You are on page 1of 26

©2004 by Economic Geology

Vol. 99, pp. 0000–0000

Economic Geology
BULLETIN OF THE SOCIETY OF ECONOMIC GEOLOGISTS

VOL. 99 August 2004 NO. 5

Characteristics and Evolution of the Hydrothermal Fluid in the


North Zone High-Grade Area, Porgera Gold Deposit, Papua New Guinea
E. RONACHER,†,* J. P. RICHARDS,
Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta, Canada T6G 2E3

M. H. REED,
Department of Geological Sciences, University of Oregon, Eugene, Oregon 97403

C. J. BRAY, E. T. C. SPOONER,
F. Gordon Smith Fluid Inclusion Laboratory, Department of Geology, University of Toronto, Toronto, Ontario, Canada M5S 3B1

AND P. D. ADAMS**
Porgera Joint Venture, Porgera, Papua New Guinea

Abstract
The ~20-million-ounce (Moz) Porgera gold deposit, Papua New Guinea, is hosted by 6-m.y.-old alkalic in-
trusions and Cretaceous sedimentary rocks in which the intrusions were emplaced. Gold-bearing veins occur
in three stages: (1) magnetite-sulfide-carbonate ± quartz veins with minor gold (prestage I), (2) base metal-sul-
fide-carbonate ± quartz ± Au veins (stage I), and (3) quartz-roscoelite-pyrite-gold veins and breccias (stage II).
Stage II veins are economically the most significant.
Quartz-roscoelite-pyrite-gold veins form high-grade zones associated with the Roamane fault (a late normal
fault that crosscuts the intrusive complex) and in the footwall to the fault (the North zone). The North zone
mineralization is the main focus of this study. The quartz-roscoelite-pyrite-gold assemblage occurs in three tex-
turally distinct styles: (1) thin (1–5 mm) veinlets in which roscoelite-pyrite-gold are more abundant than quartz
and in which roscoelite and gold also occur in the wall rock; (2) veins (5 mm to 10 cm) in which roscoelite-
pyrite-gold with minor quartz form a band at the vein edges, followed by coarse-grained quartz and the vein
centers commonly filled with anhydrite and carbonate; and (3) breccia veins and breccias in which wall-rock
fragments are rimmed by roscoelite-pyrite-gold and minor quartz, followed by vuggy quartz infilling.
Fluid inclusions from quartz in these veins and breccias are mostly liquid rich, and average salinities in in-
dividual samples range from 7.5 ± 1.0 to 9.6 ± 0.2 wt percent NaCl equiv. In five of 27 samples, an additional
cluster of salinities between 4.4 and 6.2 wt percent NaCl equiv was observed. These relatively low salinity in-
clusions occur toward the vein center and are less abundant than high-salinity inclusions that occur toward the
vein margins. Three samples exhibit a continuous salinity trend from 4.5 to 10.2 wt percent NaCl equiv. For
samples where CO2 analyses were available average corrected salinities range from 5.1 to 8.0 wt percent NaCl
equiv. Average homogenization temperatures (Th) of individual samples range from 127° ± 12° to 167° ± 25°C.
The average Th of the low-salinity inclusions (145° ± 9°C) is marginally lower but overlaps with that of all high-
salinity inclusions (152° ± 17°C).
Gas chromatographic analyses showed that the high-salinity fluid contains up to 2 mol percent CO2, 0.11 mol
percent CH4, 0.065 mol percent N2, and traces of C2H4, C2H6, and COS. Concentrations of Cl– (310–609
mM/l), Br– (0.28–0.75 mM/l), Li+ (1.25–8.80 mM/l), Na+ (462–1126 mM/l), K+ (0–81 mM/l), Mg2+ (0–7.0
mM/l), and Ca2+ (0–185 mM/l) were determined by ion chromatography.

† Corresponding author: e-mail, e_ronacher@hotmail.com


*Present address: Department of Geology and Geological Engineering, Colorado School of Mines, Golden, Colorado 80401.
**Present address: Placer Granny Smith Pty., Ltd., Laverton, Western Australia 6440, Australia.

©2004 by Economic Geology, Vol. 99, pp. 843–867


0361-0128/01/3437/843-25 $6.00 843
844 RONACHER ET AL.

The δ18O values of quartz range from 13.9 to 18.3 per mil, and calculated δ18OH2O values range from –1.2 to
4.1 per mil. The δDH2O values lie between –77 and –52 per mil. The calculated isotopic composition of the hy-
drothermal fluid lies between that of magmatic and that of meteoric water. The δ18O values of carbonates range
from 15.2 to 16.6 per mil. Carbon isotopes were analyzed on carbonates (δ13C = –3.3 to –2.4‰), altered and
unaltered sedimentary rocks (–5.4 to –4.0 and –1.5 to 0.0‰, respectively), and organic carbon in shales (–23.6
to –16.8‰). The δ34Spyrite values range from –9.4 to 6.1 per mil, and δ34Sanhydrite values range from 12.4 to 17.5
per mil.
The petrographic and analytical results suggest that an ascending fluid interacted with the sedimentary rocks
and/or fluid hosted by the sedimentary rocks at depth (rather than at the site of ore deposition). This interac-
tion is suggested by the presence of organic-derived volatiles and high NH+4 concentrations in inclusion fluid.
In some samples, two fluids were involved in stage II vein formation. Vapor-rich fluid inclusions indicate that
the hydrothermal fluid boiled locally, and continuous salinity trends suggest that fluid mixing occurred in some
stage II veins.
The analytical data and paragenetic information were used to estimate element concentrations in the Porg-
era hydrothermal fluid that were used for thermodynamic reaction path modeling using the software
CHILLER. This fluid was subjected to boiling, mixing with sedimentary formation water, cooling, and fluid-
rock reaction. During boiling, quartz, pyrite, gold, and K-feldspar formed; mixing resulted in the deposition of
pyrite, gold, and mica; reaction with diorite produced the observed wall-rock alteration assemblage; and cool-
ing formed quartz plus minor mica, pyrite, gold, and kaolinite.
Observations, analytical data, and modeling results suggest that more than one process was involved in stage
II vein formation in different locations and at different times. Boiling, mixing, and fluid-rock reaction, all ac-
companied by cooling, occurred intermittently over the entire depth extent of the North zone, and the size of
Porgera may be a result of all of these processes occurring at the deposit.

Introduction responsible for the formation of high-grade zones. Most


THE PORGERA gold deposit is located in the Pliocene fold and recently, Ronacher et al. (2000a, b) documented coexisting
thrust belt that forms the highlands of Papua New Guinea vapor- and liquid-rich fluid inclusions in stage II quartz and
(Fig. 1). With 10.3 million ounces (Moz) of gold recovered suggested that boiling was, at least locally, responsible for ore
between the start of production in 1990 and June 2001, the formation. In this study, we use the term boiling to refer to
mine is a significant gold producer in the southwest Pacific vapor phase separation from a saline, gas-rich fluid (cf. Trues-
region. Reserves and resources in June 2001 amount to 113 dell, 1985, p. 7), resulting in the coexistence of two separate
million tons (Mt) of ore at a grade of 3.5 g/t. The total con- phases. By itself, this does not constrain whether the system
tained gold at Porgera is thus ~20 Moz. was open or closed or whether boiling occurred adiabatically
The deposit is hosted by 6-m.y.-old alkalic intrusions and or nonadiabatically.
Cretaceous sedimentary rocks into which the magmas were This study focuses on a high-grade area, the North zone,
emplaced. Three types of gold-bearing vein assemblages which is hosted by a diorite body and its surrounding sedimen-
occur: (1) magnetite-sulfide-carbonate veins with minor gold tary rocks in the footwall of the Roamane fault (Fig. 2). Quartz-
(prestage I), (2) base metal-sulfide ± gold-carbonate veins roscoelite-pyrite-gold veins were examined by petrography,
(stage I), and (3) quartz-roscoelite-pyrite-gold veins (stage II). fluid inclusion microthermometry, and bulk crush-leach gas
Stage II hosts the highest gold grades and is economically and ion chromatography, and stable isotope geochemistry (O,
the most significant. The discovery of bonanza gold grades in
stage II veins hosted by a late normal fault, the Roamane fault
zone, and its splays proved the economic potential of the de- 0° 140°E
posit in 1983. In the high-grade domain referred to as zone Porgera
VII, up to 1,000 g/t Au occurred in quartz-roscoelite breccias;
this zone is now mined out. Much research has focused on the Australia
quartz-roscoelite-pyrite-gold veins, and various different Bismark Sea
mechanisms for stage II ore formation have been proposed. Fold Porgera
Handley and Bradshaw (1986) and Handley and Henry &
(1990) suggested that boiling or fluid mixing could have oc- Irian thru
Jaya Papua st
curred but considered that the bonanza-type zones were most Solomon Sea
be

New Guinea
likely a result of boiling, by analogy with other gold deposit
lt

types having this style of mineralization. Richards and Kerrich


(1993) agreed with this interpretation and argued, on the Coral Sea
basis of mineralogical and textural grounds, that vapor phase 10°S
N
separation was responsible for stage II ore formation, a model Australia
further supported by fluid inclusion gas data (Richards et al., 200 km
140°E 150°E
1997). However, Wall et al. (1995) and Cameron (1998) con-
cluded from studies of vanadium chemistry and mineralogical FIG. 1. Simplified map of Papua New Guinea, showing the location of the
observations that mixing of an oxidized magmatic fluid and a 6 m.y.-old Porgera gold deposit in the Pliocene fold and thrust belt. Modified
reduced fluid expelled from carbonaceous sediments was after Richards et al. (1990) and Smith (1990).

0361-0128/98/000/000-00 $6.00 844


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 845

11400mN
11000mN
N

11800mN
S

“Zone VII”
2400asl 2400asl

inferred structure


ne
lt

Zo
au
eF

th
or
an

“N
2000asl 2000asl
am
Ro

?
?
11000mN

11800mN
11400mN

Diorite Altered Black Shales Sample location Gold grades:

Fault 1.5-2.5 g/t


Gabbro Black Shales
2.5-3.5 g/t
Feldspar Porphyry Calcareous Sediments Inferred Structure >10 g/t

FIG. 2. Schematic north-south cross section through the Porgera mine, showing a dioritic intrusion in the center and
smaller gabbroic and feldspar porphyry bodies. The intrusions are hosted by altered black shales that grade into unaltered
black shales. White dots represent samples in which vapor- and liquid-rich fluid inclusions were observed. Black dots repre-
sent samples in which only liquid-rich inclusions were found, and crosses represent samples that showed dilution trends in
the fluid inclusion record. Note that one sample showed a dilution trend and coexisting vapor- and liquid-rich inclusions.
Samples were collected from drill core (except the sample from zone VII) but traces of the densely spaced drill holes were
omitted for clarity. Most of the samples were collected along a high-grade ore zone, the North zone, that strikes parallel to
the Roamane fault, a steeply dipping normal fault. Contour lines refer to average gold grades modeled by the Porgera Joint
Venture staff (grades averaged over 10 × 10 × 15 m blocks from assays of 2 m of drill core; individual assay results of up to
several hundred grams per ton were obtained locally). a.s.l. is the elevation above sea level (in meters), and the northing
refers to the Porgera mine grid.

H, C, S) to characterize the hydrothermal fluid from which which was emplaced at an estimated paleodepth of ~2.0 to 2.5
they formed. These data were used to numerically simulate km (Davies, 1983; Fleming et al., 1986). In the absence of a
mineral precipitation based on thermodynamic modeling, classification scheme for subvolcanic alkalic rocks, Richards
using the software CHILLER (Reed, 1982, 1998) to deter- (1990a) used the classification of volcanic rocks of LeBas et al.
mine the mechanism(s) of vein formation. Suggested mecha- (1986) and determined that the hypabyssal intrusions were
nisms involved in high-grade ore formation are discussed in characterized by alkali basaltic, hawaiitic, and mugearitic
the light of previous hypotheses of ore deposition at Porgera. compositions. Texturally, the plutons are gabbros, porphyritic
diorites, and feldspar porphyries, intruded as stocks (≤500 m
Geology of the Porgera Gold Deposit diam) and dikes.
The Porgera Intrusive Complex was emplaced into Creta-
Porgera Intrusive Complex ceous sedimentary rocks of the Chim Formation (Davies,
The Porgera gold deposit is hosted by a suite of mafic, sodic- 1983) in the latest Miocene (~6 Ma; Richards and Mc-
alkalic intrusions known as the Porgera Intrusive Complex, Dougall, 1990; Ronacher et al., 2002). The sedimentary rocks

0361-0128/98/000/000-00 $6.00 845


846 RONACHER ET AL.

are mainly carbonaceous mudstones and calcareous silt- veins. Here, the dominant sulfide minerals are pyrite, spha-
stones, which were deposited on the shelf of the Australasian lerite, and galena, with rare chalcopyrite, arsenopyrite, marc-
plate margin (Davies, 1983; Gunson et al., 2000). Intrusion of asite, freibergite, and proustite/pyrargyrite (Richards et al.,
the magmas occurred slightly prior to, or during, the uplift of 1991). Gold occurs as invisible gold in pyrite or as micro-
the Papuan fold and thrust belt that formed as a result of the scopic inclusions of native gold in pyrite. The gangue consists
collision of the Indo-Australian plate with the Bismarck Sea of complex Mn-Ca-Mg carbonates and subordinate quartz.
and Caroline oceanic plates (Hill and Gleadow, 1989; Hill and The veins and the disseminated ore are accompanied by in-
Raza, 1999). tense phyllic alteration consisting of fine-grained muscovite/
The mineralizing event occurred immediately after the em- illite and carbonate. Richards and Kerrich (1993) determined
placement of the Porgera Intrusive Complex (Ronacher et al., an average homogenization temperature (Th) of fluid inclu-
2002) and was accompanied locally by intense phyllic alter- sions in stage I quartz of 299° ± 33°C (n = 274) and a salinity
ation that affected the shallow-level intrusions. A late steeply of 9.5 ± 1.8 wt percent NaCl equiv. After correcting for esti-
dipping normal fault, the Roamane fault, crosscuts all intru- mated pressure, they calculated a trapping temperature (Tt)
sions and hosts the richest gold bearing veins and breccias. of ~325°C.
The economically most significant vein assemblage, stage
Mineralization II, is characterized by quartz-roscoelite[K(V,Al)2(Si3AlO10)
Magnetite-sulfide-carbonate veins (prestage I) occur in the (OH)2]-pyrite-gold veins and breccia veins (Fig. 3). A list of
central and deep parts of the mine. Magnetite is the dominant all ore minerals described from stage II is shown in Table 1.
mineral and contains lenses of pyrite, pyrrhotite, chalcopy- Some pyrite was reported to be arsenian, and arsenopyrite oc-
rite, and rare inclusions of gold in pyrite. The main non- curs locally in addition to pyrite; chalcopyrite is rare. Gold oc-
metallic mineral is calcite, and quartz is present in minor con- curs dominantly as native gold and electrum. The Ag-Au tel-
centrations. This vein type is economically not significant. lurides, petzite, krennerite, calaverite, and the Ag telluride,
Thin alteration selvages around the veins contain fine-grained hessite, were described from the deposit as well as col-
biotite and actinolite, and the age of the alteration biotite has oradoite (HgTe) and altaite (PbTe; Handley and Bradshaw,
been determined to be 5.98 ± 0.13 Ma (Ronacher et al., 1986; Handley and Henry, 1990; Richards et al., 1991), but
2002). The magnetite-sulfide-carbonate veins are crosscut by only minor hessite was observed in the samples analyzed for
all other vein types (Ronacher et al., 1999). this study. Tellurides appear to have been most abundant in
More gold is observed in the stage I base metal-sulfide ± zone VII, the richest zone of the mine hosted by the Roamane
gold-carbonate veins than in the magnetite-sulfide-carbonate fault zone. In this area, proustite was also observed in rare

a b rosc-py-Au
rosc-Au
rosc-py-Au

qtz qtz

rosc-py-Au

qtz anh

FIG. 3. Drill core photographs of typical stage II quartz-roscoelite-pyrite-gold veins. (a). Four-mm-wide quartz veinlets
with roscoelite, pyrite, and gold rimming the vein edges (sample P99-061B). (b). Breccia vein with roscoelite, pyrite, and gold
rimming parts of the breccia clasts (sample P99-016). (c). Stage II vein with a band of roscoelite, pyrite, and gold forming
the vein edge, followed by a layer of quartz and anhydrite filling the vein center (sample P99-093). Scale in centimeters. Ab-
breviations: anh = anhydrite, Au = gold, py = pyrite, qtz = quartz, rosc = roscoelite.

0361-0128/98/000/000-00 $6.00 846


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 847

TABLE 1. Ore and Gangue Minerals in Stage II Veins, Porgera (in approximate order of decreasing abundance)

Early Late

Quartz SiO2 xx xxx


Roscoelite K(V,Al)2Si3AlO10(OH)2 xxx
Pyrite FeS2 xxx xx
Anhydrite CaSO4 xxx
Ca-Fe-Mg carbonate (Ca,Fe,Mg)CO3 xx xx
Gold, locally tellurides Au (~20% Ag) xx
hessite Ag2Te x
petzite Ag3AuTe2 x
krennerite (Au,Ag)Te2 x
calaverite AuTe2 x
coloradoite HgTe x
altaite PbTe x
Tetrahedrite (Cu,Ag,)10(Fe,Zn)2Sb3S13 xx
Arsenopyrite FeAsS xx
Barite BaSO4 x
Adularia KAlSi3O8 x
Pyrargyrite/prousite Ag3SbS3/Ag3AsS3 x
Chalcopyrite CuFeS2 x
Sphalerite ZnS xx
Marcasite FeS2 x
Hematite Fe2O3 x
Galena PbS x
Acanthite Ag2S x

Notes: Mineral names in italics not observed in this study (cf. Fleming et al., 1986; Handley and Henry, 1990; Richards et al., 1991; Richards and Kerrich,
1993); x = rare, xx = moderately abundant, xxx = abundant

cases as a late mineral intergrown with coarse-grained quartz, fine-grained muscovite/illite, carbonate, and pyrite, and rare
and minor barite was reported to occur locally with roscoelite adularia. The halos are typically only 5 mm wide (e.g., Figs.
and quartz (Richards and Kerrich, 1993). In some veins from 3a-b; 4a-c).
the North zone, tetrahedrite is intergrown with quartz, and 2. Breccias and breccia veins contain angular clasts of wall
qualitative electron microprobe analyses showed that this rock rimmed by bands of roscoelite, pyrite, gold, and minor
mineral is rich in Ag and contains traces of Fe and Zn. All quartz. The clasts are supported by a matrix of quartz. Wall-
other minerals listed in Table 1 occur only locally and in trace rock alteration adjacent to the breccia veins may widen to sev-
amounts. eral tens of centimeters (e.g., Fig. 3b).
Stage II veins in the high-grade zones within the Roamane 3. Thin (≤1-cm) veinlets are also present with roscoelite,
fault zone are texturally distinct from the North zone stage II pyrite, and gold in the wall-rock alteration and quartz in the
veins. The veins in the Roamane fault zone are typically vein (e.g., Fig. 4d-f, h).
banded; the outermost band consists of fine-grained quartz
with minor pyrite, barite, and carbonate, followed inward by No change in stage II vein and alteration mineralogy with
a band of brown-green roscoelite intimately intergrown with depth or along strike nor any systematic variation in vein style
pyrite, gold, minor quartz, and locally barite. This band is suc- with depth were observed. In contrast, all vein types occur
ceeded by a layer of coarse-grained comb quartz. This se- along the entire depth extent of the North zone.
quence is repeated in some cases, and the vuggy vein centers Fluid inclusions in the coarse-grained quartz of stage II
are typically either empty or filled with Ca- Fe-Mg carbonate veins from the Roamane fault zone have average Th values of
or anhydrite (Richards and Kerrich, 1993; Ronacher et al., 146° ± 13°C (n = 519; Richards and Kerrich, 1993); inclusions
2000a). Massive anhydrite fills late meter-wide veins within from the North zone have similar average Th values of 149° ±
the Roamane fault zone; however, these veins contain no 15°C (n = 137; Ronacher et al., 2001). Richards and Kerrich
roscoelite and little gold. (1993) determined a bimodal distribution of salinities with
In the North zone, stage II veins have the following three modes of 4.2 ± 0.3 (n = 104) and 7.8 ± 0.7 wt percent NaCl
different textural styles: equiv (n = 346); similar results were obtained in the present
study except that the majority of inclusions fall in the high-
1. Banded veins have a layer of roscoelite, pyrite, gold, and salinity group (see below).
minor quartz rimming the wall rock and roscoelite is abun-
dant as rosettes in the veins. This band is followed by a layer Sampling and Analytical Methods
of quartz, and locally the vuggy center is filled with anhydrite Samples of quartz-roscoelite-pyrite-gold veins were col-
or carbonate. The veins are typically not wider than 1 cm and lected from the high-grade North zone, hosted by a diorite
locally, the quartz-rich center is missing and the veinlet con- body and its surrounding sedimentary rocks in the footwall of
sists of roscoelite, pyrite, gold, and minor quartz only; wall- the Roamane fault (Fig. 2). The ore shoot forms a steeply dip-
rock alteration around these veins consists of thin halos of ping tabular body parallel to the Roamane fault, but at depth

0361-0128/98/000/000-00 $6.00 847


848 RONACHER ET AL.

a 100µm b 100µm c 100µm


quartz
quartz
quartz

wall rock

wall rock
d 100µm e 100µm f

quartz
carb

fg qtz
qtz

py-Au quartz

100µm

g 100µm h

rosc-py-Au-qtz

FIG. 4. Photomicrographs of stage II veins. (a). and (b). Wall rock rimmed by a layer of roscoelite, pyrite, and gold with
minor quartz and followed by coarse-grained comb quartz (samples P98-1157 and P98-618). (c). Band of roscoelite, pyrite,
gold, and minor quartz forming the vein margin (sample P98-932). (d). Roscoelite and pyrite ± gold in wall rock; the first
layer in the vein consists of quartz which is followed by a band of fine-grained quartz, pyrite, and carbonate, with coarse-
grained quartz in the vein center (sample P98-1466). (e). Same as (d) photographed with crossed polars; the bright colors of
quartz under polarized light are due to the thickness of the section (~80–100 µm; fluid inclusion section). (f). Roscoelite and
pyrite ± gold in the wall rock (sample P98-618). (g). Intergrown quartz, pyrite, and roscoelite forming the vein edge (sample
P98-1234A). (h). Gold in pyrite in wall rock (sample P97-16; photo from Richards, 1998, Porgera ’97 drill core samples: Porg-
era, Papua New Guinea, Porgera Joint Venture, unpub. company report, 20 p.). Abbreviations: Au = gold, py = pyrite, qtz =
quartz, rosc = roscoelite.

0361-0128/98/000/000-00 $6.00 848


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 849

it flattens toward the fault; therefore, it is interpreted to be a method of Clayton and Mayeda (1963) and a Finnigan MAT
splay of the Roamane fault. It extends 600 m downdip, has as 252 multiple collector mass spectrometer. The results are
strike extent of ~400 m, and is about 50 m wide. The ore reported relative to VSMOW, and the analytical precision
shoot consists primarily of stage II quartz-roscoelite-pyrite- and accuracy are both ±0.2 per mil. Isotopic fractionation
gold veins, veinlets, and breccia veins with abundant gold, lo- between quartz and water was calculated using the equa-
cally up to several hundred grams per ton. Drill core samples tion of Clayton et al. (1972) at temperatures defined by
were taken along a large part of the ore shoot with the aim of fluid inclusion homogenization. Oxygen isotope analyses on
determining variations in ore and gangue mineral assem- Ca-Fe-Mg carbonates were conducted following dissolu-
blages, microthermometric properties, and geochemical and tion in 100 percent H3PO4 at 50°C and using an acid frac-
isotopic compositions of fluid inclusions in quartz. tionation factor for ankerite of 1.01060 (Kontak and Ker-
rich, 1997). The analytical precision is ±0.2 per mil, and
Fluid inclusion microthermometry accuracy is ±0.5 per mil.
Fluid inclusion microthermometry was conducted on a The hydrogen isotope composition of fluid inclusion water
Linkam THMSG 600 heating and freezing stage calibrated was determined by thermal decrepitation (heating to
using synthetic fluid inclusions manufactured by Syn Flinc. ~1,200°C) of fluid inclusions in handpicked, cleaned quartz.
Measurements below 31.1°C, the critical point of CO2, are CO2 and H2O were trapped in a U tube cooled by liquid ni-
accurate to ±0.1°C, and measurements above this tempera- trogen. All other gases (e.g., CH4, N2) were pumped away by
ture are accurate to ±1°C. Salinities were calculated using the cooling the U tube with ethanol. The water was reduced to H2
equation of Bodnar (1993) for the H2O-NaCl system. by passing it over depleted uranium before transferring it to
the mass spectrometer. The results are reported relative to
Gas and ion chromatography VSMOW, with accuracy within ±3 per mil.
Bulk crush-leach combined gas and ion chromatography was Carbon isotopes were analyzed on Ca-Fe-Mg carbonate
conducted on 19 samples of quartz from quartz-roscoelite- from stage II veins. CO2 was extracted from ~20 g of these
pyrite-gold veins. Large quartz grains (~1 mm3) were hand- carbonates using 100 percent H3PO4 at 50°C. The results are
picked under a binocular microscope, soaked in aqua regia, fol- reported relative to VPDB, with an accuracy of ±0.2 per mil.
lowed by a brief rinse with hydrofluoric acid to remove surface Organic carbon was analyzed by crushing the rock, remov-
impurities. The samples were then washed with pure water, and ing carbonates by rinsing in dilute acid (10% HCl), and then
3 to 5 g of this quartz was cleaned using electrolytic cleaning running the dried residue through a combustion furnace at
cells according to the technique outlined by Roedder (1958). 1,020°C in the presence of pure oxygen. The furnace is linked
The chromatographic analyses were conducted at the F.G. to an SIRA II mass spectrometer. The accuracy of the analy-
Smith Fluid Inclusion Laboratory at the University of Toronto. ses is ±0.5 per mil.
For gas chromatography, a 1- to 2-g aliquot was placed into a The carbon isotope composition of CO2 in fluid inclusions
stainless steel cylinder and flushed with He overnight. The sam- was analyzed on one sample by crushing the quartz in a ball
ple was crushed in one step, and the gas released was analyzed mill with special gas ports, from which the gas was extracted
using a Hewlett Packard HP 5890A gas chromatograph fitted with a syringe and transferred into an HP 6890 gas chro-
with thermal conductivity and photo-ionization detectors. Stan- matograph. The gas was carried by an He carrier gas to the
dard gas mixtures were injected for calibration. Details of the mass spectrometer (Micromass IsoChrom mass spectrome-
method are outlined in Bray et al. (1991). ter) where its isotopic composition was analyzed with an ac-
The crushed residue was leached with ultra-pure water, di- curacy of ±0.5 per mil.
luted to 10 ml of leachate, and analyzed using a Dionex Sulfur isotopes were analyzed using a continuous-flow iso-
2000i/SP ion chromatograph with an Ion Pac AG9-SC 4-mm tope ratio mass spectrometer (Carlo Erba NA 1500 elemen-
guard column and an Ion Pac AS9-SC 4-mm analytical col- tal analyzer interfaced to a magnetic-sector mass analyzer).
umn. The technique is described in detail by Channer and The sulfide sample was oxidized with O2(g) and transported to
Spooner (1992), and Channer and Spooner (1994) discuss ac- the mass spectrometer in an He carrier gas. The results are
curacy, precision, and detection limits. The analytical preci- reported relative to CDT with an accuracy of ±0.7 per mil.
sion for the gas species was determined using standard gas
mixtures and is better than 10 percent; precision is better Results
than 15 percent for ionic species based on 35 standard runs
(Channer and Spooner, 1994). Inclusion-free Brazilian quartz Fluid inclusion microthermometry
cleaned in the same manner as the samples was used as a Fluid inclusion microthermometry was conducted on 27
blank. Analyses were rejected if the blank value exceeded 25 samples of stage II quartz-roscoelite-pyrite-gold veins from
percent of the measured value. the high-grade North zone and one sample from the high-
The concentrations of the ionic species are calculated by grade zone VII in the Roamane fault zone. The dominant host
combining the gas and ion chromatographic data and calcu- of the fluid inclusions is the coarse-grained comb quartz that
lating the concentrations of ionic species in the amount of succeeds the gold- and pyrite-bearing roscoelite bands. In-
H2O analyzed (cf. Channer and Spooner 1999). clusions in minor quartz intergrown with roscoelite are rare,
small, and most are unworkable. Only primary and pseu-
Light stable isotopes dosecondary inclusions were selected for analysis according
Oxygen isotope analyses were conducted on 10 to 20 mg of to the criteria of Roedder (1984); secondary inclusions are
cleaned, handpicked quartz, using the bromine pentafluoride mostly small, unworkable, and volumetrically not significant

0361-0128/98/000/000-00 $6.00 849


850 RONACHER ET AL.

(<5%; see also Richards and Kerrich, 1993, p. 1029). Exam- CO2, CO2 vapor, and minor amounts of liquid H2O. No eu-
ples of representative groups of fluid inclusions are shown in tectic melting was observed in any of the samples.
Figure 5. Table 2 lists average homogenization temperatures (Th) and
The fluid inclusions are typically two-phase, liquid-rich, ice melting temperatures of the two-phase aqueous inclu-
and irregular in shape, characteristic of low-temperature sions. A plot of Th versus salinity (uncorrected for CO2) for all
trapping conditions (Bodnar et al., 1985). In seven of the 27 samples from the North zone is shown in Figure 6; Figure 7
analyzed samples from the North zone, vapor-rich inclusions shows data for representative individual samples. The Th val-
coexist with the dominant liquid-rich inclusions (Fig. 5c). ues do not increase with depth but are similar (within error)
These vapor- and liquid-rich assemblages include heteroge- along the entire depth extent of the North zone. Similarly,
neously trapped three-phase inclusions consisting of liquid there is no increase or decrease in salinity in this high-grade
zone.
CO2 concentrations known from gas chromatography may
a 10µm be used to correct salinity estimates from ice melting tem-
peratures but, because CO2 concentrations are not known for
all samples and because the CO2 concentrations are variable,
a general correction for all samples would be inappropriate
(although corrections were calculated for those samples
where CO2 measurements are available; see below). Appar-
ent salinity values used to construct Figures 6 and 7 are there-
fore uncorrected.
The absence of liquid CO2 in the liquid-rich aqueous in-
clusions constrains the CO2 content to <2.2-m CO2 (Heden-
quist and Henley, 1985). CO2 concentrations of ≤2.2-m CO2
can depress the ice melting point by up to 4°C in pure water
and thus increase apparent salinity by up to 6.5 wt percent
NaCl equiv (Hedenquist and Henley, 1985, eq 4). The maxi-
mum concentration of CO2 in the samples analyzed in this
study is 1-m CO2. However, this equates to a freezing point
depression of only ~1.9°C and a maximum increase in appar-
b ent salinity of ~3.2 wt percent NaCl equiv.
All samples show a dominant grouping of apparent salini-
ties at ~7 to 10 wt percent NaCl equiv, and five of the 27 sam-
ples from the North zone show an additional cluster between
4 and 5 wt percent NaCl equiv. The number of individual in-
clusions containing the lower salinity fluid is small, and the
low-salinity inclusions are mostly located in the vein centers,
whereas the high salinities occur toward the vein margins.
Both groups exhibit similar average Th values (145° ± 9°C, n
10µm = 77, for the low-salinity group; 152° ± 17°C, n = 995, for the
high-salinity group). Three samples show a continuous range
of salinities from ~4 to ~10 wt percent NaCl equiv (P98-1164,
P99-145, P98-611), possibly representing a dilution trend. Av-
erage Th values for all samples range from 127° ± 12° to 167°
± 25°C (Table 2; see below for a discussion of trapping
temperatures).
Vapor-rich inclusions were typically completely filled with
vapor, and phase changes were not visible. In the three-phase
heterogeneously trapped inclusions from the North zone
(P98-1576, P98-701, P98-1320, P98-066C, P99-197, P98-
1835, P98-611), Tm(CO2) could be measured (avg Tm(CO2) =
–56.6° ± 0.1°C, n = 27; Th(CO2) = 26.9 ± 1.9°C, n = 26), and
clathrate melting (Tm(clathrate)) ranged from 7.0° to 8.8°C (avg
7.4° ± 0.5°C, n = 26; estimated by observing instantaneous
bubble movement; Table 3). Salinities calculated from
these Tm(clathrate) data using the software Flincor (Brown and
Hagemann, 1994) range from 2.4 to 5.7 wt percent NaCl
FIG. 5. Photomicrographs of fluid inclusion assemblages in quartz from equiv (avg 4.6 ± 0.9 wt % NaCl equiv) and give an estimate of
quartz-roscoelite-pyrite-gold veins. (a). Group of primary two-phase inclu-
sions (sample P99-197). (b). Primary fluid inclusion in growth zone (sample
the salinity of the heterogeneously trapped aqueous phase
P98-701). (c) Liquid- and vapor-rich inclusions in growth zone (sample P99- in these inclusions. Heterogeneous entrapment is further
1320). indicated by the anomalously high total homogenization

0361-0128/98/000/000-00 $6.00 850


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 851

TABLE 2. Average Microthermometric Data of Liquid-Rich Fluid Inclusions Hosted by Stage II Quartz from Samples from the Footwall Diorite

Sample no. Elevation (m a.s.l.) Th1 (°C) Tm(ice)1 (°C) Salinity1 (wt % NaCl equiv) N Au2 (ppm)

Roamane fault zone


P98-2204* 2,390 142 ± 4, 157 ± 16 –2.8 ± 0.2, –5.5 ± 0.3 4.7 ± 0.2, 8.6 ± 0.4 10, 27 n.a.3
North zone
P99-120 2,335 140 ± 4, 140 ± 14 –3.8 ± 0.1 ,–5.3 ± 0.2 6.2 ± 0.2, 8.2 ± 0.3 7, 7 <0.1
P98-1576* 2,330 144 ± 9 –5.0 ± 0.6 7.9 ± 0.8 18 4
P98-992 2,315 157 ± 18 –6.3 ± 0.2 9.6 ± 0.2 9 15
P98-1466 2,190 145 ± 12 –5.4 ± 0.4 8.4 ± 0.5 18 2
P98-1459 2,190 154 ± 13 –5.1 ± 0.4 8.1 ± 0.6 36 1
P98-1600 2,150 139 ± 22 –5.8 ± 0.3 9.0 ± 0.4 13 5
P98-1448 2,120 149 ± 12 –5.5 ± 0.2 8.5 ± 0.3 18 7
P98-1413 2,120 127 ± 12 –5.3 ± 0.3 8.3 ± 0.5 37 60
P98-701* 2,080 137 ± 13 –5.3 ± 0.8 8.3 ± 1.1 40 11
P98-1379 2,030 150 ± 10 –5.0 ± 1.6 8.2 ± 0.7 52 10
P98-1320* 2,030 155 ± 15 –4.9 ± 1.4 7.9 ± 0.5 53 8
P98-1631 2,020 145 ± 11 –4.8 ± 0.7 7.5 ± 1.0 34 3
P98-846 2,000 148 ± 20 –5.6 ± 0.3 8.6 ± 0.4 25 9
P98-1111 1,980 132 ± 11 –4.9 ± 0.2 7.8 ± 0.3 6 3
P99-145 1,980 160 ± 10 –2.7 to –5.5 4.5 to 9.2 48 5
P99-178 1,945 150 ± 12, 146 ± 14 –3.0 ± 0.1, –4.9 ± 0.3 5.0 ± 0.0, 7.8 ± 0.4 3, 69 51
P98-611* 1,930 156 ± 11 –3.0 to –5.6 5.0 to 8.7 101 9
P98-612B 1,930 142 ± 19 –5.3 ± 0.5 8.3 ± 0.7 38 11
P98-1162 1,915 161 ± 11 –5.2 ± 0.6 8.2 ± 0.7 12 9
P98-1537 1,910 144 ± 13, 153 ± 9 –2.8 ± 0.1, –4.2 ± 0.4 4.8 ± 0.1, 6.8 ± 0.6 10, 49 28
P98-1164 1,900 163 ± 19 –2.7 to –6.5 4.5 to 10.2 96 9
P99-066C* 1,880 155 ± 16 –5.6 ± 0.5 8.7 ± 0.6 68 2
P99-095C 1,875 161 ± 13 –5.4 ± 0.3 8.4 ± 0.4 47 15
P99-197* 1,855 142 ± 2, 149 ± 13 –2.9 ± 0.1, –5.2 ± 0.3 4.8 ± 0.0, 8.1 ± 0.4 4, 36 <0.1
P98-1170 1,850 153 ± 13 –5.3 ± 1.4 8.5 ± 0.4 71 <0.1
P98-1835* 1,845 167 ± 25 –5.3 ± 0.3 8.3 ± 0.4 27 <0.1
P99-231 1,820 147 ± 7, 163 ± 12 –2.7 ± 0.3, –5.7 ± 0.5 4.4 ± 0.3, 8.8 ± 0.4 13, 5 70

Notes: Samples are listed in order of increasing depth; asterisk (*) denotes samples where vapor-rich inclusions were observed; m a.s.l. = meters above sea
level; n = number of fluid inclusions analyzed (two values represent the number of inclusions in each of the two salinity groups)
1
Where two averages are indicated, the numbers correspond to separate low- and high-salinity groups
2
Gold assayed over a 2-m interval
3
No gold assay data are available for this sample, but it was collected from a high-grade stope within the Roamane fault zone

temperatures of approximately 300°C and the fact that some H2O vapor bubble, suggesting that it may be an immiscible
inclusions homogenize to the liquid and others to the vapor hydrocarbon that was trapped heterogeneously. In sample
phase; still others decrepitated before total homogenization. P124, the presumed hydrocarbon occurred in an assemblage
In some fluid inclusions in samples P98-2204 and P124 of low-salinity inclusions; the H2O vapor bubble homoge-
from the Roamane fault zone, a phase other than CO2 was ob- nized at 135°C and the organic-rich bubble at 180°C,
served. This phase formed a second bubble in addition to the whereas in sample P98-2204 the hydrocarbon occurred in an
assemblage of high-salinity inclusions, and the H2O vapor
bubble homogenized after the hydrocarbon bubble.
Tmice (°C)
-2.0 -3.0 -4.0 -5.0 -6.0 -7.0 Gas and ion chromatography
250
Gaseous species detected include H2O, CO2, CH4, N2,
C2H4, C2H6 (including C2H2), and COS (Table 4), and ionic
200 species detected include Cl–, Br–, Li+, Na+, K+, NH+4, Mg2+,
and Ca2+ (Table 5). Sulfur concentrations were measured as
Th (°C)

SO2–4 , but this analysis includes the actual concentration of


150 SO2–4 in the fluid plus an unknown amount of reduced S,
which may be partially or fully oxidized in the chromatograph
100
column. Thus, H2S, if present, would not have been detected,
and the measured SO2– 4 value gives a minimum estimate of
the total amount of sulfur in the fluid. Those samples in which
50 vapor-rich inclusions were observed are distinguished with an
3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 asterisk after the sample number in Table 4 and were excluded
wt % NaCl equiv from calculations of the gas content of the hydrothermal fluid
FIG. 6. Apparent salinity (in wt % NaCl equiv) vs. homogenization tem- (i.e., the gas content prior to ore deposition) because boiling
perature (Th) of all samples from the North zone. results in a concentration of gases in the vapor phase.

0361-0128/98/000/000-00 $6.00 851


852 RONACHER ET AL.

(a) P98-1170 Tmice(°C ) (b) P98-1448 Tmice(°C )


-3.0 -4.0 -5.0 -6.0 -7.0 -3.0 -4.0 -5.0 -6.0 -7.0
200 200

150 150

Th(°C)
100 100

50 50
4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0
(c) P98-1164 (d) P99-145
-3.0 -4.0 -5.0 -6.0 -7.0 -3.0 -4.0 -5.0 -6.0 -7.0
200 200
Th(°C)

150 150

100 100

50 50
4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0

(e) P98-1537 (f) P99-231


-3.0 -4.0 -5.0 -6.0 -7.0 -3.0 -4.0 -5.0 -6.0 -7.0
200 200

150 150
Th(°C)

100 100

50 50
4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0

wt % NaCl equiv wt % NaCl equiv


FIG. 7. Apparent salinity (in wt % NaCl equiv) vs. homogenization temperature (Th) and ice melting temperature (Tm(ice))
of representative samples of quartz-roscoelite-pyrite-gold veins. Three distinct patterns can be recognized. The examples in
(a) sample P98-1170 and (b) sample P98-1448 belong to a single group with salinities between ~8 and 9 wt percent NaCl
equiv, examples in (c) sample P98-1164 and (d) sample P99-145 show a range of salinities from ~4 to ~10 wt percent NaCl
equiv, and examples (e) sample P98-1537 and (f) sample P99-231 show a bimodal salinity distribution with one cluster at ~4
to 5 and a second cluster at ~7 and ~9 wt percent NaCl equiv, respectively.

Q+/Q– in Table 5 is the ratio of total cationic to anionic for carbonate species can be made using the bicarbonate ion.
charge (Channer and Spooner, 1992). Deviation from the ideal However, nearly all of the corrected Q+/Q– values are less
value of 1 is due to the presence of ions that cannot be mea- than 1, suggesting that using HCO–3 alone may result in over-
sured with this technique (e.g., CO2– – – 2+
3 , HCO3, HS , Fe , Mn )
2+ correction. This seems likely because the wall-rock alteration
+ – indicates that the fluid was moderately acidic and therefore
as well as analytical error and sulfide oxidation. Q /Q is >1
for all samples indicating a deficiency of anions. A correction most likely contained a substantial amount of dissolved car-
bonate as H2CO3 (pKa for H2CO3 is 6.7 at 150°C).
TABLE 3. Average Microthermometric Data for Light stable isotopes
Fluid Inclusion Assemblages Containing CO2 in Stage II Quartz
from Samples from the Footwall Diorite Oxygen and hydrogen: The δ18O values for quartz from
stage II veins range from 13.9 to 18.3 per mil (Table 6). Val-
Sample no. Tm(CO2) n Tm(clathrate) n Th(CO2) n
ues for water (δ18OH2O) in equilibrium with quartz were cal-
P98-611 –56.6 ± 0.1 12 7.8 ± 0.4 12 27.9 ± 0.6 12 culated using the fractionation equation of Clayton et al.
P98-1835 –56.6 ± 0.1 1 7.8 ± 0.1 1 26.8 ± 0.1 1 (1972) and the average Th values measured in each sample.
P98-701 –56.6 ± 0.1 1 7.1 ± 0.1 1 29.0 ± 0.1 1 Because of the observed evidence for boiling, pressure cor-
P98-1576 –56.6 ± 0.1 11 7.6 ± 0.3 10 25.3 ± 1.9 11 rections are unnecessary and Th represents trapping temper-
P99-066 –56.6 ± 0.1 1 6.2 ± 0.1 1 30.0 ± 0.1 1
P99-197 –56.6 ± 0.1 1 8.5 ± 0.1 1 29.0 ± 0.1 1 ature (Tt; see below). Calculated δ18OH2O values range from
1.4 to 4.1 per mil.
Notes: n = number of fluid inclusions analyzed; standard deviation for Oxygen isotope ratios of stage II carbonates intergrown
samples with one measurement only is the analytical error with coarse-grained quartz (as opposed to the more abundant

0361-0128/98/000/000-00 $6.00 852


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 853

TABLE 4. Results of Gas Chromatographic Analysis of Fluid Inclusions in Quartz

Sample no. N21 CH4 CO2 H2O C2H4 C2H62 COS


(mol %) (mol %) (mol %) (mol %) (mol %) (mol %) (mol %)

Roamane fault zone


P98-2204* 0.04553 0.07431 1.820 98.06 n.d. n.d. n.d.
P98-2204A3* 0.11011 0.02255 1.108 98.76 n.d. n.d. n.d.
North zone
P99-120 0.01314 0.03965 1.815 98.12 n.d. 0.00209 0.00793
P98-992A 0.03115 0.10982 1.187 98.67 n.d. n.d. n.d.
P98-1466 0.02624 0.03537 2.004 97.93 n.d. n.d. n.d.
P98-1459 0.01426 0.03249 1.415 98.54 n.d. n.d. n.d.
P98-1600 0.00938 0.02073 1.393 98.58 n.d. n.d. n.d.
P98-1448 0.01479 0.03818 1.627 98.32 n.d. n.d. n.d.
P98-1320A* 0.04923 0.08114 0.621 99.25 n.d. n.d. n.d.
P99-145 0.04043 0.09273 1.384 98.48 n.d. n.d. 0.00214
P99-178 0.06471 0.12656 0.718 99.09 0.00012 0.00112 n.d.
P99-178X3 0.05208 0.10255 0.575 99.27 n.d. 0.00063 n.d.
P98-611* 0.03041 0.04844 1.547 98.37 n.d. n.d. n.d.
P98-612B 0.03737 0.03717 1.444 98.48 n.d. n.d. n.d.
P98-1162 0.01551 0.03573 1.407 98.54 n.d. n.d. 0.00069
P98-1537 0.03908 0.09279 0.850 99.02 n.d. n.d. n.d.
P99-066C* 0.02413 0.04093 1.106 98.83 0.00003 0.00090 n.d.
P99-095B 0.06388 0.03546 1.846 98.05 n.d. n.d. n.d.
P99-197* 0.04013 0.05695 1.227 98.67 n.d. 0.00050 0.00243
P99-197X3* 0.04017 0.05789 1.167 98.73 0.00005 0.00061 0.00175
P98-1170 0.02100 0.04539 1.198 98.73 n.d. n.d. 0.00275
P99-231C 0.02351 0.09177 0.948 98.93 n.d. n.d. 0.00166

Notes: * = samples contain liquid- and vapor-rich fluid inclusions, for all other samples, the values correspond to the gas dissolved in the aqueous liquid
phase (see text); n.d. = not detected
1
The N2 peak is a composite peak of N2 + Ar + CO + O2
2
C2H6 values are combined values of C2H6 and C2H2
3
Two samples from the same quartz vein were analyzed; differences in concentrations may be due to crusher efficiency and abundance of fluid inclusions
in the sample or to the proportion of liquid- versus vapor-rich inclusions

TABLE 5. Ion Chromatographic Analyses (in millimoles per liter) of Fluid Inclusions in Quartz

Sample no. Cl–


Br– ΣS1 Li+ Na+ NH+4 K+ Mg2+ Ca2+ Q+/Q– Q+/Q–corr2

Roamane fault zone


P98-2204 633 1.62 n.a. n.d. 2008 0 828 n.a. n.a. 4.47 0.28
P98-2204A3 298 0.40 75.1 2.41 448 27 153 n.d. 0.0 1.95 0.85
North zone
P99-120 529 0.42 99.8 1.72 584 56 81 2.8 184.9 2.20 0.61
P98-992A 429 0.63 26.0 4.40 620 0 0 0.0 0.0 1.30 0.09
P99-066C 609 0.68 89.6 5.77 876 55 47 0.0 28.3 1.94 1.12
P98-1537 347 0.50 74.4 5.51 479 37 0 0.0 0.0 1.71 0.90
P98-1320A 575 0.69 42.3 3.08 776 16 100 2.5 15.7 1.63 1.09
P98-1448 485 0.71 89.8 5.16 884 23 0 0.0 0.0 1.68 0.73
P98-1466 538 0.75 96.9 4.58 996 18 30 0.0 0.0 1.65 0.67
P98-1170 539 0.52 n.a. 1.58 829 0 372 n.a. n.a. 2.23 1.03
P99-145 353 0.36 0.0 6.04 462 12 121 0.0 0.0 2.00 0.65
P99-178 474 0.65 18.5 6.47 654 43 28 7.0 0.0 2.20 1.27
P99-178X3 279 0.53 n.a. 4.85 545 n.a. 0 n.a. n.a. 1.97 0.94
P98-611 335 0.31 42.1 8.80 473 20 371 0.0 0.0 2.50 0.85
P98-1600 420 0.56 69.9 5.44 780 14 24 0.0 0.0 1.70 0.73
P98-1162 446 0.65 71.1 3.40 780 12 0 0.0 0.0 1.54 0.68
P98-1459 356 0.36 44.6 3.89 557 30 0 0.0 0.0 1.94 0.72
P98-612B 495 0.54 76.6 3.32 904 14 281 0.0 0.0 2.05 0.94
P99-095B 510 0.63 226.4 2.97 1126 10 28 0.0 0.0 1.31 0.65
P99-197 399 0.50 90.2 4.32 731 23 0 0.0 0.0 1.66 0.78
P99-197X 3 310 0.28 n.a. 2.60 613 n.a. n.a. n.a. n.a. 1.98 0.66
P99-231C 421 0.51 32.6 1.25 659 0 0 0.0 0.0 1.36 0.67

Notes: corr = corrected, n.a. = not analyzed, n.d. = not detected


1
Values of ΣS were measured as SO2– 2–
4 , but this analysis includes the actual concentration of SO4 in the fluid plus an unknown amount of reduced S,
which may be partially or fully oxidized in the chromatograph column; thus, H2S, if present, would not have been detected, and the measured SO2– 4 value
gives a minimum estimate of the total amount of sulfur in the fluid
2 –
Q+/Q– is the cation/anion ratio and Q+/Q–corr is the cation/anion ratio corrected for the presence of CO2 dissolved as HCO3 (after Channer and Spooner,
1992; see text for details)
3
Two samples from the same quartz vein were analyzed; differences in concentrations may be due to crusher efficiency and relative abundance of low-
versus high-salinity fluid inclusions in the sample

0361-0128/98/000/000-00 $6.00 853


854 RONACHER ET AL.

TABLE 6. Oxygen and Hydrogen Isotope Data for Fluid Inclusion Water from quartz (in per mil relative to V-SMOW)

Sample no. δ18Oquartz Average Th (°C) δ18OH2O1 δ18OH2O2 δDH2O

Roamane fault zone3


P2.Q1 17.7 136 3.2 –46

P17.Q1 18.1 141 4.0 62
P18.Q1 18.9 144 5.0 –48
P39.Q1 18.7 139 4.4 –52
P45.Q1 18.8 148 5.2 –42
P46.Q1 17.1 130 2.0 –34
P123.Q1 18.5 134 3.8 –57
P124.Q1 18.4 144 4.5 –45
P98-2204 18.3 131 3.3 5.2 –64

North zone
P98-1576 16.4 144 0.6 2.7 –63
P98-981 16.5 No workable fluid inclusions
P98-992 15.2 157 0.6 2.5 –67
P98-1466 14.9 145 –0.8 1.3 –55
P98-1600 15.1 139 –1.1 1.1 –57
P98-1448 13.9 149 –1.4 0.6 –77
P98-1413 16.5 127 –0.9 1.5
P98-701 15.8 137 –0.6 1.2
P98-1631 17.0 145 1.3 3.4 –64
P98-1379 14.8 150 –0.5 1.6 –58
P98-846 15.9 148 0.5 2.6 –53
P98-1111 16.4 132 –0.5 1.8
P98-611 14.6 156 –0.1 1.9 –57
P98-1162 14.5 161 0.2 2.1 –52
P98-1537 13.9 151 –1.2 0.8 –81
P98-1164 17.6 163 3.5 5.4 –64
P98-1170 16.9 153 1.2 3.2 –66
P98-1833 16.7 No workable fluid inclusions –76
P98-1835 17.9 167 4.1 5.9

1 Calculated assuming average Th ≈ Tt; fractionation equation of Clayton et al. (1972)


2 Calculated using average Th +25°C; fractionation equation of Clayton et al. (1972)
3 The values for the Roamane fault zone are from Richards and Kerrich (1993)

vug-filling carbonate) range from 15.2 to 16.6 per mil (avg isotope ratios of the sedimentary host rocks were determined
15.9 ± 0.5‰, n = 5), and δ18O of the fluid was calculated on a carbonaceous mudstone (δ34S = 2.1‰), on a calc-aren-
using the equation of Zheng (1999) for ankerite and assum- ite (δ34S = 5.9‰), and on an altered carbonaceous mudstone
ing an equilibrium temperature of 160°C. The resulting (δ34S = –12.7‰). Petrographic observations show that essen-
δ18OH2O values range from 2.2 to 3.6 per mil (avg 2.9 ± tially all sulfur in these rocks is in pyrite.
0.5‰, n = 5).
Hydrogen isotope ratios of water in fluid inclusions in stage Interpretation of Analytical Data
II quartz range from –81 to –52 per mil.
Carbon: Carbon isotopes were analyzed in stage II vein car- Fluid inclusion microthermometry
bonate intergrown with coarse-grained quartz, in organic car- No pressure correction was applied to the Th values because
bon and carbonate from the sedimentary country rocks, and of local evidence for boiling (i.e., similar average Th values for
in one sample of CO2 from bulk inclusion fluid (Table 7). fluid inclusion groups with and without coexisting vapor-rich in-
Stage II carbonates have a narrow δ13C range of –3.3 to –2.4 clusions, see Table 2). The highest average Th in stage II quartz
per mil, with an average value of –2.7 per mil (± 0.3‰, n = is 167° ± 25°C (Th of individual inclusions is up to 200°C), and
6). The δ13C values for carbonate in unaltered mudstones and this sample (P98-1835) contains vapor-rich, heterogeneously
siltstones are between –1.5 and 0.0 per mil (avg –0.9 ± 0.6‰, trapped inclusions, indicating that boiling took place at this tem-
n = 4), whereas carbonate in altered rocks has δ13C values of perature (Th ≈ Tt). Ronacher et al. (2000a) estimated the trap-
–5.4 to –4.0 per mil (avg –4.6 ± 0.5‰, n = 4). Organic carbon ping pressure of a group of fluid inclusions to be between ~250
in the sedimentary rocks has δ13C values from –23.6 to –16.8 and 340 bars. Given the estimated depth of ore deposition of 2.0
per mil (avg –20.7 ± 2.2 ‰, n = 6). The δ13CCO2 of CO2 in to 2.5 km (Davies, 1983; Fleming et al., 1986), hydrostatic pres-
fluid inclusions from sample P98-2204 is –9.5 per mil. sure at these depths is ~200 to 250 bars, slightly below the
Sulfur: The δ34S values of sulfur in pyrite range from –9.4 calculated pressure of Ronacher et al. (2000a) and suggesting
to 6.1 per mil, and values for late, vug-filling anhydrite range overpressuring of hydrothermal fluid. This is consistent with
from 12.4 to 17.5 per mil (n = 8; Table 8). Whole-rock sulfur hydraulic fracturing and brecciation in the veins.

0361-0128/98/000/000-00 $6.00 854


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 855

TABLE 7. Carbon Isotope Values for Carbonate and Organic Carbon from Various Rocks from Porgera, Stage II Vein Carbonate,
and for CO2 in Inclusion Fluids (in ‰ relative to V-PDB)

Sample no. Description δ13Ccarb δ13Corg-shale δ13CCO2 δ18Ocarb δ18Ofluid

P98-1538 Stage II carbonate –2.4 16.6 3.6


P98-1532 Stage II carbonate –2.6 15.5 2.5
P98-1325 Stage II carbonate –2.5 16.2 3.2
P98-1537 Stage II carbonate –2.7 16.0 3.0
P98-1344 Stage II carbonate –3.3 15.2 2.2
P144 Carbonaceous mudstone (Om Fm) –1.5 –20.0
RJR-69 Altered (qtz, carbonate, pyrite) Chim Fm1 –4.6 –22.9
RJR-75 Altered (qtz, carbonate, pyrite) Chim Fm1 –5.4 –20.3
P139 Mudstone (Chim Fm) –0.8 –20.5
P141 Siltstone 0.0 –16.8
RJR-68 Altered (qtz, carbonate, pyrite) Chim Fm1 –4.0 –23.6
RJR-74 Calc-arenite (disseminated py; altered?; Chim Fm) –4.3
RJR-73 Calc-arenite (Chim Fm) –1.4
P143 Limestone (Nipa Gr)2 1.2
P98-2204 Inclusion fluid –9.5

Notes: δ18Ofluid calculated using the equation of Zheng (1999) for ankerite and a temperature of 160°C; abbreviations: carb = carbonate, Fm = formation,
Gr = group, org = organic, qtz = quartz, py = pyrite
1 Carbonaceous mudstone
2 Nipa Gr limestone overlies the Porgera deposit; analysis from Richards and Kerrich (1993)

Apparent fluid inclusion salinities, shown in Table 2 and in Richards and Kerrich (1993) in which most samples have bi-
Figures 6 and 7, exhibit three patterns: (1) most salinities modal populations with modes at 4.3 ± 0.4 (n = 119) and 7.8
cluster between ~7 and ~9 wt percent NaCl equiv (Fig. 7a-b), ± 0.7 wt percent NaCl equiv (n = 439).
(2) a bimodal distribution of ~4 to ~5 and ~7 to ~9 wt percent Because CO2 concentrations are not known for all samples,
NaCl equiv was observed in several samples (Fig. 7e-f), and a general correction of the apparent salinities is not possible.
(3) some samples display a full range of salinities from ~4 to However, for those samples where the CO2 concentration is
~10 wt percent NaCl equiv (Fig. 7c-d). The majority of sam- known salinities were corrected according to Hedenquist and
ples from the North zone (19 of 27) fall in group one, whereas Henley (1985): Tm = –(KNa+ × mNa+ + KCl– × mCl– + KCO2 ×
six samples have a bimodal distribution of salinities, and three mCO2), where Tm is the melting point of ice, KNa+ and KCl– are
have a range of salinities. The fluid inclusion salinities in the 1.72 Kelvin/mol, and KCO2 = 1.86 Kelvin/mol, and mCO2 is the
sample from the Roamane fault zone (P98-2204) are distrib- molality of CO2. Corrections applied to the data in Table 9 are
uted bimodally, consistent with data previously reported by between 0.9 and 2.5 wt percent NaCl equiv.
Calculations of salinity based on Na+ and Cl– concentra-
tions obtained from ion chromatography (Table 9) are con-
TABLE 8. δ34S Data for Pyrite and Late, Vug-Filling Anhydrite/Gypsum sidered less reliable than CO2-corrected ice melting temper-
from the North Zone and the Roamane Fault Zone
(in ‰ relative to V-CDT) atures because of the unknown concentrations of additional
anions such as bicarbonate (e.g., Richards et al., 1997).
Sample no. Mineral Area δ34S (‰) Three samples from the North zone (P98-611, P98-1164,
P98-145) show a range of salinities from ~4.5 to ~10.0 wt per-
P98-2204 Stage II pyrite Roamane fault zone 6.0
P98-1466 Stage II pyrite North zone –8.2
cent NaCl equiv, suggesting fluid mixing between a high- and
P98-1459 Stage II pyrite North zone 2.5 a low-salinity fluid of about the same temperature. The range
P98-1600 Stage II pyrite North zone –6.7 of salinities is therefore interpreted to represent a dilution
P98-145 Stage II pyrite North zone 1.7 trend. In the five other samples where two fluids were ob-
P98-1162 Stage II pyrite North zone 6.1 served (P98-120, P98-1537, P98-178, P98-197, P98-231), the
P98-1170 Stage II pyrite North zone –9.4
P99-197 Stage II pyrite North zone 1.0 salinity trend was not continuous but bimodal. Hence, in situ
P97-2 Anhydrite, vug filling North zone 12.4 mixing cannot be inferred in these samples.
P97-6 Anhydrite, vug filling North zone 17.5 In samples that contain both high- and low-salinity fluid in-
P97-16 Gypsum, vug filling North zone 14.7 clusions, the high-salinity inclusions occur close to the vein
P97-93 Anhydrite, vug filling Roamane fault zone 16.0
P97-94 Anhydrite, vug filling Roamane fault zone 16.0
edge and the low-salinity inclusions occur toward the vein
P97-95 Anhydrite, vug filling Roamane fault zone 17.0 center, suggesting that these fluids were not present coevally
P97-96 Anhydrite, vug filling Roamane fault zone 17.2 in these veins. However, the different fluid inclusion popula-
P98-2206 Anhydrite, vug filling Roamane fault zone 17.2 tions do not correlate with a particular vein type. For exam-
P144 Mudstone (Om Formation) Host rocks1 2.1 ple, two of the three samples that show a dilution trend are
RJR-145 Calc-arenite Host rocks1 5.9
RJR-68 Mudstone quartz veins with a roscoelite-pyrite-quartz band at the vein
(Chim Formation) Altered host rocks1 –12.7 edges (P98-611, P98-1164), and the third sample (P99-145) is
a breccia vein with angular clasts of altered mudstone in a
1 Sulfur in pyrite matrix of quartz and minor carbonate. Likewise, the samples

0361-0128/98/000/000-00 $6.00 855


856 RONACHER ET AL.

TABLE 9. Average Inclusion Fluid Salinity Corrected for the Presence of CO2 in the Fluid Inclusions and
Salinity Estimated from Cl– and Na+ Concentrations Analyzed by Ion Chromatography

Salinity based on Cl– Salinity based on Na+


Salinity1 (uncorrected; CO2 Salinity corrected for concentration3 concentration4
2
Sample no. wt % NaCl equiv) (mol %) CO2 (wt % NaCl equiv) (wt % NaCl equiv) (wt % NaCl equiv)

P99-120 6.2, 8.2 1.815 5.6 3.2 5.4


P98-992 9.6 1.187 8.0 2.6 6.0
P98-2204A 4.7, 8.6 1.108 5.9 1.8 3.7
P98-1466 8.4 2.004 5.5 3.2 6.7
P98-1459 8.1 1.415 6.1 2.2 5.1
P98-1600 9.0 1.393 7.0 2.5 5.6
P98-1448 8.5 1.627 6.1 2.9 7.5
P98-1320 7.9 0.621 7.0 3.5 4.6
P99-145 7.1 1.384 5.0 2.1 6.0
P99-178 5.0, 7.8 0.718 6.7 2.9 5.1
P98-612B 8.3 1.444 6.2 3.0 5.8
P98-611 7.4 1.547 5.1 2.1 6.1
P98-1162 8.2 1.407 6.1 2.7 5.3
P98-1537 4.8, 6.8 0.850 5.5 2.1 4.1
P99-066C 8.7 1.106 7.1 3.7 6.9
P99-095C 8.4 1.846 5.7 3.1 5.9
P99-197 4.8, 8.1 1.227 6.3 2.4 4.6
P98-1170 8.5 1.198 6.8 3.2 4.9
P99-231 4.4, 8.8 1.182 7.0 2.7 4.9

Note that only the high-salinity values were corrected for CO2 because clathrate was not observed in the low-salinity inclusions
1 Salinity calculated from ice melting point (Table 2); the two values for salinity represent the low- and the high-salinity group
2 Salinity calculated according to the method outlined by Hedenquist and Henley (1985)
3 Salinity calculated assuming all Cl– analyzed by ion chromatography is present in the fluid as NaCl
4 Salinity calculated assuming all Na+ analyzed by ion chromatography is present in the fluid as NaCl

exhibiting a bimodal salinity distribution have a variety of tex- inclusions are the same, within error. The average CO2 con-
tures. Samples P98-1537 and P99-197 are breccia veins with tent of samples containing liquid-rich inclusions is 1.31 ± 0.40
quartz and carbonate matrix and thin (1-mm) bands of mol percent (n = 16) and that of samples containing rare
roscoelite-pyrite-quartz rimming the wall rock (visible gold was vapor-rich inclusions in addition to liquid-rich inclusions is
found in sample P99-197). Sample P98-120 is an ~2-cm-thick 1.13 ± 0.30 mol percent (n = 4). This similarity suggests that
quartz vein with thin (~2-mm) roscoelite bands. Sample P99- the contribution of gas from gas-rich inclusions did not affect
178 is an 8-cm-thick quartz vein with thin (~1-mm) roscoelite the analyses, and this is consistent with the small number of
bands in quartz (this sample also contained visible gold). Sam- gas-rich inclusions observed. However, certain gases, particu-
ple P99-231 consists of a network of quartz veinlets and veins larly C2H6, appear to be more abundant in samples with a bi-
(1 mm to 1 cm thick), where the wider veins contain small modal distribution of salinities. For example, three of the five
clasts of black shale rimmed by roscoelite and abundant visible samples that show bimodal salinity distribution also contain
gold. The high-salinity fluid inclusions occur in all vein types. ethane (0.0005–0.00209 mol % C2H6), whereas only one of 11
Whereas there are clear dilution trends that are interpreted samples with high salinities only contains C2H6 (0.0009 mol
to be the result of fluid mixing, vapor-rich fluid inclusions co- %). No clear correlation between gas content and salinity pat-
existing with liquid-rich inclusions in some samples (Fig. 5c) terns or vein textures could be observed for the other gases.
indicate that boiling also took place during the formation of at
least some stage II quartz veins. Although gold was observed Ion chromatography
to be closely intergrown with quartz containing coexisting liq- Br–/Cl– ratios of Porgera fluids range from 0.8 × 10–3 to 1.9
uid- and vapor-rich fluid inclusions in one sample from a × 10–3 with one outlier of 2.6 × 10–3 (avg value of 1.2 ± 0.2
high-grade area close to the Roamane fault zone (Ronacher et × 10–3, n = 22). These values overlap with the Br–/Cl– ratio of
al., 2000b), this relationship was not observed in the samples modern seawater (1.5 × 10–3) and lie just outside the value
from the North zone. In the former case, a direct link be- reported for magmatic water (~0.9 × 10–3) by Böhlke and
tween boiling and gold precipitation was suggested Irwin (1992).
(Ronacher et al., 2000b), but this correlation is unclear for the I– was not determined in the present study, however, Fig-
North zone. Both boiling and mixing occurred locally over the ure 8 shows a Br–/Cl– versus I–/Cl– diagram with data for
entire depth extent of the North zone and it is not possible to Porgera from Richards et al. (1997) in comparison to data for
clearly distinguish zones where boiling or mixing dominated. seawater, Broadlands and Wairakei, New Zealand, and Yel-
lowstone from Böhlke and Irwin (1992). The Br–/Cl– ratios of
Fluid inclusion gas contents most samples from Porgera are similar to the Br–/Cl– ratios
The gas contents of samples containing liquid-rich inclu- for geothermal waters from Broadlands and Wairakei as well
sions only and samples with coexisting liquid- and vapor-rich as to seawater, but they show much higher I–/Cl– ratios than

0361-0128/98/000/000-00 $6.00 856


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 857

3 10
North Zone
SMOW
2.5 Roamane Fault Zone
-10
Br/Cl (x10 -3)

2 felsic magmatic
-30

δDH2O (%o )
L water
W
1.5 Porgera M
Broadlands -50
1
Wairakei Porgera
Seawater -70 magmatic
0.5
Yellowstone MORB water
-90 fluid
0
Porgera mine water
0 50 100 150 200 250 300
-6 -110
I/Cl (x10 ) -15 -10 -5 0 5 10
– – – –
FIG. 8. Br /Cl vs. I /Cl in fluid inclusion water (from Richards et al.,
δ OH2O (%o )
– – – – 18
1997). The Br /Cl of seawater falls within the range of Br /Cl ratios of the
– –
Porgera fluid, but the two fluids have distinctly different I /Cl ratios. Values
from Broadlands and Wairakei (New Zealand) and Yellowstone (United FIG. 9. Oxygen and hydrogen isotope data of water in equilibrium with
States) are shown for reference (Böhlke and Irwin, 1992). stage II quartz. δ18OH2O was calculated from δ18Oquartz using the fractionation
equation of Clayton et al. (1972) and measured homogenization tempera-
tures (assumed to be the trapping temperatures; see text for details). The iso-
topic composition of present-day Porgera mine water is plotted for reference
seawater, and higher I–/Cl– ratios than the New Zealand geo- (data from Richards and Kerrich, 1993). The filled circles represent data
from the North zone (this study), whereas the open circles represent data
thermal waters. Böhlke and Irwin (1992) attributed the high from the Roamane fault zone (Richards and Kerrich, 1993). Note that for the
I–/Cl– values (up to ~200 × 10–6) in some waters to reaction samples from Richards and Kerrich (1993) a 25°C temperature correction
with organic-rich material. Similarly, Kendrick et al. (2001) was applied to the Th to calculate δ18OH2O from δ18Oquartz because no clear ev-
attributed I–/Cl– values of 121 × 10–6 in porphyry Cu deposits idence for boiling was observed in the fluid inclusion record from these sam-
to mixing of magmatic water with a sedimentary formation ples. Ranges for MORB fluid and felsic magmatic arc water from Taylor
(1986). Ranges for Porgera magmatic water are from Richards and Kerrich
water. A similar explanation may apply to Porgera. (1993). MWL = meteoric water line.
The Br–/Cl– ratio of seawater (1.5 × 10–3) falls within the
range of Br–/Cl– ratios of Porgera fluid (avg 1.3 ± 0.3 × 10–3),
which may indicate a contribution from a connate seawater
because the sedimentary rocks that surround the Porgera In-
trusive Complex are Cretaceous to Tertiary marine shelf sed-
iments (Davies, 1983). Alternatively, the same signature could range, i.e., –0.5 and 3.1 per mil, respectively. Likewise, there
be a result of interaction of a fluid with shales; the Br–/Cl– ra- is no relationship between δ18OH2O or δDH2O and vein tex-
tios of modern shales analyzed by Walter et al. (1989) and tures. The δ18OH2O values of breccias range from –1.2 to 3.3
Kharaka et al. (1977) range from 0.9 to 2.6 × 10–3, and the per mil, and the δ18OH2O values of quartz-roscoelite-pyrite-
Porgera values fall within this range. gold veins range from –0.1 to 3.5 per mil.
Carbon isotopes: The average carbonate carbon isotope
Stable isotopes composition of altered sedimentary rocks is –4.6 per mil,
Oxygen and hydrogen isotopes: Oxygen isotope compositions which is similar to that of stage I carbonates (from –7.1 to
of water in equilibrium with quartz fall between the values –4.5‰). Richards and Kerrich (1993) interpreted these val-
typical of meteoric water and Porgera magmatic water ues to reflect a magmatic origin for the fluid that altered the
(Richards and Kerrich, 1993; Fig. 9). Hydrogen isotope com- sedimentary rocks and formed the stage I carbonate veins.
positions of stage II water range from slightly lower than Unaltered sedimentary rocks show an average carbonate car-
Porgera magmatic water to within the range of Porgera mag- bon isotope composition of –0.9 ± 0.6 per mil (n = 4). The av-
matic water. For comparison, Richards and Kerrich’s (1993) erage value of stage II carbonate is –2.7 ± 0.3 per mil (n = 6)
data from the Roamane fault zone are also plotted in Figure and is intermediate between the magmatic and sedimentary
9. values, suggesting that mixing between magmatic and sedi-
Both the oxygen and hydrogen isotope compositions of mentary carbon may have occurred.
water in fluid inclusions from the samples from the North The δ13C value of CO2 in fluid inclusions from one sample
zone are lower than those from the Roamane fault zone of stage II quartz was –9.5 per mil. This value is similar to
(Richards and Kerrich, 1993). This difference is not an arti- some active geothermal systems where CO2 is thought to be
fact of the different Th values used to calculate δ18OH2O and is magmatic (e.g., –6.9 to –9.1‰ at Broadlands; Lyon and Hul-
interpreted to reflect a real difference in the isotopic compo- ston, 1984). The carbonate isotope composition of stage II
sition of the fluid from which stage II veins formed in the carbonates is consistent with a δ13C value of –9.5 per mil and
Roamane fault zone and in the North zone. the fractionation between CO2 and carbonate (~6‰ at 150°C
There is no correlation between oxygen and hydrogen iso- for siderite, which is considered to be a proxy for Porgera
tope data and fluid inclusion salinity. For example, samples stage II Ca-Fe-Mg carbonates; Hoefs, 1997). The δ13C of Fe-
P98-611 and P98-1164 both show a dilution trend but corre- rich carbonate in equilibrium with a fluid having a δ13CCO2 of
sponding δ18OH2O values are at the opposite ends of the –9.5 is –3.5 per mil, similar to the analyzed average value of

0361-0128/98/000/000-00 $6.00 857


858 RONACHER ET AL.

–2.7 ± 0.3 per mil. However, other sources of light carbon are heterogeneously during boiling because final total homoge-
discussed below. nization (Th(total)) occurred to the vapor in some and to the liq-
Sulfur isotopes: The variable δ34S values for pyrite from the uid phase in others; Th(total) of some of the inclusions could not
North zone suggest that isotopic equilibrium between H2S in be measured because of decrepitation, and where it could be
the hydrothermal fluid and pyrite was not reached, possibly measured, Th(total) exceeded Th of the liquid-rich inclusions by
because of the slow reaction kinetics at temperatures below more than 120°C. Such three-phase, CO2-rich inclusions ap-
200°C (Ohmoto and Rye, 1979; Ohmoto and Goldhaber, pear to be rare in epithermal deposits, although they have
1997). The negative δ34S values for pyrite may be due to a been described (Belkin et al., 1985). High CO2 concentra-
change in oxidation state of the hydrothermal fluid (Ohmoto tions in fluid inclusions were reported from the Cripple
and Rye, 1979; Ohmoto and Goldhaber, 1997). For example, Creek district, Nevada (Thompson et al., 1985), the Bessie G
oxidation accompanying boiling may result in formation of deposit, Colorado (Saunders and May, 1986), and from the
sulfides with negative δ34S values from heavier sulfur (Drum- epithermal gold deposits in the Baguio district, Philippines
mond and Ohmoto, 1985; McKibben and Eldridge, 1990). (0.06–0.73 mol % CO2, Sawkins et al., 1979). The distribution
The sedimentary host rocks are characterized by positive δ34S of CO2-rich fluid inclusions suggests that hydrothermal sys-
values (2.1 and 5.9‰) except for the altered mudstone tems in alkalic rocks are richer in CO2 than are systems in
(–12.7‰), indicating a significant change in either the S calc-alkaline rocks (Mutschler and Mooney, 1993; Richards,
source or the oxidation state of sulfur in altered relative to un- 1995; Jensen and Barton, 2000).
altered mudstones. The importance of CO2 in increasing the pressure at which
The δ34S values of late anhydrite (12.4–17.5‰) are similar boiling occurs was stressed by Barton and Chou (1993) who
to the sulfur isotope values of sulfate in seawater, suggesting found that even small, undetected concentrations of CO2 can
that evolved seawater descending in faults during the waning significantly increase the depth of boiling and can thus have
stages of the hydrothermal system could have caused the for- an impact on the depth of epithermal-type ore deposition
mation of the late anhydrite. Porgera was formed slightly (Wilkinson, 2001). CO2 was not the only gas contributing to
prior to or during the onset of tectonic uplift or synuplift, the boiling point depression at Porgera; CH4 and N2 also
which makes the involvement of seawater plausible (Hill and occur in the inclusion fluid, and these components may have
Gleadow, 1989; Hill and Raza, 1999). an even greater effect than CO2 in deepening the onset of
boiling because of their greater volatility (Price, 1981; Duan
Constraints on Fluid Sources and et al., 1992). Duan et al. (1992) plotted CO2-CH4 solubility in
Hydrothermal Processes a 5 wt percent NaCl equiv fluid at 149°C and 345 bars from
the data of Price (1981) and showed that CH4 decreases the
Bimodal salinity distribution and mixing solubility of CO2 significantly. The P-T-salinity conditions of
The fluid inclusions in stage II quartz provide evidence the model of Duan et al. (1992) resemble the conditions at
for two fluids with different apparent salinities. This may be Porgera, and Richards et al. (1997) superimposed the Porgera
explained by the existence of two fluids of different salini- CH4 and CO2 data (from the Roamane fault zone) on the
ties but similar Th that had access to the veins at different curve of Duan et al. (1992). Richard et al. (1997) showed that
times during the evolution of the hydrothermal system. most of the data points fall close to the curve, suggesting that
This is further supported by the observation that the low- the Porgera fluids may have been saturated with gas (Fig. 10).
salinity inclusions occur toward the vein center, whereas We added the new gas data from the North zone to the plot
the high-salinity inclusions cluster near the vein edge. The of Richards et al.(1997), and most of the gas data from the
higher salinity fluid could have been a deep fluid that had North zone (circles in Fig. 10) fall slightly above the CO2-CH4
cooled upon ascent, whereas the lower salinity fluid may solubility curve of Duan et al. (1992), which suggests that the
have been a formation water at the ambient temperature at conditions at which the saturation curve was modeled may
2.0- to 2.5-km depth. The formation water appears to have not be correct for the North zone. Higher pressures may have
been ~150°C at this depth, suggesting a geothermal gradi- prevailed in the North zone where boiling was less common
ent of ~50°C/km. Such a gradient may be realistic for the than in the Roamane fault zone. Because the Roamane fault
aureole of a shallow-level intrusive complex. Alternatively, is a major fault, the pressure may have been lower (i.e., closer
Richards and Kerrich (1993) suggested that both fluid to hydrostatic) than in the North zone, which is a smaller
sources were deep. splay. Richards and Kerrich (1993) suggested that the brec-
The presence of a range of apparent salinities in a minority cias observed dominantly in the Roamane fault zone are con-
of samples also suggests that the two fluids may have mixed. sistent with boiling in this area. In addition, they interpreted
The fact that a continuous salinity trend is present in at least the fine-grained silica margins that they observed in stage II
some of the samples suggests that this process may have oc- veins from the Roamane fault zone as having formed during
curred locally. rapid precipitation due to boiling.
The possibility of CO2 degassing as cause of the higher ap-
parent salinity is excluded because the CO2-corrected salinity Possible fluid source(s)
values are still higher than the low-salinity values. Br/Cl and I/Cl data for Porgera fluids (Richards et al., 1997;
this study) can be interpreted in a number of ways. Porgera
Boiling fluids have Br/Cl ratios that are similar to that of modern sea-
The three-phase inclusions containing liquid CO2, CO2 water, but the I/Cl ratios are higher. These fluids bear a closer
vapor, and liquid H2O are interpreted to have been trapped resemblance to geothermal fluids and particularly those fluids

0361-0128/98/000/000-00 $6.00 858


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 859

0.40 comparison, solutions from oceanic spreading centers such as


liquid-rich only inclusions
the East Pacific Rise contain <0.01 mmol/kg (Von Damm et
liquid-rich and vapor-rich inclusions
al., 1985 a, b; Viets et al., 1996; Cruse and Seewald, 2001).
0.30 data from Richards et al. (1997) Von Damm et al. (1985b) attributed the higher NH+4 concen-
CH4 (mole%)

from the Roamane Fault Zone


trations to interaction of the hydrothermal solution with or-
0.20 ganic sediments in the Guayamas basin. A similar source
seems likely for the Porgera fluids.
A further indication of interaction of the hydrothermal fluid
0.10 with the sedimentary country rocks is the presence of higher
hydrocarbon species, such as C2H4 and C2H6 (including
0.00 C2H2; see Table 4) in the fluid inclusions. Although these
0.0 0.5 1.0 1.5 2.0
species are present in minor amounts, they are most likely of
organic origin and are thus useful for tracing fluid sources
CO2 (mole%) (e.g., Norman et al., 1997; Brathwaite and Faure, 2002). Liq-
FIG. 10. Solubility of CO2 and CH4 in a 5 wt percent NaCl equiv solution uid hydrocarbons also were observed in fluid inclusions in
at 150°C and 345 bars (from Price, 1981, and Duan et al., 1992). Data from some samples from the North zone, and these may have been
Porgera are overlain on the curve. Black circles represent samples that con- derived from the organic-rich sedimentary rocks into which
tain liquid-rich inclusions only, and open circles represent samples that con-
tain vapor- in addition to liquid-rich inclusions. Triangles represent data from
the Porgera Intrusive Complex was emplaced and which
Richards et al. (1997). partly host mineralized veins. The hydrocarbons occur in
both the high- and low-salinity inclusions. The latter could
represent formation waters in the organic-rich sedimentary
rocks.
that have reacted with organic matter. However, some contri- Richards and Kerrich (1993) explained the negative
bution from connate fluids of seawater composition cannot be δ34Spyrite (–14.0 to –1.6‰) values in the Roamane fault zone
ruled out. by oxidation of magmatic sulfur due to boiling. The difference
CH4 and N2 concentrations in the Porgera fluid (0.0225 to in the range of δ34Spyrite values between the North zone and
0.1226 and 0.0094–0.1375 mol %, respectively) appear to be the Roamane fault zone suggests that the precipitation
high compared to data from other epithermal deposits (e.g., process operating in the North zone was either different or
0.0002–0.0100 mol % CH4, and from 0.0096–0.0724 mol % only partly the same as in the Roamane fault zone, with addi-
N2 at Acupan, Philippines, Sawkins et al., 1979; 0.013 mol % tional processes operating in the North zone. The sulfur
CH4 and 0.029 mol % N2 at Ladolam, Lihir Island, Papua source for stage II pyrite remains ambiguous, and both mag-
New Guinea, Müller et al., 2002). A potential source of CH4 matic sulfur and sedimentary sulfur may have contributed to
and N2 is the organic-rich sedimentary rocks into which the the system.
Porgera Intrusive Complex was emplaced. A contribution of Richards et al. (1991) analyzed the lead isotope composi-
methane gas from the sediments might also have been im- tions of igneous rocks (Porgera Intrusive Complex), sedimen-
portant for ore formation. Richards et al. (1997) showed that tary host rocks of the Chim and Om Formations, and gold and
CH4 was critical for boiling to occur under the pressure and sulfides of stages I and II mineralization. They found that the
temperature conditions during stage II vein formation. lead isotope compositions of gold and sulfides fall between
Although the δ13C of CO2 in fluid inclusions (–9.5‰) is the lead isotope compositions of the igneous and sedimentary
similar to that of magmatic CO2 in modern geothermal sys- rocks and concluded tentatively that the hydrothermal miner-
tems, the slightly lower value for the Porgera sample may be als incorporated a mixture of lead from the Chim Formation,
due to a contribution of light carbon from pyrolysis or hy- the Om Formation, and igneous lead.
drolysis of organic matter or graphite (e.g., 2C + 2H2O → CO2 The δ18OH2O and δDH2O values for the Porgera hydrother-
+ CH4; Ohmoto and Rye, 1979). Given that the average mal fluid from quartz, carbonate, and fluid inclusion water
δ13Corganic of the Porgera shales is –20.7 ± 2.2 per mil (n = 6) suggest the presence of both a magmatic and a meteoric
and that the expected fractionation between graphite and water component. Samples from the Roamane fault zone
CO2 at temperatures below 300°C is between 12 and 14 per tend to record higher δ18OH2O and δDH2O values, which may
mil, the analyzed δ13CCO2 value of –9.5 per mil could reflect a indicate different proportions of magmatic and meteoric
contribution of light carbon from the shales. water in the two areas.
NH+4 concentrations in the Porgera inclusion fluids are high The analyzed fluid chemistry and the observed salinity dis-
(up to 61 mmol/kg) relative to waters from geothermal sys- tributions suggest that two distinct fluids were involved in the
tems (e.g., 0.1–1.0 mmol/kg at Fushime, Japan: Akaku et al., hydrothermal system. One end member (salinity of ~6 wt %
1991; 0.05 mmol/kg at Wairakei, New Zealand: White et al., NaCl equiv, Th ≈ 160°C) may have been a magmatically de-
1963). The Porgera ammonium concentrations are also high rived fluid ascending from depth, whereas the second fluid
relative to brines, such as Searles Brine Lake, California (~1 (salinity of ~3 wt % NaCl equiv, Th ≈ 150°C) could have been
mmol/kg, White et al., 1963). Slightly higher values were formation water hosted by the country rock surrounding the
recorded from sedimentary formation waters (2–4 mmol/kg, Porgera Intrusive Complex. These two fluids appear locally to
Merino, 1975), and even higher values were detected in hy- have mixed at the level of vein formation as suggested by Th-
drothermal solutions from the Guayamas basin, Gulf of Cali- salinity plots of samples P98-1164 and P99-145, but in other
fornia (up to 15.6 mmol/kg NH4+: Von Damm et al., 1985b). In samples (e.g., P98-1537 and P99-231) the fluid inclusion

0361-0128/98/000/000-00 $6.00 859


860 RONACHER ET AL.

record shows no evidence for mixing. In another group of The initial temperature of the solution at the start of reac-
samples (e.g., P98-1576 and P99-066C), the high-salinity tion was set at 170°C, corresponding to the highest average
fluid was undergoing boiling, as indicated by heterogeneously fluid inclusion homogenization temperature measured in
trapped fluid inclusions. Thus, it appears that both boiling stage II quartz.
and fluid mixing were involved in vein formation at Porgera.
Composition of the mixing solution
Numerical Modeling To simulate fluid mixing, a separate fluid corresponding to
Numerical modeling was conducted to test the effective- sedimentary formation water was chosen because fluid inclu-
ness of boiling, fluid mixing, fluid-rock reaction, and simple sion data and ion chromatography indicate the involvement of
cooling for vein deposition in the North zone of the Porgera such a fluid (see above). We produced a deep ground water
deposit. The computer program CHILLER (Reed, 1982, that had equilibrated with the sedimentary rocks that host the
1998) was used to model boiling, mixing, cooling, and wall- Porgera Intrusive Complex, i.e., siltstones and mudstones and
rock reaction. minor sandstones typical of shelf sediments (Davies, 1983;
Gunson et al., 2000). The known composition of a mudstone
Starting solution composition and modeling conditions (Richards, 1990b) was equilibrated with seawater (simplified
The composition of the inferred high-salinity fluid at Porg- from Berner and Berner, 1996) to simulate formation water
era was estimated using gas and ion chromatography data (Table 11).
from fluid inclusions and the mineral paragenesis observed in
stage II veins (Table 10). Average values of Cl–, Na+, K+, and Assumptions and limitations of the method
NH4+ were used. The value for K+ is approximately what The bulk fluid inclusion samples were taken from quartz
Richards et al. (1997) measured. The component species overlying the roscoelite-pyrite-gold-quartz band, and there-
NH4+ incorporates all N species, and therefore the gas analy- fore the fluid inclusions may not represent the exact compo-
sis of N2 was added to the measured concentrations of NH+4. sition of the initial solution from which the ore precipitated
The component species HCO3– was calculated, using the con- because it was trapped after the main ore deposition. Never-
centrations of CO2 and CH4 determined by gas chromatogra- theless, it is the closest approximation available because few
phy (Reed, 1998). The concentrations of Fe2+, Ca2+, and Mg2+ workable fluid inclusions occur in the roscoelite band.
were fixed by near saturation with siderite, calcite, and mag- Roscoelite precipitation cannot be modeled at the present
nesite, respectively, because Ca-Fe-Mg carbonate was ob- time because thermodynamic data for aqueous V-bearing
served in stage II veins. Al3+ was determined by saturating the species and minerals are not included in the Soltherm data-
solution with muscovite at 170°C. The pH was buffered by base used by CHILLER. However, we assumed that precipi-
muscovite under conditions that were slightly undersaturated tation of muscovite occurs under similar conditions as
with respect to K-feldspar. Saturation with quartz was used to roscoelite because of the chemical similarity of the two micas
fix SiO2, and saturation with pyrite was used to fix the total [KAl2(Si3AlO10)(OH)2 versus K(V,Al)2(Si3AlO10)(OH)2];
sulfur concentration. The fluid was also saturated with gold. therefore, muscovite was used as a proxy for roscoelite.
Concentrations of other elements (e.g., Zn, Cu, and Pb) were Cameron (1998) modeled the stability of roscoelite and con-
low and poorly constrained and so were not added to the cluded that at 200°C roscoelite was stable between pH values
starting solution. of 4 and 12 and log fO2 values between –60 and –35, thus over-
lapping the muscovite field.
TABLE 10. Composition of the Starting Solution Used Although the starting solution consists of a limited number
for Numerical Modeling of species and is thus only an approximation, all principal
T (°C) 170
pH 5.00 TABLE 11. Chemical Composition of the Formation Water Used for the
Log fO2 –42.59 CHILLER Fluid Mixing Run
Component species Concentration (total molality) T (°C) 150
pH 6.1
Cl– 7.3 × 10–1 Log fO2 –44.16
SO2–4 1.2 × 10–2
HCO3– 6.5 × 10–1 Component species Concentration (total molality)
HS– 1 –1.1 × 10–2
SiO2 2.4 × 10–3 Cl–
5.2 × 10–1
Al3+ 7.3 × 10–8 SO2– 3.6 × 10–2
4
Ca2+ 2.8 × 10–5 HCO3– 2.3 × 10–2
Mg2+ 5.5 × 10–6 HS– 3.3 × 10–6
Fe2+ 2.9 × 10–8 SiO2 1.7 × 10–3
K+ 3.0 × 10–2 Al3+ 2.9 × 10–7
Na+ 7.1 × 10–1 Ca2+ 3.2 × 10–3
NH4+ 2.9 × 10–2 Mg2+ 1.4 × 10–4
Au+ 5.6 × 10–9 Fe2+ 1.1 × 10–6
K+ 2.8 × 10–3
1 The negative value of HS– is a reflection of the relatively higher abun- Na+ 5.9 × 10–1
dance of SO2– –
4 than HS in this fluid; for details on how CHILLER calculates Au+ 2.3 × 10–12
these negative values see Reed (1998, p. 112)

0361-0128/98/000/000-00 $6.00 860


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 861

minerals occurring in stage II veins can be modeled with this Boiling


fluid (if muscovite is accepted as an analogue for roscoelite).
The program CHILLER assumes that all reactions reach (a) -1
equilibrium. Although kinetics most likely influences some qtz
-3
mineral precipitation reactions, including gold precipitation

log grams
(Brown, 1989), the program has proved useful in replicating -5
mineral assemblages in the geothermal and epithermal envi- py micro
ronments (Reed and Spycher, 1985; Spycher and Reed, 1989; -7
Saunders and Schoenly, 1995; Simmons and Browne, 2000;
-9
Cooke and McPhail, 2001), for Carlin-type (Hofstra et al.,
Au
1991), MVT (Plumlee et al., 1994), greisen (Halter et al., -11
1998), VHMS (Reed, 1983), and laterite deposits (Getahun et (b) 8.0
al., 1992).
Boiling 7.0

pH
Adiabatic boiling was modeled starting at 170°C because
6.0
the highest average Th observed was 167° ± 25°C. One kilo-
gram of starting solution was adiabatically boiled from 170° to
100°C. As soon as a mineral is saturated and precipitates, it is 5.0
removed from the contact with the hydrothermal fluid, simu-
(c) 30
lating flow along a path.
The mineral assemblage that precipitates upon adiabatic

logΣS6+/ΣS2-
boiling is shown in Figure 11a. The assemblage consists of 20
quartz, pyrite, and gold, which all precipitate during the ear-
liest stages of boiling. Microcline forms when a temperature
of ~135°C is reached. The pH increases from 5.0 at 170°C to 10
7.5 at 100°C (Fig. 11b) due to loss of volatiles, such as CO2 to
the vapor phase (cf. Reed, 1998).
Figure 11c shows the ratio of the sums of all oxidized and 0
reduced sulfur species in the system (ΣS6+/ΣS2–) and corre- 170 160 150 140 130 120 110 100
sponds to the redox state of the fluid (Cooke and McPhail, T (°C)
2001). The liquid becomes oxidized during boiling due to ex- FIG. 11. Results of model calculations for boiling of 1 kg of starting solu-
solution of species such as H2S, CH4, and H2. The experiment tion (see text). (a). Minerals precipitating during adiabatic boiling from 170°
yields a paragenesis similar to that observed in stage II veins, to 100°C. The reactions progress from left to right as boiling proceeds from
inasmuch as gold, pyrite, and quartz coprecipitate; however, higher to lower temperatures. The curves symbolize minerals precipitating
from the solution; a mineral, once precipitated, is removed from reacting
microcline is formed instead of muscovite. with the fluid; the lines indicate the rate of mineral precipitated per degree
of temperature change. (b). Change of pH during boiling. (c). Change in the
Fluid mixing redox state of the fluid represented by S6+/S2– (Cooke and McPhail, 2001).
The temperature of the mixing solution was set at 150°C to Abbreviations: Au = gold, micro = microcline, py = pyrite, qtz = quartz.
match the Th values of the low-salinity fluid inclusions. In this
model, 2 kg of mixing solution was progressively mixed into 1
kg of starting solution at 170°C. Minerals were removed upon
saturation.
After mixing 1 kg of formation water with the starting solu-
tion, the temperature is ~160°C, and after adding 2 kg of for- 120°C, the fluid no longer precipitates muscovite and forms
mation water the temperature is 157°C. Mixing results in the kaolinite instead, reflecting the drop in pH from 5.0 at 170°C
early precipitation of muscovite and pyrite, followed by gold to 4.6 at 100°C (Fig. 13b). The oxidation state of the fluid in-
after adding 1.4 kg of formation water. The pH changes creases upon cooling (Fig. 13c). All minerals predicted to pre-
slightly from 5.0 to 5.2, and the oxidation state of the mixture cipitate during cooling occur in stage II except kaolinite.
increases. The minerals calculated to precipitate upon mixing However, temperatures in the veins do not fall to 120°C, so
are observed in the vein assemblage. However, quartz, a dom- kaolinite is not expected to form.
inant mineral in most veins, does not form in the mixing
model. Also, gold precipitation is delayed relative to mus- Fluid–wall-rock reaction
covite and pyrite (Fig. 12). Fluid–wall-rock reaction was modeled by titrating 10 g of a
dioritic host rock of known composition (Roamane diorite;
Cooling Richards, 1990a) into 1 kg of starting solution at 170°C. In
Simple cooling of the starting solution was modeled from this case, minerals were not removed because fluid flow in
170° to 100°C, and minerals were removed after precipita- the wall rock is assumed to be slower than in veins, and thus
tion. This results in the precipitation of quartz, pyrite, mus- minerals are expected to be in contact with the hydrothermal
covite, and gold during the first 5°C of cooling (Fig. 13a). At fluid.

0361-0128/98/000/000-00 $6.00 861


862 RONACHER ET AL.

Mixing Cooling
(a)
T (°C) (a) -2
(a) 170 163 160 158 157 -2
-2 qtz
qtz

grams
-4

loggrams
-4
musc
-4
log grams

-6
py -6 py musc kaol
py musc kaol

log
-6 Au -8 Au
-8 Au
-8 -10
-10
(b) 5.5
(b) 5.5
-10
(b) 5.0

pH
5.0

pH
5.0
pH

5.5

4.5
4.5
(c)
(c)
4.5
20
2- 2-
(c) 160 20
/Σ/ΣSS
logΣS6+/ΣS2-

6+6+

120
SS
10
ΣΣ
10
log
log

80

40 0
0
170 160 150 140 130 120 110 100
170 160 150 140 130 120 110 100
0 T (°C)
T (°C)
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 FIG. 13. Modeling results of simple cooling of 1 kg of starting solution. (a).
amount of mixing solution (kg) Minerals precipitating from the solution upon cooling from 170° to 100°C.
(b). Change in pH during mixing. (b) Change in pH during cooling. (c)
FIG. 12. Modeling results for mixing of 1 kg of starting solution (170°C) Change in redox state of the fluid. Abbreviations: Au = gold, kaol = kaolinite,
with up to 2 kg of formation water (150°C). The amount of mixing solution musc = muscovite, py = pyrite, qtz = quartz.
(in kilograms) increases from left to right. (a). Minerals precipitating from
the solution during mixing. The temperature change that occurred during
mixing is indicated on the top axis of the plot. (b). Change in pH during mix- quartz that forms the center of most veins. Cooling is very ef-
ing. (c). Change in redox state of the fluid represented by S6+/S2– (Cooke and ficient in deposition of quartz, whereas only small amounts of
McPhail, 2001). Abbreviations: Au = gold, musc = muscovite, py = pyrite.
muscovite, pyrite, and gold form upon cooling, and it is there-
fore likely that the coarse-grained quartz that forms the bulk
of some veins was deposited by cooling. Boiling and mixing
The minerals calculated to precipitate include quartz, mus- were accompanied by cooling. Modeling of fluid-rock reac-
covite, pyrite, siderite, gold, and apatite followed by dolomite. tion satisfactorily generates the mineral paragenesis observed
After 5 g of rock was titrated into the starting solution, in the wall-rock alteration and in some of the mineralized
ankerite also precipitated, followed by hematite at 7 g (Fig. veins at Porgera (i.e., veinlets with roscoelite-pyrite-gold in
14). The observed alteration assemblage consists of carbon- the wall-rock alteration). However, because only a limited
ate-sericite-pyrite with minor quartz and gold, which is simi- amount of wall rock can interact with the hydrothermal fluid
lar to the alteration paragenesis. Hematite, clay minerals, and due to shielding of fresh rock by a layer of altered rock and
chlorite are also observed in small amounts in the altered wall mineral deposition at vein walls, this process, although ob-
rock, and apatite is present as a stable residual phase in the served locally, probably played a minor role in ore formation
diorite. This numerical experiment also served as a test for at Porgera.
the quality of the starting solution composition because the The modeling results suggest that the individual bands of
minerals simulated to precipitate could be compared to the stage II veins may have formed by different processes. The
observed alteration assemblage. roscoelite-pyrite-gold band may have formed by mixing or
Boiling alone is unlikely to have been responsible for stage fluid-rock reaction (which forms the largest amount of mus-
II vein formation because muscovite (i.e., roscoelite) does not covite of all processes), whereas some of the early quartz may
form in the boiling run. Fluid mixing may account only for have formed by boiling, and the bulk of the coarse-grained
roscoelite-pyrite-gold formation, but mixing does not seem to quartz in the vein centers may have formed by cooling. It is
be responsible for the precipitation of the coarse-grained also likely that a combination of processes occurred.

0361-0128/98/000/000-00 $6.00 862


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 863

Fluid-wall rock reaction by groups of higher and lower salinity inclusions. Vapor-rich
(a) -1 fluid inclusions coexisting with liquid-rich inclusions in some
-2
qtz samples indicate that boiling occurred, at least locally, and in
musc ank
-3 py the Roamane fault zone a direct link has been made between
log grams

-4 sid
apa boiling and gold deposition (Ronacher et al., 2000b). In the
-5

nontr
North zone samples, vapor-rich fluid inclusions were not ob-
-6
dolo
hem served in quartz intergrown with gold, but their presence in
-7
-8 Au quartz immediately overlying the gold-rich layer nevertheless
-9 demonstrates that conditions were favorable for boiling when
-10 this quartz was deposited. Organic compounds in the fluid
(b) 5.5 (e.g., C2H2, C2H4, C2H6) and liquid hydrocarbons in rare fluid
inclusions indicate a reaction of the fluid with organic-rich
sedimentary rocks at depth in the system. Also, the gas and
ion compositions of stage II fluids suggest the involvement of
a formation water (e.g., CH4, N2, I–/Cl–; NH4, Br–/Cl–, pres-
pH

5.0
ence of higher hydrocarbons, and, in rare cases, liquid hydro-
carbon). The carbon isotope signatures of stage II carbonates
(~–2.5‰) are different from the magmatic value of stage I
4.5 carbonates (~–5‰; Richards and Kerrich, 1993) and suggest
(c) 0.1 a contribution from a source other than magmatic water,
probably the sedimentary rocks. The δ18O and δD data sug-
0.08
logΣS6+/Σ S2-

gest a significant contribution of meteoric water to the hy-


0.06 drothermal fluid (Richards and Kerrich, 1993; this study).
Numerical simulations indicate that the most efficient mech-
0.04
anisms to precipitate quartz are boiling and cooling, whereas
0.02 muscovite (used as a proxy for roscoelite) is formed most ef-
ficiently by fluid-rock reaction and fluid mixing, and gold is
0 more likely deposited by boiling and fluid-rock reaction. Pet-
0 2 4 6 8 10 rographic observations showed that gold also occurs in the
diorite added (grams) wall rock adjacent to veins, implying fluid-rock reaction
FIG. 14. (a). Results of model calculations for reaction of the starting so- and/or wall-rock sulfidation.
lution with dioritic wall rock. Diorite (composition from Richards, 1990b) is These data indicate that several different processes were
titrated into 1 kg of solution. The amount of titrated diorite increases from 0 involved in ore formation. Vapor-rich fluid inclusions in some
to 10 g of diorite titrated as shown on the abscissa. The lines in this plot in-
dicate cumulative mineral precipitation because the minerals are not re-
samples suggest boiling, continuous salinity trends in Th-
moved from the solution (see text). (b). pH evolution during fluid-rock reac- salinity plots for other samples suggest fluid mixing, and gold
tion. (c). Redox state of the fluid. Abbreviations: ank = ankerite, apa = in the wall rock suggests fluid-rock reaction as a precipitation
apatite, Au = gold, dolo = dolomite, hem = hematite, musc = muscovite, ntr mechanism. Because these different processes are not mutu-
= nontronite, py = pyrite, qtz = quartz, sid = siderite. ally exclusive, are geologically reasonable, and can be related
to particular samples, they can be combined and integrated
Discussion and Conclusion into a model for ore formation at Porgera. The observations
Both boiling and fluid mixing have been suggested as ore- indicate that boiling, mixing, and fluid-rock reaction, accom-
forming mechanisms at Porgera. Arguments for boiling were panied by cooling, all occurred intermittently along the North
based on textural observations of abundant breccias and brec- zone. In several cases, more than one process has been docu-
cia veins in the Roamane fault zone and vein margins consist- mented in individual samples, suggesting that a combination
ing of fine-grained quartz interpreted to be recrystallized of processes may have been involved in vein deposition, op-
chalcedony (Richards and Kerrich, 1993). In addition, vapor- erating in the same location but at different times, forming
rich fluid inclusions coexisting with liquid-rich inclusions the individual bands of the quartz-roscoelite-pyrite-gold
were observed in stage II quartz intergrown with gold in one veins.
sample from near the Roamane fault zone (Ronacher et al., We suggest that the ore-forming fluid exsolved from the in-
2000b). In contrast, Wall et al. (1995) and Cameron et al. ferred parent pluton at depth, interacted with the country
(1995), based on considerations of mobility of V3+ versus V5+ rocks, mudstones and siltstones of the Om and Chim Forma-
in a hydrothermal fluid, argued that precipitation of tions (Davies, 1983), and mixed with the deep ground water
roscoelite was a result of reduction (rather than oxidation hosted by these rocks. As this mixed fluid rose to shallower
caused by boiling) and that mixing with a reduced fluid was a levels, it reacted with the dioritic and gabbroic wall rocks and
likely mechanism of ore formation (see also Cameron, 1998). started to precipitate minerals in the order predicted by the
The data on quartz-roscoelite-pyrite-gold veins in this study fluid-rock simulation. Fractured rock resulting from normal
have been used to readdress the question of ore deposition at movement along the Roamane fault may have allowed more
Porgera. The crucial results from the North zone are summa- formation water to interact with the ascending fluid, and lo-
rized here and are integrated to generate a model of ore for- cally, mineral precipitation may have cemented the available
mation for this zone. The presence of two fluids is indicated open space in the faults, resulting in overpressuring of the

0361-0128/98/000/000-00 $6.00 863


864 RONACHER ET AL.

fluid, followed by hydrofracturing, decompression, breccia- The findings of our study and the above evidence from the
tion, and boiling. Boiling may have been prevalent in the literature emphasize that epithermal deposits are both
highest grade areas, such as zone VII located directly in the rapidly evolving and are dynamic systems (cf. Cooke and
Roamane fault zone, where quartz intergrown with gold con- McPhail, 2001), and a variety of depositional mechanisms are
tains both liquid and vapor inclusions (Ronacher et al. 2000b). only a logical consequence of the fast-changing nature of
In the North zone of the footwall diorite, in contrast, where these systems.
gold occurs partly in thin veinlets intergrown with abundant
roscoelite and partly in the wall rock adjacent to thin quartz-
roscoelite-pyrite-gold veinlets and breccia veins, fluid-wall Is Porgera unique? Implications for exploration
rock reaction may have been an additional ore precipitation for Porgera-type deposits
mechanism. Elsewhere in the system, fluid mixing may have As pointed out by Sillitoe (2002), some characteristics of
been the dominant ore deposition mechanism as indicated by Porgera do not appear to be typical of epithermal deposits.
the fluid inclusion salinity trends observed in some samples For example the inferred depth of ore formation of ~2 to ~2.5
from the North zone. Given the size of the Porgera deposit, it km is greater than the typical depth of ore formation in ep-
is not surprising that more than one mechanism caused ore ithermal deposits. In addition, the host rocks of the Porgera
formation. This interpretation is similar to that of Cooke and Intrusive Complex consist of sedimentary rocks dominantly of
McPhail (2001) who concluded that phase separation, fluid the Chim Formation (siltstones and mudstones) and, perhaps
mixing, cooling, and fluid-rock reaction all contributed to vein more importantly, the Porgera Intrusive Complex is underlain
formation of the Acupan deposit, Baguio district, Philippines, by carbonaceous mudstones and siltstones of the Om Forma-
and that these processes may have occurred simultaneously tion, which is a source of oil and gas (Davies, 1983). Hence,
(Cooke and Bloom, 1990; Cooke et al., 1996). fluids exsolved from the inferred parent pluton at a ~6-km
depth (Richards, 1990a) passed through ~4 km of hydrocar-
Porgera in the context of alkalic rock-related gold deposits bon-rich sedimentary rocks. The interaction with these rocks
Gold deposits associated with alkalic rocks have been the is argued to have enriched the fluids in CH4 and N2, causing
topic of a number of reviews (e.g., Mutschler and Mooney, these fluids to boil at depths of ~2.0 to 2.5 km. It remains un-
1993; Richards, 1995; Jensen and Barton, 2000; Müller and clear, however, whether the interaction with organic-rich sed-
Groves, 2000) in which the similarity between these deposits imentary rocks had other consequences, such as providing or-
was stressed. In contrast, Sillitoe (2002) emphasized the ganic ligands that may have played a role in transport not only
uniqueness of deposits like Porgera, Emperor, and Cripple of precious metals but also of Al and Na (Gize, 2000; Shock,
Creek. The following section compares depositional mecha- 2000).
nisms inferred to have played a role at Porgera with the ore- The inferred paleodepth of ore formation at Porgera indi-
forming processes in other alkalic systems and comments on cates that exploration for similar deposits could extend to
the potential uniqueness of Porgera and the implications for depths greater than those assumed for the epithermal envi-
exploration. ronment, especially in the vicinity of carbonaceous sedi-
A variety of depositional processes have been suggested for mentary rocks and alkalic, CO2-rich intrusions. Under these
alkalic-rock related, low-temperature vein deposits (e.g., conditions, even areas where some degree of erosion has
Jensen and Barton, 2000), including fluid phase separation, been documented may still be prospective for epithermal-
fluid mixing, cooling, and fluid-rock reaction. Ahmad et al. type deposits.
(1987) suggested, based on the fluid inclusion record, that ore We suggest that Porgera is a giant deposit because of the in-
deposition at the Emperor mine, Fiji, was dominantly due to terplay and coincidence of a number of different factors.
phase separation. However, Kwak (1990) interpreted his own These include (1) a tectonic setting prone to alkalic magma-
fluid inclusion data to indicate that cooling due to fluid mixing tism (i.e., back-arc environment, postsubduction, or exten-
and wall-rock reaction were responsible for ore formation. sional setting: Richards, 1995; Sillitoe, 2002); (2) a structural
Poliquin and Simmons (1998), in another study on the Em- setting favorable for magma ascent (e.g., intersection of deep
peror mine, obtained results similar to those of Ahmad et al. crustal structures: Corbett et al., 1995); (3) a structural setting
(1987) and concluded that phase separation was involved in ore favorable for focused ore precipitation (e.g. Roamane fault
formation. They noted that phase separation over a long verti- and its splays); (4) conditions favorable for phase separation,
cal interval (~500 m) with no vertical temperature gradient was possibly due to interaction with the host sedimentary rocks;
“a puzzling feature of the ore-forming hydrothermal condi- (5) conditions favorable for fluid mixing and cooling in dila-
tions of this deposit” (Poliquin and Simmons, 1998, p. A302). tional zones; and (6) the large size of the parent pluton indi-
Thompson et al. (1985) suggested that cooling may have cated by a 5-km-diam aeromagnetic anomaly around Porgera.
been the dominant ore deposition mechanism at Cripple Similar deposits may form in the absence of one or more of
Creek, and Jensen and Barton (2000) stressed that, locally, these conditions. For example, the Mount Kare deposit, 15
there is evidence for fluid-rock reaction at Cripple Creek. km southwest of Porgera, located along the same structural
Zhang and Spry (1994) and Spry and Thieben (1998), based trend and associated with similar alkalic intrusions, may be
on isotopic and microthermometric observations, suggested smaller because of the much smaller size of the intrusive
that mixing was involved in the formation of the Gies gold-sil- complex. However, it seems likely that the large size of Porg-
ver telluride deposit, Montana (except for the bonanza ore era can be attributed to the unusual coincidence of all of
zones that may have formed by phase separation; P. Spry, these conditions in one deposit (see also Richards, 1997; Sil-
pers. commun., 2002). litoe, 1997).

0361-0128/98/000/000-00 $6.00 864


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 865

Acknowledgments Cameron, G.H., 1998, The hydrothermal evolution and genesis of the Porg-
era gold deposit, Papua New Guinea: Unpublished Ph.D. thesis, Canberra,
We thank the Porgera Joint Venture for funding this study Australian National University, 137 p.
and for permission to publish our results. In particular, we Cameron, G.H., Wall, V.J., Walshe, J.L., and Heinrich, C.A., 1995, Gold min-
thank Mike Johnston, Rod Henham, Shawn Crispin, Sonia eralization at the Porgera gold mine, Papua New Guinea, in response to
fluid mixing [abs.]: Pacific Rim Congress ‘95, Australian Institute of Min-
Konopa, Tony Burgess, and Martin Torovi for their support ing and Metallurgy, Auckland, New Zealand, Nov. 19–22, 1995, Proceed-
on the Porgera mine site. ings, p. 99–100.
E.R. was financially supported in part by the Austrian Channer, D.M.DeR., and Spooner, E.T.C., 1992, Analysis of fluid inclusion
Academy of Sciences and by the Izaac Walton Killam Memo- leachates from quartz by ion chromatography: Geochimica et Cosmochim-
rial Foundation, which is gratefully acknowledged. E.T.C.S. ica Acta, v. 56, p. 249–259.
——1994, Combined gas and ion chromatographic analysis of fluid inclu-
and C.J.B. thank the Natural Sciences and Engineering Re- sions: Application to Archean granite pegmatite and gold-quartz vein fluids:
search Council and the Department of Geology, University of Geochimica et Cosmochimica Acta, v. 58, p. 1101–1118.
Toronto (Steven D. Scott, Chair) for support of the F. Gordon ——1999, Integrated cation-anion/volatile fluid inclusion analysis by gas and
Smith Fluid Inclusion Laboratory. ion chromatography; methodology and examples: Chemical Geology, v.
Stable isotope analyses were conducted at the University 154, p. 59–82.
Clayton, R.N, and Mayeda, T.K., 1963, The use of bromine pentafluoride in
of Alberta, the University of Calgary, the University of the extraction of oxygen from oxides and silicates for isotopic analyses:
Saskatchewan, and the University of Waterloo, and we thank Geochimica et Cosmochimica Acta, v. 27, p. 43–52.
Karlis Muehlenbachs, Stephen Taylor, Chris Holmes, and Clayton, R.N., O’Neil, J.R., and Mayeda, T.K., 1972, Oxygen isotope ex-
Bob Drimmie. change between quartz and water: Journal of Geophysical Research, v. 77,
p. 3057–3067.
Reviews of an earlier draft of this paper by Andrew Camp- Cooke, D.R., and Bloom, M.S., 1990, Epithermal and subjacent porphyry
bell, Sarah Gleeson, Karlis Muehlenbachs, and Tom Chacko mineralization, Acupan, Baguio district, Philippines: A fluid inclusion and
significantly improved the manuscript. We thank Stuart Sim- paragenetic study: Journal of Geochemical Exploration, v. 35, p. 297–340.
mons, Paul Spry, and Bruce Mountain for their thorough and Cooke, D.R, and McPhail, D.C., 2001, Epithermal Au-Ag-Te mineralization,
insightful reviews, which were greatly appreciated. Mark Acupan, Baguio district, Philippines: Numerical simulations of mineral de-
position: ECONOMIC GEOLOGY, v. 96, p. 109–132.
Hannington’s comments improved the style of the manuscript Cooke, D.R., McPhail, D.C., and Bloom, M.S., 1996, Epithermal gold min-
significantly. eralization, Acupan, Baguio district, Philippines: Geology, mineralization,
alteration, and the thermochemical environment of ore deposition: ECO-
September 10, 2002; March 15, 2004 NOMIC GEOLOGY, v. 91, p. 243–272.
Corbett, G., Leach, T., Stewart, R., and Fulton, B., 1995, The Porgera gold
deposit: Structure, alteration and mineralisation: Pacific Rim Congress ’95,
REFERENCES Australasian Institute of Mining and Metallurgy, Auckland, New Zealand,
Ahmad, M., Solomon, M., and Walshe, J.L., 1987, Mineralogical and geo- Nov. 19–22, 1995, Proceedings, p. 151–156.
chemical studies of the Emperor gold telluride deposit, Fiji: ECONOMIC Cruse, A.M., and Seewald, J.S., 2001, Metal mobility in sediment-covered
GEOLOGY, v. 82, p. 345–370. ridge-crest hydrothermal systems: Experimental and theoretical con-
Akaku, K., Reen, M.H., Yagi, M., Kai, K., and Yasuda, Y., 1991, Chemical and straints: Geochimica et Cosmochimica Acta, v. 65, p. 3233–3247.
physical processes occurring in the Fushime geothermal system, Kyushu, Davies, H.L., 1983, Wabag, Papua New Guinea, 1:250,000 Geological Series
Japan: Geochemical Journal, v. 25, p. 315–333. – Explanatory Notes, Port Moresby, Geological Survey of Papua New
Barton, P., and Chou, I-Ming, 1993, Refinement of the evaluation of the role Guinea, 84 p.
of CO2 in modifying estimates of the pressure of epithermal mineralization: Drummond, S.E., and Ohmoto, H., 1985, Chemical evolution and mineral
ECONOMIC GEOLOGY, v. 88, p. 873–884. deposition in boiling hydrothermal systems: ECONOMIC GEOLOGY, v. 80, p.
Belkin, H., De Vivo, B., Gianelli, G., and Lattanzi, P., 1985, Fluid inclusions 126–147.
from the geothermal fields of Tuscany, Italy: Geothermics, v. 14, p. 59–72. Duan, Z., Møller, N., and Weare, J.H., 1992, An equation of state for the
Berner, E.K., and Berner R.A., 1996, Global environment, water, air and geo- CH4-CO2-H2O system: II. Mixtures from 50 to 1000°C and 0 to 1000 bar:
chemical cycles: Upper Saddle River, N.J., Prentice Hall, 376 p. Geochimica et Cosmochimica Acta, v. 56, p. 2619-2631.
Bodnar, R.J., 1993, Revised equation and table for determining the freezing Fleming, A.W., Handley, G.A., Williams, A.L., Hills, A.L., and Corbett, G.J,
point depression of H2O-NaCl solutions: Geochimica et Cosmochimica 1986, The Porgera gold deposit, Papua New Guinea: ECONOMIC GEOLOGY,
Acta, v. 57, p. 683–684. v. 81, p. 660–680.
Bodnar, R.J., Reynolds, T.J., and Kuehn, C.A., 1985, Fluid-inclusion system- Getahun, A., Reed, M.H., and Retallack, G.J., 1992, A numerical simulation
atics in epithermal systems: Reviews in Economic Geology, v. 2, p. 73–97. of the chemical development of laterite and china clay [abs.]: Geological
Böhlke, J.K., and Irwin, J.J., 1992, Laser microprobe analyses of Cl, Br, I, Society of America Abstracts with Programs, v. 24, no. 5, p. 27.
and K in fluid inclusions: Implications for sources of salinity in some an- Gize, A.P., 2000, The organic geochemistry of gold, platinum, uranium, and
cient hydrothermal fluids: Geochimica et Cosmochimica Acta, v. 56, p. mercury deposits: Reviews in Economic Geology, v. 9, p. 217–250.
203–225. Gunson, M., Hall, G., and Johnston, M., 2000, Foraminiferal coloration index
Brathwaite, R.L., and Faure, K, 2002, The Waihi epithermal gold-silver-base as a guide to hydrothermal gradients around the Porgera intrusive complex,
metal sulfide-quartz vein system, New Zealand: Temperature and salinity Papua New Guinea: ECONOMIC GEOLOGY, v. 95, p. 271–282.
controls on electrum and sulfide deposition: ECONOMIC GEOLOGY, v. 97, p. Halter, W.E., Williams-Jones, A.E., and Kontak, D.J., 1998, Modeling fluid-
269–290. rock interaction during greisenization at the East Kemptville tin deposit:
Bray, C.J., Spooner, E.T.C., and Thomas, A.V., 1991, Fluid inclusion volatile Chemical Geology, v. 150, p. 1–17.
analysis by heated crushing, on-line gas chromatography; applications to Handley, G.A., and Bradshaw, P.M.D., 1986, The Porgera gold deposit,
Archean fluids: Journal of Geochemical Exploration, v. 42, p. 167–193. Papua New Guinea, in MacDonald, A.J., ed., Gold ’86: International Sym-
Brown, K.L., 1989, Kinetics of gold precipitation from experimental hy- posium on the Geology of Gold Deposits: Willowdale, Ontario, Konsult In-
drothermal sulfide solutions: ECONOMIC GEOLOGY MONOGRAPH 6, p. ternational, p. 416–424.
320–327. Handley, G.A., and Henry, D.D., 1990, Porgera gold deposit, in Hughes,
Brown, P.E., and Hagemann, S.G., 1994, Macflincor: A computer program F.E., ed., Geology of the mineral deposits of Australia and Papua New
for fluid inclusion data reduction and manipulation, in De Vivo, B., and Guinea: Melbourne, Australasian Institute of Mining and Metallurgy, p.
Frezzotti, M.L., eds., 1994, Fluid inclusions in minerals: Methods and ap- 1717–1724.
plications: Short Course Notes, Virginia Polytechnical Institute, Virgina Hedenquist, J.W., and Henley, R.W., 1985, The importance of CO2 on
Tech Press, p. 231–250. freezing point measurements of fluid inclusions: Evidence from active

0361-0128/98/000/000-00 $6.00 865


866 RONACHER ET AL.

geothermal systems and implications for epithermal ore deposition: ECO- ——1983, Seawater-basalt reaction and the origin of greenstones and related
NOMIC GEOLOGY, v. 80, p. 1379–1406. ore deposits: ECONOMIC GEOLOGY, v. 78, p. 466–485.
Hill, K.C., and Gleadow, A.J.W., 1989, Uplift and thermal history of the ——1998, Calculation of simultaneous chemical equilibria in aqueous-min-
Papuan fold belt, Papua New Guinea: Austalian Journal of Earth Sciences, eral-gas systems and its application to modeling hydrothermal processes:
v. 36, p. 515–539. Reviews in Economic Geology, v. 10, p. 109–124.
Hill, K.C., and Raza, A., 1999, Arc-continent collision in Papua New Guinea: Reed, M.H., and Spycher, N.F., 1985, Boiling, cooling, and oxidation in ep-
Constraints from fission track thermochronology: Tectonics, v. 18, p. 950–966. ithermal systems: A numerical modeling approach: Reviews in Economic
Hoefs, J., 1997, Stable isotope geochemistry: Berlin, Springer, 201 p. Geology, v. 2, p. 249–272.
Hofstra, A.H., Leventhal, J.S., Northrop, H.R., Landis, G.P., Rye, R.O., Richards, J.P., 1990a, Petrology and geochemistry of alkalic intrusives at the
Birak, D.J., and Dahl, A.R., 1991, Genesis of sediment-hosted dissemi- Porgera gold deposit, Papua New Guinea: Journal Geochemical Explo-
nated-gold deposits by fluid mixing and sulfidization: Chemical-reaction- ration, v. 35, p. 141–199.
path modeling of ore-depositional processes documented in the Jerritt ——1990b, The Porgera gold deposit, Papua New Guinea: Geology, geo-
Canyon district, Nevada: Geology, v. 19, p. 36–40. chemistry and geochronology: Unpublished Ph.D. thesis, Canberrra, Aus-
Jensen, E.P., and Barton, M.D., 2000, Gold deposits related to alkaline mag- tralian National University, 106 p.
matism: Reviews in Economic Geology, v. 13, p. 279–314. ——1995, Alkalic-type epithermal gold deposits: A review: Mineralogical As-
Kendrick, M.A., Burgess, R., Pattrick, R.A.D., and Turner, A., 2001, Fluid in- sociation of Canada Short Course Series, v. 23, p. 367–400.
clusion noble gas and halogen evidence on the origin of Cu-porphyry min- ——1997, Controls on scale of Porgera-type porphyry/epithermal gold de-
eralising fluids: Geochimica et Cosmochimica Acta, v. 65, p. 2651–2668. posits associated with mafic, alkalic magmatism: Transactions of the Insti-
Kharaka, Y., Callender, E., and Carothers, W., 1977, Geochemistry of geo- tution of Mining and Metallurgy, v. 106, p. B1–B8.
pressured geothermal waters from the Texas Gulf Coast: Geopressured Ge- Richards, J.P., and Kerrich, R., 1993, The Porgera gold mine, Papua New
othermal Energy Conference, Lafayette, Louisiana, USA, Nov. 16–18, Guinea: Magmatic hydrothermal to epithermal evolution of an alkalic-type
1977, Proceedings, v. 3, p. Gl.121–Gl.165. precious metal deposit: ECONOMIC GEOLOGY, v. 88, p. 1017–1052.
Kontak, D.J., and Kerrich, R., 1997, An isotopic (C, O, Sr) study of the vein Richards, J.P., and McDougall, I, 1990, Geochronology of the Porgera gold
gold deposits in the Meguma terrane, Nova Scotia: Implication for source deposit, Papua New Guinea: Resolving the effects of excess argon on K-Ar
and reservoirs: ECONOMIC GEOLOGY, v. 92, p. 161–180. and 40Ar/39Ar age estimates for magmatism and mineralization: Geochimica
Kwak, T.A.P., 1990, Geochemical and temperature controls on ore mineral- et Cosmochimica Acta, v. 54, p. 1397–1415.
ization at the Emperor gold mine, Vatukoula, Fiji: Journal of Geochemical Richards, J.P., Chappell, B.W., and McCulloch, M.T., 1990, Intraplate-type
Exploration, v. 36, p. 297–337. magmatism in continent-island-arc collision zone: Porgera intrusive com-
Le Bas, M.J., Le Maitre, R.W., Streckeisen, A., and Zanettin, B., 1986, A plex, Papua New Guinea: Geology, v. 18, p. 958–961.
chemical classification of volcanic rocks based on the total alkali-silica dia- Richards, J.P., McCulloch, M.T., Chappell, B.W., and Kerrich, R., 1991,
gram: Journal of Petrology, v. 27, p. 745–750. Sources of metals in the Porgera gold deposit, Papua New Guinea: Evi-
Lyon, G.L., and Hulston, J.R., 1984, Carbon and hydrogen isotopic compo- dence form alteration, isotope, and noble metal geochemistry: Geochimica
sition of New Zealand geothermal gases: Geochimica et Cosmochimica et Cosmochimica Acta, v. 55, p. 565–580.
Acta, v. 48, p. 1161–1171. Richards, J.P., Bray, C.J., Channer, D.M.DeR., and Spooner, E.T.C., 1997,
McKibben, M.A., and Eldridge, C.S., 1990, Radical sulfur isotope zonation Fluid chemistry and processes at the Porgera gold deposit, Papua New
of pyrite accompanying boiling and epithermal gold deposition: ECONOMIC Guinea: Mineralium Deposita, v. 32, p. 119–132.
GEOLOGY, v. 85, p. 1917–1925. Roedder, E., 1958, Technique for the extraction and partial chemical analy-
Merino, E., 1975, Diagenesis in Tertiary sandstones from Kettleman north sis of fluid-filled inclusions from minerals: ECONOMIC GEOLOGY, v. 53, p.
dome, California; II, Interstitial solutions: distribution of aqueous species 235–269.
at 100 degrees C and chemical relation to the diagenetic mineralogy: ——1984, Fluid inclusions: Reviews in Mineralogy, v. 12, 646 p.
Geochimica et Cosmochimica Acta, v. 39, p. 1629–1645. Ronacher, E., Richards, J.P., and Johnston, M.D., 1999, Environments of ore
Müller, D., and Groves, D.I., 2000, Potassic igneous rocks and associated deposition at the Porgera gold deposit, Papua New Guinea: Proceedings of
gold-copper mineralization, 3rd ed.: Berlin, Springer, 252 p. the 5th Biennial SGA meeting: Mineral Deposits: Processes and processing,
Müller, D., Kaminski, K., Uhlig, S., Graupner, T., Herzig, P.M., and Hunt, S., London, Balkema, p. 95–98.
2002, The transition from porphyry- to epithermal-style gold mineraliza- ——2000a, Evidence for phase separation in high-grade ore zones at the
tion at Ladolam, Lihir Island, Papua New Guinea: A reconnaissance study: Porgera gold deposit, Papua New Guinea: Mineralium Deposita, v. 35, p.
Mineralium Deposita, v. 37, p. 61–74. 683–688.
Mutschler, F.E., and Mooney, T.C., 1993, Precious-metal deposits related to ——2000b, Evidence for boiling during ore deposition at Porgera, Papua
alkalic igneous rocks: Provisional classification, grade-tonnage data and ex- New Guinea [abs.]: Geological Society of America Program with Abstracts,
ploration frontiers: Geological Association of Canada Special Paper, v. 40, v. 32, no. 7, p. A280.
p. 479–520. Ronacher, E., Richards, J.P., Villeneuve, M.E., and Johnston, M.D., 2002,
Norman, D.I., Moore, J.N., and Musgrave, J., 1997, Gaseous species in fluid Short life-span of the hydrothermal system at Porgera, Papua New Guinea:
inclusions: Workshop on Geothermal Reservoir Engineering, 22nd, Stan- Laser 40Ar/39Ar dates for roscoelite, biotite, and hornblende: Mineralium
ford University, Stanford, California, Proceedings, p. 419–426. Deposita, v. 37, p. 75–86.
Ohmoto, H., and Goldhaber, M.B., 1997, Sulfur and carbon isotopes, in Saunders, J.A., and May, E.R., 1986, Bessie G: A high-grade epithermal
Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed: New gold telluride deposit, La Plata County, Colorado, U.S.A., in Macdonald,
York, Wiley, p. 517–612. A.J., ed., Gold ‘86: Willowdale, Ontario, Konsult International, p.
Ohmoto, H, and Rye, R.O., 1979, Isotopes of sulfur and carbon: Geochem- 436–444.
istry of hydrothermal ore deposits, 2nd ed., in Barnes, H.L., ed., Geochem- Saunders, J.A., and Schoenly, P.A., 1995, Boiling, colloid nucleation and ag-
istry of hydrothermal ore deposits: New York, Wiley, p. 509–567. gregation, and the genesis of bonanza Au-Ag ores of the Sleeper deposit,
Plumlee, G.S., Leach, D.L., Hofstra, A.H., Landis, G.P., Rowan, E.L., and Nevada: Mineralium Deposita, v. 30, p. 199–210.
Viets, J.G., 1994, Chemical reaction path modeling of ore deposition in Sawkins, F.J., O’Neil, J.R., and Thompson, J.M., 1979, Fluid inclusion and
Mississippi Valley-type Pb-Zn deposits of the Ozark region, U.S. midconti- geochemical studies of vein gold deposits, Baguio district, Philippines:
nent: ECONOMIC GEOLOGY, v. 89, p. 1361–1383. ECONOMIC GEOLOGY, v. 74, p. 1420–1434.
Poliquin, M.J., and Simmons, S.F., 1998, Mineralogical and fluid inclusion Shock, E.L., 2000, Thermodynamic response of organic compounds in geo-
studies of the R-1 zone, Emperor gold mine, Fiji [abs.]: Geological Society chemical processes of sedimentary basins: Reviews in Economic Geology,
of America Program with Abstracts, v. 30, no. 7, p. A302. v. 9, p. 105–117.
Price, L.C., 1981, Methane solubility in brines with application to the geo- Sillitoe, R.H., 1997, Characteristics and controls of the largest porphyry cop-
pressured resource: Geopressured and Geothermal Energy Conference, per-gold and epithermal gold deposits in the circum-Pacific region: Aus-
5th, Louisiana State University, Baton Rouge, Proceedings, p. 205–214. tralian Journal of Earth Sciences, v. 44, p. 373–388.
Reed, M.H., 1982, Calculation of multicomponent chemical equilibria and ——2002, Some metallogenic features of gold and copper deposits related to
reaction processes in systems involving minerals, gases and an aqueous alkaline rocks and consequences for exploration: Mineralium Deposita, v.
phase: Geochimica et Cosmochimica Acta., v. 46, p. 513–528. 37, p. 4–13.

0361-0128/98/000/000-00 $6.00 866


NORTH ZONE HIGH-GRADE AREA, PORGERA DEPOSIT, PNG 867

Simmons, S.F., and Browne, P.R.L., 2000, Hydrothermal minerals and pre- Von Damm, K.L., Edmond, J.M., Grant, B., and Measures, C.I., 1985a,
cious metals in the Broadlands-Ohaaki geothermal system: Implications for Chemistry of submarine hydrothermal solutions at the 21°N, East Pacific
understanding low-sulfidation epithermal deposits: ECONOMIC GEOLOGY, v. Rise: Geochimica et Cosmochimica Acta, v. 49, p. 2197–2220.
95, p. 971–1000. Von Damm, K.L., Edmond, J.M., Measures, C.I., and Grant, B., 1985b,
Smith, R.I., 1990, Tertiary plate tectonic setting and evolution of Papua New Chemistry of submarine hydrothermal solutions at Guayamas Basin, Gulf
Guinea: of Papua New Guinea Petroleum Convention, 1st, Port Moresby, of California: Geochimica et Cosmochimica Acta, v. 49, p. 2221–2237.
Proceedings, p. 229–244. Wall, V.J., Munroe, S.M., Cameron, G.H., Heinrich, C.A., Cox, S.F., and Wal-
Spry, P.G., and Thieben, S. E., 1998, The origin of gold-silver telluride de- she, J.L., 1995, Porgera: Plumbing and processes of gold mineralization, in
posits of the central Montana alkalic belt [abs.]: Geological Society of Clark, A.H., ed., Giant ore deposits – II. Proceedings of the Second Giant
America Program with Abstracts, v. 30, no.7, p. A301. Ore Deposits Workshop: Kingston, Ontario, Queen’s University, p.
Spycher, N.F., and Reed, M.H., 1989, Evolution of a Broadlands-type ep- 649–667.
ithermal ore fluid along alternative P-T paths: Implications for the trans- Walter L.M., Stueber, A.M, and Huston, T.J., 1989, Cl/Br ratios of Illinois
port and deposition of base, precious, and volatile metals: ECONOMIC GE- basin brines: Constraints on origin of fluids [abs.]: Geological Society of
OLOGY, v. 84, p. 328–359. America Program with Abstracts, v. 21, no. 6, p. A316.
Taylor, B.E., 1986, Magmatic volatiles: Isotopic variation of C, H, and S: Re- White, D.E., Hem, J.D., and Waring, G.A., 1963, Chemical composition of
views in Mineralogy, v. 16, p. 185–225. subsurface waters: U.S. Geological Survey Professional Paper 440–F, p.
Thompson, T.B., Trippel, A.D., and Dwelley, P.C., 1985, Mineralized veins 1–67.
and breccias of the Cripple Creek district, Colorado: ECONOMIC GEOLOGY, Wilkinson, J.J., 2001, Fluid inclusions in hydrothermal ore deposits: Lithos,
v. 80, p. 1669–1688. v. 55, p. 229–272.
Truesdell, A.H., 1985, Introduction to chemical calculations: Reviews in Eco- Zhang, X., and Spry, P.G., 1994, Petrological, mineralogical, fluid inclusion,
nomic Geology, v. 1, p. 1–8. and stable isotope studies of the Gies gold-silver telluride deposit, Judith
Viets, J.G., Hofstra, A.H., and Emsbo, P., 1996, Solute compositions of fluid Mountains, Montana: ECONOMIC GEOLOGY, v. 89, p. 602–627.
inclusions in sphalerite from North American and European Mississippi Zheng, Y.-F., 1999, Oxygen isotope fractionation in carbonate and sulfate
Valley-type ore deposits: Ore fluids derived from evaporated seawater: So- minerals: Geochemical Journal, v. 33, p. 109–126.
ciety of Economic Geologists Special Publication 4, p. 465–482.

0361-0128/98/000/000-00 $6.00 867


0361-0128/98/000/000-00 $6.00 868

You might also like