Sulfur sources of buckshot pyrite in the Auriferous

advertisement
Sulfur sources of buckshot pyrite in the Auriferous
Conglomerates of the Mesoarchean Witwatersrand and
Ventersdorp Supergroups, South Africa
The MIT Faculty has made this article openly available. Please share
how this access benefits you. Your story matters.
Citation
Guy, B. M., S. Ono, J. Gutzmer, Y. Lin, and N. J. Beukes. “Sulfur
Sources of Sedimentary ‘buckshot’ Pyrite in the Auriferous
Conglomerates of the Mesoarchean Witwatersrand and
Ventersdorp Supergroups, Kaapvaal Craton, South Africa.” Miner
Deposita 49, no. 6 (April 9, 2014): 751–775.
As Published
http://dx.doi.org/10.1007/s00126-014-0518-3
Publisher
Springer-Verlag
Version
Author's final manuscript
Accessed
Wed May 25 22:46:33 EDT 2016
Citable Link
http://hdl.handle.net/1721.1/98891
Terms of Use
Creative Commons Attribution-Noncommercial-Share Alike
Detailed Terms
http://creativecommons.org/licenses/by-nc-sa/4.0/
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
Sulfur Sources of Buckshot Pyrite in the Auriferous Conglomerates of the Mesoarchean Witwatersrand
and Ventersdorp Supergroups, South Africa
B. M. Guya,*, S. Onob, J. Gutzmera,c, Y. Linb,1, A. Bauerb and N. J. Beukesa
a
Department of Geology, University of Johannesburg, Auckland Park, 2006, South Africa
email: nbeukes@uj.ac.za
b
Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, 77
Massachusetts Avenue, Cambridge, Massachusetts, 02139, USA
email: sono@mit.edu, annbauer@mit.edu
c
Helmholtz-Institute Freiberg for Resource Technology, Halsbrücker Str. 34, 09599 Freiberg, Germany.
email: j.gutzmer@hzdr.de
*
Present address: Mineral Services, SGS South Africa, 58 Melville Street, Booysens, Johannesburg, 2091, South
Africa.
email: bradley.guy@sgs.com
1
Present address: School of Earth and Space Sciences, University of Science and Technology of China, Hefei,
230026, P. R. China
email: yinglin@ustc.edu.cn
Abstract
Large rounded pyrite grains (>1 mm), commonly referred to as ‘buckshot’ pyrites, are a characteristic feature of
the auriferous conglomerates (reefs) in the Witwatersrand and Ventersdorp supergroups. Detailed petrographic
analyses of the reefs indicated that the vast majority of the buckshot pyrites are of reworked sedimentary origin,
i.e., that the pyrite originally formed in the sedimentary environment during sedimentation and diagenesis.
Forty-five of these reworked sedimentary pyrite grains from the Main, Vaal, Basal, Kalkoenkrans, Beatrix and
Ventersdorp Contact reefs were analyzed for their multiple sulfur isotope compositions (δ34S, Δ33S and Δ36S) to
determine the source of the pyrite sulfur. In addition, four epigenetic pyrite samples (pyrite formed after
sedimentation and lithification) from the Middelvlei and the Ventersdorp Contact reefs were measured for
comparison. The δ34S, Δ33S and Δ36S values of all 45 reworked sedimentary pyrite grains indicate clear
signatures of mass-dependent and mass-independent fractionation and range from −6.8 to +13.8‰, −1.7 to
+1.7‰ and −3.9 to +0.9‰, respectively. In contrast, the four epigenetic pyrite samples display a very limited
range of δ34S, Δ33S and Δ36S values (+0.7 to +4.0‰, −0.3 to +0.03‰ and −0.3 to +0.1‰, respectively).
Despite the clear signatures of mass-independent sulfur isotope fractionation, very few data points plot along
the primary Archean photochemical array suggesting a weak photolytic control over the data set. Instead, other
factors command a greater degree of influence such as pyrite paragenesis, the prevailing depositional
environment and non-photolytic sulfur sources. In relation to pyrite paragenesis, reworked syngenetic
sedimentary pyrite grains (pyrite originally precipitated along the sediment-water interface) are characterized by
negative δ34S and Δ33S values, suggesting open system conditions with respect to sulfate supply and the
presence of microbial sulfate reducers. On the contrary, most reworked diagenetic sedimentary pyrite grains
(pyrite originally precipitated below the sediment-water interface) show positive δ34S and negative Δ33S values,
suggesting closed system conditions. Negligible Δ33S anomalies from epigenetic pyrite suggest that the sulfur
was sourced from a mass-dependent or isotopically homogenized metamorphic/hydrothermal sulfur reservoir.
Contrasting sulfur isotope compositions were also observed from different depositional environments, namely
fluvial conglomerates and marine-modified fluvial conglomerates. The bulk of the pyrite grains from fluvial
conglomerates are characterized by a wide range of δ34S values (−6.2 to +4.8‰) and small Δ33S values (±0.3‰).
This signature likely represents a crustal sulfate reservoir derived from either volcanic degassing or from
weathering of sulfide minerals in the hinterland. Reworked sedimentary pyrite grains from marine-modified
fluvial conglomerates share similar isotope compositions, but also produce a positive Δ33S/δ34S array that
overlaps with the composition of Archean barite, suggesting the introduction of marine sulfur. These results
demonstrate multiple sources of sulfur, including atmospheric, crustal and marine sulfur reservoirs. The
prevalence of the mass-dependent crustal sulfur isotope signature suggests that sulfate concentrations were
probably much higher in terrestrial settings in comparison to marine environments, which were sulfate-deficient.
However, the optimum conditions for forming sedimentary pyrite with a crustal sulfur isotope signature were
probably not during fluvial progradation but rather during the early phases of flooding of low angle
unconformities, i.e., during retrogradational fluvial deposition, coupled in some cases with marine
transgressions, immediately following inflection points of maximum rate of relative sea level fall.
1
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
Introduction
Pyrite in the gold-bearing conglomerates of the upper Witwatersrand Supergroup has traditionally been
classified into three groups, namely ‘compact rounded’, ‘porous rounded’ and hydrothermal (Ramdohr 1958;
Saager 1970; Utter 1978; Hirdes and Saager 1983; England et al. 2002). The use of descriptive terminology for
the first two groups is primarily a reflection of genetic uncertainty, i.e., whether the rounded pyrite grains
represent clastic components of the conglomerate or whether they represent post-depositional pseudomorphs
after rounded Fe-oxide minerals (Barnicoat et al. 1997; Law and Phillips 2005). Despite this uncertainty, most
authors regard ‘compact rounded’ pyrite and ‘porous rounded’ pyrite as fragments of pyrite that were derived
from the erosion of granitic-greenstone and sedimentary sources, respectively (Hallbauer 1986; England et al.
2002).
Reworked pyrite grains of a sedimentary origin have been documented by most Witwatersrand researchers,
i.e., the so-called ‘buckshot’ pyrites, which are generally large rounded pyrite grains (>1 mm in size). However,
the bulk of the literature has focused on compact rounded pyrite (<1 mm in size) and hydrothermal pyrite
because of their significance in gold exploration in the Witwatersrand Supergroup (McLean and Fleet 1989;
Phillips and Myers 1989; Myers et al. 1993; Barnicoat et al. 1997; Fleet 1998; Kirk et al. 2001; England et al.
2002; Frimmel et al. 2005). Consequently, reworked sedimentary pyrite grains are routinely overlooked, despite
the fact that they comprise the bulk of total pyrite in several auriferous conglomerates of the Witwatersrand
Supergroup (Ramdohr 1958; Utter 1978; Tucker 1980; Hirdes and Saager 1983; Hallbauer 1986; Hutchinson
and Viljoen 1987; England et al. 2002) and that they can contain substantial contents of solid-solution gold
and/or nano-particle gold (Utter 1978; Koglin et al. 2010; Large et al. 2013). Considering the scientific and
economic importance of this world class Au-U deposit, it is somewhat surprising that virtually no one has
questioned how the substantial quantities of sedimentary pyrite could have formed if the Archean atmosphere
was highly depleted in oxygen (Farquhar et al. 2000), implying that sulfide weathering would have been
negligible and that both continental runoff and ocean water would have essentially been sulfate-free (Habicht et
al. 2002; Frimmel 2005).
Previous authors have reported relatively small δ34S variations from Witwatersrand pyrite (−0.5 to +6.5‰)
(Hoefs et al. 1968; Hattori et al. 1983; Palmer 1986; Strauss and Beukes 1991; Zhao 1998). The small range of
δ34Spy values from barren and mineralized horizons, and between different paragenetic associations, has been
used to suggest a magmatic source for the sulfur. However, recent small-scale analyses of the ‘porous rounded’
pyrite (laser ablation and micro-drilling) have revealed a much wider range of δ34Spy values (−17 to +16‰),
indicating that a portion of Witwatersrand pyrite might have formed through microbial sulfate reduction
(England et al. 2002; Hofmann et al. 2009; Large et al. 2013). This information, supplemented with multiple
sulfur isotope data (e.g., Hofmann et al. 2009), provides further evidence for a sedimentary origin as first
suggested by the petrographic studies of Ramdohr (1958), Saager (1970), Utter (1978), Dimroth (1979), Hirdes
and Saager (1983) and Hallbauer (1986).
In this paper, the results of a combined petrographic and multiple sulfur isotope investigation (δ 34S, Δ33S and
36
Δ S) are presented in order to constrain the source of the sulfur for reworked sedimentary pyrite grains of the
upper Witwatersrand and lower Ventersdorp supergroups (~2.9 to 2.7 Ga). In the first part of the investigation,
pyrite is classified according to textural and paragenetic constraints, where it will be argued that the vast
majority of the buckshot pyrite grains were derived from a sedimentary source. In the second part of the
investigation, it is shown that the bulk of the sulfur isotope compositions are controlled by terrestrial geological
processes with only minor atmospheric input. Finally, a model for sedimentary pyrite formation in the upper
Witwatersrand and lower Ventersdorp supergroups is presented.
Geological setting
The Witwatersrand Supergroup represents a ~7 km thick siliciclastic sequence that was deposited between 2.985
and 2.780 Ga over the central parts of the Archean Kaapvaal Craton (Fig. 1a) (Kositcin et al. 2003; Kositcin and
Krapez 2004; SACS 2006). The succession unconformably overlies a granitoid-greenstone basement (3.48 to
3.03 Ga) and the volcano-sedimentary rocks of the 3.09 to 3.07 Ga Dominion Group (Armstrong et al. 1991;
Robb et al. 1992; Poujol et al. 2003). The Witwatersrand Supergroup has experienced only low-grade
greenschist facies regional metamorphism, although certain areas have experienced higher grades (e.g., the
northwestern parts of the basin, the overturned collar of the Vredefort Dome and the northeastern part of the
Evander gold field) (Gibson and Wallmach 1995; Frimmel 1997; Phillips and Law 2000; Law and Phillips
2005). Stratigraphically, the Witwatersrand Supergroup is divided into the lower West Rand Group (~5 km
thick) and the upper Central Rand Group (~2 km thick) (Fig. 1b).
2
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
Fig. 1.
The West Rand Group is composed of arenite and shale, with minor banded iron formation (BIF),
conglomerate, diamictite and lava. Marine shelf depofacies are prominent in the lower parts of the group,
whereas fluvial braidplain and distal marine shelf depofacies characterize the middle to upper parts (Watchorn
1981; Tankard et al. 1982; Beukes 1995). Most authors agree that the lower Hospital Hill Subgroup was
deposited in a passive margin setting, with an open ocean to the southeast. However, there is little consensus
regarding the tectonic setting of the middle and upper parts of the West Rand Group. Whilst some authors have
argued for a passive margin setting for the entire West Rand Group (Kositcin and Krapez 2004; Koglin et al.
2010), others have suggested that the change to a foreland basin setting occurred at the junction between the
Hospital Hill and Government subgroups (Beukes 1995; Coward et al. 1995; Schmitz et al. 2004). Alternatively,
Catuneanu (2001) proposed a foreland basin setting for the entire Witwatersrand Supergroup, where the West
Rand and Central Rand groups represent the underfilled and overfilled stages, respectively. The maximum age
of the West Rand Group is 2985 ± 14 Ma (detrital zircon grains, Kositcin and Krapez 2004) and the minimum
age for most of the group is 2914 ± 8 Ma (the age of the lavas in the Crown Formation, Armstrong et al. 1991).
In contrast to the West Rand Group, the Central Rand Group contains a greater proportion of arenaceous and
rudaceous lithotypes, with minor argillaceous and volcanic units such as the Booysens Formation and the Bird
Lava, respectively. The Central Rand Group is dominated by fluvial braidplain sedimentary rocks that were
deposited in a restricted retroarc foreland basin setting (Catuneanu 2001; Kositcin and Krapez 2004; Schmitz et
al. 2004; Frimmel et al. 2005). Fluvial sedimentation was controlled by deep-seated thrust faults along the basin
margin, which led to synsedimentary folding and progressive uplift, source area rejuvenation, erosion and
recycling of older sediments (Myers et al. 1990; McCarthy 1992; Coward et al. 1995). Marine incursions are
recorded at various times during deposition of the Central Rand Group in the form of laterally continuous
mudstones (Booysens Formation) and transgressive shallow marine quartz arenites. A number of prominent
gold-bearing conglomerates are overlain by transgressive shallow marine quartz arenites (e.g., the Main, Carbon
Leader, Vaal, Basal and Denny’s reefs; Fig. 2). Marine modification of these fluvial conglomerates, via
winnowing and wave reworking, likely played a crucial role in upgrading their gold and heavy mineral content
(Minter 1972; Verrezen 1987; Watchorn and O’Brien 1991; Beukes and Nelson 1995; Catuneanu and Biddulph
2001). The maximum age of the Central Rand Group is 2902 ± 13 Ma (detrital zircon grains, Kositcin and
Krapez 2004) and the minimum age is 2780 ± 3 Ma (authigenic xenotime, Kositcin et al. 2003). However, based
on stratigraphic correlations between the Mozaan Group and the Witwatersrand Supergroup, and indications that
the two successions were deposited in the same basin, i.e., the Greater Witwatersrand basin (Beukes and
Cairncross 1991), it is possible that Central Rand Group sedimentation came to an end around 2837 Ma
(Gutzmer et al. 1999).
The auriferous Ventersdorp Contact Reef (VCR) overlies the Witwatersrand Supergroup with an erosional
contact. Deposition probably took place around 2780 Ma based on the erosional event that preceded the
correlative Derdepoort, Dennilton and Kanye-Lobatse volcanic successions (Beukes and Nelson 1995; Hartzer
2000). The reef forms part of the Venterspost Formation and represents terraced fluvial and debris-flow deposits
(Krapez 1985; Frimmel et al. 2005).
Methods and samples
Forty-four pyrite-bearing conglomerate samples of the Central Rand Group and Ventersdorp Contact Reef were
collected from fresh drill-core and underground mine workings from the East Rand, Central Rand, West Rand,
Carletonville, Klerksdorp and Welkom gold fields (Fig. 1b). Sampled reefs include the North, Carbon Leader,
Main Reef Leader, Main, Middelvlei, Basal, Vaal, Saaiplaas, Kalkoenkrans, Kimberley, Beatrix, Kloof, Elsburg
and the Ventersdorp Contact reefs (Fig. 2). From these samples, polished blocks were prepared and examined
using reflected light and scanning electron microscopy (SEM). Minerals were identified in the SEM by energydispersive spectroscopy (EDS). Petrographic studies indicated that seven conglomerates contained reworked
sedimentary pyrite grains suitable for micro-drilling (i.e., grains larger than 500 μm) (Fig. 2):
Fluvial conglomerates
a.
b.
c.
d.
Beatrix Reef, Welkom gold field (Minter et al. 1988; Genis 1990)
Kalkoenkrans Reef, Welkom gold field (H Brouwer 2013, pers. comm.)
Middelvlei Reef, Carletonville gold field (Myers et al. 1993)
Ventersdorp Contact Reef, Carletonville gold field (Krapez 1985)
3
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
Marine-modified fluvial conglomerates
e.
f.
g.
Vaal Reef, Klerksdorp gold field (Catuneanu and Biddulph 2001)
Basal Reef, Welkom gold field (Frimmel et al. 2005)
Main Reef, Carletonville gold field (Beukes and Nelson 1995)
From these seven reefs, 45 reworked sedimentary pyrite grains and four epigenetic pyrite samples were
selected for multiple sulfur isotope analyses (33S/32S, 34S/32S and 36S/32S). Sulfur was extracted by conventional
Cr reduction and Ag2S precipitation (Canfield et al. 1986). Ag2S was then heated to 300°C, in the presence of
F2, to form SF6. The SF6 (3 to 6 μmol) was purified by a gas chromatography (GC) system similar to that
described in Ono et al. (2006), and introduced to an isotope ratio mass spectrometer using a dual-inlet mode for
measurements of masses 127, 128, 129, and 131. The mass spectrometry measurement consists of three runs,
with each run consisting of eight cycles of reference versus sample injection. Reproducibility for complete
analysis, from fluorination, GC purification, and isotope ratio analysis are 0.1, 0.2 and 0.4‰ (1σ) for δ33S, δ34S
and δ36S, respectively, and 0.01 and 0.1‰ (2σ) for Δ33S and Δ36S, respectively. Sulfur isotope values are
reported relative to the Vienna Canyon Diablo Troilite Standard (V-CDT) by defining IAEA S-1 to be: δ33S =
−0.055‰, δ34S = −0.300‰ and δ36S = −1.14‰.
Fig. 2.
Classification of pyrite
The auriferous conglomerates contain a wide variety of paragenetic and morphological types of pyrite, which
are typically concentrated on footwall contacts along with other detrital heavy minerals and gold (Ramdohr
1958; Saager 1970; Utter 1978; Hirdes 1979; Hirdes and Saager 1983; Hallbauer 1986; Fleet 1998; England et
al. 2002). In this contribution, pyrite is classified according to its position in the paragenetic sequence (e.g.,
Hallbauer 1986), where each paragenetic association is further subdivided according to textural constraints. For
ease of reference, the genetic classification system employed by Guy et al. (2010) for pyrite in the West Rand
Group is utilized in the current contribution (Table 1). Application of this classification to the conglomerates of
the Central Rand Group and Ventersdorp Supergroup required two minor adjustments, namely:
207
208
209
1) the inclusion of a syngenetic (synsedimentary) pyrite association (SYN) and
2) the allocation of certain diagenetic pyrite textures (e.g., banded and bladed pyrite), which were not
observed in the West Rand Group, into the nodular and concretionary category (DIA-5).
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
It should be noted that the syngenetic and diagenetic pyrite grains in the auriferous conglomerates occur as
reworked fragments of earlier sedimentary pyrite aggregates, and not as components formed in situ as reported
for the West Rand Group (Guy et al. 2010). Furthermore, many of the delicate pyrite types that were observed in
the shales of the West Rand Group were not observed in the conglomerates of the Central Rand Group and
lower Ventersdorp Supergroup. These include finely disseminated pyrite (DIA-3), pyrite aggregates (DIA-4),
dendritic skeletal pyrite (DIA-7), pyrite laths (DIA-8), phyllosilicate pyrite (DIA-9) and pyrite whiskers (DIA10). Chert pebbles with bands of finely disseminated pyrite, as documented by Hirdes and Saager (1983), are
not included in this classification system since they were not observed during the petrographic examinations.
However, this particular type of pyrite will be revisited in the discussion.
Pyrite descriptions
In the descriptions that follow, each type of pyrite will be discussed, with specific reference to pyrite texture,
stratigraphic occurrence and other salient aspects, including mechanisms of formation and a review of previous
literature (where applicable).
Detrital pyrite
The detrital pyrite association (DET) pertains to pyrite that was initially eroded and transported from an external
source area, and then incorporated into the auriferous conglomerates. There are two types of detrital pyrite:
compact rounded pyrite (DET-1) and pyrite in composite lithoclasts (DET-2) (Table 1).
Table 1.
4
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
Compact rounded pyrite (DET-1) represents inclusion-free, non-porous euhedral pyrite crystals (usually as
single crystals; Fleet 1998) that have been rounded to various degrees (Fig. 3a). The average grain size ranges
from 0.1 to 1.5 mm in diameter, although grains as small as 10 μm and as large as 5 mm have been documented
(Feather and Koen 1975; Utter 1978). The majority of compact rounded pyrite grains identified in this study
were less than 0.5 mm in diameter. Detrital grains of compact pyrite and cobaltite are common in the Main,
Main Reef Leader, Carbon Leader and Elsburg reefs. Conversely, detrital grains of compact pyrite are less
common in hydrothermally altered reefs (e.g., the Middelvlei Reef) and in reefs dominated by reworked
sedimentary pyrite grains (e.g., the Beatrix, Kalkoenkrans and Ventersdorp Contact reefs). Most authors agree
that compact rounded pyrite was derived from an Archean granite-greenstone hinterland (Utter 1978; Hirdes and
Saager 1983). However, recent chemical and Pb-Pb isotopic studies have shown that compact rounded pyrite
could also represent rounded fragments of diagenetic and/or metamorphic pyrite (Guy et al. 2010; Large et al.
2013). Thus, compact rounded pyrite could have been derived from a variety of geological source areas. It is
also acknowledged that an unknown percentage of compact rounded pyrite grains may represent pseudomorphs
after detrital iron-oxide/iron-titanium-oxide minerals (Ramdohr 1958; Saager 1970) but also rutile, chlorite,
shale, chert, quartz and other lithic fragments (Myers et al. 1993; Zhou et al. 1995). Note that compact rounded
pyrite grains were not analyzed for their multiple sulfur isotope compositions since they were not the focus of
this investigation (due to small grain sizes and uncertainties relating to provenance).
The second type of detrital pyrite relates to pyrite situated in lithoclasts (DET-2). The designation of this
type of pyrite into the detrital paragenetic association is appropriate since the host rock or mineral particle from
which it came was lithified, eroded and transported to its current position (Hallbauer 1986; Fig. 7A of England
et al. 2002). The term is generally applied when the provenance of the pyrite-bearing lithoclast is equivocal.
Reworked syngenetic pyrite
The reworked syngenetic pyrite association (SYN) refers to pyrite that originally formed in the water column or
at the sediment-water interface (i.e., synsedimentary pyrite). This contrasts with diagenetic pyrite that formed
below the sediment-water interface. There are two types of reworked syngenetic pyrite present in the auriferous
conglomerates, namely oolitic pyrite (SYN-1) and fine-grained micro-porous pyrite (SYN-2) (Table 1).
Oolitic pyrite, or concentrically laminated pyrite (SYN-1), are spherical to near-spherical pyrite grains (up to
7 mm in size, Hallbauer 1986) that contain fine concentric pyrite laminae ranging between 5 and 50 µm in size.
Some examples may possess a core of radiating pyrite crystals (Fig. 3b, c; Figs. 5A and 15A of England et al.
2002). Oolitic pyrite has been documented from the Vaal, Kimberley, Elsburg, Composite, Ventersdorp Contact
and Black reefs (Ramdohr 1958; Utter 1978; Hirdes 1979; Tucker 1980; Hallbauer 1986; England et al. 2002).
In this study, oolitic pyrite grains were observed in the Carbon Leader, Middelvlei, Elsburg and Ventersdorp
Contact reefs. Petrographic studies reveal the presence of galena, chalcopyrite and gold inclusions within the
cortical laminae of the pyrite oolites (Fig. 3d). A replacement origin of precursor carbonate or ironstone ooids
has been suggested by previous authors (e.g., England et al. 2002). However, these explanations are unlikely
considering that carbonate detritus has not been reported from the auriferous conglomerates and that
phyllosilicate minerals occur within the pyrite laminae (Tucker 1980; Agangi et al. submitted; this study).
Furthermore, variable δ34S signatures across the cortical envelope would argue against post-depositional
sulfidation (England et al. 2002; Schieber and Riciputi 2005). Schieber and Riciputi (2005) envisaged that
oolitic pyrite formed via the accretion of iron sulfide laminae under shallow burial conditions, which were
subject to periods of intermittent reworking, grain abrasion and reburial.
Only one example of fine-grained, micro-porous pyrite (SYN-2) was observed, namely in the Ventersdorp
Contact Reef (Fig. 3e, f). The micro-porous pyrite occurs in a laminated pyrite grain. Individual laminae are
generally less than 500 µm thick. The pyrite is very fine-grained and contains numerous microscopic pores
(~0.5 µm in size), as well as inclusions of galena. These features, together with the absence of siliciclastic
mineral inclusions, suggest that pyrite nucleation was rapid and that precipitation occurred from highly
supersaturated fluids (Scott et al. 2009). The most interesting feature, however, is the occurrence of microspherical structures (Fig. 3f). These micro-spherical structures, which are composed of quartz, are orientated
parallel to the pyrite laminae and exhibit a uniform size range between 10 to 20 µm. Although speculative, they
may represent pseudomorphs after gaseous fenestrae associated with microbes or decaying organic matter.
Reworked diagenetic pyrite
Diagenetic pyrite (DIA) refers to pyrite that formed below the sediment surface after sedimentation but before
lithification and metamorphism. Six major types of reworked diagenetic pyrite were observed, namely
5
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
framboidal pyrite (DIA-1), small euhedral pyrite crystals (DIA-2), pyrite nodules and concretions (DIA-5),
euhedral pyrite crystals (DIA-6), pyrite overgrowths (DIA-11) and replacive pyrite (DIA-12) (Table 1). Of
these, pyrite nodules and concretions are the most abundant.
Framboidal pyrite (DIA-1) was observed in samples from the Vaal, Beatrix and Kalkoenkrans reefs (Fig.
3g). The individual pyrite framboids are relatively large (10 to 20 µm) and together form sizeable aggregates (1
to 2 mm in size). Thin compact overgrowths were commonly observed around the framboids. Putative atoll
structures, which are thought to reflect the diagenetic transformation of framboidal pyrite (England and Ostwald
1993), were observed in the Kalkoenkrans Reef (Fig. 3h). An early diagenetic origin for the framboidal pyrite is
inferred by the large diameter of the pyrite framboids (~10 µm), as syngenetic framboids that form in euxinic
water columns are generally much smaller (<5 µm) (Wilkin et al. 1996; Wilkin et al. 1997). However, the
uniform size range of the framboids, their close packing and lack of clastic mineral inclusions may point
towards an earlier stage of formation (i.e., syngenetic). Pyrite framboids in carbonaceous chert have also been
reported by Hirdes (1979) and Hirdes and Saager (1983) in the Evander gold field. It is interesting to speculate
on the source area of the above-mentioned pyrite-chert pebbles, since a greenstone belt provenance has been
inferred for much of the detritus in the Kimberley Reef (e.g., a high proportion of BIF pebbles, chromite grains
and platinum group minerals; Hirdes and Saager 1983).
Small (<30 µm) euhedral pyrite crystals (DIA-2) were observed in reworked and rounded carbonaceous
lithoclasts in the Basal, Vaal, Elsburg and Ventersdorp Contact reefs (Fig. 3i). In situ analogues of this type of
pyrite were also observed in carbonaceous shale units from the West Rand Group (Guy et al. 2010) and
generally form during the early stages of diagenesis (Taylor and Macquakor 2000).
Pyrite nodules and concretions (DIA-5) constitute the bulk of the reworked and rounded diagenetic pyrite in
the auriferous conglomerates (Saager 1970). Volumetrically, pyrite nodules are subordinate to pyrite
concretions. Pyrite nodules have been interpreted to form during the earliest stages of diagenesis in shallow
burial depths where the sediment is characterized by a high degree of plasticity, high fluid content and an
isotropic stress field (Selles-Martinez 1996). The defining criterion for pyrite nodules is the general lack of
siliciclastic mineral inclusions. Three broad subtypes of pyrite nodules are recognized: a) dendritic, b) bladed,
radial and c) banded.
Dendritic pyrite nodules are characterized by cone-in-cone or branching structures, and are common in the
Basal (Fig. 3j, k) and Steyn reefs (Fig. 6A of England et al. 2002). Similar in situ dendritic pyrite nodules have
been reported from the West Rand Group (Guy et al. 2010) and from Phanerozoic black shale successions
(Woodland 1975; Carstens 1985; Carstens 1986; Ruppert et al. 2005).
Bladed and radial pyrite nodules were observed in the Kalkoenkrans, Basal and Ventersdorp Contact reefs
(Fig. 3b, l, m), but have also been reported from the Kimberley, Steyn, Composite and Elsburg reefs (Tucker
1980; Hirdes and Saager 1983; Hallbauer and von Gehlen 1983; England et al. 2002; Reimold et al. 2004).
Although pyrite can develop extremely elongated habits (Ohfuji and Rickard 2005), previous authors have
suggested that the bladed and radial forms of pyrite were initially derived from marcasite that had converted to
pyrite during greenschist facies metamorphism (Hallbauer and von Gehlen 1983; Hallbauer 1986; England et al.
2002; Falconer 2003; Falconer et al. 2006). For example, the radial pyrite of Hirdes and Saager (1983, Plates
10E and F) displays a striking resemblance to the radial marcasite of Falconer et al. (2006, Fig. 6D). In this
study, marcasite ancestry is supported by the observation of a conspicuous texture that resembles the
recrystallized marcasite of Falconer et al. (2006) – compare Fig. 3n and Fig. 54 of Ramdohr (1958) with Fig. 6F
of Falconer et al. (2006). Interestingly, abiotic dissolution or microbial tunneling is documented from one of
these samples (Fig. 3o). Alternative explanations for the bladed and radial textures include derivation from
massive sulfide deposits and replacement of precursor evaporitic gypsum (England et al. 2002). However, the
lack of massive sulfide detritus (e.g., detrital base metal sulfides) and evaporitic gypsum in the Witwatersrand
Supergroup, coupled with the recognition of sedimentary equivalents, supports a sedimentary provenance.
Banded pyrite nodules are characterized by various textures associated with growth banding, i.e., the
banding does not reflect pyritized sedimentary layers (see replacive pyrite, DIA-12). Banded pyrite nodules
have been recorded from the Kimberley, Steyn and Black reefs (Ramdohr 1958; Hirdes and Saager 1983;
England et al. 2002). The ‘chevron-textured’ and ‘vein-fill’ pyrite of England et al. (2002) probably represents
two varieties of banded pyrite (Fig. 6E of Falconer et al. 2006). In this study, only a small number of banded
pyrite nodules were identified from the Kalkoenkrans, Basal and Ventersdorp Contact reefs (Fig. 4a) and only
one example of colloform pyrite was observed (Fig. 4b). Interestingly, Ketzer et al. (2003) observed similar
colloform pyrites in a paralic sandstone bed overlying coal layers. Hirdes and Saager (1983) and England et al.
(2002) suggested that banded pyrite was derived from sulfidic-banded iron formations and volcanogenic
massive sulfides, respectively. However, we prefer an early diagenetic origin due to the lack of massive sulfide
6
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
detritus in the auriferous conglomerates and the recognition of comparable early diagenetic pyrite textures in
shale, sandstone and conglomerate of different ages (Frizzo et al. 1991; Ketzer et al. 2003; Falconer et al. 2006).
Pyrite concretions, in contrast to pyrite nodules, incorporate the surrounding host rock material by filling in
the pore spaces between siliciclastic grains (Selles-Martinez 1996). Concretions that form during early
diagenesis are characterized by few point contacts between quartz grains, whereas those concretions that form
during the latter stages of diagenesis, replace the siliciclastic framework (Selles-Martinez 1996). In this study,
pyrite concretions were observed in a range of sedimentary rock types, which include sandstone, siltstone and
mudstone. Pyrite concretions that formed in sandstone were observed in the Main, Kimberley and Ventersdorp
Contact reefs (Fig. 4c). Previous authors have documented this type of pyrite in the proximal facies of several
Witwatersrand and Ventersdorp reefs (e.g., quartz-pyrite aggregates of Ramdohr 1958; Hirdes and Saager 1983;
Hallbauer 1986). Although Ramdohr (1958) advocated a replacement origin (i.e., of magnetite and hematite
quartzite), comparable pyrite and marcasite concretions in recently deposited sandstone and conglomerate
suggest a sedimentary origin (Ketzer et al. 2003; Falconer 2003). The bulk of the pyrite concretions in the
auriferous conglomerates are characterized by silt-sized quartz grain inclusions and not clay- or sand-sized
quartz grain inclusions (Fig. 4d, e). Mineral inclusions such as quartz, chlorite, muscovite, feldspar, rutile and
chromite infer a sedimentary protolith. Note that concretions are characterized by mineral inclusions and not
round/elliptical pores. (Hence, the frequent usage of the term ‘porous pyrite’ for concretionary pyrite is a
misnomer.) Interestingly, polygonal syneresis cracks were observed in one concretion from the Vaal Reef. The
cracks subsequently became filled with compact pyrite (Fig. 4f). Pyrite concretions with negligible rounding
reflect a local source, but not an in situ origin as suggested by England et al. (2002), since the mineralogy and
small size of the inclusions (fine-grained phyllosilicate minerals) is not consistent with their formation in quartzpebble conglomerates (Fig. 4e and Table 2).
Diagenetic euhedral pyrite crystals (DIA-6) are relatively uncommon in the auriferous conglomerates of the
Central Rand Group and Ventersdorp Supergroup. However, it is possible that many of the euhedral pyrite
crystals have been rounded to form of compact rounded pyrite grains (e.g., Guy et al. 2010). Only one example
was observed from the Carbon Leader Reef as part of a larger reworked composite grain (Fig. 4g; compare with
an in situ example in supplementary Fig. S1j).
Diagenetic pyrite overgrowths (DIA-11) are also uncommon in the auriferous conglomerates of the Central
Rand Group and Ventersdorp Supergroup. Nevertheless, they were observed as thin coatings (~10 µm) around
framboidal pyrite and small euhedral pyrite crystals in the Basal and Kalkoenkrans reefs (Fig. 3g). Diagenetic
pyrite overgrowths are generally much smaller than epigenetic overgrowths and occur in sediments that show
little evidence of metamorphic/hydrothermal fluid flow.
Replacive pyrite (DIA-12) was observed in a small number of sedimentary lithoclasts in the Kalkoenkrans
and Beatrix reefs (Fig. 4h, i, j). There are two types, namely variably-pyritized black shale clasts and
‘mineralized bacteria’. Interestingly, Tucker (1980) found that the former type was common in samples with
high gold grades. Historically, the term ‘mineralized bacteria’ has been applied to pyrite morphologies that
exhibit small microspherical structures (10-45 µm in size). These structures were interpreted to represent
pyritized bacterial colonies or the remains of primitive life forms (Ramdohr 1958; Schidlowski 1965; Saager
1970). However, following the work of numerous authors (e.g., Scott et al. 2009 and references therein), it has
become apparent that a large proportion of the ‘mineralized bacteria’ textures represent framboidal pyrite or
structures derived from framboidal pyrite (England and Ostwald 1993). Nevertheless, not all of these textures
represent pyrite framboids or atoll structures. In one sample from the Beatrix Reef, small 5 to 20 µm-sized
elliptical microstructures and wrinkled pyrite laminae were observed (Fig. 4h, i; Figs. 31 and 32 of Saager
1970). These features are consistent with the morphology of bacterial debris (Fig. 3A of Lott et al. 1999) and
may represent the pyritized remains of ancient organic material (Bubela and Cloud 1983; Schieber 2002). SEMEDS analyses of the minerals inside the elliptical structures indicate the presence of quartz, galena, chalcopyrite,
Fe-Mg-Mn carbonates, pyrophyllite and muscovite.
Epigenetic pyrite
The epigenetic pyrite association (EPI) consists of pyrite that formed after sedimentation and lithification,
during regional metamorphism and local hydrothermal events. There are five types of epigenetic pyrite:
euhedral crystals (EPI-1), overgrowths (EPI-2), veinlets (EPI-3), replacive pyrite (EPI-4) and pyrrhotite and
base metal sulfides (EPI-5) (Table 1). Although epigenetic pyrite is not the primary focus of the investigation,
some important observations are noted below.
7
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
Euhedral pyrite and pyrite overgrowths (EPI-1 and EPI-2) (both porous and compact) comprise the largest
proportion of epigenetic pyrite in the auriferous quartz-pebble conglomerates (Fig. 4k, l). The crystals are
generally coarse-grained and can reach several millimeters in size (Feather and Koen 1975). They commonly
occur in aggregates, and overgrow compact rounded pyrite grains and reworked sedimentary pyrite grains. They
are particularly abundant in reefs that contain an appreciable amount of reworked diagenetic pyrite (e.g., the
Basal and Ventersdorp Contact reefs). Hirdes and Saager (1983) and Hallbauer and von Gehlen (1983) observed
that epigenetic euhedral pyrite crystals in the Kimberley and Ventersdorp Contact reefs were commonly
associated with local heat sources, such as dykes and lavas of the Ventersdorp Supergroup. Similar occurrences
were reported from the Promise Formation in the West Rand Group, where epigenetic euhedral pyrite crystals
were closely associated with diabase sills and early diagenetic pyrite (Guy et al. 2010; see also Wagner and
Boyce 2006). Rounded fragments of putative epigenetic porous pyrite were observed in the Beatrix,
Kalkoenkrans, Elsburg, North and Vaal reefs (Fig. 4k), but also in the Basal Reef (Fig. 47 of Saager 1970) and
in the Coronation Formation, West Rand Group (Fig. 7I of Guy et al. 2010). In situ analogues of this type of
pyrite have been reported from a number of metamorphosed sedimentary successions (Large et al. 2009; Guy et
al. 2010; Thomas et al. 2011).
Pyrite veinlets (EPI-3) are relatively uncommon in comparison to euhedral pyrite crystals and pyrite
overgrowths. The vast majority of veinlet-associated pyrite occurs in micro-fractures within quartz and chert
pebbles (Fig. 4m). However, one sample consisting of brittle and granular pyrite was collected from a major
auriferous and carbonaceous vein from the Klerksdorp gold field (Cloete 2009).
Only a small number of replacement textures (EPI-4) were observed during this study (e.g., Reimer and
Mossman 1990; Fleet 1998). These include pyrite replacing quartz grains, shale fragments and Fe-Ti oxide
minerals from the Basal, Kalkoenkrans and Ventersdorp Contact reefs (Fig. 4n). In each case, however, the
replacement was partial and the precursor mineral was recognizable. Previous researchers have documented
pyrite and pyrrhotite replacing hematite, magnetite, titaniferous magnetite, ilmenite, rutile, phyllosilicates, chert,
quartz pebbles and other lithic rock fragments (Ramdohr 1958; Saager 1970; Hirdes and Saager 1983; Phillips et
al. 1987; Myers et al. 1993; Phillips and Dong 1994; Zhou et al. 1995; Barnicoat et al. 1997; Phillips and Law
2000). However, some of these examples probably reflect textural varieties of diagenetic pyrite (e.g., the quartzpyrite aggregates of Ramdohr 1958).
Epigenetic pyrrhotite and base metal sulfides (EPI-5) were rarely observed in this study (Fig. 4o). However,
the scarcity of pyrrhotite and base metal sulfides probably reflects the sampling bias towards unaltered reef
sections that contained well-preserved sedimentary pyrite grains. Zhou et al. (1995) reported that pyrrhotite was
common in the northern parts of the basin (West Rand and Carletonville gold fields) but less common in the
East Rand gold field and absent in the Evander and Welkom gold fields. In addition, Zhou et al. (1995) also
documented pyrrhotite in areas that were unaffected by contact metamorphism and suggested a regional
metamorphic influence. However, a close relationship between pyrrhotite and carbonaceous lithologies in the
Witwatersrand Supergroup (Tweedie 1968; Guy 2012) suggests that the pyrite to pyrrhotite transformation may
also be linked to the presence of organic carbon and is not only dependent on regional metamorphism (Ferry
1981; Thomas et al. 2011).
Fig. 3.
Fig. 4.
Multiple sulfur isotope data
The δ34S, Δ33S and Δ36S values of the entire sample set range from −6.8 to +13.8‰, −1.7 to +1.7‰ and −3.9 to
+0.9‰, respectively (Figs. 5, 6 and Table 2). Sulfur isotope compositions vary according to paragenetic
association, pyrite type and depositional environment.
Variation according to paragenetic association
Reworked syngenetic pyrite grains display negative δ34S values (−6.2 to −1.5‰) and small negative Δ33S
anomalies (average −0.3‰) (Fig. 5a). In contrast, the majority of reworked diagenetic pyrite grains display more
positive δ34S values (most between ~0 and +5‰) with variable Δ33S anomalies (−1.7 to +1.7‰) (Fig. 5a). Not
all diagenetic samples exhibit positive δ34S values, as a few samples are characterized by negative δ34S values
(down to −6.8‰). In addition, reworked diagenetic pyrite grains display much larger Δ36S offsets in comparison
8
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
to reworked syngenetic pyrite grains (relative to the Archean Δ36S/Δ33S array of −1 ± 0.2). Epigenetic pyrites,
which include both in situ samples and putative reworked fragments (Fig. 4k), are characterized by highly
attenuated δ34S, Δ33S, and Δ36S values (average +2.5, −0.1 and −0.1‰, respectively) (Figs. 5a and 6a).
Variation according to pyrite type
Connections between sulfur isotope composition and textural pyrite type are most pronounced for oolitic pyrite
(SYN-1), fine-grained, micro-porous pyrite (SYN-2), dendritic pyrite nodules (DIA-5), pyrite concretions with
sand-sized quartz inclusions (DIA-5) and reworked epigenetic porous pyrite aggregates (EPI-1) (Fig. 5b).
However, many other diagenetic pyrite types, such as small euhedral crystals (DIA-2) and concretions with finegrained mineral inclusions (DIA-5), lack appreciable systematic isotope variation (Fig. 5b).
Oolitic pyrite (SYN-1) yields negative δ34S excursions (down to −6.2‰) and small negative Δ33S anomalies
(average −0.4‰) (Fig. 6b, d, i). The single fine-grained, micro-porous pyrite sample (SYN-2) from the VCR
exhibits negative δ34S and negligible Δ33S values (−5.2 and −0.07‰, respectively) (Fig. 6e). Dendritic pyrite
nodules (DIA-5) are marked by small negative δ34S values (average −1.3‰) but relatively large negative Δ33S
anomalies (average −1.4‰). Pyrite concretions with sand-sized quartz grain inclusions (DIA-5) display the most
variable δ34S and Δ33S compositions (−6.6 to +13.8‰ and −0.6 to +1.7‰, respectively) (Fig. 6h). Dendritic
pyrite nodules and pyrite concretions with sand-sized quartz grain inclusions also exhibit some of the largest
offsets in Δ36S (~3‰). Reworked porous pyrite aggregates (EPI-1) display limited variations in δ34S and Δ33S,
and average +3.2 and −0.02‰, respectively.
Variation according to depositional environment
Excluding two outliers, reworked sedimentary pyrite grains from the fluvial-dominated Beatrix, Kalkoenkrans
and Ventersdorp Contact reefs display low magnitude Δ33S (±0.3‰) and variable δ34S values (−6.2 to +4.8‰)
(Fig. 5c). Conversely, pyrite grains from the marine-reworked conglomerates of the Main, Vaal and Basal reefs
are characterized by a wider range of Δ33S anomalies (−1.7 to +0.9‰) and a Δ33S/δ34S array that intercepts the
Δ33S and δ34S axes at ~−1.5 and +12‰, respectively (Fig. 5c).
Fig. 5.
Fig. 6.
Table 2.
Discussion
Detailed petrographic analysis of the buckshot pyrite from the gold-bearing conglomerates of the upper
Witwatersrand and lower Ventersdorp supergroups suggest that the majority of the large pyrite grains were
derived from a sedimentary source rock. In the discussion that follows, we expand on this premise and use the
multiple sulfur isotope data to provide further constraints on their origin.
Sedimentary provenance for buckshot pyrite
A number of source terranes have been proposed to account for the abundance and startling morphological
variety of the large rounded pyrite grains in the auriferous conglomerates of the Witwatersrand and Ventersdorp
supergroups. These include volcanogenic massive sulfide deposits, carbonate platforms, evaporites (Hutchinson
and Viljoen 1987; England et al. 2002) and sulfidic-banded iron formations (Hallbauer and von Gehlen 1983;
Hirdes and Saager 1983; Hutchinson and Viljoen 1987; Koglin et al. 2010). However, the requisite bulk rock
detritus from these terranes is conspicuously absent and evidence for pre-burial sulfidation is lacking, e.g.,
detrital clasts of partially pyritized carbonates were not observed (Myers et al. 1993). Similarly, it is also
unlikely that the large quantity of pyrite in the auriferous conglomerates was sourced from Paleoarchean granitegreenstone belts, as these typically contain little pyrite (Köppel and Saager 1974; Phillips and Powell 2011).
Rather, it may be suggested that the majority of the buckshot pyrite was derived from an intrabasinal
sedimentary setting, i.e., that they are of endoclastic origin and were not transported over very long distances
(Dimroth 1979; Hallbauer 1986; Large et al. 2013). In support of this interpretation, virtually every single one of
the pyrite morphologies observed in the Witwatersrand conglomerates possesses an in situ modern or ancient
textural equivalent from a sedimentary setting (Fig. 7). Additional lines of evidence for an intrabasinal
sedimentary source includes immature grain shapes, fragile internal textures, siliciclastic mineral inclusions,
9
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
presence of shale clasts and trace element compositions (Minter 1976; Dimroth 1979; Hallbauer and von Gehlen
1983; Hirdes and Saager 1983; Hallbauer 1986; Phillips and Law 2000; Agangi et al. submitted; Large et al.
2013). Possible source rocks of this affinity include sediments of the stratigraphically lower West Rand Group
and contemporaneous Central Rand Group.
Note that at least two types of rounded pyrite were probably derived from an extrabasinal source area. These
include rounded fragments of epigenetic porous euhedral pyrite and pyrite-bearing chert clasts. The former
probably reflects a metamorphic provenance and the latter, a greenstone belt provenance (Hirdes and Saager
1983; Phillips and Dong 1994).
Fig. 7.
Geological context of multiple sulfur isotope signatures
Paragenetic association
Links between pyrite paragenesis and multiple sulfur isotope compositions are most pronounced for sedimentary
rocks of Archean age – but less so for post-Archean sedimentary rocks that were deposited in more sulfate-rich
settings (Kamber and Whitehouse 2007; Partridge et al. 2008; Ono et al. 2009b; Farquhar et al. 2013). This
statement holds true for the reworked sedimentary pyrite grains in the Mesoarchean Witwatersrand and
Ventersdorp supergroups. More specifically, reworked syngenetic pyrite grains are characterized by sulfur
isotope values that plot to the left of the Archean Reference Array (ARA), whereas most of the sulfur isotope
values from reworked diagenetic pyrite grains plot to the right of the ARA (Fig. 5a). Ono et al. (2009a)
attributed this variation to reflect the openness of the system, where samples that plot to the left of the ARA
represent open system conditions (i.e., abundant sulfate supply) and those that plot to the right of the ARA
represent closed system conditions below the sediment-water interface (i.e., restricted sulfate supply).
Interestingly, very little data plots along the ARA in the Δ33S-δ34S plot (Fig. 5a), suggesting either microbial
reworking of the primary atmospheric signal (Johnston, 2011), changes in atmospheric photochemistry
(Farquhar et al. 2007, Domagal-Goldman et al. 2008), existence of non-photolytic sulfur reservoirs (Guy et al.
2012; Roerdink et al. 2012) or a combination thereof. Similarly, the bulk of the multiple sulfur isotope data in
the Δ36S-Δ33S plots below the ARA (Fig. 5a). Whereas syngenetic pyrite displays minor offsets, some diagenetic
pyrites are characterized by impressive Δ36S offsets of ~3‰. Deviation from this array can be explained by
overprinting of mass-dependent isotope effects (Δ36S/Δ33S of −6.8; Ueno et al. 2008; Ono et al. 2009b) and/or
isotope self-shielding due to increased volcanic SO2 loading (Δ36S/Δ33S of −4.6; Ono et al. 2013). Despite these
offsets, the Δ36S/Δ33S ratio of syngenetic pyrite (−1.0) is consistent with the main Archean S-MIF reaction (Fig.
5a) (Kaufman et al. 2007; Ono et al. 2009b; Ono et al. 2013).
Pyrite type
Five morphological types of pyrite reveal unique sulfur isotopic compositions (Fig. 5b). Syngenetic oolitic
pyrite (SYN-1) from the Ventersdorp Contact Reef yields relatively large negative δ34S values (−1.5 to −6.2‰)
and weakly negative Δ33S values (~−0.4‰) (Fig. 6b, d, i). In addition to relatively large negative δ34S values,
England et al. (2002) observed core-to-rim δ34S gradients in oolitic pyrite from the same reef (see their Fig. 15).
Similar δ34S gradients have been documented in modern oolitic pyrite laminae (Fig. 10 of Schieber and Riciputi
2005) and point to changes in pore water sulfate concentrations; a feature probably caused by microbial sulfate
reduction in an environment subject to short-lived high-energy pulses and sedimentary reworking. In support of
a microbial origin, Schieber and Riciputi (2005) found thin sheet-like structures on the cortices of modern pyrite
ooids that resembled organic biofilms. They suggested that the biofilms might have facilitated in the
precipitation of sulfide minerals by providing an organic substrate for sulfate-reducing bacteria and for acting as
a growth template.
Although only one example of syngenetic fine-grained, micro-porous pyrite (SYN-2) was observed, the δ34S
value of −5.2‰ is similar to that of oolitic pyrite (SYN-1) (Fig. 5b). Indeed, the unique textural characteristics
of the sample are suggestive of elevated sulfur concentrations that may have enabled microbial discrimination of
sulfur isotopes.
In contrast to the oolitic and micro-porous pyrite varieties, diagenetic dendritic pyrite nodules from the Basal
Reef (DIA-5) display small δ34S values (~−0.4‰) and relatively large negative Δ33S values (~−1.6‰), which
plot near the ARA (Fig. 5b). Dendritic nodules from the West Rand Group are also characterized by large
negative Δ33S values that plot near the ARA (−3.1 to −4.0‰, Guy et al. 2012). This implies that these
morphological types of pyrite are adept at acquiring large Δ33S fractionations with minimal mass-dependent
reworking (Ono et al. 2003; Ono et al. 2009a; Johnston 2011). Although the reasons for this relationship are
10
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
unclear, it may be related to rapid disequilibrium precipitation (as indicated by the dendritic growth textures) in
a closed system (e.g., Carstens 1985; Carstens 1986; Staude et al. 2007).
Diagenetic pyrite concretions (DIA-5) with sand-sized quartz grain inclusions reveal a wide range in sulfur
isotope compositions, both in relation to δ34S and Δ33S (−6.6 to +13.8‰ and −0.6 to +1.7‰, respectively) (Fig.
5b). These results contrast with the limited fractionation range of pyrite concretions with fine-grained mineral
inclusions. The variability of the sulfur isotope signatures could reflect the porous and permeable nature of the
host sandstone, i.e., high hydraulic conductivities and open system diagenetic conditions.
Three examples of reworked epigenetic porous pyrite aggregates (EPI-1) from the Beatrix and Kalkoenkrans
reefs yield remarkably uniform δ34S and Δ33S compositions (average +3.2 and −0.02‰, respectively) (Fig. 5b).
Based on the internal texture of the pyrite, the presence of silicate mineral inclusions and mass-dependent Δ33S
values, it could be speculated that these porous aggregates were derived from a source area consisting of
metamorphosed sedimentary rocks.
Depositional environment
Links between depositional facies and sulfur isotope compositions have contributed to a better understanding of
sulfur cycles in a broader depositional context (Kamber and Whitehouse 2007; Ono et al. 2009a; Guy et al.
2012). Similarly, in the auriferous conglomerates of the Witwatersrand and Ventersdorp supergroups, there are
tangible differences between the sulfur isotope ratios of pyrite from fluvial conglomerates and marine-modified
fluvial conglomerates (Fig. 5c).
Reworked sedimentary pyrite grains collected from fluvial conglomerates (i.e., the Kalkoenkrans, Beatrix
and Ventersdorp Contact reefs) generally yield small Δ33S values (±0.3‰) and variable δ34S values (−6.2 and
+4.8‰) (Fig. 5c). With respect to δ34S-Δ33S systematics, two major trends were observed. The first trend is
characterized by relatively large negative δ34S values (down to −6.2‰) and small negative Δ33S anomalies
(~−0.4‰). Such values are indicative of microbial sulfate reduction (Ono et al. 2003; Farquhar et al. 2013) and
probably reflect elevated concentrations of sulfate in abandoned fluvial channels, stagnant pools and ephemeral
ponds (Fig. 8a). The second trend is characterized by negligible Δ33S anomalies and δ34S values between ~0 and
+5‰. This range of δ34S values is typical for crustal (riverine) sulfate today and may signify reservoir effects
associated with microbial sulfate reduction (Turchyn et al. 2013) (solid line ‘MSR’, Fig. 8a). Alternatively, it
may represent mixing between two sulfur reservoirs, i.e., a juvenile volcanogenic reservoir and a putative crustal
mass-dependent reservoir (dashed line ‘V’ to ‘C’, Fig. 8a). Nevertheless, the δ34S composition of this reservoir
is comparable to the inferred mass-dependent reservoir observed from fluvial deposits in the middle and upper
West Rand Group (‘C’ in Fig. 8b) (Guy et al. 2012). This reservoir was initially interpreted to represent
continentally-derived sulfate due to the close association of dampened Δ33S anomalies with early diagenetic
pyrite, fluvial braidplain deposits and a foreland basin tectonic setting (Guy et al. 2012). Sulfate was probably
supplied to the sedimentary environment via oxidation of sulfide minerals (Reinhard et al. 2009; Lefticariu et al.
2010; Wacey et al. 2011; Stueeken et al. 2012) and/or degassing from subaerial volcanoes (Kump and Barley
2007; Gaillard et al. 2011; Binder and Keppler 2011; Philippot et al. 2012). Evidence for a terrestrial sulfur flux
can be found in other parts of the Kaapvaal and Zimbabwe cratons during the Mesoarchean and Neoarchean
eras. Reworked pyrite grains of sedimentary origin have been documented from the ~3.1 Ga Dominion Reef
(Rantzsch et al. 2011), the ~3.0 Ga Mozaan Contact Reef (Saager et al. 1986; Hofmann et al. 2009), the ~2.9 Ga
conglomerates of the Uitkyk Formation (Saager and Muff 1986; de Wit et al. 1993; Kröner et al. 2000), the ~2.7
Ga Ventersdorp Contact Reef (Hallbauer 1986; England et al. 2002), the ~2.7 Ga Belingwe greenstone belt
(Grassineau et al. 2001; Grassineau et al. 2002; Hofmann et al. 2009) and the ~2.6 Ga Black Reef (Barton and
Hallbauer 1996; Hofmann et al. 2009). In contrast, reworked sedimentary pyrite grains are scarce/absent in the
chert-dominated conglomerates of the ~3.5 to 3.2 Ga Barberton greenstone belt (Hessler and Lowe 2006; de Wit
et al. 2011). A similar narrative applies to the temporal distribution of chert and chert-pyrite clasts in Archean
conglomerates on the Kaapvaal Craton (Figs. 3.11C and 4.5D in Guy 2012; Plate 12 of Hirdes and Saager 1983;
Fig. 1 of Phillips and Dong 1994).
Fig. 8.
Reworked sedimentary pyrite grains from marine transgressive conglomerates (i.e., the Main, Vaal and
Basal reefs) typically yield larger Δ33S values in comparison to pyrite grains from fluvial conglomerates (Fig.
5c). With respect to δ34S-Δ33S systematics, three trends were observed. The first trend relates to the small Δ33S
anomalies that overlap with the sulfur isotope compositions of reworked sedimentary pyrite from fluvial
conglomerates. This trend probably reflects the incorporation of pyrite from underlying fluvial channels, i.e.,
pyrite with a crustal signature (Fig. 5c). The second trend is characterized by a positive Δ33S/δ34S array that
intercepts the Δ33S and δ34S axes at ~−1.5 and +12‰, respectively (dashed line ‘A’ to ‘M’, Fig. 8a). These endmembers have been observed in previous studies and were interpreted to represent two sulfate reservoirs,
11
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
namely a mass-independent reservoir derived from photochemical processes (‘A’ in Fig. 8a) (Farquhar et al.
2000, Ono et al. 2003; Ueno et al. 2008) and a mass-dependent reservoir derived from seafloor hydrothermal
systems (‘M’ in Fig. 8a) (Bao et al. 2007; Ueno et al. 2008). The occurrence of a similar mass-dependent sulfur
reservoir in the marine-dominated parts of the West Rand Group, Witwatersrand Supergroup (Fig. 8b; Guy et al.
2012) and the close compositional fit with Archean barite (Fig. 8c; Bao et al. 2007; Ueno et al. 2008; Roerdink
et al. 2012) suggests that this array represents a marine sulfate signature (‘M’) (Fig. 8a). If valid, the
preservation of this array is consistent with low levels of sulfate in the Archean ocean (Habicht et al. 2002; Ono
et al. 2009a; Jamieson et al. 2012). The third trend is related to the data points that plot to the left of the ARA in
the second quadrant (Fig. 8a). Previous studies have interpreted such data as evidence for S0 disproportionation
(Philippot et al. 2007) and oxidation of photochemical sulfur (Kaufman et al. 2007). Alternatively, the data
points may also point to mixing between polysulphides derived from an elemental sulfur precursor (Δ33S>0) and
hydrogen sulphide derived from microbial sulfate reducers (δ34S<0 and Δ33S<0) (e.g., Farquhar et al. 2013).
Conceptual models for sedimentary pyrite formation
Establishing the original depositional setting and source of buckshot pyrite in the auriferous conglomerates
has proven problematic since the pyrite segregations have been dislodged from their enclosing sediment. Indeed,
published accounts of in situ sedimentary pyrite in the auriferous conglomerates are exceedingly scarce (see
Tucker 1980 for two notable exceptions). One explanation could be that the sedimentary pyrite grains were
derived from the erosion and reworking of pyrite-bearing strata from the stratigraphically lower West Rand
Group. This could have happened during deposition of the Central Rand Group in a shrinking foreland basin
with strata of the West Rand Group being exposed along the tectonic margins of the basin. However, such an
explanation may only be partly applicable since there is a clear difference in sulfur isotope populations between
fluvial and marine-modified conglomerates (Fig. 5c). If the sedimentary buckshot pyrites were exclusively
derived from erosion of older strata, then the fluvial and marine-modified conglomerates should have displayed
similar sulfur isotope patterns. However, because we see only partial overlap of the data (Fig. 5c), it can be
concluded that a major proportion of the buckshot pyrites were formed and reworked during deposition of the
two aforementioned conglomerate facies, i.e., they are syndepositional and strictly endoclastic in origin.
Reworked sedimentary pyrite grains with a crustal sulfur isotope signature are interpreted to have formed in
a braided fluvial depositional environment, i.e., on the floor of abandoned fluvial channels, stagnant pools and
interchannel flats (Tucker 1980; Hallbauer and von Gehlen 1983). In this continental setting, Tucker (1980)
suggested that sedimentary pyrite would have formed during periods of quiescence in abandoned channels under
low-energy flow conditions. However, during short-lived high-energy pulses, the pyrite would be ripped up and
re-deposited. Reworked sedimentary pyrite grains with a marine sulfur isotope signature are interpreted to have
formed in sedimentary rocks associated with marine transgressions.
There are, however, many caveats to this model. For instance, it is important to realize that all major
auriferous conglomerates in the Witwatersrand Supergroup overlie laterally extensive and very flat erosional
unconformities. The conglomerates thus represent retrogradational braided fluvial complexes, sometimes in
combination with marine transgressions. It is envisaged that conditions for sedimentary pyrite formation were
most favorable when accommodation space in the basin started to increase following degradation and the
development of the erosional unconformities (Fig. 9). It is suggested that under such conditions, sedimentary
pyrite (carrying the crustal sulfur isotope signature) formed in flooded abandoned fluvial channels and on
interchannel flood plains, whereas sedimentary pyrite (carrying the marine sulfur isotope signature) formed in
paralic marine embayments (Tucker 1980; Hallbauer and von Gehlen 1983; Saager and Muff 1986; Eriksson et
al. 1998). As relative sea levels increased, these deposits were reworked by wave action to form the marinemodified conglomerates. (Fig. 9). This succession of events would help to explain the sheet-like architecture of
the major auriferous conglomerates, the pyrite-rich matrix and the deposition of thin quartz arenites overlying
many of the auriferous conglomerates (Verrezen, 1987; Catuneanu and Biddulph 2001). Note that the formation
of sedimentary pyrite extended into deeper water settings, as evidenced from the Booysens Formation in the
Evander gold field, where pyrite layers in carbonaceous shale can be traced for over 23 km (e.g., the upper and
lower pyrite markers) (Tweedie 1968; Bailey et al. 1990). The pyrite layers are individually composed of pyrite
seams, nodules and concretions that exhibit concentric and radiating growths of pyrite (Tweedie 1968; personal
observation). Erosion of such strata, via fluvial or marine processes, would also help to explain the textural
variations of pyrite in the auriferous conglomerates (see supplementary Fig. S1 for examples of pyrite textures
in black shale) and perhaps provide a source for some of the gold (see Fig. 7.3 of Guy 2012 for an example of a
gold-bearing diagenetic pyrite nodule in West Rand Group shale).
In contrast to the retrogradational fluvial conglomerates, it is informative to note that the thick successions of
progradational braided fluvial deposits and associated discontinuous lenticular conglomerates formed during
periods of decreasing accommodation space in the basin, prior to the development of low angle erosional
unconformities (Fig. 9). These progradational fluvial deposits are typically characterized by dispersed
12
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
concentrations of buckshot pyrite (e.g., Catuneanu and Biddulph 2001). This observation supports the
conclusion that sedimentary pyrite formed preferentially during periods of increased rates of relative sealevel
rise and drowning of peneplained land surfaces. However, this is not to say that sedimentary pyrite did form not
in progradational fluvial sediments. During degradation and unconformity development, such dispersed
sedimentary pyrite accumulations may have become reworked and concentrated in heavy mineral lags or
pediment deposits along unconformity surfaces.
Fig. 9.
Conclusion
Petrographic and multiple sulfur isotope analyses were carried out on buckshot pyrite grains from auriferous
conglomerates in the Witwatersrand and Ventersdorp supergroups to determine the origin of the pyrite sulfur.
Textural and geochemical constraints indicate that the majority of the buckshot pyrite grains were derived from
a sedimentary provenance and that the wide range of pyrite morphologies are characteristic of sedimentary
pyrite, and not multiple source areas (Dimroth 1979). Forty-five of these reworked sedimentary pyrite grains
were analyzed for their multiple sulfur isotope compositions. The δ34S, Δ33S and Δ36S values range from −6.8 to
+13.8‰, −1.7 to +1.7‰ and −3.9 to +0.9‰, respectively. Relationships between δ34S and Δ33S point to the
presence of three sulfur reservoirs: 1) mass-independent photochemical SO42- and S0 (data points on the ARA),
2) mass-dependent marine SO42- (positive Δ33S/δ34S array in the fourth quadrant) and 3) mass-dependent crustal
SO42- (δ34S = ~0 to +5‰, Δ33S = ~0‰). Of these three, the latter is the most prominent reservoir and implies a
significant contribution of continentally-derived sulfur for most of the pyrite grains in the auriferous quartz
pebble conglomerates. The emergence of a continentally-derived crustal sulfur reservoir may be linked to the
colonization of land by sulfide-oxidizing microbes (Wacey et al. 2011; Stueeken et al. 2012; Turchyn et al.
2013) and/or the advent of modern plate tectonic processes and oxidized volcanic gases (Chaussidon et al. 1989;
Ohmoto and Goldhaber 1997; Kump and Barley 2007; Smithies et al. 2007; Frimmel 2008; Binder and Keppler
2011; Gaillard et al. 2011; Schneiderhan et al. 2011; Philippot et al. 2012; Roerdink et al. 2012). In contrast, the
preservation of a marine sulfur isotope signature (Fig. 8a) is consistent with low levels of sulfate in the Archean
ocean (Habicht et al. 2002; Jamieson et al. 2012) since elevated concentrations of sulfate would have erased the
photochemical signature – producing a Δ33S trend resembling the modern sulfur cycle (Bekker et al. 2004; Guo
et al. 2009; Johnston 2011). The conclusion that most of the large rounded pyrite grains in the auriferous
conglomerates are of sedimentary and endoclastic origin has important ramifications regarding the cycling of
sulfur during the Archean, but also perhaps for the metallogenesis of gold in the conglomerates. More detailed
sulfur isotope studies, in the context of depositional systems tracts, are required to test the model presented here.
Acknowledgments
This work is supported by a research grant to NJB by the National Research Foundation, Pretoria (GUN
2053184) and a NASA Exobiology grant, NNX07AU12G to SO. Danie Thompson and Dawie Strydom of
Taung Gold, Louis Cloete of AngloGold Ashanti and particularly Hans Brouwer of Gold Fields (now Sibanye
Gold) are gratefully acknowledged for providing access to drill-core and underground mine samples. This
manuscript benefited greatly from discussions with Donna Falconer and Eva Stueeken. The editor and two
anonymous reviewers are thanked for their constructive comments on an earlier version of the manuscript.
References
Agangi A, Hofmann A, Wohlgemuth-Ueberwasser CC. Pyrite zoning as a record of mineralization in the
Ventersdorp Contact Reef, Witwatersrand Basin, South Africa. Economic Geology (submitted)
Armstrong RA, Compston W, Retief EA, Williams LS, Welke HJ (1991) Zircon ion microprobe studies bearing
on the age and evolution of the Witwatersrand basin. Precambrian Research 53:243–266
Bailey AC, Law JDM, Cadle AB, Phillips GN (1990) The Zandfontein Quartzite Formation, a marine deposit in
the Central Rand Group, Witwatersrand Supergroup. South African Journal of Geology 93:135–146
Bao H, Rumble D, Lowe DR (2007) The five stable isotope compositions of Fig Tree barites: implications on
sulfur cycle in ca 3.2 Ga oceans. Geochimica et Cosmochimica Acta 71:4868–4879
Barnicoat AC, Henderson IHC, Knipe RJ, Yardley BWD, Napier RW, Fox NPC, Kenyon AK, Muntingh DJ,
Strydom D, Winkler KS, Lawrence SR, Cornford C (1997) Hydrothermal gold mineralization in the
Witwatersrand basin. Nature 386:820–824
13
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
Barton ES, Hallbauer DK (1996) Trace-element and U-Pb isotope compositions of pyrite types in the
Proterozoic Black Reef, Transvaal Sequence, South Africa: Implications on genesis and age. Chemical
Geology 133:173–199
Bekker A, Holland HD, Wang PL, Rumble D, Stein HJ, Hannah JL, Coetzee LL, Beukes NJ (2004) Dating the
rise of atmospheric oxygen. Nature 427:117–120
Beukes NJ (1995) Stratigraphy and basin analysis of the West Rand Group with special reference to prospective
areas for placer gold deposits. (Unpubl. Report), Rand Afrikaans University (University of Johannesburg),
117p
Beukes NJ, Cairncross B (1991). A lithostratigraphic–sedimentological reference profile for the late Mozaan
Group, Pongola Sequence: application to sequence stratigraphy and correlation with the Witwatersrand
Supergroup. South African Journal of Geology 94: 44–69
Beukes NJ, Nelson JP (1995) Sea-level fluctuation and basin subsidence controls on the setting of auriferous
paleoplacers in the Archean Witwatersrand Supergroup: a genetic and sequence stratigraphic approach.
Extended Abstract, South African Geocongress, pp 860–863
Binder B, Keppler H (2011) The oxidation state of sulfur in magmatic fluids. Earth and Planetary Science
Letters 301:190–198
Bubela B, Cloud P (1983) Sulfide mineralization of microbial cells. Journal of Australian Geology and
Geophysics 8:355–357
Canfield DE, Raiswell R, Westrich JT, Reaves CM, Berner RA (1986) The use of chromium reduction in the
analysis of reduced inorganic sulfur in sediments and shales. Chemical Geology 54:149–155
Carstens H (1985) Early diagenetic cone-in-cone structures in pyrite concretions. Journal of Sedimentary
Research 55:105–108
Carstens H (1986) Displacive growth of authigenic pyrite. Journal of Sedimentary Research 56:252–257
Catuneanu O (2001) Flexural partitioning in the Late Archaean Witwatersrand foreland system, South Africa.
Sedimentary Geology 141–142:95–112
Catuneanu O, Biddulph MN (2001) Sequence stratigraphy of the Vaal Reef facies associations in the
Witwatersrand foredeep, South Africa. Sedimentary Geology 141–142:113–130
Chaussidon M, Albarède F, Sheppard SMF (1989) Sulphur isotope variations in the mantle from ion microprobe
analyses of micro-sulphide inclusions. Earth and Planetary Science Letters 92:144–156
Cloete LM (2009) Characterization of a recently discovered zone of intense hydrothermal alteration,
deformation and unusual Au mineralization at Anglogold Ashanti’s Kopanang gold mine. MSc Thesis
(Unpubl.), University of Johannesburg, South Africa, 91p
Coward MP, Spencer RM, Spencer CE (1995) Development of the Witwatersrand Basin, South Africa. In:
Coward MP, Ries AC (eds) Early Precambrian Processes. Geological Society, London, Special
Publications 95, pp 243–269
de Wit MJ, Armstrong RA, Kamo SL, Erlank AJ (1993) Gold-bearing sediments in the Pietersburg Greenstone
Belt: Age equivalents of the Witwatersrand Supergroup sediments, South Africa. Economic Geology
88:1242–1252
de Wit MJ, Furnes H, Robins B (2011) Geology and tectonostratigraphy of the Onverwacht Suite, Barberton
Greenstone Belt, South Africa. Precambrian Research 186:1–27
Dimroth E (1979) Significance of diagenesis for the origin of Witwatersrand-type uraniferous conglomerates.
Philosophical Transactions of the Royal Society of London. A291:277–287
Domagal-Goldman SD, Kasting JF, Johnston DT, Farquhar J (2008) Organic haze, glaciations and multiple
sulfur isotopes in the Mid-Archean Era. Earth and Planetary Science Letters 269: 29–40
England BM, Ostwald J (1993) Framboid-derived structures in some Tasman fold belt base-metal sulphide
deposits, New South Wales, Australia. Ore Geology Reviews 7:381–412
England GL, Rasmussen B, Krapez B, Groves DI (2002) Paleoenvironmental significance of rounded pyrite in
siliciclastic sequences of the Late Archaean Witwatersrand Basin: oxygen-deficient atmosphere or
hydrothermal evolution? Sedimentology 49:1133–1156
Eriksson PG, Condie, KC, Tirsgaard H, Mueller WU, Altermann W, Miall AD, Aspler LB, Catuneanu O,
Chiarenzelli JR (1998) Precambrian clastic sedimentation systems. Sedimentary Geology 120:5–53
Falconer DM (2003) Sediment-hosted gold and sulphide mineralization, Belle Brook, Southland, New Zealand.
MSc Thesis (Unpubl.), University of Otago, New Zealand, 373p
Falconer DM, Craw D, Youngson JH, Faure K (2006) Gold and sulphide minerals in Tertiary quartz pebble
conglomerate gold placers, Southland, New Zealand. Ore Geology Reviews 28:525–545
Farquhar J, Bao H, Thiemans M (2000) Atmospheric influence of Earth’s earliest sulfur cycle. Science
289:756–758
Farquhar J, Cliff J, Zerkle AL, Kamyshny A, Poulton SW, Claire M, Adams D, Harms B (2013) Pathways for
Neoarchean pyrite formation constrained by mass-independent sulfur isotopes. Proceedings of the National
Academy of Sciences USA. doi: 10.1073/pnas.1218851110
14
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
Farquhar J, Peters M, Johnston DT, Strauss H, Masterson A, Wiechert U, Kaufman AJ (2007) Isotopic evidence
for Mesoarchaean anoxia and changing atmospheric sulphur chemistry. Nature 449: 706–709
Feather CE, Koen GM (1975) The mineralogy of the Witwatersrand Reefs. Minerals Science and Engineering
7:189–224
Ferry JM (1981) Petrology of graphitic sulfide-rich schists from south-central Maine: An example of
desulfidation during prograde regional metamorphism. American Mineralogist 66:908–930
Fleet ME (1998) Detrital pyrite in Witwatersrand gold reefs: X-ray diffraction evidence and implications for
atmospheric evolution. Terra Nova 10:302–306
Frimmel HE (1997) Chlorite thermometry in the Witwatersrand basin: constraints on the Paleoproterozoic
geotherm in the Kaapvaal Craton, South Africa. The Journal of Geology 105:601–615
Frimmel HE (2005) Archaean atmospheric evolution: evidence from the Witwatersrand Gold Fields, South
Africa. Earth Science Reviews 70:1–46
Frimmel HE (2008) Earth’s continental crustal gold endowment. Earth and Planetary Science Letters 267:45–55
Frimmel HE, Groves DI, Kirk J, Ruiz J, Chesley J, Minter WEL (2005) The formation and preservation of the
Witwatersrand Gold Fields, the World's largest gold province. In: Hedenquist JW, Thompson JFH,
Goldfarb RJ, Richards JP (eds) Economic Geology 100th Anniversary Volume. Society of Economic
Geologists, Littleton, pp 769–797
Frizzo P, Rampazzo G, Molinaroli E (1991) Authigenic iron sulphides in recent sediments of the Venice Lagoon
(Northern Italy). European Journal of Mineralogy 3:603–612
Gaillard F, Scaillet B, Arndt NT (2011) Atmospheric oxygenation caused by a change in volcanic degassing
pressure. Nature 478:229–232
Genis J (1990) The Sedimentology and Depositional Environment of the Beatrix Reef, Witwatersrand
Supergroup. MSc Thesis (Unpubl.), University of the Witwatersrand, Johannesburg, South Africa
Gibson RL, Wallmach T (1995) Low pressure-high temperature metamorphism in the Vredefort Dome, South
Africa–anticlockwise pressure-temperature path followed by rapid decompression. Geological Journal
30:121–135
Grassineau NV, Nisbet EG, Bickle MJ, Fowler CMR, Lowry D, Mattey DP, Abell P, Martin A (2001) Antiquity
of the biological sulphur cycle: evidence from sulphur and carbon isotopes in 2700 million year old rocks
of the Belingwe belt, Zimbabwe. Proceedings of the Royal Society of London Series B: Biological
Sciences B268: 113–119
Grassineau NV, Nisbet EG, Fowler CMR, Bickle MJ, Lowry D, Chapman HJ, Mattey DP, Abell P, Yong J,
Martin A (2002) Stable isotopes in the Archaean Belingwe belt, Zimbabwe: Evidence for a diverse
microbial mat ecology. In: Fowler CMR, Ebinger CJ, Hawkesworth CJ (eds) The Early Earth: Physical,
Chemical and Biological Development. Geological Society, London, Special Publication 199, pp 309–328
Guo Q, Strauss H, Schröder S, Gutzmer J, Wing BA, Baker MA, Kaufman AJ, Kim ST, Farquhar J (2009)
Reconstructing Earth’s surface oxidation across the Archean-Proterozoic transition. Geology 37:399–402
Gutzmer J, Nhleko N, Beukes NJ, Pickard A, Barley ME (1999) Geochemistry and ion microprobe (SHRIMP)
age of a quartz porphyry sill in the Mozaan Group: geochronological implication for the Pongola and the
Witwatersrand Supergroups. South African Journal of Geology 102:139–146
Guy BM (2012) Pyrite in the Mesoarchean Witwatersrand Supergroup, South Africa. PhD Thesis (Unpubl.),
University of Johannesburg, South Africa, 492p
Guy BM, Beukes NJ, Gutzmer J (2010) Paleoenvironmental controls on the texture and chemical composition
of pyrite from non-conglomeratic sedimentary rocks of the Mesoarchean Witwatersrand Supergroup,
South Africa. South African Journal of Geology 113:195–228
Guy BM, Ono S, Gutzmer J, Kaufman AJ, Lin Y, Fogel ML, Beukes NJ (2012) A multiple sulfur and organic
carbon isotope record from non-conglomeratic sedimentary rocks of the Mesoarchean Witwatersrand
Supergroup, South Africa. Precambrian Research 216–219:208–231
Habicht KS, Gade M, Thamdrup B, Berg P, Canfield DE (2002) Calibration of sulfate levels in the Archean
ocean. Science 298:2372–2374
Hallbauer DK (1986) The mineralogy and geochemistry of Witwatersrand pyrite, gold, uranium and
carbonaceous matter. In: Anhaeusser CR, Maske S (eds) Mineral Deposits of Southern Africa. Geological
Society of South Africa, Johannesburg, pp 731–752
Hallbauer DK, von Gehlen K (1983) The Witwatersrand pyrites and metamorphism. Mineralogical Magazine
47:473–479
Hartzer FJ (2000). Geology of Transvaal inliers in the Bushveld Complex. Council for Geoscience 88: 222p.
Hattori K, Campbell FA, Krouse HR (1983) Sulphur isotope abundances in Aphebian clastic rocks; implications
for the coeval atmosphere. Nature 302:323–326
Hessler AM, Lowe DR (2006) Weathering and sediment generation in the Archean: An integrated study of the
evolution of siliciclastic sedimentary rocks of the 3.2 Ga Moodies Group, Barberton Greenstone Belt,
South Africa. Precambrian Research 151:185–210
15
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
Hirdes W (1979) The Proterozoic gold-uranium Kimberley Reef placers in the Evander and East Rand
Goldfields Witwatersrand, South Africa: different facies and their source area aspects. PhD Thesis
(Unpubl.), University of Heidelberg, Germany, 199p
Hirdes W, Saager R (1983) The Proterozoic Kimberley Reef placer in the Evander Goldfields, Witwatersrand,
South Africa. Monograph Series on Mineral Deposits 20, Gebrudewr Borntraeger, 101p
Hoefs J, Nielsen H, Schidlowski M (1968) Sulfur isotope abundances in pyrite from the Witwatersrand
Conglomerates. Economic Geology 63:975–977
Hofmann A, Bekker A, Rouxel O, Rumble D, Master S (2009) Multiple sulphur and iron isotope composition of
detrital pyrite in Archaean sedimentary rocks: a new tool for provenance analysis. Earth and Planetary
Science Letters 286:436–445
Hutchinson RW, Viljoen RP (1987) Re-evaluation of gold source in Witwatersrand ores. South African Journal
of Geology 91:157–173
Jamieson JW, Wing BA, Farquhar J, Hannington MD (2012) Neoarchaean seawater sulphate concentrations
from sulphur isotopes in massive sulphide ore. Nature Geoscience 6:61–64
Johnston DT (2011) Multiple sulfur isotopes and the evolution of Earth’s surface sulfur cycle. Earth Science
Reviews 106:161–183
Kamber BS, Whitehouse MJ (2007) Micro-scale sulphur isotope evidence for sulphur cycling in the late
Archean shallow ocean. Geobiology 5:5–17
Kaufman AJ, Johnston DT, Farquhar J, Masterson AL, Lyons TW, Bates S, Anbar AD, Arnold GL, Garvin J,
Buick R (2007) Late Archean Biospheric Oxygenation and Atmospheric Evolution. Science 317:1900–
1903
Ketzer JM, Holz M, Morad S, Al-Aasm IA (2003) Sequence stratigraphic distribution of diagenetic alterations
in coal-bearing paralic sandstones: evidence from the Rio Bonito Formation (early Permian) southern
Brazil. Sedimentology 50:855–877
Kirk J, Ruiz J, Chesley J, Titley S, Walshe J (2001) A detrital model for the origin of gold and sulfides in the
Witwatersrand basin based on Re-Os isotopes. Geochimica et Cosmochimica Acta 65:2149–2159
Koglin N, Frimmel HE, Minter WEL, Brätz H (2010) Trace-element characteristics of different pyrite types in
Mesoarchaean to Palaeoproterozoic placer deposits. Mineralium Deposita 45:259–280
Köppel VH, Saager R (1974) Lead isotope evidence on the detrital origin of Witwatersrand pyrites and its
bearing on the provenance of the Witwatersrand gold. Economic Geology 69:318–331
Kositcin N, Krapez B (2004) SHRIMP U-Pb detrital zircon geochronology of the Late Archaean Witwatersrand
Basin of South Africa: Relation between zircon provenance age spectra and basin evolution. Precambrian
Research 129:141–168
Kositcin N, McNaughton NJ, Griffin BJ, Fletcher IR, Groves DI, Rasmussen B (2003) Textural and
geochemical discrimination between xenotime of different origin in the Archaean Witwatersrand Basin,
South Africa. Geochimica et Cosmochimica Acta 67:709–731
Krapez B (1985) The Ventersdorp Contact Placer: a gold-pyrite placer of stream and debris-flow origins from
the Archaean Witwatersrand Basin of South Africa. Sedimentology 32:223–234
Kröner A, Jaeckel P, Brandl G (2000) Single zircon ages for felsic to intermediate rocks from the Pietersburg
and Giyani greenstone belts and bordering granitoid orthogneisses, northern Kaapvaal Craton, South
Africa. Journal of African Earth Sciences 30:773–793
Kump LR, Barley ME (2007) Increased subaerial volcanism and the rise of atmospheric oxygen 2.5 billion
years ago. Nature 448:1033–1036
Large RR, Danyushevsky L, Hollit C, Maslennikov V, Meffre S, Gilbert S, Bull SW, Scott R, Emsbo P, Thomas
H, Singh B, Foster J (2009) Gold and trace element zonation in pyrite using a laser imaging technique:
implications for the timing of gold in orogenic and Carlin-style sediment-hosted deposits. Economic
Geology 104:635–668
Large RR, Meffre S, Burnett R, Guy B, Bull S, Gilbert S, Goemann K, Danyushevsky L (2013) Evidence for an
intra-basinal source and multiple concentration processes in the formation of the Carbon Leader Reef,
Witwatersrand Supergroup, South Africa. Economic Geology 108: 1–28
Law JDM, Phillips GN (2005) Hydrothermal replacement model for Witwatersrand gold. In: Hedenquist JW,
Thompson JFH, Goldfarb RJ, Richards JP (eds) Economic Geology 100th Anniversary Volume. Society of
Economic Geologists, Littleton, pp 799–811
Lefticariu L, Pratt LA, LaVerne JA, Schimmelmann A (2010) Anoxic pyrite oxidation by water radiolysis
products – A potential source of biosustaining energy. Earth and Planetary Science Letters 292:57–67
Lott DA, Coveney RM, Murowchick JB, Grauch RI (1999) Sedimentary Exhalative Nickel-Molybdenum Ores
in South China. Economic Geology 94:1051–1066
McCarthy TS (1992) Syn-sedimentary deformation in the Witwatersrand basin. In: A short course reviewing
recent developments in the understanding of the Witwatersrand basin v 1. Economic Geology Research
Unit, University of the Witwatersrand, pp 85–94
16
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
McLean PJ, Fleet ME (1989) Detrital pyrite in the Witwatersrand gold fields of South Africa: evidence from
truncated growth banding. Economic Geology 84:2008–2011
Minter WEL (1972) Sedimentology of the Vaal Reef in the Klerksdorp Area. PhD Thesis (Unpubl.), University
of the Witwatersrand, South Africa, 112p
Minter WEL (1976) Detrital Gold, Uranium and Pyrite Concentrations Related to Sedimentology in the
Precambrian Vaal Reef Placer, Witwatersrand, South Africa. Economic Geology 71:157–176
Minter WEL, Feather CE, Glatthar CW (1988) Sedimentological and Mineralogical Aspects of the Newly
Discovered Witwatersrand Placer Deposit That Reflect Proterozoic Weathering, Welkom Gold Field,
South Africa. Economic Geology 83:481–491
Myers RE, McCarthy TS, Stanistreet IG (1990) A tectono-sedimentary reconstruction of the development and
evolution of the Witwatersrand basin with particular emphasis on the Central Rand Group. South African
Journal of Geology 93:180–201
Myers RE, Zhou T, Phillips GN (1993) Sulphidation in the Witwatersrand Goldfields: evidence from the
Middelvlei Reef. Mineralogical Magazine 57:395–405
Ohfuji H, Rickard D (2005) Experimental syntheses of framboids–a review. Earth Science Reviews 71:147–170
Ohmoto H, Goldhaber MB (1997) Sulfur and Carbon Isotopes. In: Barnes HL (ed) Geochemistry of
Hydrothermal Ore Deposits. John Wiley and Sons, New York, pp 571–611
Ono S, Beukes NJ, Rumble D (2009b) Origin of two distinct multiple-sulfur isotope compositions of pyrite in
the 2.5 Ga Klein Naute Formation, Griqualand West Basin, South Africa. Precambrian Research 169:48–
57
Ono S, Eigenbrode JL, Pavlov AA, Kharecha P, Rumble D, Kasting JF, Freeman KH (2003) New insights into
Archean sulfur cycle from mass-independent sulfur isotope records from the Hamersley Basin, Australia.
Earth and Planetary Science Letters 213:15–30
Ono S, Kaufman AJ, Farquhar J, Sumner DY, Beukes NJ (2009a) Lithofacies control on multiple-sulfur isotope
records and the Neoarchean sulfur cycles. Precambrian Research 169:58–67
Ono S, Wing B, Johnston DT, Farquhar J, Rumble D (2006) Mass-dependent fractionation of quadruple sulfur
isotope system as a new tracer of sulfur biogeochemical cycles. Geochimica et Cosmochimica Acta
70:2238–2252
Ono S, Whitehill AR, Lyons JR (2013) Contribution of isotopologue self-shielding to sulfur mass-independent
fraction during sulfur dioxide photolysis. Journal of Geophysical Research: Atmospheres 118:1–11
Palmer JA (1986) Paleoweathering in the Witwatersrand and Ventersdorp Supergroups. MSc Thesis (Unpubl.),
University of the Witwatersrand, South Africa, 166p
Partridge MA, Golding SD, Baublys KA, Young E (2008) Pyrite paragenesis and multiple sulfur isotope
distribution in late Archean and early Paleoproterozoic Hamersley Basin sediments. Earth and Planetary
Science Letters 272:41–49
Philippot P, Van Zuilen M, Lepot K, Thomazo C, Farquhar J, Van Kranendonk MJ (2007) Early Archaean
microorganisms preferred elemental sulphur, not sulfate. Science 317:1534–1537
Philippot P, Van Zuilen M, Rollion-Bard C (2012) Variations in atmospheric sulphur chemistry on early Earth
linked to volcanic activity. Nature Geoscience 5:668–674
Phillips GN, Dong G (1994) Chert-plus-pyrite-pebbles in the Witwatersrand Goldfields. International Geology
Review 36:65–71
Phillips GN, Law JDM (2000) Witwatersrand gold fields: Geology, genesis and exploration. Society of
Economic Geologists (SEG) Reviews 13:439–500
Phillips GN, Myers RE (1989) Witwatersrand gold fields: Part II. An origin for Witwatersrand gold during
metamorphism and associated alteration. Economic Geology Monograph 6:598–608
Phillips GN, Myers RE, Palmer JA (1987) Problems with the placer model for Witwatersrand gold. Geology
15:1027–1030
Phillips GN, Powell R (2011) Origin of Witwatersrand gold: a metamorphic devolatilization-hydrothermal
replacement model. Applied Earth Science 120:112–129
Poujol M, Robb LJ, Anhaeusser CR, Gericke B (2003) A review of the geochronological constraints on the
evolution of the Kaapvaal Craton, South Africa. Precambrian Research 127:181–213
Ramdohr P (1958) New observations on the ores of the Witwatersrand in South Africa and their genetic
significance. Transactions of the Geological Society of South Africa 61:1–50 (Annexure)
Rantzsch U, Gauert CDK, Van der Westhuizen WA, Duhamel I, Cuney M, Beukes GJ (2011) Mineral chemical
study of U-bearing minerals from the Dominion Reefs, South Africa. Mineralium Deposita 46:187–196
Reimer TO, Mossman DJ (1990) Sulphidation of Witwatersrand black sands: from enigma to myth. Geology
18:426–429
Reimold WU, Przybylowicz WJ, Gibson RL (2004) Quantitative major and trace elemental mapping by PIXE of
concretionary pyrite from the Witwatersrand Basin, South Africa. X-Ray Spectrometry 33:189–203
17
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
1069
1070
Reinhard CT, Raiswell R, Scott C, Anbar AD, Lyons TW (2009) A late Archean sulfidic sea stimulated by early
oxidative weathering of the continents. Science 326:713–716
Robb LJ, Davis D, Kamo SL, Meyer FM (1992) Ages of altered granites adjoining the Witwatersrand basin with
implications for the origin of gold and uranium. Nature 357:677–680
Roerdink DL, Mason PRD, Farquhar J, Reimer T (2012) Multiple sulfur isotopes in Paleoarchean barites
identify an important role for microbial sulfate reduction in the early marine environment. Earth and
Planetary Science Letters 331–332:177–186
Ruppert LF, Hower JC, Eble CF (2005) Arsenic-bearing pyrite and marcasite in the Fire Clay coal bed, middle
Pennsylvanian Breathitt Formation, eastern Kentucky. International Journal of Coal Geology 63:27–35
Saager R (1970) Structures in pyrite from the Basal Reef in the Orange Free State Goldfield. Transactions of the
Geological Society of South Africa 73:29–46
Saager R, Muff R (1986) The auriferous placer at Mount Robert, Pietersburg Greenstone Belt. In: Anhaeusser
CR, Maske S (eds) Mineral Deposits of Southern Africa. Geological Society of South Africa,
Johannesburg, 213–220
Saager R, Stupp HD, Utter T, Matthey HO (1986) Geological and mineralogical notes on placer occurrences in
some conglomerates of the Pongola sequence. In: Anhaeusser CR, Maske S (eds) Mineral Deposits of
Southern Africa. Geological Society of South Africa, Johannesburg, pp 473–487
SACS (South African Committee for Stratigraphy) (2006) A revised stratigraphic framework for the
Witwatersrand Supergroup. Council of Geoscience Lithostratigraphic Series 42:1–7
Schidlowski M (1965) Probable life forms from the Precambrian of the Witwatersrand System (South Africa).
Nature 205:895–896
Schieber J (2002) Sedimentary pyrite: A window into the microbial past. Geology 30:531–534
Schieber J, Riciputi L (2005) Pyrite-marcasite coated grains in the Ordovician Winnipeg Formation Canada: an
intertwined record of surface conditions, stratigraphic condensation, geochemical ‘‘reworking,’’ and
microbial activity. Journal of Sedimentary Research 75:905–918
Schmitz MD, Bowring SA, de Wit MJ, Gartz V (2004) Subduction and terrane collision stabilize the western
Kaapvaal craton tectosphere 2.9 billion years ago. Earth and Planetary Science Letters 222:363–376
Schneiderhan E, Zimmerman U, Gutzmer J, Mezger K, Armstrong R (2011) Sedimentary Provenance of the
Neoarchean Ventersdorp Supergroup, South Africa: Shedding Light on the Evolution of the Kaapvaal
Craton during the Neoarchean. The Journal of Geology 119:575–596
Scott RJ, Meffre S, Woodhead J, Gilbert SE, Berry RF, Emsbo P (2009) Development of framboidal pyrite
during diagenesis, low-grade regional metamorphism, and hydrothermal alteration. Economic Geology
104:1143–1168
Selles-Martinez J (1996) Concretion morphology, classification and genesis. Earth Science Reviews 41:177–210
Smithies RH, van Kranendonk MJ, Champion DC (2007) The Mesoarchean emergence of modern subduction.
Gondwana Research 11:50–68
Staude S, Wagner T, Markl G (2007) Mineralogy, mineral compositions and fluid evolution at the Wenzel
Hydrothermal Deposit, Southern Germany: Implications for the formation of the Kongsberg-type silver
deposits. The Canadian Mineralogist 45:1147–1176
Strauss H, Beukes NJ (1991) A geochemical study of carbon and sulfur in sedimentary rocks from the
Witwatersrand and Ventersdorp Supergroups and its bearing on the depositional environment. (Unpubl.
Progress Report), Rand Afrikaans University (University of Johannesburg), 37p
Stueeken EE, Catling DC, Buick R (2012) Contributions to late Archaean sulphur cycling by life on land.
Nature Geoscience 5:722–725
Tankard AJ, Jackson MPA, Eriksson KA, Hobday DK, Hunter DR, Minter WEL (1982) Crustal Evolution of
Southern Africa, 3.8 Billion Years of Earth History. Springer-Verlag, New York, 523p
Taylor KG and Macquaker JHS (2000) Early diagenetic pyrite morphology in a mudstone-dominated
succession: the Lower Jurassic Cleveland Ironstone Formation, eastern England. Sedimentary Geology
131: 77–86
Thomas HV, Large RR, Bull SW, Maslennikov V, Berry RF, Fraser R, Froud S, Moye R (2011) Pyrite and
pyrrhotite textures and composition in sediments, laminated quartz veins, and reefs at Bendigo Gold Mine
Australia: insights for ore genesis. Economic Geology 106:1–31
Tucker RF (1980) The sedimentology and mineralogy of the Composite Reef on Cooke Section, Randfontein
Estates Gold Mine, Witwatersrand, South Africa. MSc Thesis (Unpubl.), University of the Witwatersrand,
South Africa, 355p
Turchyn AV, Tipper ET, Galy A, Lo JK, Bickle MJ (2013) Isotope evidence for secondary sulfide precipitation
along the Marsyandi River, Nepal, Himalayas. Earth and Planetary Science Letters 374: 36–46
Tweedie KAM (1968) The stratigraphy and sedimentary structures of the Kimberley Shales in the Evander gold
field, Eastern Transvaal, South Africa. Transactions of the Geological Society of South Africa 71:235–256
18
1071
1072
1073
1074
1075
1076
1077
1078
1079
1080
1081
1082
1083
1084
1085
1086
1087
1088
1089
1090
1091
1092
1093
1094
1095
1096
1097
1098
1099
1100
1101
1102
1103
1104
1105
1106
1107
1108
1109
1110
1111
1112
Ueno Y, Ono S, Rumble D, Maruyama S (2008) Quadruple sulfur isotope analysis of ca 3.5 Ga Dresser
Formation: new evidence for microbial sulfate reduction in the Early Archean. Geochimica et
Cosmochimica Acta 72:5675–5691
Utter T (1978) Morphology and geochemistry of different pyrite types from the Upper Witwatersrand System of
the Klerksdorp Goldfield, South Africa. Geologische Rundschau 67:774–804
Verrezen J (1987) Sedimentology of the Vaal Reef paleoplacer in the western portion of Vaal Reefs Mine. MSc
thesis (Unpubl.), Rand Afrikaans University (University of Johannesburg), South Africa, 194p
Wacey D, Saunders M, Brasier MD, Kilburn MR (2011) Earliest microbially mediated pyrite oxidation in ~3.4
billion-year-old sediments. Earth and Planetary Science Letters 301:393–402
Wagner T, Boyce AJ (2006) Pyrite metamorphism in the Devonian Hunsrück Slate of Germany: Insights from
laser microprobe sulfur isotope analysis and thermodynamic modeling. American Journal of Science
306:525–552
Walraven F, Grobler DF, Key RM (1996) Age equivalence of the Plantation Porphyry and the Kanye Volcanic
Formation, southeastern Botswana. South African Journal of Geology 99: 23–31
Watchorn MB (1981) The stratigraphy and sedimentology of the West Rand Basin in the Western Transvaal.
PhD Thesis (Unpubl.), University of the Witwatersrand, South Africa, 154p
Watchorn MB, O’Brien MF (1991) The significance of marine modification in some Witwatersrand placers – an
example from the Lower Witwatersrand, West Rand Group. South African Journal of Geology 94:333–
339
Wiese RG, Fyfe WS (1986) Occurrence of iron sulfides in Ohio coals. International Journal of Coal Geology
6:251–276
Wilkin RT, Arthur MA, Dean WE (1997) History of water-column anoxia in the Black Sea indicated by pyrite
framboid size distributions. Earth and Planetary Science Letters 148:517–525
Wilkin RT, Barnes HL, Brantley SL (1996) The size distribution of framboidal pyrite in modern sediments: an
indicator of redox conditions. Geochimica et Cosmochimica Acta 60:3897–3912
Woodland BG (1975) Pyritic cone-in-cone concretions Fieldiana, Geology 33:125–139
Zhao B (1998) A mineralogical and geochemical study of alteration associated with the Ventersdorp Contact
Reef in the Witwatersrand Basin. PhD Thesis (Unpubl.), University of the Witwatersrand, South Africa,
297p
Zhou T, Phillips GN, Dong G, Myers RE (1995) Pyrrhotite in the Witwatersrand Gold Fields, South Africa.
Economic Geology 90:2361–2369
1113
1114
1115
1116
1117
1118
1119
1120
1121
1122
1123
1124
1125
1126
1127
1128
1129
Fig. 2. Simplified stratigraphic column of the upper Witwatersrand Supergroup and sampled reefs. Column 1:
Groups. Column 2: Subgroups. Column 3: Formations. Column 4: Genetic successions after Beukes and
Nelson (1995). Column 5: Lithology. Column 6: Member/bed/reef. Sampled reefs for sulfur isotope analyses in
bold. Column 7: SHRIMP U-Pb or Pb-Pb ages after Frimmel et al. (2005), where (1) zircon (Armstrong et al.
1991), (2) zircon and xenotime (Kositcin and Krapez 2004), (3) authigenic xenotime (Kositcin et al. 2003) and
(4) zircon (Gutzmer et al. 1999).
Table 1. Paragenetic and morphological classification of pyrite in the auriferous conglomerates of the
Witwatersrand and Ventersdorp supergroups.
Table 2. Multiple sulfur isotope data of reworked sedimentary pyrite grains from the Witwatersrand and
Ventersdorp supergroups.
Fig. 1. a Geographic positions of the Witwatersrand and Pongola supergroups on the Archean Kaapvaal Craton.
b Geological map of the Witwatersrand Supergroup, illustrating major gold fields, paleoslopes, major faults and
the lower West Rand and upper Central Rand groups.
Fig. 3. SEM-BSE and reflected light (RF) photomicrographs of detrital pyrite and reworked syngenetic and
diagenetic pyrite. Detrital pyrite. a Compact rounded pyrite grains (DET-1) (see arrows). Reworked
syngenetic pyrite. b Oolitic or concentrically laminated pyrite (SYN-1) with a radiating pyrite core. c Oolitic
pyrite (SYN-1) with an abraded rim. d Oolitic pyrite laminae (SYN-1). Bright mineral is galena. e, f Finegrained, micro-porous pyrite (SYN-2). Note scale in f. Reworked diagenetic pyrite. g Framboidal pyrite (DIA1) with compact pyrite overgrowths (DIA-11). h Putative atoll structures after framboidal pyrite (DIA-1). i
Small euhedral crystals (DIA-2) in a shale clast. j Dendritic pyrite nodule (DIA-5) (cone-in-cone). k Branching
dendritic pyrite nodule (DIA-5). l Bladed, radiating pyrite nodule (DIA-5). m Bladed pyrite rim on concretion
(DIA-5). n Pyrite pseudomorphs after marcasite spheroids? (DIA-5). o Abiotic dissolution/microbial tunneling.
Refer to Table 2 for numbers 1 to 7 (top left corner of selected photomicrographs).
19
1130
1131
1132
1133
1134
1135
1136
1137
1138
1139
1140
1141
1142
1143
1144
1145
1146
1147
1148
1149
1150
1151
1152
1153
1154
1155
1156
1157
1158
1159
1160
1161
1162
1163
1164
1165
1166
1167
1168
1169
1170
1171
1172
1173
1174
1175
1176
1177
1178
1179
1180
1181
1182
1183
1184
Fig. 4. SEM-BSE and reflected light (RF) photomicrographs of reworked diagenetic pyrite and epigenetic
pyrite. Reworked diagenetic pyrite. a Banded pyrite nodule (DIA-5). b Colloform banded pyrite nodule (DIA5). Note truncation. c Quartz-pyrite concretion (DIA-5). Note compact cement and sand-sized quartz grains. d
Rounded pyrite concretion (DIA-5). e Fine-grained pyrite concretion (DIA-5). Note lack of rounding. Vertical
line is polishing scratch. f Polygonal syneresis cracks in a rounded pyrite concretion (DIA-5). g Zoned euhedral
pyrite crystals. h, i ‘Mineralized bacteria’ (DIA-12). Note wavy contortions of pyrite laminae. j Pyritized shale
clast (DIA-12). Epigenetic pyrite. k Detrital fragment of porous euhedral pyrite (EPI-1, 2). l Compact euhedral
pyrite crystals (EPI-1) and overgrowths (EPI-2). m Pyrite veinlet in quartz pebble (EPI-3). n Partial pyrite
replacement of a quartz grain (EPI-4). o Pyrrhotite enveloping a compact rounded pyrite grain (EPI-5). Refer to
Table 2 for numbers 8 to 13 (top left corner of selected photomicrographs).
Fig. 5. Plots of Δ33S vs. δ34S (‰) and Δ36S vs. Δ33S (‰) for a paragenetic association, b pyrite type and c
depositional environment.
Fig. 6. Hand specimen (4.7 cm in length) from the Ventersdorp Contact Reef. a Pyrite veinlet (EPI-3). b Oolitic
pyrite or concentrically laminated pyrite (SYN-1). c Fine-grained pyrite concretion (DIA-5). d Oolitic pyrite or
concentrically laminated pyrite (SYN-1). e Fine-grained, micro-porous pyrite (SYN-2). f Pyrite concretion
(DIA-5). g Putative replacive pyrite (DIA-12/EPI-4). h Quartz-pyrite concretion. i Oolitic pyrite (SYN-1) with
core of radiating pyrite. Values in permil (‰). Scale bar corresponds to 500 µm.
Fig. 7. Schematic representation of the pyrite types identified in the auriferous conglomerates (left) and
examples of their in situ textural equivalents (right). References: (1) Guy et al. (2010). (2) Black shale specimen
from the Transvaal Supergroup. (3) Carstens (1985, 1986). (4) Schieber and Riciputi (2005). (5) Ketzer et al.
(2003). (6) Falconer et al. (2006). (7) Frizzo et al. (1991). (8) Woodland (1975) (9) Wiese and Fyfe (1986). (10)
Large et al. (2009). (11) Ruppert et al. (2005).
Fig. 8. Plots of Δ33S vs. δ34S (‰) for a conglomerates of the Central Rand Group and b non-conglomeratic
sedimentary rocks from the West Rand Group (Guy et al. 2012). Dashed lines indicate reservoir mixing,
whereas solid lines indicate microbial processing (Johnston 2011). c Comparative Δ33S-δ34S data sets of barite
(Bao et al. 2007; Ueno et al. 2008; Roerdink et al. 2012) and reworked sedimentary pyrite from the
Witwatersrand and Ventersdorp supergroups (Hofmann et al. 2009).
Fig. 9. Conceptual models for sedimentary pyrite formation in the upper Witwatersrand and lower Ventersdorp
supergroups. a Highstand systems tract characterized by high accommodation space, rapid sedimentation and
progradation of braided fluvial systems. Dispersed sedimentary pyrite may have formed in the fluvial and
marine depositional environments. b Late highstand systems tract characterized by increasing degradation and
sedimentary bypassing. Sedimentary pyrite with a crustal signature may have formed in abandoned channels and
ephemeral ponds. c These sedimentary pyrite grains would have been concentrated into heavy mineral lags
during the development of low angle unconformities (e.g., Beatrix, Kalkoenkrans and Ventersdorp Contact
reefs). d Sedimentary pyrite formation (both with crustal and marine signatures) would have followed the
development of the unconformities due to very flat land surfaces and flooding of river channels and paralic
marine embayments. e These pyrites were then reworked into marine-modified conglomerates during marine
transgressions (e.g., Main, Basal and Vaal reefs).
Fig. S1. Variations in pyrite texture in black shale (Transvaal Supergroup). a Radiating pyrite spheroids (DIA5). Note framboid core (DIA-1). b Putative atoll structures (DIA-1). c Spheroidal pyrite nodules (DIA-5).
Compare with Fig. 3n. d Bladed, radiating pyrite (DIA-5). e Banded pyrite (DIA-5). f Fine-grained radiating
pyrite (DIA-5). g, h Pyrite concretion (DIA-5). Note differential pyritization. i, j Euhedral pyrite crystals (DIA6). Note zoning in j and compare zoning with Fig. 4g.
20
1
2
3
4
5
6
7
8
9
S1
Download