON. GEOL., v. 81, p. 1760-1774.
Robert, F., and Brown, A. C., 1986a, Archean gold-bearing quartz
veins at the Sigma mine, Abitibi greenstone belt, Quebec: Part
I. Geologic relations and formation of the vein system: ECON.
GEOL., v. 81, p. 578-592.
-- 1986b, Archean gold-bearing quartz veins at the Sigma mine,
Abitibi greenstone belt, Quebec: Part II. Vein paragenesis and
hydrothermal alteration: ECON. GEOL., v. 81, p. 593-616.
Robert, F., and Kelly, W. C., 1987, Ore-forming fluids in Archean
gold-bearing quartz veins at the Sigma mine, Abitibi greenstone
belt, Quebec, Canada: ECON. GEOL., v. 82, p. 1464-1482.
Roberts, R. G., 1987, Ore deposit models #11: Archean lode gold
deposits: Geosci. Canada, v. 14, no. 1, p. 37-52.
Roedder, E., 1979, Fluid inclusions as samples of ore fluids, in
Barnes, H. L., ed., Geochemistry of hydrothermal ore deposits:
New York, Wiley Intersci., p. 684-737.
-- 1984, Fluid inclusions: Rev. Mineralogy, v. 12, 644 p.
Romberger, S. B., 1986, The solution chemistry of gold applied
to the origin of hydrothermal deposits: Canadian Inst. Mining
Metallurgy Spec. Vol. 38, p. 168-186.
Rusmore, M. E., 1987, Geology of the Cadwallader Group and
the Intermontane-Insular superterrane boundary, southwestern
British Columbia: Canadian Jour. Earth Sci., v. 24, p. 2279-
2291.
Rusmore, M. E., Potter, C. J., and Umhoefer, P. J., 1988, Middle
Jurassic terrane accretion along the western edge of the Inter-
montane superterrane, southwestern British Columbia: Geol-
ogy, v. 16, p. 891-894.
Rye, R. O., and Ohmoto, H., 1974, Sulfur and carbon isotopes
and ore genesis: A review: ECON. GEOL., v. 69, p. 826-842.
Santosh, M., 1986, Ore fluids in the auriferous Champion reef of
Kolar, South India: ECON. GEOL., v. 81, p. 1546-1552.
Schiarizza, P., Gaba, R. G., Glover, J. K., and Garver, J. I., 1989,
Geology and mineral occurrences of the Tyaughton Creek area
(920/2, 92J/15, 16): British Columbia Ministry Energy, Mines
Petroleum Resources, Geol. Fieldwork 1988, Paper 1989-1, p.
115-130.
Seward, T. M., 1973, Thio complexes of gold and the transport
of gold in hydrothermal ore solutions: Geochim. et Cosmochim.
Acta, v. 37, p. 379-399.
-- 1984, The transport and deposition of gold in hydrothermal
systems, in Foster, R. P., ed., Gold '82: Rotterdam, A. A. Bal-
kema Pub., p. 165-181.
Shelton, K. L., So, C.-S., and Chang, J.-S., 1988, Gold-rich me-
sothermal vein deposits of the Republic of Korea: Geochemical
studies of the Jungwon gold area: ECON. GEOL., v. 83, p. 1221-
1237.
Shenberger, D. M., 1985, Gold solubility in aqueous sulfide so-
lutions: Unpub. M. Sc. thesis, Pennsylvania State Univ., 102 p.
Sibson, R. H., 1981, Fluid flow accompanying faulting: Field ev-
idence and models: Am. Geophys. Union, Maurice Ewing Ser.,
v. 4, p. 593-603.
Sibson, R. H., Robert, F., and Poulsen, K. H., 1988, High-angle
reverse faults, fluid-pressure cycling, and mesothermal gold-
quartz deposits: Geology, v. 16, p. 551-555.
Sketchley, D. A., and Sinclair, A. J., 1987, Multi-element litho-
geochemistry of alteration associated with gold-quartz miner-
alisation of the Erickson mine, Cassiar district: British Columbia
Ministry Energy, Mines Petroleum Resources, Geol. Fieldwork
1986, Paper 1987-1, p. 57-63.
Smith, T. J., Cloke, P. L., and Kesler, S. E., 1984, Geochemistry
of fluid inclusions from the McIntyre-Hollinger gold deposit,
Timmins, Ontario, Canada: ECON. GEOL., v. 79, p. 1265-1285.
Stevenson, J. S., 1958, Bridge River area, British Columbia: British
Columbia Ministry Energy, Mines Petroleum Resources, unpub.
rept., 320 p.
Studemeister, P. A., and Kilias, S., 1987, Alteration pattern and
fluid inclusions of gold-bearing quartz veins in Archean
trondhjemite near Wawa, Ontario, Canada: E(ON. GEOL., v.
82, p. 429-439.
Sugiyama, T., 1986, Short report on preliminary test of new de-
crepitation system on samples from University of British Co-
lumbia, Canada: Tokyo, Japan, Mitsubishi Metal Corp., Central
Research Inst., unpub. rept., 16 p.
Swanenberg, H. E. C., 1979, Phase equilibria in carbonic systems,
and their application to freezing studies of fluid inclusions: Cont.
Mineralogy Petrology, v. 68, p. 303-306.
Taylor, B. E., 1987, Stable isotope geochemistry of ore-forming
fluids: Mineralog. Assoc. Canada Short Course Handbook, v.
13, p. 337-418.
Taylor, B. E., and Gerlach, T. M., 1986, Mantle degassing at Long
Valley, Steamboat Springs, and the Coso Range [abs.]: EOS, v.
65, p. 1153.
Taylor, H. P., Jr., 1974, The application of oxygen and hydrogen
isotope studies to problems of hydrothermal alteration and ore
deposition: ECON. GEOL., v. 69, p. 843-883.
-- 1979, Oxygen and hydrogen isotope relationships in hy-
drothermal ore deposits, in Barnes, H. L., ed., Geochemistry
of hydrothermal ore deposits: New York, Wiley Intersci., p.
236-277.
Thompson, M. L., 1986, Petrology of the Crixas gold deposit,
Brazil: Evidence for gold associated with hydrothermal alter-
ation, subsequent to metamorphism, in Macdonald, A. J., ed.,
Gold '86: Willowdale, Ontario, Konsult Internat., p. 284-296.
Tilling, R. I., Gottfried, D., and Rowe, J. J., 1973, Gold abundance
in igneous rocks: Bearing on gold mineralization: ECON. GEOL.,
v. 68, p. 168-186.
Trommsdorff, V., and Skippen, G., 1986, Vapour loss ("boiling")
as a mechanism for fluid evolution in metamorphic rocks: Contr.
Mineralogy Petrology, v. 94, p. 317-322.
Walsh, J. F., Kesler, S. E., Duff, D., and Cloke, P. L., 1988, Fluid
inclusion geochemistry of high-grade, vein-hosted gold ore at
the Pamour mine, Porcupine camp, Ontario: ECON. GEOL., v.
83, p. 1347-1368.
Walther, J. V., and Helgeson, H. C., 1977, Calculation of the
thermodynamic properties of aqueous silica and the solubility
BRALORNE Au DEPOSIT, B. C. 353
of quartz and its polymorphs at high pressures and temperatures:
Am. Jour. Sci., v. 277, p. 1315-1351.
Weir, R. H., Jr., and Kerrick, D. M., 1987, Mineralogic, fluid
inclusion, and stable isotope studies of several gold mines in
the Mother Lode, Tuolumne and Mariposa Counties, California:
ECON. GEOL., v. 82, p. 328-344.
Wenner, D. B., and Taylor, H. P., Jr., 1971, Temperatures of
serpentinization of ultramarie rocks based on lSO/160 fraction-
ation between coexisting serpentine and magnetite: Contr.
Mineralogy Petrology, v. 32, p. 165-185.
Wheeler, J. O., and McFeely, P., 1987, Tectonic assemblage map
and correlation chart of the Canadian Cordillera and adjacent
parts of the United States of America: Canada Geol. Survey
Open-File Rept. 1.56,5 (1:2,000,000 scale map).
Winkler, H. G. F., 197 l, Petrogenesis of metamorphic rocks, 3rd
ed.: New York, Springer-Verlag, 237 p.
Wood, P. C., Burrows, D. R., Thomas, A. V., and Spooner,
E. T. C., 1986, The Hollinger-McIntyre Au-quartz vein system,
Timrains, Ontario, Canada: Geological characteristics, fluid
properties and light stable isotopes, in Macdonald, A. J., ed.,
Gold '86: Willowdale, Ontario, Konsult Internat., p..56-80.
Wood, S. A., 1987, Application of a multiphase ore mineral sol-
ubility experiment to the separation of base metal and gold
mineralization in Arebean greenstone terrains: ECON. GEOL.,
v. 82, p. 1044-1048.
Woodsworth, G. J., 1977, Pemberton (92J) map-area, British Co-
lumbia: Canada Geol. Survey Open-File Rept. 482 (1:2.50 000
scale map).
Workman, A. W., 1986, Geology of the McDermott gold deposit,
Kirkland Lake area, northeastern Ontario, Canada, in Macdon-
ald, A. J,, ed., Gold '86: Willowdale, Ontario, Konsult Internat.,
p. 184-190.

Anda mungkin juga menyukai