Sie sind auf Seite 1von 16

Mineralogy and Petrology (2006) 87: 171186

DOI 10.1007/s00710-006-0128-6

The goldvanadiumtellurium association


at the Tuvatu goldsilver prospect, Fiji:
conditions of ore deposition
P. G. Spry and N. L. Scherbarth
Department of Geological and Atmospheric Sciences, Iowa State University,
Ames, Iowa, USA
Received October 23, 2005; revised version accepted February 3, 2006
Published online June 6, 2006; # Springer-Verlag 2006
Editorial handling: N. Cook
Summary
The Tuvatu goldtelluride prospect is one of several epithermal gold systems along the
>250 km northeast trending Viti Levu lineament, Fiji, which are genetically associated
with alkalic magmatism. Vein structures contain a variety of sulfides, native elements,
sulfosalts, and tellurides. Calaverite is intimately associated with various vanadiumbearing minerals: roscoelite, karelianite, vanadian muscovite, Ti-free nolanite, vanadian
rutile, schreyerite, and an unnamed vanadium silicate. Thermodynamic calculations
for the systems VAlKSiOH (Cameron, 1998) and AuTeClSOH at estimated conditions of formation of the telluride-native gold stage at Tuvatu (250  C,
Au 1 ppb, Te 1 ppb, S 0.001 m, V 0.0001 m, and aK 0.01), show that
the stability fields of calaverite, roscoelite, and karelianite converge in pH-f O2 space
near the hematitemagnetite buffer and at neutral to slightly acid conditions. Thermodynamic and textural data suggest that these minerals were deposited together at Tuvatu
and likely explain the common coexistence of roscoelite and calaverite in epithermal
gold systems elsewhere. The presence of magnetite with up to 0.7 wt.% V2O3 in the
Navilawa Monzonite is consistent with the derivation of V from the alkalic intrusive
rocks, which are also considered to be the source of Au and Te in the Tuvatu deposit.

Introduction
The intimate spatial relationship between vanadium minerals and gold, particularly
gold-bearing tellurides, has long been known since the identification of roscoelite
[K(V3,Al,Mg)2AlSi3O10(OH)2] by Blake (1876) in the Stuckslager gold deposit,
Coloma district, California. Roscoelite and vanadium muscovite, are characterized
by the 1 M and 2 M structural types, respectively (Heinrich and Levinson, 1955) and

172

P. G. Spry and N. L. Scherbarth

by roscoelite exhibiting >17 wt.% V2O3. They are the two most common V-bearing
minerals in gold telluride deposits. Roscoelite is relatively common in epithermal
gold telluride deposits where it has been identified in the Cripple Creek deposit,
Colorado (Jensen and Barton, 2000), Boulder County deposits, Colorado (Lindgren,
1907; Kelly and Goddard, 1969; Kurtz and Hauff, 1988; Saunders, 1991), Emperor
deposit, Fiji (Ahmad et al., 1987), Spotted Horse, Maginnis, and Gies deposits, Judith
Mountains, Montana (Forrest, 1971; Zhang and Spry, 1994; Thieben and Spry, 1995),
and Porgera deposit, Papua New Guinea (Cameron, 1998; Cameron et al., 1995). In
most of these deposits, there is direct spatial association between the easy to identify green roscoelite and gold mineralization, which is usually in the form of goldbearing tellurides, particularly calaverite, and either native gold or electrum.
Although roscoelite is rare in mesothermal gold deposits, one exception being
the Boulder Reef deposit, Golden Mile, Western Australia (Simpson, 1952), the
presence of other vanadium silicates or oxides is more common than in epithermal
gold telluride deposits. For example, the giant Golden Mile district contains vanadian muscovite, tomichite, nolanite, tivanite, vanadian hematite, vanadian magnetite,
and vanadian tourmaline intergrown with gold tellurides, in particular calaverite
(Nickel, 1977; Nickel and Grey, 1982; Gatehouse et al., 1983). This assemblage of
minerals constitute the well-known green leader gold lodes of the Golden Mile
district (Nickel, 1977). Coulsonite (FeV2O4) was also reported at the Kalgoorlie deposit by Spiridinov (1978). An even more diverse array of vanadian silicates and
oxides is present in the large Hemlo gold deposit, Ontario, which contains vanadian
muscovite, roscoelite, vanadian rutile, barian tomichite, vanadian hematite, karelianite, hemloite, vanadian titanite, vanadian carfarsite, vanadian phlogopite, vanadian allanite, vanadian pumpellyite, vanadian vesuvianite, vanadian epidote-group
minerals, and vanadian grossular (Harris, 1989; Pan and Fleet, 1991, 1992). In addition, minor amounts of V (<2.5 wt.% V2O3) are present in tourmaline, chlorite,
talc, chromite, and tremolite (Harris, 1989; Pan and Fleet, 1992). As with the
Kalgoorlie deposit, one of the best indicators to gold ore at Hemlo is the presence
of light to dark green colored vanadian muscovite (Harris, 1989).
At the epithermal Emperor and Tuvatu gold telluride deposits, Fiji, dark green
roscoelite is intimately intergrown with gold-bearing tellurides (Ahmad et al.,
1987; Scherbarth and Spry, 2001; Pals and Spry, 2003). Here we report the results
of a systematic mineralogical, electron-microprobe, and scanning electron microscope study of vanadium silicates and oxides in the Tuvatu deposit, noting that the
presence of vanadium minerals other than roscoelite and=or vanadian muscovite
are rare in epithermal gold telluride deposits. The major objectives of the study are
to (1) Evaluate the reasons for the presence of vanadium silicates and oxides in the
Tuvatu deposit, and (2) Explain the intimate spatial relationship between calaverite
(AuTe2) and vanadium minerals in the deposit. The latter objective is pertinent
to understanding why the AuVTe association is so common in epithermal and
mesothermal gold telluride ores, in general.
Geological setting
The Tuvatu goldsilver telluride deposit contains reserves of 480,000 ozs of gold
and, in Fiji, is second in size to the 11.5 Moz Emperor gold telluride deposit.

The goldvanadiumtellurium association

173

They are two of several epithermal gold deposits genetically associated with
alkalic igneous rocks that are localized along the >250 km northeast trending Viti
Levu lineament, Fiji. These alkalic igneous rocks formed during the breakup of
the Vitiaz island arc in the south-east Pacific Ocean. The Tuvatu deposit is located
in the Sabeto Valley, approximately 15 km northeast of Nadi. The basal unit of
the rocks in the Tuvatu area is the 1226 Ma Nadele Breccia (andesiticbasaltic
breccias, pillow lavas, and sediments), which is a member of the Wainimala
Group (Fig. 1). The Wainimala Group is unconformably overlain by members
of the Sabeto Volcanics (interbedded andesitic volcaniclastics and flows), which
represent the basal unit of the 4.85.5 Ma Koroimavua Volcanic Group (Colley
and Flint, 1995; Hatcher, 1998). The Nadele Breccia was intruded by the 4.85 Ma
(McDougall, 1963) Navilawa Monzonite, which is interpreted to be co-magmatic
with the Sabeto Volcanics, and is composed of a micromonzonite that envelopes
coarse monzonite (Fig. 2). These monzonites, which host the Tuvatu deposit,
are cut by several basaltic-andesite dikes and are considered to be the source
of the ore-forming components including Au, Te, and V (Scherbarth and Spry,
2000, 2006).

Fig. 1. Geology of Viti Levu, Fiji (modified after Begg, 1996; Rodda, 1967). The location
of the Tuvatu prospect with respect to the Viti Levu lineament is indicated

174

P. G. Spry and N. L. Scherbarth

Fig. 2. Simplified geological map of the Tuvatu prospect

Mineralogy and conditions of ore-formation


Gold mineralization in the Tuvatu deposit is generally hosted in sub-vertical, north
south trending and north northeastsouth southwest trending veins as well as shallow south dipping veins, and appears to be intimately related to the emplacement
of the Navilawa Monzonite (A-Izzedin, 1998; Scherbarth and Spry, 2000). Hatcher
(1998) suggested that gold was deposited in three different lode types, steepdipping veins striking northeast (e.g., Nasivi and Upper Ridges lodes), shallowly
dipping veins or flatmakes representing reactivated oblique thrust faults (e.g.,
Murau lode), and irregular brecciated bodies or shatter zones (e.g., SKL lode)
that occur at the intersection of the other two lode types (Fig. 3). Gold mineralization overprints porphyry-style copper mineralization in the Navilawa Monzonite in
the northern parts of the deposit in the H and Tuvatu lodes where it occurs in
biotitemagnetiteK feldspar dikes. However, most of the gold occurs in epithermal-style veins that cover an area of approximately 1 km by 0.25 km.
Mineralogical and paragenetic studies Ashley and Andrew (1989), Hatcher
(1998), and Scherbarth and Spry (2000) suggest a five stage paragenetic

The goldvanadiumtellurium association

175

Fig. 3. Plan view of the Tuvatu prospect lode structures

sequence (Fig. 4). A magmatic stage, which is best developed in the H and
Tuvatu lodes, is followed by three primary hydrothermal stages characterized
by different alteration assemblages (potassic, propylitic, phyllic), and a late
supergene stage (or post-mineralization stage). The most important gold-bearing
stage is the phyllic stage (stage 4) where native gold is spatially associated with
base metal sulfides (galena and sphalerite), bismuthinite (Bi2S3), and various
tellurides [calaverite (AuTe2), krennerite ((Au,Ag)Te2), sylvanite ((Au,Ag)2Te4),
utzite (Ag5-xTe3), hessite (Ag2Te), altaite (PbTe), and
petzite (Ag3AuTe2), st
coloradoite (HgTe)]. Altaite is the most common telluride in the Tuvatu deposit.
Compositions of native gold, tellurides, and sulfides are given in Scherbarth and
Spry (2006).

P. G. Spry and N. L. Scherbarth: The goldvanadiumtellurium association

177

By combining the fluid inclusion date of Ashley and Andrew (1989) and
Scherbarth (2002), it can be concluded that primary fluid inclusions in stage 1
apatite formed over a wide range of temperatures (276 to >500  C), with most
homogenizing at temperatures >450  C. These fluids boiled and were hypersaline
(>50 eq. wt.% NaCl), and were followed by stage 2 variably saline, boiling fluids
(5 to >40 eq. wt.% NaCl), which formed at ca. 310  C. Stage 3 fluids were generally less saline (1 to 10 eq. wt.% NaCl) than stage 1 and 2 fluids and formed at ca.
300  C. However, in places, stage 3 fluids were also locally boiling and very saline
(up to 37 eq. wt.% NaCl). Stage 4 non-boiling, moderately saline fluids (mean
8.4 wt.% NaCl equiv) formed between approximately 325 and 100  C (mean
257  C) and accompanied telluride and base metal sulfide deposition, as well as
the formation of vanadium-bearing minerals. The pressure correction to stage 4
fluids was likely small (<10  C) with lithostatic conditions persisting through stage
1 to 3 until hydrostatic conditions dominated during stage 4. Values of 34S for
sulfides in the porphyry and epithermal veins range from 15.3 to 3.6 and
reflect an increase in the SO4=H2S ratio of a boiling magmatic fluid (Scherbarth
and Spry 2001, Scherbarth, 2002).
Oxygen isotope compositions for water in equilibrium with stage 2 orthoclase,
magnetite, and phlogopite, and stage 3 quartz and muscovite range from 4.4 to
10.2, whereas values of 18O for water in equilibrium with stage 4 quartz range
from 7.8 to 11.5 (Ashley and Andrew, 1989; Scherbarth and Spry, 2001; Scherbarth,
2002). These values, coupled with 18O and D values of fluids in equilibrium with
phlogopite of 6.5 to 9.8 and 25.9 to 9.9, respectively, overlap with waters
from the arc magmas and subduction-related volcanic vapor boxes and are consistent
with a magmatic source and permissive of a meteoric water contribution. The Tuvatu
deposit appears to have originally developed as a porphyry copper system that was
subsequently overprinted by epithermal goldsilver telluride mineralization. The
goldsilver veins are considered by us to be the late stages of the same hydrothermal
system that involves the porphyry copper-style mineralization.
Analytical procedures
Surface and underground specimens were obtained from underground locations
and drill core from several epithermal veins in the Tuvatu deposit. Samples were
examined with a dual transmitted-reflected light Olympus BX 60 petrographic
microscope and Hitachi S-2460N and JEOL JSM-35 scanning electron microscopes possessing EDAX area mapping and back-scattered imaging capabilities.
Mineral compositions were obtained using an ARL-SEMQ electron microprobe at
Iowa State University. The instrument was operated under the following conditions: 15 kV, a sample current of 10 nA, and 2 mm beam diameter. Standards used
were natural (orthoclase for K, kyanite for Al, albite for Na and Si, hornblende for
Ca, gahnite for Zn, chromite for Cr, ilmenite for Ti, scapolite for Ca) and synthetic
minerals (knebelite for Mn, fayalite for Fe, forsterite for Mg, V2O5 for V).
1

Fig. 4. Paragenetic sequence for the mineralogy of the Tuvatu prospect (derived from
Ashley and Andrew, 1989; Hatcher, 1998; Scherbarth, 2002)

22
7.361
0.000
0.996
0.003
3.209
0.000
0.010
0.220
0.001
0.010
0.011
1.437

No. of cations
Si
Ti
Al
Cr
V
Fe
Mn
Mg
Zn
Ca
Na
K

22
6.820
0.008
1.262
0.004
3.478
0.013
0.006
0.297
0.003
0.003
0.125
1.688

47.61
0.08
7.47
0.03
30.28
0.11
0.05
1.39
0.03
0.02
0.45
9.24
96.75

22
6.700
0.000
1.319
0.018
3.715
0.043
0.000
0.139
0.006
0.014
0.008
1.630

47.30
0.00
7.90
0.16
32.71
0.36
0.00
0.66
0.06
0.09
0.03
9.02
98.29

22
6.759
0.014
2.079
0.000
2.765
0.028
0.007
0.290
0.000
0.024
0.000
1.676

47.89
0.13
12.50
0.00
24.43
0.24
0.06
1.38
0.00
0.16
0.00
9.31
96.09

16
0.010
0.000
0.091
0.019
9.972
0.843
0.001
0.009
0.000
0.003
0.000
0.000

0.07
0.00
0.51
0.16
82.17
6.66
0.01
0.04
0.00
0.02
0.00
0.00
89.64

16
0.027
0.000
0.110
0.074
10.268
0.180
0.000
0.024
0.000
0.027
0.049
0.026

0.18
0.00
0.63
0.63
86.66
1.46
0.00
0.11
0.00
0.17
0.17
0.14
90.15

2
0.007
0.000
0.000
0.006
1.979
0.002
0.000
0.003
0.003
0.000
0.000
0.000

0.30
0.00
0.00
0.31
99.89
0.08
0.00
0.08
0.17
0.01
0.00
0.00
100.84

2
0.011
0.000
0.000
0.010
1.972
0.001
0.002
0.000
0.001
0.000
0.000
0.000

0.45
0.00
0.00
0.53
98.61
0.03
0.08
0.00
0.07
0.01
0.00
0.00
99.78

2
0.002
0.952
0.002
0.000
0.057
0.001
0.000
0.000
0.000
0.003
0.001
0.001

0.13
94.10
0.11
0.00
5.24
0.10
0.00
0.00
0.00
0.24
0.04
0.03
99.99

10

2
0.000
0.979
0.001
0.000
0.020
0.006
0.000
0.000
0.001
0.001
0.000
0.002

0.02
95.78
0.07
0.00
1.86
0.56
0.00
0.00
0.13
0.09
0.00
0.11
98.62

11

9
0.029
3.006
0.022
0.000
1.890
0.030
0.000
0.000
0.000
0.004
0.000
0.000

0.45
62.40
0.30
0.00
37.19
0.55
0.01
0.00
0.00
0.09
0.00
0.00
100.99

12

12
3.301
0.000
0.172
0.009
3.375
0.004
0.000
0.007
0.000
0.003
0.039
0.060

41.49
0.00
1.84
0.14
52.90
0.06
0.00
0.06
0.00
0.04
0.26
0.59
97.38

13

1 Roscoelite (00TV-23F); 2 Roscoelite (PS99TV-69F); 3 Roscoelite (00TV-55); 4 Roscoelite (PS99TV-68F); 5 Roscoelite (99TV-7); 6 Nolanite
(99TV-68a); 7 Nolanite (PS99TV-68F); 8 Karelianite (99TV-68a); 9 Karelianite (PS99TV-68F); 10 Rutile (00TV-4); 11 Rutile (00TV-100); 12
Schreyerite (00TV-4); 13 Unnamed V-silicate (99TV-68F); detection limits are <0.1 wt.%

22
6.822
0.000
1.649
0.000
3.195
0.020
0.010
0.210
0.000
0.022
0.031
1.632

52.16
0.00
5.99
0.03
28.36
0.00
0.08
1.05
0.01
0.06
0.04
7.98
95.77

Element (wt %)
46.93
SiO2
TiO2
0.00
Al2O3
9.63
Cr2O3
0.00
V2O3
27.41
FeO
0.16
MnO
0.08
MgO
0.97
ZnO
0.00
CaO
0.14
Na2O
0.11
K2O
8.78
Total
94.20

Table 1. Representative electron microprobe analyses of V-bearing silicates and oxides


178
P. G. Spry and N. L. Scherbarth

The goldvanadiumtellurium association

179

Fig. 5. Photomicrographs of a Radiating needles of roscoelite (Rc) in quartz (Qtz).


Masses and needles of intergrowths of karelianite (Kr) and nolanite (Nl) occur in roscoelite (plane polarized light); b Radiating cluster of roscoelite (Rc) needles in quartz
(Qtz). Nolanite (Nl) occurs as masses and as replacements of roscoelite needles (plane
polarized transmitted light); c Euhedral and subhedral karelianite (Kr) crystals in masses
of roscoelite (Rc). Calaverite (Ct) occurs as isolated crystals in roscoelite and as veinlets
cross cutting roscoelite (plane polarized reflected light); d Karelianite (Kr) crystals intergrown with and replaced by nolanite (Nl) within roscoelite (Rc) and quartz (Qtz). Note
the presence of a calaverite (Ct) grain within roscoelite (plane polarized reflected light);
e Karelianite (Kr) intergrown with calaverite (Ct). Nolanite (Nl) has replaced karelianite
(Kr) and roscoelite (Rc). Small bladed subhedral crystals of an unnamed vanadium
silicate (Un) occur in quartz (plane polarized reflected light); f Blades of an unnamed
vanadium silicate (Un), nolanite (Nl) and calaverite (Ct) in quartz (Qtz) (back-scattered
electron image)

180

P. G. Spry and N. L. Scherbarth

Vanadium minerals
Stage 4 calaverite is intimately associated with fine-grained masses of green
mica. EPMA of the green mica shows that it mostly consists of roscoelite (up
to 32.71 wt.% V2O3, which is among the highest reported vanadium values for
roscoelite from an epithermal AuAgTe deposit. (Table 1, Fig. 5a and b). Vanadian muscovite is the other green mica present (Fig. 6).
Five vanadium-bearing oxide minerals were also identified in various veins:
karelianite (V2O3), Ti-free nolanite ((V,Fe,Al)10O14(OH)2, with between 65.2 and
87.3 wt.% V2O3), schreyerite (V2Ti3O9), rutile (with up to 5.2 wt.% V2O3) and magnetite (with up to 0.7 wt.% V2O3). Representative microprobe analyses are reported
in Table 1. Of the vanadium-bearing oxides, karelianite is the most common. It occurs
as euhedral to subhedral crystals (Fig. 5c), elongate laths (Fig. 5d) and irregular
masses up to 2.5 mm in length. Nolanite forms as irregular masses with karelianite
(Fig. 5d) where it appears to have replaced the latter mineral. In these masses, marked
variations occur in the Fe to V ratio of nolanite. It is essentially devoid of Ti, and
contrasts in composition with nolanite from, for example, the Kalgoorlie gold deposit,
which contains minor amounts of Ti (Nickel, 1977). Elsewhere in the Tuvatu deposit,
nolanite occurs as fine elongate crystals that replaced mats (Fig. 5a) and individual
needles of roscoelite (Fig. 5b). Where it replaced roscoelite it gives the roscoelite a
brown hue rather than the more characteristic green color generally associated with
this mineral. Schreyerite, although rare, forms as fine flakes (<0.04 mm) intimately
associated with granular masses of rutile (<0.16 mm) in colloform quartz, roscoelite
and pyritohedral (single or masses) and platy pyrite grains. Magnetite is common as
coarse euhedral grains in early porphyry-style mineralization and to a lesser extent as
subhedral to euhedral minor magnetite in pyrite within epithermal veins. Magnetite
in both settings contains up to 0.7 wt.% V2O3.
Fine needles (up to 150 mm in length) of a vanadium silicate within quartz
occurs on the margins of small intergrowths of karelianite, nolanite and roscoelite

Fig. 6. Chemical variation diagram of V content (per formula


unit) versus Al content (in the
octahedral site) of vanadian mica
at Tuvatu. Note that the most
common mica at Tuvatu is roscoelite rather than vanadian muscovite

The goldvanadiumtellurium association

181

in the Upper Ridges veins. These grains were first identified in back-scattered electron images (Fig. 5d and e) and yielded qualitative analyses using an SEM that
yielded an approximate formula of VSiO3. Electron microprobe analyses yielded
the same 1:1 V:Si ratio, but totals were generally <100 wt.%. It is likely that the
mineral is VSiO3(OH), where V3 is partly replaced by Al3 (Table 1). The composition of this silicate does not correspond to that of any named mineral, and it
may therefore represent a new mineral species.
Discussion
The source of V, Au, and Te is unknown in gold telluride deposits but it is often
linked to alkalic igneous rocks spatially associated with the gold ores (e.g.,
Ahmad et al., 1987; Zhang and Spry, 1994; Pals and Spry, 2003). Jensen and
Barton (2000) pointed out that owing to the chalcophile nature of tellurium, it
is most abundant in mantle and crustal rocks, and particularly enriched in alkalic
hydrothermal systems. In alkalic igneous rocks, tellurium and gold constitutes up
to 10 ppb. Such settings also contain magnetite-bearing mafic igneous rocks that
host vanadium-bearing ores. It should be noted that the Navilawa Monzonite, Sabeto
Volcanics, and the basaltic dikes in the Tuvatu mine area contain 280387 ppm V
(mean 317 ppm V), 292355 ppm V (mean 329 ppm V), and 235465 ppm V
(mean 356 ppm V), respectively, and that V is located in magnetite and phlogopite
in these rocks (Scherbarth and Spry, 2006). These values are higher than the
average crustal abundances values for V of 151 ppm (Rudnick and Fountain,
1995) and 230 ppm (Taylor and McLennan, 1995). Like those at Tuvatu, magnetite
from alkalic rocks associated with the Golden Sunlight (Zhang and Spry, 1994)
and the Porgera gold telluride deposits (Richards, 1990) are also enriched in
vanadium.
Gold is transported in hydrothermal solutions as Au and is carried as aqueous complexes with sulfur-bearing (e.g., AuHS(aq), HAu(HS)2(aq), AuHS2  ) and
halogen-bearing ligands (e.g., AuCl2  ) (Seward, 1991; Cooke and Simmons,
2000) whereas Te is carried as a variety of complexes in the system TeOH
(e.g., H2Te(aq), HTe , Te2 2 ) (Zhang and Spry, 1994; McPhail, 1995). Although
aqueous telluro-gold species are unknown in nature, Seward (1973) speculated
that AuTe2  , Au2(Te2)0, and AuTe2 2 3 may be important in the formation of
gold telluride deposits. Saunders and May (1986) also suggested that AuTe2 
may be the gold complex in gold telluride systems. Other potential aqueous telluro-gold species that have been considered are AuHTe(aq) (McPhail, 1995) and
AuHTe2  (Cooke and McPhail, 2001), as well as structurally bound gold
species such as Fe(SAs)Au(HTe)0 and Fe(SAs)Au2Te0 (Pals et al., 2003)
However, there is an increasing body of thermodynamic, experimental, and
field evidence that Te (McPhail, 1995; Cooke and McPhail, 2001; Larocque
et al., 2006) as well as Au and Ag can also be transported in the vapor phase
(Williams-Jones et al., 2002). Although vanadium is transported in hydrothermal
solutions, it is unclear whether it can be transported in the vapor. The stability of
various species in the system VOH were calculated at 25  C by Wanty and
Goldhaber (1992) and under hydrothermal conditions up to 300  C by Cameron
(1998).

182

P. G. Spry and N. L. Scherbarth

By combining the results of the thermodynamic calculations of Zhang and Spry


(1994) in the system AuTeClSOH and those of Cameron (1998) for the
system VAlSiOH, the stability of the minerals calaverite, roscoelite, and
karelianite can be evaluated at conditions of ore-formation at the Tuvatu deposit
assuming that these minerals deposited from a hydrothermal solution. Details of
the assumptions made for the thermodynamic calculations and the sources of thermodynamic data are given in Zhang and Spry (1994) and Cameron (1998). Calculations for both systems, in large part, employed the isocoulombic method (e.g.,
Gurney, 1938) to calculate equilibrium constants of reactions. Uncertainties of up
to 1 unit in log K (equilibrium constant) may be expected using this method (e.g.,
Mountain and Wood, 1988). Note that thermodynamic data for nolanite and
schreyerite are unavailable.
The stability fields of roscoelite, karelianite, and calaverite in f O2-pH space at
300 and 250  C for Au 1 ppb, Te 1 ppb, S 0.001 m, V 0.0001 m,
and aK 0.01 are shown in Figs. 7 and 8, respectively. The temperatures chosen
are those that span the deposition of tellurides in stage 4 at Tuvatu (mean
Th 257  C, with a trapping temperature <10  C higher than the mean). Values
of Au 1 ppb and Te 1 ppb are typical of those found in modern geothermal
systems whereas those for S 0.001 m, V 0.0001 m, and aK 0.01 are appro-

Fig. 7. Log f O2-pH diagram for the system VAlSiOH superimposed onto the systems
AuVTeClSOH and FeSO at 300  C highlighting the overlapping stability fields
of calaverite with roscoelite, muscovite, and karelianite. Conditions used in construction are Au and Te 1 ppb, S 0.01 m, V 0.0001 m, aK 0.01. Details of calculations of the system AuVTeClSOH and FeSO are given in Zhang and Spry
(1994) whereas those for the system VAlSiOH are reported in Wall et al. (1995) and
Cameron (1998).

The goldvanadiumtellurium association

183

Fig. 8. Log f O2-pH diagram for the system VAlSiOH superimposed onto the systems
AuVTeClSOH and FeSO at 250  C highlighting the overlapping stability fields
of calaverite with roscoelite, muscovite, and karelianite. Conditions used in construction
are Au and Te 1 ppb, S 0.01 m, V 0.0001 m, aK 0.01

priate for compositions associated with alkalic igneous rock gold telluride systems
and were the compositions used by Cameron (1998) in his calculation of the stability of aqueous species and minerals in the system VAlSiOH associated
with the formation of the Porgera gold deposit. Values of S 0.001 m and
aK 0.01 were also proposed by Scherbarth and Spry (2006) and are consistent
with stabilities of members in the systems KAlSiO and FeSO and sulfur isotope compositions when plotted in log f O2-pH space.
At 250 and 300  C for neutral to moderately low pH conditions, V4 and V5
complexes predominate with V3 being incorporated into roscoelite and karelianite. As was shown by Wall et al. (1995) and Cameron (1998), vanadium is transported by an oxidizing fluid and is fixed as roscoelite and karelianite upon
reduction at near neutral and moderately acid conditions respectively. At neutral
to slightly acid conditions, near the hematitemagnetite buffer, calaverite is deposited from solution and overlaps the stability fields of karelianite and roscoelite. The
overlap is greater at 250  C than at 300  C suggesting that the field of overlap in
pH-f O2 space increases with decreasing temperatures. The coexistence among
karelianite, calaverite, and roscoelite at 250  C (approximate temperature of stage
4 mineralization at Tuvatu) suggests pH conditions of 4.5 and a value of log f O2 of
35 to 30. Nolanite formed after karelianite but the conditions of its formation
are unknown.
Conclusions
The elements, Au, Te and V, likely accompany each other throughout the entire ore
forming cycle at Tuvatu. Each of these elements was likely derived from the

184

P. G. Spry and N. L. Scherbarth

Navilawa Monzonite and then subsequently carried in a magmatic hydrothermal


solution in the aqueous phase and possibly in the supercritical vapor phase, in the
cases of Au and Te. Deposition from these fluids resulted in the formation of
precious metal tellurides, including calaverite, and the precipitation of various
vanadium silicates and oxides. The coexistence of calaverite, roscoelite, and karelianite suggests that the ore fluid at 250  C had a pH of 4.5 and a value of log fO2
of 35 to 30. While roscoelite is the most common vanadium phase at Tuvatu,
the incorporation of the other vanadium minerals in roscoelite warrants closer
inspection for similar phases in other epithermal ores.
Acknowledgements
This study was financed by Emperor Gold Mining Company. We thank Tony Woodward,
David A-Izzedin, Saimoni, Don Milella, and Charles Barclay for their assistance while in
Fiji. Greg Cameron is kindly thanked for providing a copy of part of his Ph.D. thesis
pertaining to the stability of vanadium in hydrothermal systems. Diana Oshun read a preliminary draft of the manuscript while Alfred Kracher and Warren Straszheim are thanked for
their assistance with the electron microprobe and SEM analyses, respectively. The manuscript
was improved by the reviews of Nigel Cook, Yuanming Pan, Johann Raith and Jim Saunders.

References
Ahmad M, Solomon M, Walshe JL (1987) Mineralogical and geochemical studies of the
Emperor gold telluride deposit, Fiji, Econ Geol 83: 345370
A-Izzedin D (1998) The Tuvatu gold project, western Viti Levu. Pacific Exploration Technology Conference (PET 98), Nadi, Fiji, (Abstracts)
Ashley PM, Andrew AS (1989) Petrographic, fluid inclusion and stable isotope investigation
of a suite of samples from the Tuvatu prospect, Fiji. Commonwealth Sci Indust Res Org
Div Expl Geosci Restricted Res Rept 50R, pp 36
Blake J (1876) Article IV-On roscoelite, a vanadium mica. Am J Sci 12: 3132
Begg G (1996) Genesis of the Emperor gold deposit, Fiji. PhD thesis, Monash Univ, Clayton,
Australia, p 466
Cameron GH (1998) The hydrothermal evolution and genesis of the Porgera gold deposit,
Papua New Guinea. PhD thesis, Australian National Univ, Canberra, Australia, p 137
Cameron GH, Wall VJ, Walshe JL, Heinrich CA (1995) Gold mineralization at the Porgera
mine, Papua New Guinea, in response fluid mixing. PACRIM 95 Conference, Auckland,
New Zealand, pp 99100 (Abstracts)
Colley H, Flint DJ (1995) Metallic mineral deposits of Fiji. Mineral Res Dept Fiji Mem 4,
192 p
Cooke DR, McPhail DC (2001) Epithermal AuAgTe mineralization, Acupan, Baguio
district, Philippines; numerical simulations of mineral deposition. Econ Geol 96:
109132
Cooke DR, Simmons SF (2000) Characteristics and genesis of epithermal gold deposits. Rev
Econ Geol 13: 221244
Forrest RA (1971) Geology and mineral deposits of the Warm Springs-Giltedge district,
Fergus County, Montana. MS thesis, Montana College Min Sci Tech, Butte, Montana,
USA, p 191
Gatehouse BM, Grey IE, Nickel EH (1983) The crystal chemistry of nolanite (V,Fe,Ti,
Al)10O14(OH)2, from Kalgoorlie, Western Australia. Am Mineral 68: 833839

The goldvanadiumtellurium association

185

Gurney RW (1938) Exchange forces and electrostatic forces between ions in solution. J
Chem Phys 6: 499505
Harris DC (1989) The mineralogy and geochemistry of the Hemlo gold deposit, Ontario.
Geol Surv Can Econ Geol Rep 38, 88 p
Hatcher R (1998) Relation of structures, alteration and mineralisation at Tuvatu gold
prospect, Viti Levu, Fiji Island. BSc Hons thesis, Queensland Inst Tech, Brisbane,
Australia, p 151
Heinrich EW, Levinson AA (1955) X-ray data on roscoelite and barium-muscovite. Am J Sci
253: 3943
Jensen EP, Barton MD (2000) Gold deposits related to alkaline magmatism. Rev Econ Geol
13: 279314
Kelly WC, Goddard EN (1969) Telluride ores of Boulder County, Colorado. Geol Soc Am
Mem 109, 237 p
Kurtz JP, Hauff PL (1988) Roscoelite in Colorado telluride ores. In: Modreski PJ (ed)
Mineralogy of precious metal deposits. Friends of Mineralogy, Golden, 20
Larocque ACL, Stimac JA, Siebe C, Greengrass K, Chapman R, Mejia SR (2006) Deposition of a high sulphidation Au assemblage from a magmatic volatile phase, Volcan
Popocatapetl, Mexico. J Volcan Geotherm Res (in press)
Lindgren WT (1907) Some gold and tungsten deposits of Boulder County, Colorado. Econ
Geol 2: 453463
McDougall I (1963) Potassium-argon ages of some rocks from Viti Levu, Fiji. Nature 198: 677
McPhail DC (1995) Thermodynamic properties of aqueous tellurium species between 25  C
and 350  C. Geochim Cosmochim Acta 59: 851866
Mountain BW, Wood SA (1988) Chemical controls on the solubility, transport, and deposition
of platinum and palladium in hydrothermal solutions: a thermodynamic approach. Econ
Geol 83: 492510
Nickel EH (1977) Mineralogy of the Green Leader gold ore at Kalgoorlie, Western
Australia. Proc Australas Inst Min Metall 263: 913
Nickel EH, Grey IE (1982) A vanadium-rich mineral assemblage associated with the gold
telluride ores at Kalgoorlie. Proc XIII Gen Meeting, Varna, Bulgaria. Int Mineral Assoc,
pp 116
Pals DW, Spry PG (2003) Telluride mineralogy of the low-sulfidation epithermal Emperor
gold deposit, Vatoukola, Fiji. Mineral Petrol 79: 285307
Pals DW, Spry PG, Chryssoulis S (2003) Invisible gold and tellurium in arsenic-rich pyrite
from the Emperor gold deposit, Fiji: implications for gold distribution and deposition.
Econ Geol 98: 493514
Pan Y, Fleet ME (1991) Vanadian allanite-(La) and vanadian allanite-(Ce) from the Hemlo
gold deposit, Ontario, Canada. Mineral Mag 55: 497507
Pan Y, Fleet ME (1992) Mineral chemistry and geochemistry of vanadian silicates in the
Hemlo gold deposit, Ontario, Canada. Contrib Mineral Petrol 109: 511525
Richards JP (1990) Petrology and geochemistry of alkalic intrusives at the Porgera gold
deposit, Papua New Guinea. J Geoch Expl 35: 141199
Rodda P (1967) Outline of the geology of Viti Levu, Fiji. New Zealand J Geol Geophys 10:
12601273
Rudnick RL, Fountain DM (1995) Nature and composition of the continental crust; a lower
perspective. Rev Geophys 33: 267309
Saunders JA (1991) Gold deposits of the Boulder County gold district, Colorado. US Geol
Surv Bull 1857-I: 137148
Saunders JA, May ER (1986) Bessie G: a high-grade epithermal gold telluride deposit,
La Plata County, Colorado, U.S.A. In: Macdonald AJ (ed) Proceedings of Gold 86,

186

P. G. Spry and N. L. Scherbarth: The goldvanadiumtellurium association

an international symposium on the geology of gold. Konsult Internat, Toronto,


pp 436444
Scherbarth NL (2002) Petrological, mineralogical, fluid inclusion and stable isotope characteristics of the Tuvatu goldsilver telluride deposit, upper Sabeto River, Fiji. Ms thesis,
Iowa State Univ, Ames, Iowa, USA, p 157
Scherbarth NL, Spry PG (2000) Mineralogical and geochemical characteristics of the Tuvatu
goldsilver telluride deposit, Upper Sabeto River, Fiji. In: Bucci LA, Mair JL (eds) Gold
in 2000. Short Course Poster Session Ext Abst Vol, November 1011, 2000, Lake Tahoe,
Nevada, Univ. Western Australia, pp 9297
Scherbarth NL, Spry PG (2001) Magmatic to epithermal evolution of the Tuvatu goldsilver
telluride system, Upper Sabeto River area, Fiji. Geol Soc Am Abstr Prog 33: A-46
Scherbarth NL, Spry PG (2006) Mineralogical, petrological, stable isotope, and fluid inclusion characteristics of the Tuvatu goldsilver telluride deposit, Fiji. Comparisons with
the Emperor deposit. Econ Geol 10 (in press)
Seward TM (1973) Thio complexes of gold and transport of gold in hydrothermal solutions.
Geochim Cosmochim Acta 37: 379399
Seward TM (1991) The hydrothermal geochemistry of gold. In: Foster RP (ed) Gold
metallogeny and exploration. Blackie, London, pp 435486
Simpson ES (1952) Minerals of Western Australia. Western Australian Government Perth,
Western Australia: 3
Spiridinov EM (1978) Titanium coulsonite from the Kalgoorlie deposit, Australia. Dokl
Akad Nauk SSSR 245: 180181
Taylor SR, McLennan SM (1995) The geochemical evolution of the continental crust. Rev
Geophys 33: 241265
Thieben SE, Spry PG (1995) The geology and geochemistry of Cretaceous-Tertiary alkaline
rock-related goldsilver telluride deposits of Montana, USA. In: Pasava J, Krbek B, Z ak
K (eds) Mineral deposits: from their origin to their environmental impacts. A. A.
Balkema, Rotterdam, pp 199202
Wall VJ, Munroe SM, Cameron GH, Heinrich CA, Cox SF, Walshe JL (1995) Porgera:
plumbing and processes of gold mineralization. In: Clark AH (ed) Giant ore deposits
II. Proceedings of the Giant Ore Deposits Workshop: Queenss University, Kingston,
Ontario, pp 649667
Wanty RB, Goldhaber MB (1992) Thermodynamics and kinetics of reactions involving
vanadium in natural systems: accumulation of vanadium in sedimentary rocks. Geochim
Cosmochim Acta 56: 14711483
Williams-Jones AE, Migdisov AA, Archibald SM, Xiao Z.-F. (2002) Vapor-transport of
metals. In: Hellman R, Wood SA (eds) Water-rock interactions, ore deposits, and environmental geochemistry: a tribute to David A. Crerar. Geochem Soc Spec Publ 7:
279305
Zhang X, Spry PG (1994) Petrological, mineralogical, fluid inclusion, and stable isotope
studies of the Gies goldsilver telluride deposit, Judith Mountains, Montana. Econ Geol
89: 602627
Authors addresses: P. G. Spry (corresponding author; e-mail: pgspry@iastate.edu), Department of Geological and Atmospheric Sciences, 253 Science I, Iowa State University, Ames,
Iowa 50011-3212, U.S.A.; N. L. Scherbarth, Tectonic Resources NL, Suite 4 100 Hay Street,
Subiaco, Western Australia 6008, Australia

Das könnte Ihnen auch gefallen