Sunteți pe pagina 1din 21

2005 Society of Economic Geologists, Inc.

Economic Geology, v. 100, pp. 15831603

Intrusion-Hosted Mineralization in the Charters Towers Goldfield,


North Queensland: New Isotopic and Fluid Inclusion Constraints on
the Timing and Origin of the Auriferous Veins
OLIVER P. KREUZER,*
Economic Geology Research Unit, School of Earth Sciences, James Cook University, Townsville, Queensland 4811, Australia

Abstract
Auriferous quartz sulfide veins of the Charters Towers goldfield are mainly hosted in oxidized I-type granitoids of the Ravenswood batholith. K-Ar and Ar-Ar isotope age data of hydrothermal muscovite from alteration
envelopes of veins at Charters Towers and the Hadleigh Castle mine (~40 km east of Charters Towers) are indistinguishable within error, suggesting broadly synchronous gold deposition between 410 and 404 Ma and
across a significant segment of the Ravenswood batholith. Published geochronological data indicate that several granitoid bodies were emplaced into the Ravenswood batholith at the time of gold mineralization. Despite
their association in space and time with igneous rocks, the gold deposits lack an obvious causative intrusion.
Moreover, published lead isotope studies indicate that the lead in the ore was not acquired from any of the exposed intrusions. A distal origin of the ore-forming fluids also may be inferred from nitrogen isotope values of
hydrothermal sericite, suggesting wall-rock interaction with metamorphic fluids or fluids that were in equilibrium with metamorphic rocks during the paragenetic stage of pyrite and arsenopyrite deposition (stage II).
Veins of the Charters Towers goldfield contain three different types of fluid inclusions, which are distinguishable by petrography and microthermometry. This study is the first to report saline fluid inclusions in sphalerite
(18.928.3 wt % NaCl equiv) and vein quartz (20.924.7 wt % NaCl equiv), trapped during the stage of gold
deposition (stage III). The range of temperatures and salinities, particularly of the saline inclusions in sphalerite, could indicate mixing between hotter, more saline fluids (e.g., deep-seated magmatic) and cooler, more
dilute solutions (e.g., modified meteoric) as the cause of gold deposition. The geological and geochemical data
are not compatible with derivation of fluids, metals, and ligands from individual plutons. Similarity of host
rocks, ore element associations, alteration assemblages, structural controls, and tectonic settings strongly suggest that the auriferous veins of the Charters Towers goldfield belong to a group of granitoid-hosted lode gold
deposits that are generally classified as orogenic.

Introduction
THE CHARTERS TOWERS goldfield in northern Queensland
was one of the largest producers of gold in Australia (Solomon
and Groves, 2000), with production in excess of 6 million
ounces (Moz) Au, 1 Moz Ag, 3,000 t Pb, and 1,000 t Cu metal
during the period of 1872 to 1918 (Levingston, 1972). Economic zones were characterized by high average grades of 34
g/t Au (Blatchford, 1953).
Peters and Golding (1989) suggested that the Late Silurian to Early Devonian gold-bearing veins precipitated
from deep-seated magmatic or metamorphic fluids. Sillitoe
(1997) questioned this origin and in a subsequent publication (Sillitoe and Thompson, 1998) assigned the veins to a
class of deposits with an Au-As-Pb-Zn-Cu association, genetically linked to their host intrusions. Groves et al. (2003)
classified the veins of the Charters Towers goldfield as orogenic gold deposits, following similar previous classifications
by Groves et al. (1998), Bierlein and Crowe (2000), and
Goldfarb et al. (2001). Many granitoid-hosted gold deposits,
such as those of the Charters Towers, Etheridge (Bain et al.,
1998), Grass Valley (Johnston, 1940), Parcoy-Pataz
(Schreiber, 1990a, b; Haeberlin et al., 2004), and Jiaodong
districts (Qiu et al., 2002; Fan et al., 2003), have similarities
to orogenic gold deposits but are spatially and, in many
E-mail address: okreuzer@els.mq.edu.au
*Present address: ARC National Key Centre for the Geochemistry and
Metallogeny of Continents (GEMOC), Department of Earth and Planetary
Sciences, Macquarie University, North Ryde, NSW 2109, Australia.

0361-0128/05/3559/1583-21 $6.00

cases, temporally associated with magmatic activity. However, isotopic tracers of fluid sources have failed to provide
definitive evidence of the origin of the ore-forming fluids
(e.g., Ridley and Diamond, 2000).
The eastern part of the main mining area of the Charters
Towers goldfield is hosted by the Millchester Creek tonalite
(Fig. 1A), which was emplaced between 426 4 and 425 4
Ma (based on U-Pb and Rb-Sr biotite whole-rock ages: table
8 in Hutton and Rienks, 1997). Although the ages do not
overlap, the tonalite could be as little as 1 m.y. older than the
auriferous veins, which are enveloped by wall-rock alteration
zones that have whole-rock K-Ar ages of 416 4 to 397 4
Ma (Morrison, 1988) and an Ar-Ar muscovite age of 414.8
1 Ma (Perkins and Kennedy, 1998). Based on this timing, Sillitoe and Thompson (1998) concluded that a genetic relationship between the host intrusion and gold deposits could not
be ruled out. However, Groves et al. (2003) argued that igneous bodies, such as the Millchester Creek tonalite, might
only have contributed heat for convection of externally derived ore-forming fluids.
Previous investigations (e.g., Peters, 1987; Morrison, 1988;
Peters and Golding, 1989; Perkins and Kennedy, 1998) focused on the main mining area (~15 km2) at Charters Towers,
but the geology, geochemistry, and timing of similar auriferous veins elsewhere in the Charters Towers goldfield were
not as well documented. This paper presents new fluid inclusion and stable isotope data that were collected from veins of
the Charters Towers and Hadleigh Castle areas (Fig. 1A).
Data collection and analysis were part of a district-scale

1583

1584

OLIVER P. KREUZER

Fig. 1. A. Simplified map illustrating the main subdivisions of the Ravenswood batholith (modified from Hutton and
Rienks, 1997). Also shown are the locations of veins of the Charters Towers goldfield (veins included in this study are labeled), volcanic-hosted massive sulfide (VHMS) and porphyry-related gold deposits and occurrences (Hutton and Rienks,
1997; Hartley and Dash, 1993; Hartley, 1996; G.W. Morrison, writ. commun., 2003; Australian map grid). B. Outcrop areas
of the Ravenswood and Lolworth batholiths (modified from Hutton and Rienks, 1997; Hutton et al., 1997). Abbreviations:
CT = Charters Towers, MI = Mingela, RAV = Ravenswood, TVL = Townsville.
0361-0128/98/000/000-00 $6.00

1584

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

research project on the prospectivity of the Charters Towers


goldfield and the geology, structure, timing, and genesis of its
lode gold deposits (Kreuzer, 2003, 2004, 2006; Kreuzer and
Alston, 2004).
In this study, data from published records on the Ravenswood batholith (Hutton and Rienks, 1997; Hutton et al.,
1997) are combined with new results from the Charters Towers and Hadleigh Castle areas to test whether fluids were derived from individual plutons or if they are part of a much
larger system of energy and mass flux (cf. Wyborn et al., 1994;
Hronsky, 2004) that postdated the host intrusions. New and
published data are used (1) to more tightly constrain the spatial and temporal relationship between magmatism and gold
mineralization, (2) to better characterize the origin of the oreforming fluids, and (3) to compare the characteristics of the
veins with those of orogenic and intrusion-related deposits in
similar granitoid-hosted gold provinces elsewhere.
The results presented in this paper extend another study by
Kreuzer (2006), in which the paragenetic relationships,
quartz, and ore textures and gold-related hydrothermal wallrock alteration of the veins are described.
Geologic Framework
Most auriferous veins in the Charters Towers goldfield are
hosted by granitoids of the Ravenswood batholith (Fig. 1A).
This composite plutonic mass crops out over >6,000 km2 of
the largely concealed Lolworth-Ravenswood terrane (Fig.
1B), which forms part of the northern Tasman fold belt (e.g.,
Scheibner and Veevers, 2000). Intrusions of the Ravenswood
batholith are mainly oxidized (magnetite-series), hornblendebearing I-type granitoids (Richards, 1980) that were emplaced into a composite basement. To the south, the basement consists of greenschist or lower grade volcanic and
sedimentary rocks of the Cambrian to Ordovician Seventy
Mile Range Group and the Kirk River Beds (e.g., Berry et al.,
1992; Hutton and Rienks, 1997). Basement rocks in the
northern part of the Ravenswood batholith are mainly greenschist to amphibolite-grade metasedimentary rocks of the
Neoproterozoic or Early Cambrian Charters Towers Metamorphics. Metaigneous rocks, such as the Bucklands Hill
diorite, occur as dikes and sills that intruded the basement
during Middle Cambrian time (e.g., Hutton and Rienks,
1997).
The intrusive history of the Ravenswood batholith has been
divided into Early to Middle Ordovician, Middle Silurian to
Middle Devonian, and late Carboniferous to Early Permian
phases of magmatic activity (Fig. 1A; Hutton and Rienks,
1997). These subdivisions are based on isotopic and geophysical data, igneous geochemistry, crosscutting relationships,
and the fabric of the rocks. Early to Middle Ordovician intrusions (Pb-Pb, U-Pb, and Rb-Sr biotite whole-rock ages of 490
6 to 463 3 Ma: table 4 in Hutton and Rienks, 1997) are
mainly granitic and generally strained or recrystallized (Tenison-Woods and Rienks, 1992; Hutton and Rienks, 1997). Following two major regional deformation events (D1 and D2:
Hutton and Rienks, 1997), Middle Silurian to Middle Devonian magmatism (K-Ar, U-Pb, and Rb-Sr biotite whole-rock
ages of 426 4 to 382 5 Ma: table 8 in Hutton and Rienks,
1997) overlapped with Late Silurian to Early Devonian gold
mineralization and regional shortening (D4: Kreuzer, 2004).
0361-0128/98/000/000-00 $6.00

1585

Middle Silurian to Middle Devonian plutons are typically


hornblende and biotite-bearing granodiorites and tonalites,
making up 60 percent of the exposed Ravenswood batholith.
Late Carboniferous to Early Permian plutons and volcanic
rocks (K-Ar and Rb-Sr biotite whole-rock ages of 311 3 to
283 9 Ma: table 9 in Hutton and Rienks, 1997) crop out primarily in the eastern and southeastern parts of the
Ravenswood batholith, accounting for only 6 percent of its
volume (Hutton and Rienks, 1997).
Little is known about the emplacement depth of the intrusions and their current exposure level. However, gravity modeling suggests that the Ravenswood batholith is a 5- to 6-kmthick tabular body, exposed at or close to its roof zone
(Hutton and Rienks, 1997).
Structural framework
Middle Ordovician to Middle Silurian deformation (D1) of
the Ravenswood batholith was likely the result of northeastsouthwest to north-southoriented regional shortening,
having produced mainly sinistral and up to 50-km-long,
east-weststriking mylonitic shear zones (Fig. 1A; Hutton
and Rienks, 1997; Kreuzer, 2004). These zones are locally
overprinted by dextral, northwest-southeaststriking structures that formed during north-southdirected regional
shortening (D2) sometime after D1 but prior to the Middle
Silurian (Hutton et al., 1997). Broad doming of the
Ravenswood batholith or extension along a north-south to
northeast-southwest axis (D3), which overlapped with emplacement of diorite dikes in the Middle Silurian (Peters,
1987; Hutton and Rienks, 1997), was terminated prior to the
Late Silurian. This age is constrained by the Late Silurian to
Early Devonian age of formation of the veins of the
Charters Towers goldfield (Morrison, 1988; Perkins and
Kennedy, 1998; this study). Gold deposition occurred during an episode of northeast-southwestdirected regional
shortening (D4), which was characterized by conditions of
low strain and the brittle reactivation of earlier structures
(Kreuzer, 2004).
Middle Silurian to Middle Devonian magmatism
Crosscutting relationships and limited radiometric data
suggest that auriferous vein formation was temporally and
spatially associated with magmatism, particularly that associated with the emplacement of the Chippendale, Deane,
Carse-O-Gowrie, and Broughton River granodiorites (Rb-Sr
whole-rock ages of 411 2 to 406 4 Ma: Hutton and Rienks,
1997; Fig. 1A).
At the regional scale, the gold mineralization and plutonism
of the Ravenswood batholith overlapped in time with the emplacement of the nearby Lolworth (K-Ar and U-Pb ages of
414382 Ma: Hutton et al., 1997; Fig. 1B) and Reedy Springs
batholiths (U-Pb and K-Ar ages of 410403 Ma: Hutton et al.,
1997). There was also temporal overlap between gold mineralization (K-Ar and Rb-Sr ages of 426398 Ma: Bain et al.,
1998) and granitoid emplacement (Rb-Sr and U-Pb ages of
431420 Ma: Withnall et al., 1997) in the Georgetown inlier.
Based on the Y-undepleted and Sr-depleted geochemical
signature of intrusions of the Ravenswood batholith (cf.
Wyborn et al., 1992), Hutton and Rienks (1997) proposed a
model of melt generation by underplating of the thickened

1585

1586

OLIVER P. KREUZER

crust by basic magmas, followed by emplacement of the


Middle Silurian to Middle Devonian intrusions into a zone of
back-arc extension. A subduction-related origin for the Middle Silurian to Middle Devonian intrusions of the
Ravenswood batholith is supported by: (1) the geochemical
diversity of the granitoids, commonly with volcanic-arc granite signatures (cf. Hutton and Crouch, 1993); (2) the abundance of cogenetic mafic intrusions, commonly with continental magmatic-arc signatures (cf. Hutton and Crouch,
1993); (3) the spatial and temporal association of Middle Silurian to Middle Devonian intrusions with a linear, over 700km-long igneous belt that is subparallel to the margin of the
Australian craton; and (4) their spatial and temporal coincidence with the deposition of thick continental margin sedimentary successions in the adjacent Hodgkinson-Broken
River fold belt (e.g., Henderson, 1987).
Auriferous Veins of the Charters Towers Goldfield
Geology
Detailed descriptions of veins of the Charters Towers
goldfield are given in Reid (1917), Levingston (1972), Peters (1987, 1990, 1993), Peters and Golding (1989), and
Kreuzer (2003, 2004, 2006). The veins are shallow to moderately east-, northwest- to northeast-, and less abundant Sdipping structures. Adjacent to and within most host structures, wall rocks have been fractured into clasts or
pulverized to rock flour, forming coherent cataclasites. The
largest veins (Brilliant and Day Dawn; Fig. 1A) have strike
lengths of 1 to 2.4 km, proven vertical extents of up to 900
m, and widths of up to 15 m and are characterized by abundant splays. Most veins of the Charters Towers goldfield,
however, have strike lengths ranging from 0.1 to 1 km, vertical extents of <400 m, and widths of 0.75 to 1 m (Reid,
1917; Peters, 1987).
Veins of the Charters Towers goldfield are mainly hosted by
medium-grained, hornblende- and/or biotite-bearing granite,
granodiorite, and tonalite, belonging to three igneous suites
defined by Peters (1987) and analyzed by Hutton and Crouch
(1993): (1) the medium- to high K calc-alkaline Hogsflesh
supersuite (Early to Middle Ordovician Hogsflesh Creek
granodiorite and Towers Hill granite), (2) the medium K calcalkaline Millchester supersuite (e.g., Middle Silurian Millchester Creek tonalite), and (3) the high K calc-alkaline
Broughton River suite (Late Silurian to Early Devonian
Broughton River granodiorite). Less than 5 percent of the
known deposits are contained within dioritic, gabbroic, and
metamorphic basement rocks, and only four of these occurrences yielded more than 5,000 oz Au (Reid, 1917; Peters,
1987).
As reported by Peters (1987), each auriferous vein is accompanied by a narrow, symmetrically zoned alteration envelope, generally up to two or three times the width of the vein.
Alteration envelopes can locally be as wide as 40 to 100 m
where they surround complex zones of auriferous quartz
veins. Peters (1987) subdivided the alteration into distal
propylitic and proximal sericitic zones that are generally not
texturally destructive. Sericitic zones appear green in color,
and the broader, reddish propylitic zones are dominated by
montmorillonite and chlorite.
0361-0128/98/000/000-00 $6.00

Mineralogy and geochemistry


Quartz is the major constituent of the auriferous veins of
the Charters Towers goldfield, illustrating euhedral buck,
comb, and modification textures (e.g., Dowling and Morrison,
1989). Sulfide minerals generally make up 10 to 20 vol percent, and locally the sulfide mineral content of the veins can
be as high as 90 percent. The most common sulfide minerals
are pyrite, galena, and sphalerite (>90% of the total sulfide
mineral volume). Chalcopyrite, arsenopyrite, and tetrahedrite-tennantite are less abundant, although gold is associated with these minerals. Pyrrhotite is scarce. Carbonate minerals and clasts or included fragments of wall rock, tourmaline
veins, or dike material each constitute 3 to 5 percent of the
volume of a typical vein (Kreuzer, 2006).
Both native gold and electrum are present in the veins,
forming a continuous compositional series with identical textures, settings, and paragenetic relationships. Gold particles
vary from 5 m to 2 mm in size, and the maximum particle dimension is generally between 40 and 80 m. Most of the gold
is spatially associated with paragenetically earlier pyrite, occurring as minute inclusions in pyrite cubes, fracture infillings, and is attached to pyrite grain margins. Gold is commonly accompanied by and precipitated with galena,
chalcopyrite, or both. Free gold is rare and mainly hosted in
fractures and vugs in quartz that is close to pyrite or other sulfide aggregates (Kreuzer, 2006) and notable amounts were
only recorded at two locations (Mary Lou reef: Marks, 1913;
Day Dawn reef: Reid, 1917). Rare petzite and hessite were
reported by Reid (1917).
Veins of the Charters Towers goldfield formed during four
paragenetic stages (Fig. 2). Stages I and II were significant in
terms of vein growth by deposition of large volumes of buck
quartz and lesser quantities of comb quartz, calcite, and
ankerite. Voluminous pyrite and arsenopyrite precipitation of
stage II was accompanied by wall-rock sulfidation and alteration. Gold was introduced during stage III, after sphalerite
but simultaneously with galena and chalcopyrite, minor pyrite,
and tetrahedrite-tennantite, and scarce petzite, hessite, and
pyrrhotite. Buck and comb quartz were less voluminous in
stage III than in stages I and II. Stage III was also characterized by modification of existing vein quartz by deformation,
overprinting, and fluid interaction (Kreuzer, 2006).
Trace element data (S.C. Dominy, James Cook University,
unpub. data) for 66 chip channel samples from the B lode at
Hadleigh Castle mine are plotted in Figure 3. These data
show that gold concentrations correlate well with silver, iron,
arsenic, and, to a lesser extent, with tellurium. Figure 3D illustrates the correlation between gold and silver and lead,
zinc and copper, which is the metal association of paragenetic
stage III (Fig. 2). Antimony correlates only weakly with gold,
whereas gold and tungsten are not correlated. There appears
to be no association between gold and molybdenum, and gold
and bismuth, which are commonly associated with gold in intrusion-related deposits (Smith et al., 1999; Mustard, 2001;
Baker, 2002).
Previous work
K-Ar dating of hydrothermal muscovite from cored granitic
wall rocks of the Day Dawn and Brilliant reefs (Fig. 1A)

1586

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

Fig. 2. Paragenetic relationships of veins of the Charters Towers goldfield


(CTGF; modified from Kreuzer, 2003, 2006). Quartz nomenclature based on
Hodgson (1989).

yielded absolute ages of 416 4, 407 3, 404 4, and 397


4 Ma (Morrison, 1988). Subsequent Ar-Ar dating by Perkins
and Kennedy (1998) provided an age of 414.8 1 Ma for hydrothermal muscovite from wall rock of the Brilliant reef.
However, the Ar-Ar age spectrum of the sample was not ideal.
Hence, Perkins and Kennedy (1998) suggested that the total
fusion age of 417 Ma may be the best estimate for the time of
gold deposition.
Drill core samples of vein quartz from the Day Dawn, Brilliant, and Queen reefs collected and analyzed by Peters
(1987) and Peters and Golding (1989) contained 80 percent
single-phase liquid inclusions (type 1 of Peters, 1987) with
maximum dimensions of <5 m and 20 percent two-phase,
liquid-rich inclusions (type 2 of Peters, 1987) with maximum
dimensions of 10 m. Analyses of 52 primary inclusions
(type 2 of Peters, 1987) in 16 separate wafers yielded final icemelting temperatures (Tm(ice)) ranging from 4.0 to 7.8C.
Homogenization temperatures (Th) ranged from 140 to
280C with the majority falling in the temperature range of
160 to 230C. From their microthermometry data, Peters
0361-0128/98/000/000-00 $6.00

1587

(1987) and Peters and Golding (1989) calculated salinities


ranging from 5.3 to 11 wt percent NaCl equiv and estimated
trapping temperatures (Tt) between 240 and 300C, based
on an expected formation depth between 2.5 and 3.5 km and
a pressure gradient of 0.25 kbar/km. According to Peters and
Golding (1989), the ore-forming fluids were likely water rich
and of low to moderate salinity. Using reconnaissance laser
Raman spectroscopy, Peters (1987) and Peters and Golding
(1989) detected CO2 in two inclusions of their type 2. Other
dark inclusions were suspected to contain CO2. Based on the
temperature range of CO2 clathrate dissociation (Tm(clath)) in
10 inclusions of Peters (1987) type 2, Peters and Golding
(1989) calculated a salinity of 7.5 wt percent NaCl equiv for
the CO2-bearing inclusions.
The three secondary inclusions of Peters (1987) type 2
yielded Tm(ice) ranging from approximately 2.5 to 3.5C
and Th ranging from approximately 175 to 280C. Salinities
of secondary inclusions of Peters (1987) type 2 were not
calculated.
Calculated oxygen and hydrogen isotope compositions of
ore-forming fluids from Day Dawn, Brilliant, Queen, and Ladybird reef ores measured by Golding and Wilson (1981),
Golding et al. (1987), Peters (1987), and Peters and Golding
(1989) overlap with both magmatic and metamorphic waters.
The 13C values of calcite crystals measured by Peters (1987)
and Peters and Golding (1989) overlap with magmatic 13C
values (8 to 4: Campbell and Larson, 1998), whereas sulfur isotope data fall within the typical range of many porphyry
systems and orogenic lode gold deposits (cf. Hofstra and
Cline, 2000). Golding and Wilson (1981) suggested that oxygen isotope values decreased with depth (15.7 at 100 m
below surface to 12.2 at 900 m below surface), and Peters
and Golding (1989) suggested that this reflected the host rock
(Hogsflesh Creek granodiorite, 18O whole rock = 6.5;
Towers Hill granite, 18O whole rock = 7.3; Millchester
Creek tonalite, 18O whole rock = 7.9).
Black et al. (1997) interpreted the similar 207Pb/204Pb and
206Pb/204Pb ratios of veins in the Charters Towers goldfield
and intrusions of the western Ravenswood batholith to indicate a temporal relationship between granitoid emplacement
and hydrothermal activity. However, they emphasized that
there is no support for the Pb having been derived from any
of the analyzed intrusions. The authors also argued that the
differing 208Pb/204Pb ratios of the veins and the granitoids
were inconsistent with Pb derivation from these intrusions,
implying a more complex source of Pb.
New Data
Ar-Ar geochronology
Two hydrothermal muscovite samples (Figs. 1A, 4) from
sericite alteration zones at the Hadleigh Castle mine were
collected underground from the 04 lode (sample Ar-01,
James Cook University collection number 69326 located at
855 m RL, end of west drive development as on 05 May
2001), a typical fault-fill vein with well-developed ribbon
banding (cf. Kreuzer, 2006), and a narrow, auriferous extension vein (sample Ar-02, James Cook University collection
number 69327 located at 855 m RL, end of west drive development as on 05 May 2001) in the hanging wall of the 04 lode.

1587

1588

OLIVER P. KREUZER

Fig. 3. Bivariate plots illustrating correlations between Au and Ag, Au and As, Au and Te (A-C), elements of paragenetic
stage III (D), and elements that are commonly enriched in magmatic-hydrothermal systems (E-H). Shading in (B), (G), and
(H) show typical As, Mo, and Bi concentrations (ppm) of granitoids of the Ravenswood batholith. Plots are based on ICPMS assay data of 66 pulp samples from the B lode, Hadleigh Castle (S.C. Dominy, James Cook University, unpub. data), with
0.1 to 489.5 g/t Au. As, Mo, and Bi background concentrations in (B), (G), and (H) are based on granitoid whole-rock geochemistry (XRF) data published in Hutton and Crouch (1993) and the OZCHEM database (http://www.ga.gov.au/oracle/
#geochem).
0361-0128/98/000/000-00 $6.00

1588

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

The samples were analyzed at the Western Australian argon


isotope facility, operated by a consortium of Curtin University
and University of Western Australia. After irradiation in the
H5 position of the nuclear reactor at McMaster University,

1589

Canada, the samples were analyzed, using the infrared laser


ablation total fusion technique (e.g., Hodges, 1998). Analytical methods and equipment were similar to those described
by Willner et al. (2004).
Radiometric data from sample Ar-01 fall within a range
from 412.2 2.4 to 405. 0 4.0 Ma; those from sample Ar-02
fall within the range of 409.8 2.4 to 400.1 4.9 Ma (Fig. 5,
Table 1). Two of the analyses showed higher than average
37Ar/39Ar ratios, suggesting possible contamination by a Cabearing phase. The inverse isochron plot for all analyses
(using an average J value of 0.020583 0.000103) gives an inverse isochron age of 405.4 3.7 Ma, a 40Ar/36Ar intercept of
295 82 (atmospheric 40Ar/36Ar air ratio), and a value of the
mean square of weighted deviates (MSWD) of 1.05. The inverse isochron age is within error of the unweighted and
weighted mean ages quoted in Table 1.
The geochronological data from samples Ar-01 and Ar-02
are indistinguishable within error and fall within the range of
previously determined K-Ar (Morrison, 1988) and Ar-Ar
(Perkins and Kennedy, 1998) data for hydrothermal muscovite from Charters Towers. Given the similar ages of the
deposits and the fact that the Ravenswood batholith to the
west of Charters Towers is concealed by younger cover sequences, deposits of the same age may have formed within an
area of at least 1,200 km2.
Fluid inclusions
Fluid inclusions were studied in two auriferous vein samples collected underground from the Maude St Ledger reef
(James Cook University collection 69328 located at the Cross
Vein 2 trial mine via the Central decline, 895 m RL: Australian map grid: 7780271 m N/423800 m E; Fig. 1A) and the
B lode (James Cook University collection 69317 located in the
Hadleigh Castle mine, 2 level, 1110 m RL: local mine grid:
9865 m N/10345 m E). A third auriferous sample from the
Great Britain prospect (James Cook University collection
69329 from diamond drill hole SHDD 001 at 79.8-m depth
downhole: local grid: 9990 m N/5080 m E) was included to
compare the fluid inclusions in granitoid-hosted veins with

Fig. 4. 40Ar/39Ar geochronology samples Ar-01 (04 lode, Hadleigh Castle


mine) and Ar-02 (gently dipping extension vein in the hanging wall of the 04
lode, above sample Ar-01). Boxes indicate the locations of samples Ar-01 (A)
and Ar-02 (B) with respect to the auriferous quartz veins. Photomicrographs
(PPL) illustrating the virtually sulfide-free parts of samples Ar-01 (C) and Ar02 (D) that were selected for 40Ar/39Ar dating. The locations of hydrothermal
muscovite analyses Ar-01/06 and Ar-02/01 are shown in (E) and (F), respectively. The stills also illustrate the high spatial resolution of the laser ablation
method (boxes indicate laser traverses), permitting analysis of individual
grains and making contamination by K- and Ca-bearing phases less likely. Abbreviations: bt = biotite, hbl = hornblende, PPL = plane-polarized light, qtz
= quartz, ser = hydrothermal muscovite.
0361-0128/98/000/000-00 $6.00

Fig. 5. Gaussian plot illustrating the distribution of the


geochronology results, mainly ranging from 410 to 404 Ma.

1589

40Ar/39Ar

Age (Ma)

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

core
rim, #01
core
core
rim, #04
core

406.98
412.15
405.10
404.95
407.42
409.57

2.26
2.36
2.75
4.00
5.75
3.18

12.30
12.47
12.23
12.23
12.31
12.38

40Ar*/39Ar

0.04
0.05
0.07
0.12
0.18
0.09

core
core
rim, #02
core
core
rim, #05

405.42
409.38
409.79
407.26
407.40
400.10

3.66
3.93
2.35
2.30
3.13
4.88

12.24
12.37
12.39
12.30
12.31
12.06

0.11
0.12
0.05
0.05
0.09
0.15

12.24
12.37
12.52
12.42
12.58
12.49

12.51
12.52
12.34
12.43
12.49
12.43

40Ar/39Ar

0.01
0.02
0.02
0.02
0.01
0.02

0.01
0.01
0.02
0.01
0.02
0.02

0.01193
0.01282
0.01278
0.01239
0.01158
0.01178

0.01136
0.01306
0.01245
0.01170
0.01217
0.01189

38Ar/39Ar

37Ar/39Ar

0.01066
0.00977
0.04606
0.01551
0.02230
0.00909

0.00603
0.00417
0.00498
0.00988
0.51263
0.35892

0.00016
0.00018
0.00022
0.00027
0.00031
0.00021

0.00001
0.00018
0.00045
0.00050
0.00029
0.00051

0.00268
0.00260
0.00350
0.00349
0.00814
0.00337

0.00122
0.00115
0.00161
0.00284
0.00439
0.00218

0.00000
0.00000
0.00045
0.00041
0.00092
0.00144

0.00072
0.00016
0.00037
0.00067
0.00061
0.00014

36Ar/39Ar

0.00000
0.00000
0.00015
0.00014
0.00029
0.00051

0.00014
0.00016
0.00023
0.00040
0.00062
0.00029

(cm3)

1.50E-11
1.55E-11
9.09E-12
9.83E-12
1.41E-11
7.96E-12

1.88E-11
1.66E-11
1.78E-11
1.00E-11
8.74E-12
1.86E-11

39Ar

1.12E-14
2.18E-14
1.41E-14
1.68E-14
1.46E-14
1.46E-14

2.17E-14
1.63E-14
2.17E-14
5.59E-15
1.21E-14
2.50E-14

100.00
100.00
98.95
99.02
97.83
96.59

98.31
99.62
99.11
98.41
98.55
99.66

%40Ar*

2 Weighted

1 Weighted

mean (n = 6): 408.2 1.2 Ma, unweighted mean (n = 6): 407.7 2.5 Ma, see Figure 2 for sample location
mean (n = 6): 407.5 1.2 Ma, unweighted mean (n = 6): 406.6 3.2 Ma, see Figure 2 for sample location

Analyses were carried out using the infrared laser ablation total fusion technique of 40Ar/39Ar dating, irradiation standard: Tinto B biotite (409.24 0.71 Ma), J value: 0.020581 0.00103, errors are

#01
#02
#03
#04
#05
#06

Sample Ar-02 (extension vein in the hanging wall of the 04 lode)2

#01
#02
#03
#04
#05
#06

Sample Ar-01 (04 lode)1

Analysis Area

TABLE 1. Measured Isotopic Ratios and 40Ar/39Ar Ages from Hydrothermal Muscovite from Granitic Wall-Rock Samples from the Hadleigh Castle Mine

1590
OLIVER P. KREUZER

those from veins in metamorphic basement rocks. Measurements were conducted on over 50 fluid inclusions in comb
and subhedral buck quartz and honey-colored, iron-poor
sphalerite (Fig. 6A, B), which precipitated during the same
paragenetic stage as the gold (Figs. 2, 7).
These samples contain two different types of fluid inclusions in varying abundance, including single-phase liquid
(type 1), and two-phase liquid-rich, low-salinity (type 2a) to
saline (type 2b) inclusions. Only those fluid inclusions in
growth zones of quartz crystals were considered to be primary

Fig. 6. Photomicrographs showing fluid inclusion types. A and B. Sphalerite-hosted inclusions illustrating clear paragenetic relationships between
primary, secondary, and pseudo-secondary inclusions (Maude St. Ledger
reef, Charters Towers sample BD; PPL). C. Arrangement of primary fluid inclusions (arrow) in growth zones of a subhedral grain of vein quartz (04 lode,
Hadleigh Castle mine; CPL). D. Secondary fluid inclusion trails (arrows)
with different orientations cutting each other and a euhedral crystal of vein
quartz (dark gray), suggesting a complex inclusion history (04 lode, Hadleigh
Castle mine; CPL). Abbreviations: CPL = cross-polarized light, L = liquid, P
= primary fluid inclusions, PPL = plane-polarized light, py = pyrite, PS =
pseudo-secondary fluid inclusions, qtz = quartz, S = secondary fluid inclusions, sp = sphalerite, V = vapor.

Fig. 7. Reflected light photomicrograph of gold, galena, and sphalerite of


paragenetic stage III, having precipitated after quartz and pyrite of stage II
(Maude St. Ledger reef, Charters Towers). Abbreviations: Au = gold, gn =
galena, py = pyrite, qtz = quartz, sp = sphalerite.

1590

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

(Fig. 6C), whereas most other quartz-hosted inclusions are


secondary (Fig. 6D).
Primary type 1 inclusions form subparallel trails, outlining
growth zones in comb quartz, whereas secondary type 1 inclusions are commonly arranged in discontinuous trails, terminating at the boundaries of subhedral to euhedral crystals
of buck quartz. Given that most buck and comb quartz
formed before gold deposition (Fig. 2), it is possible that the
single-phase inclusions represent an early fluid inclusion type.
Type 1 inclusions are also present in sphalerite where they are
restricted to grain boundaries and discontinuous fractures.
These are likely secondary, postdating gold and sulfide deposition of stage III. The composition of these single-phase inclusions is unknown; most have leaked and are empty.
Type 2a inclusions commonly occur in quartz, whereas liquid-rich, L + V, saline aqueous type 2b inclusions are more
abundant in sphalerite. Type 2a and 2b inclusions have similar petrographic characteristics and can only be discriminated
by microthermometry. Large (530 m), isolated or clustered type 2 inclusions with rectangular and round shapes
were considered primary, whereas secondary type 2 inclusions are commonly smaller (<5 m), with elongate and irregular shapes, and are arranged in discontinuous trails
(healed fractures), which terminate against grain boundaries.
Heating and freezing experiments were performed only on
least deformed grains. Inclusions showing obvious signs of
necking down or occurring in clusters with highly variable liquid/vapor ratios were excluded from analysis. Microthermometric data of the analyzed fluid inclusions are summarized in
Figures 8 to 10 and in Table 2.
Temperatures of first melting (Tfm) of secondary quartzhosted type 2a fluid inclusions from the Maude St. Ledger
reef and B lode range from 5.3 to 9.9C. Similar inclusions
in a quartz vein from the Great Britain prospect have slightly
lower Tfm values, ranging from 10.0 to 11.9C. Ice-melting
temperatures (Tm(ice)) of quartz-hosted type 2a inclusions
range from 0.1 to 6.9C, similar to those determined by
Peters (1987) and Peters and Golding (1987), who reported
Tm(ice) of 4.0 to 7.8C from their type 2 inclusions. Homogenization temperatures (Th) range from 177 to 280C,
with the majority of inclusions homogenizing into a liquid
phase between 220 to 240C (Fig. 8). The range of Th values
of type 2a inclusions is similar across the studied deposits and
overlaps with data obtained by Peters (1987) and Peters and
Golding (1987) from their type 2 inclusions. The Tfm of secondary type 2a inclusions in sphalerite from the Maude St.
Ledger and the B lode range from 7.1 to 9.5C (excluding
one value of 2.7C). Ice-melting temperatures range from
0.5 to 6.6C, whereas Th values are significantly lower
than those of quartz-hosted type 2a inclusions, ranging from
80 to 137C.
Microthermometric data of type 2b inclusions differ
markedly from those of type 2a fluid inclusions (Fig. 9). Temperatures of first melting of secondary, quartz-hosted type 2b
inclusions in the B lode range from 31.2 to 42.2C,
whereas Tm(ice) ranges from 17.9 to 23.6C. Primary, sphalerite-hosted type 2b inclusions in the two samples from the
Maude St. Ledger reef and B lode have Tfm values ranging
from 28.7 to 45.3C. The Tm(ice) ranges from 15.3 to
29.4C, and the Th values range from 121 to 254C and
0361-0128/98/000/000-00 $6.00

1591

Fig. 8. Histograms of freezing and heating data for type 2a and 2b fluid
inclusions. A. Drill core sample from the Great Britain prospect, Charters
Towers (GB 128). B. Underground mine sample from the B lode, Hadleigh
Castle (B 2-1). C. Underground mine sample from the Maude St. Ledger
reef, Charters Towers (BD). Abbreviations: Tfm = temperature of first melting, Th = homogenization temperature, Tm(ice) = temperature of final ice
melting.

thus are generally higher than those of secondary, sphaleritehosted type 2a fluid inclusions. Homogenization temperatures of quartz-hosted type 2b fluid inclusions were not
recorded in this preliminary study.
Salinities were calculated using the equation of Bodnar
(1993), which is applicable to most H2O-NaCl-KCl fluid inclusion compositions. Secondary type 2a inclusions in quartz

1591

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

1592

5-25

5-25

5-30

<20

<5
<5

Irregular

Irregular

Elongate

Irregular

Equant
Equant

Shape

P and S
S

Timing

Sphalerite

Quartz

Sphalerite

Quartz

Quartz
Sphalerite

Host mineral

Stage IIc

Stage IIc

Stage IIc or IId

Stage IIa to IId

Stage IIc to IId


Stage IIc or IId

Paragenetic stage

B 2-1: 235-251 (n = 3)
BD: 121-254 (n = 18)

B 2-1: 177-305 (n = 43)1

B 2-1: 177-305 (n = 43)1


BD: 219-256 (n = 17)
GB 128: 221-225 (n = 8)
B 2-1: 080-137 (n = 3)
BD: 105-107 (n = 3)

Th [C]2, 3

B 2-1: 28.0-28.3 (n = 3)
BD: 18.9-28.3 (n = 22)

B 2-1: 20.9-24.7 (n = 7)

B 2-1: 0.7-10.4 (n = 30)


BD: 0.2-1.6 (n = 8)
GB 128: 0.2-8.5 (n = 8)
B 2-1: 5.4-6.0 (n = 3)
BD: 0.9-10.0 (n = 3)

Salinity

H2O-NaCl-FeCl2, H2O-Na2CO3K2CO3, H2O-NaCl-MgCl2, H2OFeCl2, or H2O-MgCl2


H2O-NaCl-FeCl2, H2O-Na2CO3K2CO3, H2O-NaCl-MgCl2, H2OFeCl2, or H2O-MgCl2

H2O-(NaCl)

H2O-(NaCl)

Composition

2 Data

of type 2a and 2b inclusions


collected at the School of Earth Sciences, James Cook University, having used a Linkam DSC 600 heating-freezing stage mounted on an Olympus BX 51 system microscope
3 Heating-freezing temperatures were controlled to within 0.1C, stage calibration was carried out using the fluid standards H O-CO (25 mol % CO ) and H O (critical density), mean calibration
2
2
2
2
error was 0.3C
Note: See Figure 2 for sample locations and Figure 4 for details on paragenetic relationships of veins of the Charters Towers goldfield
Abbreviations: L = liquid, P = primary, S = secondary, V = vapor

5-10

L+V

1 Range

5-10

L+V

5-10

L+V

2b

20-25

L+V

2a

0
0

L
L

Vapor (%) Size (m)

Phases

Fig. 10. Plot of salinity vs. homogenization temperature (Th) for fluid inclusions in quartz and sphalerite. Data for sphalerite-hosted fluid inclusions
in the Maude St. Ledger reef (Charters Towers) and B lode (Hadleigh Castle mine) indicate a range of salinities and temperatures that individually may
indicate mixing. Boxes represent ranges of salinity and Th of quartz-hosted
inclusions in samples from the Maude St. Ledger reef (BD), B lode (B 2-1)
and Great Britain prospect (GB 128), and Charters Towers ores previously
investigated by Peters (1987) and Peters and Golding (1989). Quartz-hosted
type 2b inclusions are not plotted as their Th were not measured. Inset illustrates typical trends in Th-salinity space: I = immiscible fluids, II = fluid mixing, III = cooling, IV = boiling with cooling (modified from Wilkinson, 2001).
Note: 1 = plotted from fluid inclusion data of Peters (1987), Peters and Golding (1989).

Type

Fig. 9. Temperature of first melting (Tfm) vs. temperature of final ice


melting (Tm(ice)), illustrating the different properties of type 2a and 2b fluid
inclusions. Note that Tfm is approximate only as the earliest first melting may
not be visible.

TABLE 2. Summary of Fluid Inclusion Characteristics in the Three Samples from the B Lode (sample B 2-1, Hadleigh Castle Mine), Maude St. Ledger Reef
(sample BD, Charters Towers), and Great Britain Reef (sample GB 128, Great Britain Prospect, Charters Towers)

1592
OLIVER P. KREUZER

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

and secondary type 2a inclusions in sphalerite have Tfm values


that are well below the eutectic temperature (i.e., 21.2C) of
the H2O-NaCl system (cf. Roedder, 1984; Shepherd et al.,
1985; Brown, 1998). Salinities of type 2a inclusions in quartz
range from 0.2 to 10.4 wt percent NaCl equiv, similar to that
of type 2a fluid inclusions in sphalerite (0.910 wt % NaCl
equiv). The Tfm values of type 2b inclusions are well below
20.8C, so the presence of additional salts is implied (cf.
Roedder, 1984; Shepherd et al., 1985; Brown, 1998). Eutectic temperatures of type 2b inclusions range from 31.2 to
42.2C (quartz-hosted) and from 28.7 to 45.3C (sphalerite-hosted). Therefore, these fluid inclusions contain a
more complex salt system (e.g., H2O-MgCl2, H2O-FeCl2,
H2O-NaCl-MgCl2, H2O-NaCl-FeCl2, H2O-Na2CO3-K2CO3,
or H2O-NaCl-CaCl2). Calculated salinities for type 2b fluid
inclusions range from 20.9 to 24.7 wt percent NaCl equiv
(quartz-hosted) and 18.9 to 28.3 wt percent NaCl equiv
(sphalerite-hosted). Average salinities of all type 2b inclusions
are approximately 23 wt percent NaCl equiv. Salinities above
26 wt percent NaCl equiv may be overestimations as none of
the analyzed type 2b inclusions contained any daughter salts.
Alternatively, the fluids may be undersaturated with respect
to other salts if the major cations are not Na (Roedder, 1984).
When the Th values are plotted against the salinity of fluid inclusions in sphalerite (Fig. 10) they indicate a range of fluid
compositions from a hotter more saline to a cooler, more dilute fluid.

Isochores calculated for individual inclusions using the program FLINCOR (Brown, 1989) indicate pressures in the
range of 0.5 to 4.7 kbars, suggesting formation depths between ~2 and 17.5 km (Fig. 11). However, most pressure estimates fall within the range of 0.9 to 3.8 kbars (equivalent to
depths between 514 km).
The wide range of pressure estimates may reflect a decompression path related to Silurian to Devonian uplift of the
Ravenswood batholith, as proposed by Hutton and Rienks
(1997). The common occurrence of structures, such as (1) hydraulic breccia (cf. Jbrak, 1997), (2) flat-lying to shallow-dipping extensional veins, and (3) open-space filling textures
(Kreuzer, 2004, 2006), which, at formation depths >2 km, are
indicative of fluid overpressuring (e.g., Harley and
Charlesworth, 1996), suggests that veins of the Charters Towers goldfield formed under conditions of lithostatic to
supralithostatic fluid pressure (Kreuzer, 2004). When superimposed on a lithostatic fluid pressure gradient, the likely
depth range of vein formation can be further restricted to
~7.5 to 10 km, which is consistent with (1) the range of fluid
inclusion-based P-T estimates for gold deposits within greenschist facies rocks (e.g., McCuaig and Kerrich, 1998), (2) the
approximate P-T boundary conditions of abundant quartz
veining (e.g., Bons, 2001), (3) the approximate P-T boundary
conditions for greenschist facies metamorphism (e.g., Yardley, 1989), and (4) the approximate depth ranges of the seismogenic zone (e.g., Sibson, 2001), the fault-valve mechanism

Fig. 11. Isochore plots and pressure estimates for type 2a and 2b sphalerite- and quartz-hosted fluid inclusions. Trapping
pressures have been estimated by constructing isochores (each representing a sample) from fluid inclusion data, using the
program FLINCOR (Brown, 1989). Isotopic equilibrium temperatures based on quartz-mica (170200C, 200230C, and
330360C) and sphalerite-galena (310C) isotope fractionation data of Peters and Golding (1989) provided independent
trapping temperature (Tt) estimates for defining positions along the constructed isochores, as described by Roedder (1984)
and Shepherd et al. (1985). Calculations in FLINCOR were based on the H2O-NaCl system, applying the equation of state
from Brown and Lamb (1989). For the purpose of pressure calculations for type 2a and 2b inclusions components other than
H2O and NaCl were neglected. A. Drill core sample GB 128 (Great Britain prospect, Charters Towers). B. Underground
mine sample B 2-1 (B lode, Hadleigh Castle). C. Underground mine sample BD (Maude St. Ledger reef, Charters Towers).
Abbreviations: gn = galena, ms = hydrothermal muscovite, qtz = quartz, sp = sphalerite, Tt = trapping temperature.
0361-0128/98/000/000-00 $6.00

1593

1593

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

1594

Maude St. Ledger


Maude St. Ledger
Maude St. Ledger

02 lode
04 lode
04 lode
04 lode
B lode

Robinson Crusoe

Hanging-wall gdrt

CT
CT
CT

HC
HC
HC
HC
HC

HC

HC

Whole rock

Muscovite

Muscovite
Muscovite
Muscovite
Muscovite
Muscovite

Muscovite
Muscovite
Muscovite

Material1

N-HG01

N-HC06

N-HC01
N-HC02
N-HC03
N-HC04
N-HC05

N-CT01
N-CT02
N-CT03

Sample no.

69304

69298

69297
69296
69301
69303
69299

69295
69300
69302

JCU no.

Robinson Crusoe pit

Hadleigh Castle mine

Central decline

0460782 m E, 7777135 m N (AMG)

855/02 m RL
855/04 m RL (west)
920/04 m RL (west)
855/04 m RL (west)
2 level, 1110 m RL

894 to 896 m RL
894 to 896 m RL
894 to 896 m RL

Sample location

72.3

78.1

74.8
74.4
80.7
78.6
77.3

79.0
77.5
77.5

Mass2 (mg)

1.75

9.31

7.93
9.83
16.61
17.21
14.22

10.27
14.03
10.81

15N2 ()

0.40

0.89

1.14
3.28
1.19
1.40
2.46

1.16
1.25
1.06

Total N2 (g)

5.6

11.4

15.2
44.1
14.8
17.9
31.8

14.7
16.1
13.7

1 Mineral separates of samples N-CT01 to N-CT03 and N-HC01 to N-HC06 were produced at the Institute of Geological and Nuclear Sciences (Lower Hutt, New Zealand), using standard separation techniques, concentrations to >97% purity were achieved by handpicking under the binocular microscope, pulverized mineral separates were loaded into tin capsules and weighted
2 Analyses were carried out on a high-precision, continuous flow isotope ratio mass spectrometer (CF-IRMS) at the Centre for Soil Research (University of Saskatchewan, Canada), analytical precision (1) for mica separates is typically 0.2 per mil for 15N (Jia and Kerrich, 2000)
Note: see Figure 2 for sample locations
Abbreviations: AMG = Australian map grid, CT = Charters Towers, gdrt = granodiorite, HC = Hadleigh Castle, JCU no. = James Cook University collection number

Vein/intrusion

Fig. 12. Plot of N content vs. 15N values of hydrothermal muscovite


from alteration envelopes of auriferous veins at Charters Towers and
Hadleigh Castle. The graph also shows a value obtained from unaltered
hanging-wall granodiorite (bulk granitoid HC) near the Hadleigh Castle
mine. Fields of Archean and Phanerozoic hydrothermal muscovite, and
Archean igneous feldspar, biotite, and muscovite are based on Jia and Kerrich (1999, 2000).

Area

(e.g., Gaboury and Daigneault, 2000), and occurrence of coherent cataclasite (e.g., Twiss and Moores, 1992).
N (ppm)

Nitrogen isotopes
Samples of sericitized, granitic wall rocks next to auriferous
veins at Charters Towers and Hadleigh Castle were analyzed
for nitrogen isotopes (Table 3). Nitrogen in rocks and minerals has been shown to be an effective tracer of distal fluid origin (Faure, 1986; Jia and Kerrich, 1999, 2000; Glasmacher et
al., 2003; Jia et al., 2003). K-bearing silicates such as potassium feldspar, biotite, and muscovite are the most suitable
rock-forming minerals for such analyses, as K+ in these minerals can be partly replaced by NH4+ (Faure, 1986; Jia and
Kerrich, 1999, 2000). In addition, NH4+ may replace minor
amounts of Na+ in feldspars and clay minerals (Glasmacher et
al., 2003).
Although most samples had very low nitrogen concentrations (1.5 g/g), hydrothermal muscovite grains from auriferous veins at Charters Towers and Hadleigh Castle had N
contents ranging from 11 to 44 ppm and 15N values between
8 and 17 per mil. Most 15N values are >10 per mil (Fig. 12).

TABLE 3. Nitrogen Isotope Data from Hydrothermal Muscovite from Auriferous Veins at Charters Towers, Hadleigh Castle, and Robinson Crusoe,
also including a Whole-Rock Sample of Unaltered Granitoid from near the Hadleigh Castle Mine

1594
OLIVER P. KREUZER

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

These data are incompatible with meteoric surface waters


(15N <10) and mantle-derived (N = 12 ppm, 15N =
5) or granitic fluids (N = 2127 ppm, 15N = 610: Jia
and Kerrich, 2000; Jia et al., 2003), but they are compatible
with values of fluids of metamorphic origin. In particular, the
Charters Towers goldfield data overlap with N concentrations
and 15N values of Archean quartz veins (N = 2070 ppm,
15N = 1124) that are interpreted to be products of metamorphic fluids (Jia and Kerrich, 2000).
Discussion
Any genetic model for the origin of the gold in veins of the
Charters Towers goldfield must explain the similar ages (Fig.
13) and the uniformity of veins and zones of gold-related

wall-rock alteration within an area of at least 1,200 km2 and to


depths of at least 1 km (Fig. 1A), the lack of zoning with respect to individual intrusions, and that the veins postdate the
emplacement, crystallization, and brittle fracturing of their
host intrusions. Similar features, including a spatial and temporal link between gold deposition and I-type magmatism
were recognized in the Etheridge and Cape York peninsula
goldfields (Bain et al., 1998; Table 4). These characteristics
suggest that the ore-forming fluids in North Queensland were
more likely sourced from a single regional-scale, tectonothermal system of Late Silurian to Early Devonian age rather than
from single intrusions.
Veins of the Charters Towers goldfield are not obviously
genetically linked to a particular intrusion, and Figure 14

Fig. 13. Plot of median ages (with 2 ranges) of granitoids of the Ravenswood batholith (K-Ar biotite, K-Ar hornblende,
Rb-Sr whole rock, and U-Pb) and hydrothermal muscovite in alteration envelopes of auriferous veins of the Charters Towers goldfield (Ar-Ar and K-Ar), illustrating temporal overlap of granitoid emplacement and gold mineralization. 1 = Morrison (1988), 2 = Perkins and Kennedy (1998), 3 = this study, 4 = Hutton and Rienks (1997), 5 = intrusions of the Millchester
supersuite: Beasley Creek tonalite, Boatswain granodiorite, Casey Spring Creek granodiorite, Centauri granodiorite, Crescent granodiorite, Dalmore granodiorite, Emu Mill granodiorite, Five Mile Mill granodiorite, Heathfield West tonalite,
Meadowale granodiorite, Merriland tonalite, Molly Darling granodiorite, Spondulix granodiorite, Tullegorim granodiorite,
Two Mile granite, Urdera granodiorite, Wellington Springs tonalite, Wharleys tonalite, Yulga tonalite; intrusions of the
Barrabas supersuite: Kedumba granodiorite, Mount Cuthbert granodiorite; unassigned intrusions: Amity aplite, Balfes Creek
granodiorite, Box Forest quartz-diorite, Kirkton tonalite, and Scoop Hills granodiorite (Hutton and Rienks, 1997).
0361-0128/98/000/000-00 $6.00

1595

1595

1596

OLIVER P. KREUZER
TABLE 4. Properties of Auriferous Veins in the Charters Towers
Charters Towers,
Queensland

Majors Creek,
New South Wales15

Jiadong, China16

Jubilee Plunger,
Cumberland, Havelock
Large, multistage qtz veins

Dargues reef

Linglong, Sanshanado

Disseminations

Large, multistage qtz veins


and disseminations

>0.2 km, all open at depth

<0.2 km, some open at


depth
Within an area of at least
3 1 km

>1 km

Coeval Au-U-F-Mo,
younger porphyry and
epithermal3

No data

No data

20 t Au and Ag (33 g/t Au,


includes Ag bullion)

>0.06 t Au (>4 g/t),


alluvial: >40t Au

>900 t Au (3-30 g/t Au,


mainly 10 g/t Au)

>1, commonly 4
Oxidized, metaluminous
I-type granite, granodiorite,
and tonalite of the Ravenswood batholith3

No data
S-type granitoid of the
Forsayth batholith and
I-type granitoid of the
White Springs batholith

No data
Oxidized, metaluminous
I-type granodiorite of the
Bega batholith

No data
Granite and granodiorite of
the Linglong and Guojialing
suites

Mainly greenschist facies


metamorphic grade,
dominated by igneous rocks,
contact aureoles absent or
poorly developed, basement
to intrusions dominated by
shallow marine sedimentary
rocks3
426 4 to 382 5 Ma
(K-Ar, Rb-Sr, U-Pb)3

Greenschist to granulite
facies metamorphic grade,
dominated by metasedimentary and metaigneous
rocks

Mainly greenschist facies


metamorphic grade,
dominated by igneous rocks

Greenschist to granulite facies


metamorphic grade, basement
to intrusions dominated by
metavolcanic and metasedimentary rocks

424 11 to 404 11 Ma
(Rb-Sr, U-Pb)

415 4 to 399 6 Ma
(K-Ar, Rb-Sr)

416 4 to 397 4 Ma
(K-Ar)5,6, 400 5 to
412 2 Ma (Ar-Ar)
Qtz > sulfides >>
carbonates; open spacefilling textures common
Subhedral to euhedral buck
qtz, comb qtz, modified
gray qtz7
Low to moderate (>10 vol%)
Au, Ag, As, Pb, Zn, Cu
Te, Sb, Hg, W8

426 5 to 398 3 Ma
(K-Ar), 407 6 (Rb-Sr)

411 5 to 406 4 Ma
(K-Ar)

Intrusions: 165 to 125 Ma


(U-Pb), mafic to intermediate
dikes: 124 to 120 Ma (K-Ar)
123 to 121 Ma (Ar-Ar, Rb-Sr),
130 to 100 Ma (K-Ar, Rb-Sr)

Qtz > sulfides >>


carbonates; open spacefilling textures common
Euhedral buck qtz,
comb qtz

Sulfides and ser >>


carbonates and qtz; open
space-filling textures rare
Rare comb qtz

Low to moderate (>10 vol%)


Au, Ag, Pb, Zn, Cu

Moderate (>15-30 vol%)


Au, Cu, As, Bi, Mo, Pb, Te

Low to moderate (>10 vol%?)


Au, As, Pb, Zn, Cu

Au, el, py, gn, sp, cpy


(apy, po, td-tn, tell)1,9,10
No evidence for lateral
zoning9, possible vertical
zoning of Zn and Pb, possible
decrease of average Au grade
with depth (55 g/t near
surface --> 14 g/t at 900 m
below surface)1
Qtz, ser, cal, ank9,10
Ser, cal, ank, py (chl, ep)9,10

El, py, gn, sp, cpy

Py (Au, cpy, gn, Bi-sfs, Bi,


tell, td, po)
No evidence

Au, el, py, gn, sp, cpy, apy, hm


(mt, po, sch, tell, mo)
No evidence

Qtz, ser, cal


Ser, py, chl, carbonates,
clays, ep, rt, qtz, hm (K-fs)

Qtz, cal, ank, ser, K-fs, sd, chl


Qtz, ser, plag, K-fs, chl, cal
(sd, dol, bar)

Examples

Brilliant, Day Dawn, Queen

Principal mineralization
style

Large, multistage qtz veins

Vertical continuity
(maximum)
Spatial distribution of
mineralization

~1 km, many open at


depth1,2
Within an area of at least
60 30 km (possibly up to
230 70 km9)
Older VHMS (e.g.,
Highway, Reward)3, younger
porphyry Cu-Au (e.g.,
Mount Leyshon)3
>200 t Au (33g/t Au)2,4

Spatially associated
deposits and occurrences

Cummulative production/
reserves (average grade
in brackets)
Au / Ag ratio
Principal host rocks

Host terrane

Age of host rocks and/or


associated magmatism
Age of gold-related
wall-rock alteration
Vein/lode mineralogy

Principal quartz types

Sulfide content
Metal association

Main ore minerals (minor


constituents in brackets)
Metal zoning

Main vein gangue


Main wall-rock alteration
gangue (minor
constituents in brackets)

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

Etheridge, Queensland14

Within an area of at least


120 30 km

No evidence

Qtz, cal
Chl, clays, ser

1596

Within two areas of at least


150 60 km and 60 30 km

Qtz > sulfides >> carbonates(?);


open space-filling textures rare
Buck, qtz, rare comb qtz,
modified grey qtz

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

1597

Goldfield Compared with Veins in Similar Gold Provinces Worldwide


Grass Valley, California17

Parcoy-Pataz, Peru18

Intrusion-related gold deposits19

Phanerozoic orogenic gold deposits20

Empire, North Star, Dromedary

La Lima, El Gigante

Large, multistage qtz veins

Large, multistage qtz veins

Mother Lode, Bendigo, Macraes,


Dolgellau
Large, multistage qtz veins

>1 km

>1 km, many are open at depth

Dublin Gulch, Brewery Creek,


Timbarra
Sheeted qzt veins, disseminations,
skarn, breccia pipes, replacements,
greisen
<0.5 km

>1 km

Within an area of at least 6 4 km

well developed

Generally centered on causative


intrusion

Generally within large areas of at


least 30 10 km

No data

No data

W Sn, Mo, Bi

Older VHMS

>230 t Au (8 to 16 g/t Au)

~200 t Au (<20 g/t Au)

Most <20 t Au (most between


0.8 and 3 g/t Au)

Many >50 t (most between


5 and 30 g/t Au)

No data
Granodiorite and metamorphic
basement rocks

1
Oxidized I-type monzogranite and
granodiorite of the Pataz batholith

>1, commonly 10
Marine sedimentary sequences

Greenschist to amphibolite facies


metamorphic grade, basement to
intrusions dominated by metavolcanic and metasedimentary
rocks

Greenschist to lower amphibolite


facies metamorphic grade

1 to ~1
Mostly metaluminous to weakly
peraluminous intrusions of felsic
to intermediate composition,
intrusions span the boundary
between reduced ilmenite and
oxidized magnetite series
Mainly greenschist facies
metamorphic grade, contact
aureoles up to 4 km wide

127 3 Ma (K-Ar)

329 1 Ma (U-Pb)

Host rocks generally of same age


as gold deposits

Mother Lode district: 110 to


125 Ma (K-Ar, Rb-Sr)

314 to 312 Ma (Ar-Ar)

Not applicable

Host rocks generally much older than


gold deposits, gold mineralization
often bracketed by magmatic activity
Not applicable

Qtz >> carbonates > sulfides;


open space-filling textures common

Qtz > sulfides >> carbonates;


open space-filling textures common

Qtz > fs >> sulfides; open


space-filling textures common

Qtz > carbonates >> sulfides; open


space-filling textures not common

Buck qtz, comb qtz, modified


grey qtz

Buck qtz, comb qtz, modified


grey qtz

Buck qtz, comb qtz, modified


grey qtz, crustiform qtz

Buck qtz, modified grey qtz

Low (<10 vol %?)


Au, Ag, As, Pb, Zn, Cu, W

Low to moderate (>10 vol %)


Au, Ag, As, Fe, Zn Cu, Sb,
Bi, Te, W

Au, el, py, apy, gn, sp, cpy, hm


(mo, As, pot, sch, td, tell)
No evidence

Au, py, apy, gn, sp


(el, cpy, po, wf, td-tn)
No evidence

Generally low (<5 vol%)


Generally low (<5 vol%)
Proximal to intrusions: Au, Bi, W
Au, As, B, Bi, Hg, Sb, Te, W
Te, Mo, As; distal to intrusions:
Au, As, Sb Hg; Au, Ag, Pb, Zn
Cu associations typical for most distal
deposits (up to several kilometers
away from the causative intrusions)
Au, Bi, apy, py, po
Au, apy, py, po

Qtz, ser, cal, ank, chl


Ank, ser, py, chl, ep

Qtz, ser, chl, fuch, ank, dol, cal


Ser, chl, fuch, ank, dol, cal

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

Mainly lower to medium greenschist


facies metamorphic grade, dominated
by marine sediments

Strong district-scale zoning:


Au-W Sn-Ag Pb Zn

Cryptic with little change in


mineralogy or Au grade

Qtz, carbonates, K-fs, plag


Silicates (K-fs, ser), carbonates

Qtz, carbonates, K-fs, chl, ser


Carbonates (ank, dol, cal) and sulfides
(py, po, apy)

1597

1598

OLIVER P. KREUZER
TABLE 4.

Wall-rock alteration (less


common types in brackets)
Lateral zoning of wallrock alteration
Elements involved in
fluid/wall-rock reactions
Fluid parameters

Stable isotope data

Inferred depositional
mechanism
Inferred fluid source

Depositional environment
Strain field
Controls on ore zones

Regional/district
association

Inferred tectonic setting

Current classifications

Charters Towers,
Queensland

Etheridge, Queensland14

Majors Creek,
New South Wales15

Jiadong, China16

Sericitization, propylitization
(chloritization, silicification)9
Well-developed9

Propylitization, chloritization
(sericitization)
Well-developed

Propylitization, sericitization
(silicification)
Well-developed

Silification, sericitization,
potassic alteration (chloritization)
Well-developed

As, K, Na, Pb, S, Si, Sr, Zn10

Not specified

Not specified

Neutral to weakly acidic,


CO2-poor fluids of low (0.211 wt % NaCl equiv)9,10,11 to
moderate (18.9-28.3 wt %
NaCl equiv)10 salinity, Tt =
240 to 310C10
18O = 12.3 to 15.79,11,12,
D = -57 to -429,11,
13C = -4.7 to 3.69,11,
34S = -1.1 to 4.29,11,
15N = 7.9 to 17.2
Fluid mixing and
chemisorption10

CO2-bearing fluids of low


(6-10 wt % NaCl equiv)
salinity, Tt = 250 to
300C (possibly up to
350C)

Neutral to weakly alkaline,


CO2-bearing fluids of low
(4-11 wt% NaCl equiv)
salinity, Tt < 350C

Ag, Al, As, Au, Ba, Bi, Ca, Cu,


Fe, K, Mg, Na, Sb, Si, Sr, W, Zn
CO2-rich fluids of low (6-14
wt% NaCl equiv) salinity, rare
moderate salinity inclusions
(32 wt% NaCl equiv), Tt =
250 to 350C

18O = 8.4 to 15.7

18O = 6.5 to 10.9,


13C = -3.8 to -1.7,
34S = -1.5 to -0.4

18O = 7.0 to 13.1,


D = -91 to -63,
13C = -5.9 to -3.4,
34S = 3.0 to 14.0

Not specified

Desulfidation and
chemisorption

Ore stage I: fluids that were


in equilibrium with metamorphic rocks, ore stage II:
mixing of deep-seated
magmatic and significantly
modified meteoric fluids
~1 to 4 kbars, ~5 to 10 km,
low mean stress
Shortening9,13
Regional-scale: major fault
zones, district-scale: secondand third-order faults close
to regional-scale faults and
preexisting discontinuities
(e.g., geological/rheological
contacts, deposit-scale: jogs/
bends in gently to moderately
dipping segments of reverse
faults, splay faults, hydraulic
extension fractures1,9,10

Primary magmatic fluids


or extremely modified
meteoric fluids

Late-stage
orthomagmatic fluids

Fluid-wall rock interaction,


fluid immiscibility,
chemisorption
Magmatic fluids (degassing
from magma parental to mafic
and intermediate dikes)

>5 kbars, >3 km

Not specified

1 to 3 kbars, 4 to 8 km

Not specified
District-scale: faults and
fault abundance, hetereogeneities around geologic
contacts

Not specified
Deposit-scale: granitoid
roof zone, fractures

Shortening?
Regional-scale: major fault
zones, district-scale: major
fault zones and subsidiary
faults (especially where they
juxtapose geological contacts),
deposit-scale: fault jogs/bends,
tension gashes, en echelon
fractures

Partial crustal melting, crustal


heating, regional-scale granitoid emplacement, uplift of
thickened crust3, regional
shortening9,10 reactivation
of preexisting structures,
possible temporal link with
W-directed Pacific plate
subduction10
Well inboard of inferred or
recognized collisional plate
boundaries, potential continetal magmatic arc setting3,10

Partial crustal melting,


granitoid emplacement,
retrograde greenschist facies
metamorphism, uplift of
thickened crust, possible
temporal link with
W-directed Pacific plate
subduction

Late-stage magmatichydrothermal activity in the


crystallized roof zone of the
host intrusion

Partial crustal melting,


granitoid emplacement, uplift
of thickened crust, reactivation
of preexisting structures,
possible temporal link with
orogenesis and/or Pacific
plate subduction

Well inboard of inferred or


recognized collisional plate
boundaries, potential continetal magmatic arc setting?

Well inboard of inferred or


recognized collisional plate
boundaries, potential continetal magmatic arc setting?

Mesothermal, intrusionrelated, plutonic, orogenic

Mesothermal, plutonic

Mesothermal, epithermal,
granitoid-associated

Craton margin position


significantly landward relative
to subduction zone of Pacific
plate, potential magmatic arc
setting
Orogenic, intrusion-related

Sources: 1 = Reid (1917), 2 = Levingston (1972), 3 = Hutton and Rienks (1997), 4 = Blatchford (1953), 5 = Morrison (1988), 6 = Perkins and Kennedy
(1998), 7 = Dowling and Morrison (1989), 8 = G.W. Morrison (writ. commun., 2003), 9 = Peters (1987), 10 = Kreuzer (2003, 2006), 11 = Peters and
Golding (1989), 12 = Golding and Wilson (1981), Golding et al. (1987), 13 = Kreuzer (2004), 14 = Bain et al. (1998), 15 = McQueen and Perkins (1995),
Ho et al. (1995), 16 = Qui et al. (2002), Fan et al. (2003), 17 = Johnston (1940), Bhlke and Kistler (1986), 18 = Schreiber et al. (1990a,b), Haeberlin
(2002), Haeberlin et al. (2004), 19 = Thompson et al. (1999), Lang et al. (2000), Thompson and Newberry (2000), Lang and Baker (2001), Mustard (2001),
20 = Bierlein and Crowe (2000), Goldfarb et al. (2001), Groves et al. (2003), Jia et al. (2001)

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

1598

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

1599

(Cont.)
Grass Valley, California17

Parcoy-Pataz, Peru18

Intrusion-related gold deposits19

Phanerozoic orogenic gold deposits20

Sericitization, propylitization,
chloritization, silicification
Well developed

Sericitization (chloritization,
silicification, carbonatization)
Well developed

Restricted, commonly weak


(also described as cryptic)
Not/poorly developed

Carbonitization, (de)silicification,
sericitization, albitization, sulfidation
Well developed

Al, C, Ca, Fe, K, Mg, Na, Si

As, C, Ca, Fe, K, Mg, Na, S

Not specified

As, Au, C, K, Na, S, Si,

Mother Lode district: CO2-rich


fluids, Tt = 300 to 350C

Of low (5-15 wt% NaCl equiv)


to high (20-38wt% NaCl equiv)
salinity, Tt = 150 to 330C

Low salinity (<12 wt% NaCl equiv)


H2O-CO2 CH4 fluids, moderate
to high (10-65 wt% NaCl equiv)
salinity inclusions are present in
some deposits, Tt = 250 to 350C

Low salinity (<10 wt% NaCl equiv),


near neutral H2O-CO2 CH4 fluids,
Tt = 300 to 400C

18O = 11.4, D = -39

18O = 10.9 to 14.2,


D = -60 to -39,
13C = -6.5 to -3.4,
34S = -2.1 to 3.7

18O = 5 to 10,
D = -90 to -40,
13C = -3 to 7 and -22 to 0,
34S = -5 to 5

18O = 12 to 19, D = -80 to -20,


13C = -10 to 0,
34S = -7.4 to 8.1,
15N = 2.84 to 4.49

Not specified

Fluid mixing

Various mechanisms

Various mechanisms

Deep-seated fluids of unknown


origin, no obvious genetic link to
exposed intrusions

Magmatic or metamorphic fluids,


possible meteoric fluid input

Magmatic fluids

>3 km

0.5 to 2.6 kbars, 7 km

<1 to 7 km, moderate fluid flux

Shortening
District-scale: faults, fractures,
preexisting discontinuities (e.g.,
granitoid-basement contact,
deposit-scale: splay faults

Shortening
Regional-scale: major fault zones,
district-scale: major fault zones,
subsidiary faults, preexisting
discontinuities, deposit-scale:
fault jogs/bends, splay faults,
vein/fault intersections

Not specified
Deposit-scale: fault/fracture
networks and specific igneous
textural facies within the roof
zones of host plutons

Fluids derived from crustal-scale


reservoirs and sourced from
metamorphic devolatilization
reactions at depth, modified
magmatic and meteoric sources
cannot be ruled out
0.75 to 3 kbarS, 2 to 6 km, low
mean stress, high fluid flux
Shortening
Second-order faults close to
regional-scale faults (commonly
terrane boundaries), deposits mainly
hosted in steeply dipping reverse
faults

Granitoid emplacement, uplift of


thickened crust, reactivation of
preexisting structures, spatial link
with Cordilleran-type continental
margin and post-Nevadan
subduction

Granitoid emplacement, uplift of


thickened crust, reactivation of
preexisting structures, link with
Cordilleran-type continental
margin and plate subduction

Regional extension, granitoid


emplacement

Collisional and subduction-related


regimes, regional crustal heating
induced by metamorphism or
granitoid emplacement

Convergent margin?

Convergent margin

Well inboard of inferred or


recognized collisional plate
boundaries, typically sited within
old, cratonic continental crust

Mesothermal, mother lode-type

Intrusion-related, orogenic

Not applicable

Situated at inferred collisional plate


boundaries, typically fore-arc and
continental magmatic arc settings,
less commonly sited between the
magmatic arc and backarc
Not applicable

Abbreviations: ank = ankerite, apy = arsenopyrite, As = native arsenic, Au = native gold, bar = barite, Bi = native bismuth, Bi-sfs = Bismuth sulfosalts,
cal = calcite, chl = chlorite, cpy = chalcopyrite, dol = dolomite, el = electrum, ep = epidote, fs = feldspar group minerals, fuch = fuchsite, gn = galena, hm
= hematite, K-fs = potassium-feldspar, mo = molybdenite, mm = montmorillonite, mt = magnetite, plag = plagioclase, po = pyrrhotite, py = pyrite, qtz =
quartz, rt = rutile, sch = scheelite, sd = siderite, ser = hydrothermal muscovite (sericite group minerals), sp = sphalerite, td-tn = tetrahedrite-tennantite,
tell = telluride group minerals, to = tourmaline, Tt = trapping temperature, wf = wolframite

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

1599

1600

OLIVER P. KREUZER

Fig. 14. Conceptual diagram showing the relationship between compositions of granitoids from eastern Australia (field shown encloses approximately
3,500 analyses), oxidation and fractionation state, and main metal assemblages of related mineralization from Blevin et al. (1996). The diagram illustrates that mineralization associated with the mainly oxidized, unfractionated
Ravenswood batholith melts should be dominated by Cu-Au or W. Wholerock geochemistry data from Hutton and Crouch (1993) and the OZCHEM
database (http://www.ga.gov.au/oracle/#geochem).

illustrates that the unfractionated and dominantly oxidized intrusions that host the veins are more similar to intrusions associated with copper-gold and tungsten deposits rather than
gold-only deposits. Veins of the Charters Towers goldfield
are significantly enriched in tellurium (100 to >1,000 )
compared to average granitoid values (G.W. Morrison, writ.
commun., 2003), suggesting a magmatic contribution to the
ore-forming fluids (e.g., Cooke and McPhail, 2001). A magmatic contribution also is consistent with the results of published isotope studies, although the Pb isotope data rule out
lead derivation from any exposed granitoids of the
Ravenswood batholith (Black et al., 1997). Sulfur isotope
data are consistent with either mantle or metamorphic origins of the ore-forming fluids (Peters and Golding, 1989).
Nitrogen contents and 15N values of hydrothermal muscovite from zones of gold-related wall-rock alteration overlap
with those measured from orogenic lode gold deposits (cf. Jia
and Kerrich, 2000), indicating either a metamorphic origin
for the fluid or equilibration with metamorphic rocks. However, the lack of CO2-rich inclusions suggests that the fluids
that formed the veins of the Charters Towers goldfield are
different from those of most orogenic gold deposits (cf. Ridley and Diamond, 2000; Table 4). Moreover, the new fluid
inclusion data suggest trapping of aqueous saline fluids (~23
wt % NaCl equiv) as well as a cooler, more dilute fluid at the
time of gold deposition (Figs. 810, Table 2). The origin and
time of trapping of saline fluid inclusions in lode gold deposits is often ambiguous (e.g., McCuaig and Kerrich, 1998;
Ridley and Diamond, 2000), although significant salinity
0361-0128/98/000/000-00 $6.00

variations in hydrothermal systems are commonly attributed


to fluid mixing (e.g., Parry, 1998; Wilkinson, 2001). Mixing
between hotter saline fluids of potential deep-seated magmatic origin and cooler, more dilute solutions of potential
meteoric origin has been reported from lode gold deposits of
the Peruvian Parcoy-Pataz (Schreiber et al., 1990a, b; Haeberlin, 2002) and Indian Kolar districts (Mishra and Panigrahi, 1999). In this model, Late Silurian to Early Devonian
uplift of the thickened crust and denudation at the surface of
the Lolworth-Ravenswood terrane could have been coincident with metamorphic devolatilization (e.g., Kerrich and
Wyman, 1990; Phillips and Powell, 1993) at deeper crustal
levels during peak metamorphism (D1-D2) ~30 m.y. later
than the uppermost crust (Stwe, 1998). Release of metamorphic fluid into the uppermost crust may have been prevented by the low permeability and porosity of the thick mass
of freshly accumulated intrusions of the Ravenswood
batholith. Cyclic fracturing during an episode of regional
shortening (D4) would then have focused fluid flow into dilatant structures (cf. Cox et al. 2001; Sibson 2001). Cooling
and reaction of the metamorphic fluids with the granitic wall
rocks could have triggered wall-rock alteration and quartz,
pyrite, and arsenopyrite deposition of the paragenetic stages
I and II (Fig. 2). This model would account for the metamorphic 15N values measured from hydrothermal muscovite and the origin of type 1 and 2a fluid inclusions. The
lack of CO2-rich inclusions may be linked to CO2 fixation reactions through fluid-wall rock interaction or CO2 loss due to
cyclical fluid pressure fluctuations (e.g., Parry, 1998; Mishra
and Panigrahi, 1999). Saline fluids may have been released
from deep-seated zones of crustal melting, possibly associated with a new pulse of magmatic activity and the emplacement of Late Silurian to Early Devonian intrusions, such as
the Deane, Carse-O-Gowrie, Chippendale, and Broughton
River granodiorites (Figs. 1A, 13). This addition of heat into
the crust also may have promoted regional-scale hydrothermal circulation. Such deep-seated magmatic fluids, focused
into structures previously affected by stages I and II, are possible sources of gold, and mineralization may have been
caused by mixing with cooler, more dilute fluids circulating
in the upper crust. Mishra and Panigrahi (1999) and Haeberlin (2002) have previously suggested the influx of significantly modified, heated meteoric waters in orogenic lode
gold systems, and this could also have occurred during rapid
uplift of the Ravenswood batholith.
Veins of the Charters Towers goldfield have many features
in common with intrusion-related gold deposits (e.g., spatial
and temporal association with magmatism, types of host intrusions, depositional environment, fluid parameters, and
quartz textures; Ho et al., 1995; McQueen and Perkins, 1995;
Lang et al., 2000; Thompson and Newberry, 2000). However,
a detailed comparison (Table 4) illustrates that the veins have
more in common with orogenic gold deposits (e.g., Bierlein
and Crowe, 2000; Goldfarb et al., 2001; Groves et al., 2003),
and particularly a group of granitoid-hosted vein systems
commonly classified as orogenic (e.g., the lode gold deposits
of the Etheridge, Parcoy-Pataz, Grass Valley, and Jiaodong
districts). Thus, further exploration in the Charters Towers
goldfield should target features common to this group of
granitoid-hosted deposits.

1600

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES

Conclusions
The results of limited new age dating and fluid inclusion and
N isotope analyses have been compared from the main mining
area at Charters Towers and from the Hadleigh Castle mine,
approximately 40 km to the east. The new geochronological
data from Hadleigh Castle are indistinguishable within error
from those obtained at Charters Towers, suggesting broadly
synchronous auriferous vein formation between 410 and 404
Ma within a significant segment of the Ravenswood batholith
and part of a much larger system of energy and mass flux (cf.
Wyborn et al., 1994; Hronsky, 2004). Veins of the Charters
Towers goldfield contain three different fluid inclusion types
that are distinguishable by petrography, microthermometry,
and, as previously reported by Peters and Golding (1989), by
laser Raman spectroscopy. This study is the first to report
saline type 2b inclusions in sphalerite (18.928.3 wt % NaCl
equiv) and quartz (20.924.7 wt % NaCl equiv) having been
trapped during the paragenetic stage of gold deposition. The
sphalerite fluid inclusion data suggest that mixing between
hot, saline waters and cooler, more dilute fluids may have been
the mechanism of gold deposition. Although few samples were
analyzed, estimated fluid trapping pressures in combination
with geological and structural constraints indicate vein formation at pressures ranging from 0.9 to 3.8 kbars and depths between ~3.5 and 12 km.
Most samples had very low nitrogen concentrations, but the
nitrogen isotope data may be evidence of the reaction of wall
rocks of the veins with metamorphic fluids or fluids that were
in equilibrium with metamorphic rocks. This process may
have caused sericite formation, quartz deposition, and pyrite
and arsenopyrite precipitation of the paragenetic stages I and
II. These results are considered consistent with a genetic
model in which gold is sourced from deep-seated saline fluids
that mix with cooler, more diluted fluids higher in the crust.
In this model, the local intrusions are mainly passive hosts for
the gold deposits.
Acknowledgments
This study was carried out at and funded by James Cook
University (International Postgraduate Research and School
of Earth Sciences scholarships). Funding of sample preparation and analytical work by James Cook University (DMR
Scheme) and the Society of Economic Geologists (McKinstry
Award) is greatly acknowledged. Jo Wartho (Curtin University) and Myles Stocki (University of Saskatchewan) are
thanked for their assistance with Ar-Ar and N isotope analyses, respectively. Simon Dominy and Matthew Raine (James
Cook University) are thanked for supplying trace element
data from the B lode, Hadleigh Castle mine. Jim Morrison
and Nigel Storey (Citigold Corporation Ltd.), Ian Hodkinson,
Harry Mustard, and Stuart Hampton (SMC Gold Ltd.), and
Tony Alston (Glengarry Resources Ltd.) are thanked for their
support. Gregg Morrison (Klondike Exploration Services) is
thanked for discussions and supplying invaluable unpublished
data. Tim Baker (James Cook University), Mike Etheridge
(Macquarie University), and Stephen Peters (USGS), Economic Geology reviewers Frank Bierlein (Monash University), David Cooke (University of Tasmania), and Terry Mernagh (Geoscience Australia), and the editor of Economic
0361-0128/98/000/000-00 $6.00

1601

Geology, Mark Hannington (University of Ottawa) are


thanked for having helped to significantly improve the quality
of the original manuscript.
November 13, 2003; December 2, 2005
REFERENCES
Bain, J.H.C., Withnall, I.W., Black, L.P., Etminan, H., Golding, S.D., and
Sun, S.-S., 1998, Towards an understanding of the age and origin of gold
mineralization in the Etheridge goldfield, Georgetown region, north
Queensland: Australian Journal of Earth Sciences, v. 45, p. 247263.
Baker, T., 2002, Emplacement depth and carbon dioxide-rich fluid inclusions
in intrusion-related gold deposits: ECONOMIC GEOLOGY, v. 97, p.
11111117.
Berry, R.F., Huston, D.L., Stolz, A.J., Hill, A.P., Beams, S.D., Kuronen, U.,
and Taube, A., 1992, Stratigraphy, structure, and volcanic-hosted mineraliZation of the Mount Windsor subprovince, north Queensland, Australia:
ECONOMIC GEOLOGY, v. 87, p. 739763.
Bierlein, F.P., and Crowe, D.E., 2000, Phanerozoic orogenic lode gold: Reviews in Economic Geology, v. 13, p. 103139.
Black, L.P., Carr, G.R., and Sun, S-S., 1997, Applied isotope geochronology
and geochemistry: Australian Geological Survey Organization Bulletin 240,
and Queensland Department of Mines and Energy Queensland Geology 9,
p. 429447.
Blatchford, A., 1953, Charters Towers goldfield: Empire Mining and Metallurgical Congress, 5th, Australasian Institute of Mining and Metallurgy,
Publications, v. 1, p. 796806.
Blevin, P.L., Chappell, B.W., and Allen, C.M., 1996, Intrusive metallogenic
provinces in eastern Australia based on granite source and composition:
Transactions of the Royal Society of Edinburgh, v. 87, p. 281290.
Bodnar, R.J., 1993, Revised equation and table for determining the freezing
point depression of H2O-NaCl solutions: Geochimica et Cosmochimica
Acta, v. 57, p. 683684.
Bhlke, J.K., and Kistler, R.W., 1986, Rb-Sr, K-Ar, and stable isotope evidence for the ages and sources of fluid components of gold-bearing quartz
veins in the northern Sierra Nevada Foothills metamorphic belt, California:
ECONOMIC GEOLOGY, v. 81, p. 296322.
Bons, P.D., 2001, The formation of large quartz veins by rapid ascent of fluids in mobile hydrofractures: Tectonophysics, v. 336, p. 117.
Brown, P.E., 1989, FLINCOR: A microcomputer program for the reduction
and investigation of fluid inclusion data: American Mineralogist, v. 74, p.
13901393.
1998, Fluid inclusion modeling for hydrothermal systems: Reviews in
Economic Geology, v. 10, p. 151171.
Brown, P.E., and Lamb, W.M., 1989. P-V-T properties of fluids in the system
H2O-CO2NaCl: New graphical presentations and implications for fluid inclusion studies: Geochimica et Cosmochimica Acta, v. 53, p. 12091221.
Campbell, A.R., and Larson, P.B., 1998, Introduction to stable isotope applications in hydrothermal systems: Reviews in Economic Geology, v. 10, p.
173193.
Cooke, D.R., and McPhail, D.C., 2001, Epithermal Au-Ag-Te mineralization, Acupan, Baguio district, Philippines: Numerical simulations of mineral deposition: ECONOMIC GEOLOGY, v. 96, p. 109131.
Cox, S.F., Knackstedt, M.A., and Braun, J., 2001, Principles of structural control on permeability and fluid flow in hydrothermal systems: Reviews in
Economic Geology, v. 14, p. 124.
Dowling, K., and Morrison, G.W., 1989, Application of quartz textures to the
classification of gold deposits using North Queensland: ECONOMIC GEOLOGY MONOGRAPH 6, p. 342355.
Fan, H.R., Zhai, M.G., Xie, Y.H., and Yang, J.H., 2003, Ore-forming fluids associated with granite-hosted gold mineralization at the Sanshandao deposit,
Jiaodong gold province, China: Mineralium Deposita, v. 38, p. 739750
Faure, G., 1986, Principles of isotope geology, 2nd ed.: New York, Wiley and
Sons, 589 p.
Gaboury, D., and Daigneault, R., 2000, Flat vein formation in a transitional
crustal setting by self-induced fluid pressure equilibriuman example
from the Gant Dormant gold mine, Canada: Ore Geology Reviews, v. 17,
p. 155178.
Glasmacher, U.A., Zentilli, M., and Ryan, R., 2003, Nitrogen distribution in
Lower Palaeozoic slates/phyllites of the Meguma Supergroup, Nova Scotia,
Canada: Implications for Au and Zn-Pb mineralisation and exploration:
Chemical Geology, v. 194, p. 297329.

1601

1602

OLIVER P. KREUZER

Goldfarb, R.J., Groves, D.I., and Gardoll, S., 2001, Orogenic gold and geologic time: A global synthesis: Ore Geology Reviews, v. 18, p. 175.
Golding, S.D, and Wilson, A.F., 1981, An oxygen and carbon isotopic study
of some gold deposits of eastern Australia: Proceedings of the Australasian
Institute of Mining and Metallurgy, v. 278, p.1321.
Golding, S.D., Wilson, A.F., Scott, M., Anderson, P.K., Waring, C.L., Flitcroft, M., and Rypkema, H.A., 1987, Isotopic evidence for the diverse origins of gold mineralization in Queensland: Department of Earth Sciences,
University of Queensland, Papers, v. 12, p. 6585.
Groves, D.I., Goldfarb, R.J., Gebre-Mariam, M., Hagemann, S.G., and
Robert, F., 1998, Orogenic gold deposits: A proposed classification in the
context of their crustal distribution and relationship to other gold deposit
types: Ore Geology Reviews, v. 13, p. 727.
Groves, D.I., Goldfarb, R.J., Robert, F., and Hart, C.J.R., 2003, Gold deposits in metamorphic belts: Overview of current understanding, outstanding problems, future research, and exploration significance: ECONOMIC GEOLOGY, v. 98, p. 129.
Haeberlin, Y., 2002, Geological and structural setting, age, and geochemistry
of the orogenic gold deposits of the Pataz province, eastern Andean
Cordillera, Peru: Unpublished Ph.D. thesis, University of Geneva, 182 p.
Haeberlin, Y., Moritz, R., Fontbote, L., and Cosca, M., 2004, Carboniferous
orogenic gold deposits at Pataz, eastern Andean Cordillera, Peru: Geological and structural framework, paragenesis, alteration, and 40Ar/39Ar
geochronology: ECONOMIC GEOLOGY, v. 99, p. 73112.
Harley, M., and Charlesworth, E.G., 1996, The role of fluid pressure in the
formation of bedding-parallel, thrust-hosted gold deposits, SabiePilgrims Rest goldfield, eastern Transvaal: Precambrian Research, v. 79,
p. 125140.
Hartley, J.S., 1996, Mineral occurrencesRavenswood 1:100,000 sheet area:
Queensland Geological Record, Queensland Department of Minerals and
Energy, Brisbane, v. 1996/02, 55 p.
Hartley, J.S., and Dash, P.H., 1993, Mineral occurrencesCharters Towers
1:100,000 sheet area: Queensland Geological Record, Queensland Department of Minerals and Energy, Brisbane, v. 1993/06, 53 p.
Henderson, R.A., 1987, An oblique subduction and transform faulting model
for the evolution of the Broken River province, northern Tasman orogenic
zone: Australian Journal of Earth Sciences, v. 34, p. 237249.
Ho, S.E., McQueen, K.G., McNaughton, N.J., and Groves, D.I., 1995, Lead
isotope systematics and pyrite trace element geochemistry of two granitoidassociated mesothermal gold deposits in the southeastern Lachlan fold
belt: ECONOMIC GEOLOGY, v. 90, p. 18181830.
Hodges, K.V., 1998, 40Ar/39Ar geochronology using the laser microprobe: Reviews in Economic Geology, v. 7, p. 163220.
Hodgson, C.J., 1989, The structure of shear-related, vein-type gold deposits:
A review: Ore Geology Reviews, v. 4, p. 231273.
Hofstra, A.H., and Cline, J.S., 2000, Characteristics and models for Carlintype gold deposits: Reviews in Economic Geology, v. 13, p. 163220.
Hronsky, J.M.A., 2004, The science of exploration targeting, in Muhling, J.,
Goldfarb, R., Vielreicher, N., Bierlein, F., Stumpfl, E., Groves, D., and
Kenworthy, S., eds., SEG 2004predictive mineral discovery under cover
[abs.]: Centre for Global Metallogeny, University of Western Australia Publication 33, p. 129133.
Hutton, L.J., and Crouch, S.B.S., 1993, Geochemistry and petrology of the
western Ravenswood batholith: Brisbane, Queensland Department of Minerals and Energy, Queensland Geological Record 22, 73 p.
Hutton, L.J., and Rienks, I.P., 1997, Geology of the Ravenswood batholith:
Brisbane, Queensland Department of Mines and Energy, Queensland Geology 8, 60 p.
Hutton, L.J., Draper, J.J., Rienks, I.P., Withnall, I.W., and Knutson, J., 1997,
Charters Towers region: Australian Geological Survey Organization Bulletin 240, and Queensland Department of Mines and Energy Queensland
Geology 9, p. 165224.
Jbrak, M., 1997, Hydrothermal breccias in vein-type ore deposits: A review
of mechanisms, morphology and size distribution: Ore Geology Reviews, v.
12, p. 111134.
Jia, Y., and Kerrich, R., 1999, Nitrogen isotope systematics of mesothermal
lode gold deposits: Metamorphic, granitic, meteoric water, or mantle origin: Geology, v. 27, p. 10511054.
2000, Giant quartz vein systems in accretionary orogenic belts: the evidence for a metamorphic fluid origin from 15N and 13C studies: Earth
and Planetary Science Letters, v. 184, p. 211224.
Jia, Y., Li, X., and Kerrich, R., 2001, Stable isotope (O, H, S, C, and N) systematics of quartz veins in the turbidite-hosted Central and North Debora
0361-0128/98/000/000-00 $6.00

gold deposits of the Bendigo gold field, Central Victoria, Australia: Constraints on the origin of the ore-forming fluids: ECONOMIC GEOLOGY, v. 96,
p. 705721.
Jia, Y., Kerrich, R., and Goldfarb, R.J., 2003, Metamorphic origin of oreforming fluids for orogenic gold-bearing quartz vein systems in the North
American Cordillera: Constraints from a reconnaissance study of 15N, D,
and 18O: ECONOMIC GEOLOGY, v. 98, p. 109123.
Johnston, W.D., 1940, The gold quartz veins of the Grass Valley, California:
U.S. Geological Survey Professional Paper, v. 194, 101 p.
Kerrich, R., and Wyman, D., 1990, Geodynamic setting of mesothermal gold
deposits: An association with accretionary tectonic regimes: Geology, v. 18,
p. 882885.
Kreuzer, O.P., 2003, Structure, timing and genesis of auriferous quartz veins
in the Charters Towers goldfield, north Queensland: Implications for exploration and prospectivity: Unpublished Ph.D. thesis, Queensland, Australia, James Cook University, 310 p.
2004, How to resolve the controls on mesothermal vein systems in a
goldfield characterized by sparse kinematic information and fault reactivationa structural and graphical approach: Journal of Structural Geology, v.
26, p. 10431065.
2006, Textures, paragenesis and wall-rock alteration of lode-gold deposits in the Charters Towers district, North Queensland: Implications for
the conditions of ore formation: Mineralium Deposita, v. 40, p. 639663
(DOI 10.1007/s00126-005-0010-1)
Kreuzer, O.P., and Alston, A.J., 2004, What are the chances of discovery in
the over-explored Charters Towers region of north Queensland? [abs.]:
Australian Institute of Geoscientists Bulletin 40, p. 1317.
Lang, J.R., and Baker, T., 2001, Intrusion-related gold systems: The present
level of understanding: Mineralium Deposita, v. 36, p. 477489.
Lang, J.R., Baker, T., Hart, C.J.R., and Mortensen, J.K., 2000, Intrusion-related gold systems: Society of Economic Geologists Newsletter 40, p. 1 and
615.
Levingston, K.R., 1972, Ore deposits and mines of the Charters Towers
1:250,000 sheet area, north Queensland: Brisbane, Queensland Department of Mines and Energy Report 57, 103 p.
Marks, E.O., 1913, Outside mines of the Charters Towers goldfield: Brisbane, Queensland Geological Survey, Department of Mines Publication
238, 21 p.
McCuaig, T.C., and Kerrich, R., 1998, P-T-t-deformation-fluid characteristics of lode gold deposits: Evidence from alteration systematics: Ore Geology Reviews, v. 12, p. 381453.
McQueen, K.G., and Perkins, C., 1995, The nature and origin of a granitoidrelated gold deposit at Dargues reef, Majors Creek, New South Wales:
ECONOMIC GEOLOGY, v. 90, p. 16461662.
Mishra, B., and Panigrahi, M.K., 1999, Fluid evolution in the Kolar gold
field: Evidence from fluid inclusion studies: Mineralium Deposita, v. 34, p.
173181.
Morrison, G.W., 1988, Paleozoic gold deposits of northeast Queensland:
Contributions of the Economic Geology Research Unit, v. 29, and Bicentennial Gold 88 Excursion Guide, v. 8, James Cook University, p. 1122.
Mustard, R., 2001, Granite-hosted gold mineralization at Timbarra, northern
New South Wales, Australia: Mineralium Deposita, v. 36, p. 542562.
Parry, W.T., 1998, Fault-fluid compositions from fluid-inclusion observations
and solubilities of fracture-sealing minerals: Tectonophysics, v. 290, p.
126.
Perkins, C., and Kennedy, A.K., 1998, Permo-Carboniferous gold epoch of
northeast Queensland: Australian Journal of Earth Sciences, v. 45, p.
185200.
Peters, S.G., 1987, Geology, fluid characteristics, lode controls and oreshoot
growth in mesothermal gold-quartz veins, northeastern Queensland: Unpublished Ph.D. thesis, Queensland, Australia, James Cook University, 277
p.
1990, Lode controls of the Charters Towers goldfield, northeastern
Queensland: Proceedings of the Australasian Institute of Mining and Metallurgy, v. 2, p. 5160.
1993, Formation of oreshoots in mesothermal gold-quartz vein deposits:
Examples from Queensland: Ore Geology Reviews, v. 8, p. 277301.
Peters, S.G., and Golding, S.D., 1989, Geologic, fluid inclusion, and stable
isotope studies of granitoid-hosted gold-bearing quartz veins, Charters
Towers, northeastern Australia: ECONOMIC GEOLOGY MONOGRAPH 6, p.
260273.
Phillips, G.N., and Powell, R., 1993, Link between gold provinces: ECONOMIC GEOLOGY, v. 88, p. 10841098.

1602

CHARTERS TOWERS GOLDFIELD, NORTH QUEENSLAND, FLUID INCLUSIONS AND STABLE ISOTOPES
Qiu, Y, Groves, D.I., McNaughton, N.J., Wang, L., and Zhou, T., 2002, Nature, age and tectonic setting of granitoid-hosted, orogenic gold deposits in
the Jiaodong peninsula, eastern North China craton, China: Mineralium
Deposita, v. 37, p. 283305.
Reid, J.H., 1917, The Charters Towers goldfield: Brisbane, Geological Survey of Queensland Publication 256, 232 p.
Richards, D.N.G., 1980, Palaeozoic granitoids of northeastern Australia: Geological Society of Australia, Queensland Division, Australian Geological
Convention, 3rd, Townsville, Proceedings, p. 229246.
Ridley, J.R., and Diamond, L.W., 2000, Fluid chemistry of orogenic lode gold
deposits and implications for genetic models: Reviews in Economic Geology, v. 13, p. 141162.
Roedder, E., 1984, Fluid inclusions: Reviews in Mineralogy, v. 12, 644 p.
Scheibner, E., and Veevers, J.J., 2000, Tasman fold belt system, in Veevers,
J.J., ed., Billion-year earth history of Australia and neighbors in Gondwanaland: North Ryde, NSW, Macquarie University, National Key Centre
for the Geochemical Evolution and Metallogeny of Continents (GEMOC)
Press, p. 154233.
Schreiber, D.W., Amstutz, G.C., and Fontbote, L., 1990a, The formation of
auriferous quartz sulfide veins in the Pataz region, northern Peru: A synthesis of geological, mineralogical, and geochemical data: Mineralium Deposita, v. 25 (Supplement), p. S136S140.
Schreiber, D.W., Fontbote, L., and Lochmann, D., 1990b, Geologic setting,
paragenesis and physiochemistry of gold quartz veins hosted by plutonic
rocks in the Pataz region: ECONOMIC GEOLOGY, v. 85, p. 13281347.
Shepherd, T.J., Rankin, A.H., and Alderton, D.M.H, 1985, A practical guide
to fluid inclusion studies: Glasgow, Blackie and Son, 239 p.
Sibson, R.H., 1996, Structural permeability of fluid-driven fault-fracture
meshes: Journal of Structural Geology, v. 18, p. 10311042.
2001, Seismogenic framework for hydrothermal transport and ore deposition: Reviews in Economic Geology, v. 14, p. 2550.
Sillitoe, R.H., 1997, Gold deposits and intrusive rocks, in Papunen, H., ed.,
Mineral deposits: Research and exploration, where do they meet? [abs.]:
Rotterdam, Balkema, p. 2326.
Sillitoe, R.H., and Thompson, F.H., 1998, Intrusion-related gold deposits:
Types, tetono-magmatic settings and difficulties of distinction from orogenic gold deposits: Resource Geology, v. 48, p. 237250.
Smith, M., Thompson, J.F.H., Bressler, J., Layer, P., Mortensen, J.K., and
Hidetoshi Takaoka, I.A., 1999, Geology of the Liese zone, Pogo property,
east-central Canada: Society of Economic Geologists Newsletter 38, p. 1
and 1221.

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

1603

Solomon, M., and Groves, D.I., 2000, The geology and origin of Australias
mineral deposits (reprinted with additional material): Centre for Ore Deposits Research, University of Tasmania, and Centre for Global Metallogeny, University of Western Australia Publication 32, 1002 p.
Stwe, K., 1998, Tectonic constraints on the timing and relationships of
metamorphism, fluid production and gold-bearing quartz vein emplacement: Ore Geology Reviews, v. 13, p. 219228.
Tenison-Woods, K., and Rienks, I.P., 1992, New insights into the structure
and subdivision of the Ravenswood batholitha geophysical perspective:
Exploration Geophysics v. 23, p. 353360.
Thompson, J.F.H., and Newberry, R.J., 2000, Gold deposits related to reduced granitic intrusions: Reviews in Economic Geology, v. 13, p. 377400.
Thompson, J.F.H., Sillitoe, R.H., Baker, T., Lang, J.R., and Mortensen, J.K.,
1999, Intrusion-related gold deposits associated with tungsten-tin
provinces: Mineralium Deposita, v. 34, p. 323334.
Twiss, R.J., and Moores, E.M., 1992, Structural geology: New York, Freeman, 532 p.
Wilkinson, J.J., 2001, Fluid inclusions in hydrothermal ore deposits: Lithos,
v. 55, p. 229272.
Willner, A.P., Wartho, J-A., Kramm, U., and Puchkov, V.N., 2004, Laser
40Ar/39Ar ages of single detrital white mica grains related to the exhumation
of Neoproterozoic and Late Devonian high pressure rocks in the Southern
Urals (Russia): Geological Magazine, v. 141, p. 161172.
Withnall, I.W., Draper, J.J., Mackenzie, D.E., Knutson, J., Blewett, R.S.,
Hutton, L.J., Bultitude, R.J., Wellman, P., McConachie, B.A., Bain, J.H.C.,
Donchak, P.J.T., Lang, S.C., Domagala, J., Symonds, P.A., and Rienks, I.P.,
1997, Review of geological provinces and basins of North Queensland: Australian Geological Survey Organization Bulletin 240, and, Queensland Department of Mines and Energy Queensland Geology 9, p. 449528.
Wyborn, L.A.I., Wyborn, D., Warren, R.G., and Drummond, B.J., 1992, Proterozoic granite types in Australia: Implications for lower crust composition, structure and evolution: Transactions of the Royal Society of Edinburgh, Earth Sciences, v. 83, p. 201209.
Wyborn, L.A.I., Heinrich, C.A., and Jaques, A.L., 1994, Australian Proterozoic mineral systems: Essential ingredients and mappable criteria: Australian Institute of Mining and Metallurgy Publication Series, v. 5, p.
109115.
Yardley, B.W.D., 1989, An introduction to metamorphic petrology: Harlow,
Longman, 248 p.

1603

S-ar putea să vă placă și