Accepted Manuscript Petrography, sulfide mineral chemistry, and sulfur isotope evidence for a hydrothermal imprint on Musina copper deposits, Limpopo Province, South Africa: Evidence for a breccia pipe origin? Jeff B. Chaumba, Humbulani R. Mundalamo, Jason S. Ogola, J.A. Cox, C.J. Fleisher PII:
S1464-343X(16)30149-2
DOI:
10.1016/j.jafrearsci.2016.05.003
Reference:
AES 2562
To appear in:
Journal of African Earth Sciences
Received Date: 19 February 2016 Revised Date:
5 May 2016
Accepted Date: 6 May 2016
Please cite this article as: Chaumba, J.B., Mundalamo, H.R., Ogola, J.S., Cox, J.A., Fleisher, C.J., Petrography, sulfide mineral chemistry, and sulfur isotope evidence for a hydrothermal imprint on Musina copper deposits, Limpopo Province, South Africa: Evidence for a breccia pipe origin?, Journal of African Earth Sciences (2016), doi: 10.1016/j.jafrearsci.2016.05.003. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT 1 1
Petrography, sulfide mineral chemistry, and sulfur isotope evidence for a hydrothermal
2
imprint on Musina copper deposits, Limpopo Province, South Africa: Evidence for a
3
breccia pipe origin?
RI PT
4 Jeff B. Chaumba
6
Department of Geology & Geography, University of North Carolina at Pembroke,
7
Pembroke, NC 28372, USA,
[email protected]
8
Humbulani R. Mundalamo, Jason S. Ogola
9
Department of Mining and Environmental Geology, University of Venda, Private Bag
10
X5050, Thohoyandou, Limpopo Province, 0950, South Africa
11
J.A. Cox and C.J. Fleisher
12
Department of Geology, University of Georgia, Athens, GA 30602, USA.
M AN U
SC
5
14
ABSTRACT
15
TE D
13
The Musina copper deposits are located in the Central Zone of the Limpopo orogenic belt
17
in Limpopo Province, South Africa. We carried out a petrographic, sulfide composition,
18
and δ34S study on samples from Artonvilla and Campbell copper deposits and a country
19
rock granitic gneiss to Artonvilla Mine to place some constrains on the origin of these
20
deposits.
AC C
21
EP
16
The assemblages at both Artonvilla and Campbell Mines of brecciated quartz,
22
potassium feldspar, muscovite, chlorite, calcite, and amphibole are consistent with
23
sericitic alteration. Quartz, amphibole, feldspars, and micas often display angular textures
ACCEPTED MANUSCRIPT 2 which are consistent with breccias. Sulfur concentrations in pyrite from Artonvilla Mine
25
plot in a narrow range, from 50.2 wt. % to 55.7 wt. %. With the exception of a positive
26
correlation between Fe and Cu, no well defined correlations are shown by data from the
27
Musina copper deposits. The occurrence of sulfides both as inclusions in, or as interstitial
28
phases in silicates, suggests that hydrothermal alteration that affected these deposits most
29
likely helped concentrate the mineralization at the Musina copper deposits. Sulfur
30
concentrations in chalcopyrite samples investigated vary widely whereas the copper
31
concentrations in chalcopyrite are not unusually higher compared to those from
32
chalcopyrite from other tectonic settings, probably indicating that either the Cu in the
33
Musina copper deposits occurs in native form, and/or that it is hosted by other phases.
34
This observation lends support to the Cu having been concentrated during a later
35
hydrothermal event.
SC
M AN U
One sample from Artonvilla Mine (AtCal01) yielded pyrite δ34S values of 3.1and
TE D
36
RI PT
24
3.6 ‰ and chalcopyrite from the same sample yielded a value of 3.9 ‰. A country rock
38
granitic gneiss to Artonvilla Mine yielded a δ34Spyrite value of 8.2 ‰. For Campbell Mine
39
samples, one quartz vein sample has a δ34Spyrite value of 0.5‰ whereas chalcopyrite
40
samples drilled from different areas within the same sample yielded values of 0.4 and 0.7
41
‰. The same sample also yielded a δ34Sbornite value of 0.4‰. Another Campbell Mine
42
quartz vein sample yielded a chalcopyrite δ34S value of -0.3 ‰. Sulfur isotope
43
thermometry for one Campbell Mine quartz vein sample with coexisting sulfides yielded
44
a ∆34Schalcopyrite-bornite value of 359oC that is consistent with the stability of this mineral
45
pair. Thus, δ34S values from Campbell Mine are consistent with an igneous source for the
46
sulfur. Based on a simple two-end member isotope mixing model, contamination of the
AC C
EP
37
ACCEPTED MANUSCRIPT 3 sulfur by sulfur derived from granitic country rocks likely occurred at Artonvilla Mine.
48
Based on findings from this study and by other previous investigators, it is concluded that
49
features displayed by the Musina copper deposits are consistent with a breccia pipe origin
50
for the Musina copper deposits.
RI PT
47
51 1. Introduction
53
Although a breccia pipe origin has not yet been suggested as a source for the Musina
54
copper deposits, results presented here warrant a re-interpretation of the origin of these
55
deposits and we hypothesize that the copper deposits at Musina originated as breccia
56
pipes. A breccia pipe is a steeply dipping to subhorizontal, subcylindrical body of broken
57
rock that is long in one dimension but relatively short in two dimensions, and generally
58
forms at the intersection of tabular features such as dikes or faults (e.g., Guilbert and
59
Park, 1986). Breccia is used here as defined in the economic geology context of “a once
60
solid rock which is now broken and exhibits fragment rotation” (Taylor, 2009). Although
61
brecciation may redistribute and overprint earlier formed orebodies, Taylor and Pollard
62
(1983) and Taylor (2009) observed that breccia pipes also have a tendency to provide
63
pathways for fluids to permeate through large volumes of rock resulting in high-grade
64
deposits. The El Teniente, one of largest Cu-Mo deposits in central Chile (a deposit
65
largely classified as a porphyry Cu-Mo deposit, e.g., Klemm et al., 2007), was also
66
interpreted as a breccia deposit by Skewes et al. (2002). Other important breccia deposits
67
include the Oyu Tolgoi in Mongolia (e.g., Perelló et al., 2001), the Cananea in Mexico
68
(e.g., Bushnell, 1988), and the Los Bronces in Chile (e.g., Warnaars et al., 1985; Vargas
69
et al., 1999). Breccia pipes commonly occur in porphyry systems (e.g., Skewes et al.,
AC C
EP
TE D
M AN U
SC
52
ACCEPTED MANUSCRIPT 4 2003; Taylor, 2009) in small porphyritic intrusions thought to be apophyses to larger
71
plutons that act as centers to large-scale hydrothermal systems. Breccias contain
72
significant parts of orebodies in many porphyry deposits, for example, at Cananea in
73
Mexico (e.g., Bushnell, 1988), Los Bronces-Río Blanco in Chile (e.g., Warnaars et al.,
74
1985; Vargas et al., 1999), and Oyu Tolgoi in Mongolia (e.g., Perelló et al., 2001). The
75
character and origin of these breccias in magmatic hydrothermal environments vary
76
widely. Intrusive, magmatic hydrothermal, hydromagmatic breccias, and tectonic
77
breccias are distinguished (e.g., Sillitoe, 1985; Zhang et al., 2007; Bertelli and Baker,
78
2010; Cas et al., 2011; Li et al., 2012), and different types of breccias can occur in a
79
single deposit.
SC
M AN U
80
RI PT
70
The Musina copper district, located near Musina in the Limpopo Province, northern South Africa (Figs. 1, 2), is comprised of five mines which are, from west to
82
east: Campbell, Harper, Messina, Spence, and Artonvilla (Söhnge, 1945; van Graan,
83
1964; Jacobsen and McCarthy, 1976; Bahnemann, 1986). These mines were exploited for
84
copper for most of the twentieth century (1906-1991; e.g., Bahnemann, 1986; Wilson,
85
1998). In the Musina copper deposits, the metal is hosted in chalcopyrite, bornite, and
86
chalcocite, in addition to occurring as native copper in the central parts of the ore bodies
87
(e.g., Jacobsen and McCarthy, 1976; Bahnemann, 1986). The Musina copper deposits
88
occur as a group of nine pipe-like breccias of varying shapes; they also occur as
89
disseminated replacement deposits, vein/fissure deposits, and associated hydrothermal
90
copper deposits that were emplaced within high-grade metamorphic rocks of the
91
Limpopo orogenic belt (e.g., van Graan, 1964; McCarthy and Jacobsen, 1976; Jacobsen
92
et al., 1976; Ryan et al., 1983; Bahnemann, 1986).
AC C
EP
TE D
81
ACCEPTED MANUSCRIPT 5 Despite this long history of exploitation of copper in the Musina area, the origin
94
of these deposits is still controversial (e.g., Sawkins and Rye, 1979, 1980; McCathy and
95
Jacobsen, 1980; Wilson, 1998). Sawkins and Rye (1979) analyzed five samples from
96
Campbell Mine and obtained δ34Schalcopyrite values ranging from -0.4 to 0 ‰, two samples
97
from Campbell Mine with δ34Sbornite values of -0.3 ‰ and 2 ‰. They also obtained a
98
δ34Schalcopyrite value of 0.1 ‰ for a sample from Messina Fault and another one from
99
Spence Mine, and a δ34Sbornite value of -0.1 ‰ for a sample from Spence Mine, as well as
100
two samples from Harper Mine with δ34Schalcocite values of -2.1 and -1.0 ‰. Sawkins and
101
Rye (1979) interpreted these values to indicate an igneous source for the sulfur in
102
hydrothermal fluids.
M AN U
SC
RI PT
93
Studies of both sulfide mineral compositions and sulfur isotopes are potentially
104
useful tools in helping shed light on the genesis of sulfide mineralization (e.g., Ohmoto
105
and Rye, 1979; Ohmoto and Goldhaber, 1997; Seal, 2006). In this study, we report on
106
petrographic and sulfide mineral composition data, additional δ34S data on sulfides from
107
both Artonvilla (the easternmost) and Campbell (the westernmost) Mines, and a bulk-
108
rock δ34S value from country rocks to the Artonvilla Mine to try and place some
109
constrains on the origin of the Musina copper deposits. No country rock δ34S analyses for
110
these deposits have been determined before, and such results may be useful in helping
111
constrain the amount of crustal contamination. We specifically aim to document for the
112
first time sulfide mineral compositions from these deposits and evaluate if significant
113
copper mineralization is hosted in base metal sulfides, and attempt to constrain the
114
Musina copper mineralization temperatures from sulfur isotope thermometry.
115
AC C
EP
TE D
103
ACCEPTED MANUSCRIPT 6 2. Previous work and tectonic setting
117
The Limpopo orogenic belt is a Precambrian orogenic belt (e.g., Holzer et al., 1999;
118
Barton et al., 2006; Khosa et al., 2013) located between rocks of the Zimbabwean craton
119
to the north and those of the Kaapvaal craton to the south (Fig. 1). Based on
120
magnetotelluric data, Khosa et al. (2013) recently proposed that the 20-30 km wide
121
composite Sunnyside-Palala-Tshipise-Shear Zone represents a collisional suture zone
122
between the Kaapvaal and Zimbabwe cratons. According to Khosa et al. (2013), this
123
suturing resulted from the collision of the Zimbabwe and Kaapvaal cratons at ~2.6 Ga
124
and caused formation of high-grade granulites of the Limpopo orogenic belt. The
125
Limpopo orogeny evolved between 2.7 and 2.04 Ga, with its individual terranes having
126
experienced different tectonic histories under high-grade granulite facies metamorphism
127
(e.g., Barton et al., 2006).
SC
M AN U
Folded rocks form the dominant structure in the Musina area whose dimensions
TE D
128
RI PT
116
vary from hand specimen to regional scale (Figs. 2, 3). The Musina copper deposits,
130
located within the Central Zone of the Limpopo orogenic belt, occur within multiply
131
folded orthogneisses of varying composition (Bahnemann, 1986). These orthogneisses
132
have been subdivided into granodioritic and tonalitic varieties which form the basement
133
rocks in the area, overlain by metasediments and leucogneisses, garnetiferous schists,
134
amphibole and pyroxene granulite, metaquartzites, calc-silicates, and magnetic iron
135
formations (Bahnemann, 1986). Both the Singelele granitic gneiss and the Sand River
136
gneiss, dated at 2647 Ma and 3282 Ma respectively (Zeh et al., 2007), crop out in the
137
vicinity of the Musina copper deposits (Figs. 2, 3). The Sand River gneiss is interpreted
138
to have formed from melting of older Archean crust derived from a depleted mantle
AC C
EP
129
ACCEPTED MANUSCRIPT 7 source at 3.65 Ga (Zeh et al., 2007). The Singelele granitic gneiss, and the 2612 Ma Bulai
140
Pluton which crops out to the northwest of Musina (Figs. 2, 3), are both thought to have
141
been derived from mixtures of different proportions of reworked 3.65 Ga Paleoarchean
142
crust and juvenile magmas extracted from depleted mantle during the Neoarchean at 2.65
143
Ga (Zeh et al., 2007).
144
RI PT
139
The Bulai pluton, a calc-alkaline and dominantly porphyritic granite, was also intruded into the Central Zone of the Limpopo orogeny (Zeh et al., 2007). The Bulai
146
pluton is interpreted to represent a magmatic complex of variably deformed charnockites
147
and granites, and contains xenoliths of highly deformed metamorphic country rocks (Zeh
148
et al., 2007; Millonig et al., 2008). Further to the west of the Musina deposits, the
149
Mahalapye Complex, which forms the southwestern-most terrane of the exposed Central
150
Zone of the Limpopo orogenic belt in Botswana, is composed of granodiorites, quartz-
151
monzonites, migmatitic rocks, and Central Zone gneisses (e.g., Aldiss, 1991, Millonig et
152
al., 2010). An overall increase in the former melt fraction in the Central Zone migmatitic
153
rocks has been observed to occur from east to west (Van Breemen & Dodson, 1972;
154
Chavagnac et al., 2001). According to McCourt et al. (2004), the migmatitic gneisses are
155
crosscut by granite dykes and are also intruded by large volumes of granite plutons that
156
comprise both the syn-kinematic Mokgware Granite and the post-kinematic Mahalapye
157
Granite. A number of terranes in the Limpopo orogenic belt, such as the Franscistown
158
terrane in Botswana, are interpreted to represent continental margin arc complexes
159
(McCourt et al., 2004).
160 161
AC C
EP
TE D
M AN U
SC
145
At the Musina copper deposits, copper mining is thought to have commenced a couple of centuries ago when indigenous Africans used stone hammers and iron tools to
ACCEPTED MANUSCRIPT 8 mine the ore (e.g., Söhnge,1945; van Graan, 1964; Bahnemann, 1986) and ceased in 1991
163
(Wilson, 1998). Nine breccia pipes and associated hydrothermal copper deposits are
164
aligned along a northeast trend (e.g., Jacobsen et al., 1976) which forms a southwestern
165
projection of the 184-179 Ma Karoo igneous province (e.g., Cox, 1988, 1992; Duncan et
166
al, 1997, Jourdan et al., 2007). Both barren and three mineralized breccia pipes (Jacobsen
167
et al., 1976) which intrude rocks of the Limpopo orogenic belt (e.g., Barton et al., 2006)
168
occur in an area measuring 20 km long by 1 km wide that strikes northeast (e.g.,
169
McCarthy and Jacobsen, 1976; Jacobsen et al., 1976). Bahnemann (1986) observed that
170
the Musina copper mines occur at the intersection of faults, folds, fractures, and
171
lithological contacts (Figs. 2, 3). The Messina Fault and Dowe Tokwe wrench fault (Fig.
172
3) have been interpreted to be steep mylonite zones ranging from 2-40 m in width and are
173
cut by younger mafic dikes indicating reactivation along these faults (e.g. Bahnemann,
174
1986).
SC
M AN U
TE D
175
RI PT
162
Based on fluid inclusion studies (Sawkins, 1977) and stable isotope data (Sawkins and Rye, 1979), Sawkins and co-workers (Sawkins, 1977; Sawkins and Rye, 1979)
177
proposed a model involving formation of the Musina orebodies from circulating meteoric
178
waters whereby Cu and S were extracted from now eroded overlying Karoo basalts by
179
meteoric waters (with dissolved salts sourced in playa lakes which migrated downward
180
along faults). Based on oxygen isotope data of quartz veins with δ18O values of 9.8-14.1
181
‰, with the higher values interpreted to have occurred later at relatively lower
182
temperatures, Sawkins and Rye (1979) argued that such 18O-enriched values resulted
183
from low-temperature interaction of quartz with hydrothermal fluids that had negative
184
δ18OH2O values. As these meteoric fluids returned to surface they produced breccia and
AC C
EP
176
ACCEPTED MANUSCRIPT 9 replacement ore bodies (Sawkins and Rye, 1979). However, McCathy and Jacobsen
186
(1980) argued that the quartz that Sawkins and Rye (1979) studied is late in the
187
paragenetic sequence, and also that Sawkins and Rye (1979) did not analyze early stage
188
quartz on the basis that it was unsuitable for study. Further, McCarthy and Jacobsen
189
(1980) interpreted δ34S values {-3.4 to 0.4 ‰} presented by Sawkins and Rye (1979) to
190
lend support to a magmatic source for the sulfur in the Musina copper deposits.
A number of workers (e.g., Mihalik et al., 1974; Jacobsen and McCarthy, 1976;
SC
191
RI PT
185
McCarthy and Jacobsen, 1976; Jacobsen et al., 1976) proposed a magmatic hydrothermal
193
origin for the Musina copper deposits. According to Jacobsen and co-workers, copper-
194
bearing solutions were derived from a deep-seated alkaline magma of the Karoo igneous
195
suite that migrated upward along suitable structures and at several places became trapped,
196
enabling reactions between fluids rich in sodium and aluminum, and host granulites.
197
These reactions produced the ore deposits (e.g., McCathy and Jacobsen, 1980).
TE D
M AN U
192
Paragenetic studies conducted by Jacobsen et al. (1976), Jacobsen and McCarthy
199
(1976), and Sawkins (1977) observed a progressive increase in Cu:Fe and Fe:S ratios in
200
copper minerals with time, and specularite deposition relatively late in the paragenetic
201
sequence. Such observations imply a progressive increase in fO2 with a corresponding
202
decrease in fS2 occurred during the mineralization period (Sawkins and Rye, 1979). In the
203
Musina copper deposits paragenetic sequence, minerals such as albite, sericite, zoisite,
204
and quartz occur as replacement minerals, whereas specularite, sulfides, chlorite, and
205
calcite occur as vein filling minerals (Jacobsen et al., 1976). It was observed by Jacobsen
206
et al. (1976) that minerals such as albite and zoisite occur as both replacement (during
AC C
EP
198
ACCEPTED MANUSCRIPT 10 207
brecciation at ~375oC) and vein filling minerals, with the vein filling minerals inferred to
208
have been formed at temperatures of 160-210oC (Jacobsen et al.,, 1976).
209
The largest mining district is the Campbell Mine area (Figs. 2, 3), and here several lodes such as the Esperanza, Snake Lode, Line Lode, West Lode, Chain Lode,
211
and Maverick Lode, were mined for and produced a lot of copper (e.g., Bahnemann,
212
1986). These lodes were projected vertically transverse to the Messina Fault in a
213
southwest direction (Fig 4a; Bahnemann, 1986). What is striking to us from an
214
examination of Fig. 4 is that the main mineralization here is linked to the brecciated West
215
Lode which continues at depth. The Esperanza Lode, which appears to be linked to a
216
breccia at depth, is one of the most richly mineralized lodes not only in the Campbell
217
Mine area but in the whole Musina copper district (Bahnemann, 1986).
SC
M AN U
218
RI PT
210
Although not the subject of the present investigation, the Harper Mine, the next mine located approximately 3 km northeast of Campbell Mine, is shown in Fig. 4b to
220
demonstrate that the association of ore bodies with breccias is widespread in the Musina
221
copper deposits. On a longitudinal projection parallel to the Messina Fault in a north-west
222
direction, it can be noticed that a breccia pipe and the ore body are closely associated at
223
Harper Mine (Fig. 4b). Indeed, the occurrence of breccias and associated ore zones has
224
been documented at all mines in the Musina copper district (e.g., Van Graan, 1964;
225
Jacobsen et al., 1976; Jacobsen and McCarthy, 1976; Bahnemann, 1986).
EP
AC C
226
TE D
219
The observation that the Musina copper deposits are located along the Messina
227
Fault, as well as their preferential location where this fault intersects the contact between
228
granitic gneisses and metaquartzites, is probably an indication that the mineralization was
229
an interplay of a number of factors such as lithological contacts, faults, and folds
ACCEPTED MANUSCRIPT 11 (Bahnemann, 1986). Bahnemann (1986) interpreted these breccias to be the result of
231
tension that was created near lithological contacts by stress distribution during faulting
232
and this probably aided copper mineralization. Associated with breccia formation was
233
hydrothermal replacement and fissure mineralization (Bahnemann, 1986), implying that
234
the Musina breccias were cogenetic with the hydrothermal alteration.
RI PT
230
At the Musina copper deposits, copper mineralization is confined to a cordierite-
236
biotite-garnet gneiss unit which is hydrothermally altered (e.g., Jacobsen and McCarthy,
237
1976). According to Jacobsen and McCarthy (1976), mineral zonation occurs in four
238
concentric zones and increases in intensity towards the central part of the ore bodies.
239
Sericitization of plagioclase, sillimanite, cordierite, biotite, and quartz together with
240
pyrite marks the outermost first alteration zone. Biotite, sillimanite, and cordierite are
241
altered to chlorite whereas perthite and plagioclase are partially replaced by albite; garnet
242
is partially chloritized and pyrite is replaced by chalcopyrite to comprise the next
243
(second) zone further inwards (Jacobsen and McCarthy, 1976). Chalcopyrite is in turn
244
replaced by bornite where increasing alteration occurs in the third zone, the next
245
alteration zone inwards. Hematite is reported by Jacobsen and McCarthy (1976) to first
246
appear together with bornite in this third alteration zone. Where still more intense
247
alteration has destroyed original minerals to form the central and fourth alteration zone of
248
chlorite, epidote, and albite; associated ore minerals in this fourth alteration zone are
249
chalcocite and bornite (Jacobsen and McCarthy, 1976). Hematite and goethite also occur
250
in this fourth and most intensely altered zone (Jacobsen and McCarthy, 1976). Further, in
251
this fourth alteration zone, an assemblage of chalcocite, native copper, and goethite may
AC C
EP
TE D
M AN U
SC
235
ACCEPTED MANUSCRIPT 12 252
occur in a host rock that has been altered to gray-green clay (Jacobsen and McCarthy,
253
1976). Söhnge (1946) reported that as of June 30th, 1939, ore reserves at the Musina
255
copper deposits were estimated at 2.81 million tons assaying 2.09% Cu. In 1960, ore
256
reserves at these deposits were estimated at 6.37 million tons at 1.61% copper (Pelletier,
257
1964). In 1991, the last mine to operate in the Musina area, Messina-No. 5 shaft, stopped
258
production, bringing to an end 88 years of modern copper mining in the region. At the
259
time of closure, a total of 42 Mt of ore and 0.75 Mt of copper had been produced from the
260
Musina copper mines, with ore production peaking at 1.1 Mt/year in the 1970s (Wilson,
261
1998).
262
M AN U
SC
RI PT
254
3. Samples and analytical techniques
264
Since it is no longer possible to enter the old mine workings as they are currently sealed-
265
off, samples were collected from near closed mine shafts and dumps. Although our
266
efforts to obtain drill core data from the time when the mines were in operation were not
267
successful, our results compare very well to those from previous investigators and thus
268
validate our interpretations based on these dump samples. Samples from both Campbell
269
Mine and Artonvilla Mine utilized in this study are shown in Table 1 and their
270
approximate locations are shown in Fig. 2. Thus, rock and ore samples may not
271
necessarily conform to the geological map as they were collected from the rock dumps
272
representing the host rocks of the two deposits. Determining the actual levels from which
273
the samples originated from is now nearly impossible. However, since our samples were
AC C
EP
TE D
263
ACCEPTED MANUSCRIPT 13 274
obtained from the top part of the dumps, they are most probably from the deepest levels
275
of the workings since they are likely to have been mined last, if that was the case.
276
Examples of brecciated samples investigated in this study are shown Fig. 5. Slabbing of the samples was carried out in an effort to bring out the textures as
278
recommended by Taylor (2009). For samples from Artonvilla Mine, some breccias are
279
composed of mostly micas, amphibole and minor calcite and chlorite, potassium feldspar,
280
and angular quartz which range in grain size from 2 mm to >4 cm (Fig. 5a). The interval
281
between the amphibole and potassium feldspar clasts in Artonvilla Mine breccias is filled
282
in with mostly quartz (Fig 5a) and minor calcite in some cases (Table 1). The clear and
283
angular textured quartz is interpreted to be infill quartz, occupying the spaces between the
284
now altered broken rocks (Fig. 5a). This texture has some resemblance to a jigsaw fit-
285
stockwork texture (e.g., Zhang et al., 2007). For samples from Campbell Mine, very
286
coarse-grained and sometimes rounded (milled) quartz, angular quartz, amphibole, and
287
potassium feldspar clasts range in size from 1 mm to >7 cm (Figs. 5b, c), with quartz
288
forming the majority of the breccia clasts. Disseminated mineralization at Campbell Mine
289
is in the form of chalcopyrite, pyrite, and bornite which are hosted mostly in quartz veins
290
(Figs. 5b, c; Table 1), the quartz veins also being part of infill textures (e.g., Taylor,
291
2009). Sulfides in sample CpQv02 from Campbell Mine are very coarse-grained,
292
sometimes reaching sizes of up to 2 cm. All these lithologies are hydrothermally altered
293
represented by minerals such as chlorite and amphibole. The textures described for the
294
Campbell Mine breccias such as rounded clasts (e.g., Fig. 6) composed of different
295
lithologies are comparable to breccia pipe textures from Grasberg mine in Papua,
296
Indonesia described by Taylor (2009). At Campbell Mine, brecciated macroscopic quartz,
AC C
EP
TE D
M AN U
SC
RI PT
277
ACCEPTED MANUSCRIPT 14 micas, and amphibole grains occur in association with coarse-grained copper sulfide
298
mineralization (Figs. 5d). Another example of the brecciated sample from Campbell
299
Mine is shown Fig. 5d. Here, brecciated macroscopic quartz and amphibole grains occur
300
in association with coarse-grained copper sulphide mineralization (Figs. 5d). Pyrite is
301
observed mostly in association with sericite. These breccia textures described for the
302
Musina copper deposits which show rounded (milled) fragments are consistent with
303
breccia pipes of ‘broken rock’ described by Taylor (2009).
SC
Thin sections were examined first under a petrographic microscope, followed by
M AN U
304
RI PT
297
analyzing the major mineral phases and sulfides by electron microprobe on polished thin
306
sections. Mineral compositions of the sulfides (chalcopyrite and pyrite) were obtained
307
with a JEOL JXA-8530F Hyperprobe housed at Fayetteville State University, North
308
Carolina, USA (e.g., Chaumba et al., 2015). For sulfide analyses, the Hyperprobe was
309
operated at a beam current of 30 nA, an accelerating voltage of 30 kV, and a minimum
310
beam diameter of 1 µm. Compositions were analyzed firstly by X-ray energy dispersive
311
spectroscopy (EDS) for qualitative and semi-quantitative analysis, and then by X-ray
312
wavelength dispersive spectroscopy (WDS) for quantitative analysis using Smithsonian
313
standards. At least three spots were collected for each phase. Fifteen-second counting
314
times were used on peak and background measurements. For the calculation of the
315
oxides, the ZAF matrix correction system of Armstrong (1988) was used.
EP
AC C
316
TE D
305
Minerals for δ34S isotope analyses were separated via manual drilling using a
317
diamond drill bit at University of Georgia, Athens, GA, USA. Between 4 and 6 mg of
318
pyrite, chalcopyrite, and bornite were combined with elemental copper, quartz, and
319
vanadium pentoxide (V2O5) and finely crushed. An offline V2O5 combustion line with
ACCEPTED MANUSCRIPT 15 variable cryogenic temperature trap was used, employing a modified method from
321
Yanagisawa and Sakai (1983). The sample was then heated in a furnace at 1050oC to
322
produce SO2 and other gases. Water was removed cryogenically using a dry ice-ethanol
323
slush, and the remaining sample was collected in a variable temperature cryogenic trap.
324
After raising the temperature from -190 oC to -145 oC to remove CO2, the temperature
325
was raised again to -90 oC to convert solid SO2 to gaseous SO2.
The extraction line cryogenically isolates the SO2 gas from non-condensable
SC
326
RI PT
320
gases, CO2, and H2O using a dry ice and ethanol slush and a variable temperature trap.
328
The SO2 gas was measured using a calibrated mercury manometer, converted, and
329
reported as actual SO2 yield. SO2 was then collected in a Pyrex breakseal and analyzed
330
using a Finnigan MAT 252 dual inlet mass spectrometer. An accelerating potential of
331
8kV was used, and each sample was measured for 8 seconds for 8 standard-sample cycles
332
against the appropriate SO2 reference gases. Sulfur samples were measured with a 1000-
333
2000mV signal intensity, with an open Variable Ion Source Conductance (VISC) ‘sulfur
334
window’. An error of ±0.2 ‰ (2σ) was determined based upon replicate analysis of
335
Esperanza and ZS495 standards. Results are reported relative to the Vienna Canyon
336
Diablo Troilite Standard (VCDT) defining δ34S as:
338 339 340
TE D
EP
δ34S = {(34S/32sample - 34S/32reference)/( 34S/32reference)} x1000.
AC C
337
M AN U
327
The absolute 34S/32S ratio for VCDT is 4.50045 × 10−3 (Ault and Jensen 1963).
4. Results
341 342
4.1 Petrography
ACCEPTED MANUSCRIPT 16 343 Photomicrographs for samples from Artonvilla Mine area are shown in Fig. 6. Breccia
345
textures are not clearly visible in thin section mainly due to the very coarse-grained
346
nature of the breccias, and also probably due to overprinting events which take advantage
347
of earlier fractured zones (e.g., Taylor, 2009). Some Artonvilla Mine samples are
348
extensively hydrothermally altered such that only amphibole and micas are visible in
349
plane polarized light (Figs. 6a-d). Evidence of a breccia texture is evident in some thin
350
sections as angular points defined by the contacts of micas and amphiboles (Fig. 6b, c). In
351
some instances, some clasts (now altered to micas) in amphiboles are rounded, hinting to
352
milled (rounded) breccias (Fig. 6d). Other samples have high proportions of disseminated
353
sulfides (~10%) which occur together with sericite, muscovite, calcite, and feldspar (Figs.
354
6e, f). In some thin sections, evidence of pathways of hydrothermal fluids can be seen in
355
the form of numerous sulfide stringers and bifurcating sulfides with muscovite crystals
356
occurring between and interfingering with the sulfides (Figs. 6e, f). The sulfides also
357
contain numerous very fine- to fine-grained inclusions of muscovite (Figs. 6e, f). Granitic
358
gneisses, country rock samples to the Artonvilla Mine, which are dominated by altered
359
plagioclase, multiply-twinned plagioclase feldspar, alkali feldspar (Fig. 6g), occur in
360
association with coarse-grained biotite and sillimanite (Fig. 6h).
SC
M AN U
TE D
EP
AC C
361
RI PT
344
Campbell Mine samples show similarly extensive hydrothermal alteration of
362
silicate minerals (Fig. 7). Brecciated muscovite and potassium feldspar (being replaced
363
by sericite) are the dominant phases, with disseminated sulfides occurring as accessory
364
phases (Fig. 7a). The sulfides similarly enclose numerous very fine-grained sericite
365
crystals but the sulfides are not as abundant as in the Artonvilla Mine suite of samples.
ACCEPTED MANUSCRIPT 17 Where it is not sericitized, the alkali feldspar which occurs in association with the
367
mineralization is strongly hydrothermally altered (Fig. 7b). In some patches of Campbell
368
Mine deposit thin sections, sericite and brecciated potassium feldspar are the only
369
minerals present (Figs. 7c, d) whereas in other patches amphibole and feldspar and
370
altered alkali feldspar, chlorite, sericite, and quartz (Figs. 7e, f, respectively)
371
predominate. In certain cases, only isolated remnants of alkali feldspars remain from the
372
sericitization process (Fig. 7d). The lack of clear breccia textures in some thin sections
373
studied is probably due to extensive late-stage hydrothermal overprinting and alteration,
374
as discussed later.
M AN U
SC
RI PT
366
375 376
4.2 Sulfide mineral occurrence and composition
377
In backscattered electron images (BEIs), pyrite is lighter compared to muscovite, sericite,
379
and potassium feldspar in Artonvilla Mine samples (Figs. 8a, b). Pyrite grain size is
380
highly variable and ranges from less than 0.5 mm across to more than 2 cm in length,
381
occurring as subhedral to polygonal in shape (Figs. 8a, b). As shown in Figs. 8a and b,
382
pyrite neither occurs as inclusions nor as interstitial phases between silicate minerals, but
383
tends to occur as massive, polygonal grains with characteristic pyrite dodecahedral
384
shapes and penetration twinning (e.g., Wenk and Bulakh, 2004). The bifurcating nature
385
of the muscovite and sericite (darker (less dense) portions in Fig. 8a), consistent with
386
hydrothermal alteration, is also visible in BEIs from Artonvilla Mine samples (Fig. 8a).
387
This bifurcating nature of sericite in the Fig. 8a, although it also occurs in the center of
388
the feldspars, is concentrated more near the edges of the feldspars and in feldspars near
AC C
EP
TE D
378
ACCEPTED MANUSCRIPT 18 veins and veinlike textures, an indication that sericitization was triggered by external
390
hydrothermal which gained access through hydraulic fracturing fluids (e.g., Que and
391
Allen, 1996). Pyrite commonly occurs in association with muscovite and sericite (Figs.
392
8a, b). In addition, pyrite also occurs as numerous polygonal grains often less than 10 µm
393
across enclosed in coarser-grained potassium feldspar crystals of >500 µm in size (Fig.
394
8b).
Chalcopyrite is lighter compared to both amphibole and potassium feldspar
SC
395
RI PT
389
crystals as shown in a BEI in Fig. 8c. In samples from Campbell Mine, chalcopyrite is
397
very fine-grained and tends to occur as anhedral inclusions in brecciated amphibole
398
crystals (Fig. 8c). Further, chalcopyrite also occurs as numerous anhedral crystals often
399
less than 10 µm across enclosed in brecciated and relatively coarser-grained potassium
400
feldspar crystals of >500 µm in size (Fig. 8d). In contrast, in some thin sections,
401
plagioclase is almost completely enclosed by anhedral chalcopyrite of not more than 50
402
µm across (Fig. 8e). Moreover, chalcopyrite grains in plagioclase can occur as inclusions
403
within very coarse-grained amphibole crystals (Fig. 8e). In some cases, interstitial
404
curvilinear to anhedral chalcopyrite grains, typically <50µm, can also occur enclosed
405
within relatively coarser-grained zoned plagioclase crystals of >300 µm in size (Fig. 8f).
407 408
TE D
EP
AC C
406
M AN U
396
4.3 Pyrite
Table 2 lists representative pyrite chemical analyses from Artonvilla Mine, with
409
the rest of the results in Appendix 1. Sulfur concentrations in pyrite plot in a narrow
410
range, from 50.2 wt. % to 55.7 wt. % (Table 2, Fig. 9a). In some oxidized sulfides, O
411
replaces S (e.g., Locmelis et al., 2010) but this was not the case in our fresh samples. The
ACCEPTED MANUSCRIPT 19 majority of pyrite S concentrations (64 out of 85 analyses) fall within the 52-54 wt. % S
413
range (Fig. 9a). Very few pyrite S analyses fall in the ranges 50-51.99 wt. % S and 54-
414
54.99 wt. S (Fig. 9a). Nickel contents in Artonvilla Mine samples range from values at or
415
below detection limit to 0.45 wt. % (Table 2, Fig. 9b), and in a plot of wt. % Ni versus
416
wt. % Fe, no correlation exists between the two (Fig. 9b). The Artonvilla Mine country
417
rock gneiss sample, however, has lower Ni concentrations which range from values at or
418
below detection limit to 0.01 wt. % Ni (Fig. 9b). Iron concentrations for both Artonvilla
419
Mine and Artonvilla Mine country rock gneiss samples are comparable and they range
420
from 43-46.3 wt. % (Fig. 9b). Lead concentrations for Artonvilla Mine samples have
421
values at or below detection limit of ~0.03 wt. % to 0.43 wt. %, with the Artonvilla Mine
422
country rock gneiss values falling within the Artonvilla Mine samples (Table 2, Fig. 9c).
423
A plot of wt. % Pb versus wt. % Fe shows that Pb concentrations tend to cluster between
424
values at detection limit and 0.1 wt. % Pb whereas Fe concentrations tend to cluster
425
between 44.5 wt. % Fe and ~46.3 wt. % Fe (Fig. 9c).
426
428
4.4 Chalcopyrite
EP
427
TE D
M AN U
SC
RI PT
412
Table 3 lists representative analyses of chalcopyrite samples from Campbell Mine, with the rest of the results in Appendix 2. Chalcopyrite was not observed in the
430
Artonvilla Mine thin sections investigated. Chalcopyrite S concentrations are more or less
431
homogeneous with the majority of S analyses plotting in the 33-34.99 wt. % S range (51
432
out of 63 analyses) (Table 3, Fig. 9d). The majority of chalcopyrite S analyses, however,
433
fall in the 33-33.99 wt. % range (Fig. 9d). Very few chalcopyrite analyses plot in the 30-
434
32.99 wt. % S range and in the 35-38.99 wt. % S range (Fig. 9d). In both Campbell Mine
AC C
429
ACCEPTED MANUSCRIPT 20 amphibolites and quartz veins, Fe concentrations range from 26.7 wt. % to 29.8 wt. %
436
(Table 3, Fig. 9e). For both Campbell Mine amphibolite and quartz vein samples, Pb
437
concentrations range from values at or below detection limit to 0.2 wt. % Pb, and show
438
no correlation with Fe (Fig. 9e).
439
RI PT
435
In amphibolites, copper (Cu) concentrations in chalcopyrite plot in a very narrow range, from 32.5 wt. % to 33.9 wt. % (Table 3, Fig. 9f) whereas in quartz vein samples,
441
Cu concentrations in chalcopyrite range from 31.5 wt. % to 34.1 wt. % (Table 3, Fig. 9f).
442
In both Campbell Mine amphibolite and quartz vein samples, a good positive correlation
443
exists between Cu and Fe (Fig. 9f).
M AN U
444
SC
440
4.5 Sulfur isotopes
446
δ34S results from the present study are shown in Table 4. One sample from Artonvilla
447
Mine, AtCal01, has δ34Spyrite values of 3.1and 3.6 ‰ and chalcopyrite from the same
448
sample yielded a δ34S value of 3.9 ‰ (Table 4, Fig. 10a). A country rock granitic gneiss
449
sample to the Artonvilla Mine yielded a δ34Spyrite value of 8.2 ‰ (Table 4, Fig. 10a).
TE D
445
For samples from Campbell Mine, one quartz vein sample (CpQv01) has a
451
δ34Spyrite value of 0.5‰ whereas chalcopyrite samples drilled from different areas within
452
the same sample yielded δ34Schalcopyrite values of 0.4 and 0.7 ‰ (Table 4, Fig. 10a). The
453
same sample (CpQv01) also yielded a δ34Sbornite value of 0.4‰ as shown in Table 4.
454
Sample CpQv02, another quartz vein, yielded one δ34Schalcopyrite value of -0.3 ‰ (Table 4,
455
Fig. 10a). Sample CpAm04, an amphibolite, has chalcopyrite samples drilled from
456
different areas within the same sample yielding δ34S values of -1.1 and 0.3 ‰ (Table 4,
AC C
EP
450
ACCEPTED MANUSCRIPT 21 457
Fig. 10a). From Figure 10a, it can be seen that Campbell Mine samples all plot within a
458
range of -1.5 – 1.5 ‰.
459 5. Discussion
461
Based on both hand specimen descriptions and petrography, it is clear that samples from
462
both the Artonvilla and Campbell Mines are brecciated (Fig. 5). The lack of void spaces
463
in the samples under investigation rule out collapse breccias as collapse breccias are
464
characterized by voids comprising 20-50% of the rock (e.g., Taylor, 2009). The
465
occurrence of rounded (milled) clasts and lack of voids lend support to an intrusive
466
breccia origin. An intrusive breccia origin encompasses breccias that are characterized by
467
rounded fragments exhibiting evidence of upward transport and forceful injection to form
468
a pipe (e.g., Taylor, 2009). It should be noted that other types of breccias such as
469
hydrothermal, magmatic-hydrothermal, fluidized, explosive, phreatic, phreatomagmatic,
470
or hydromagmatic, e.g., are included in this intrusive breccia definition. At Musina, the
471
occurrence of a jigsaw fit-stockwork texture, where the interval between clasts is marked
472
by fractures or quartz veins is consistent with fluidized magmatic-hydrothermal breccias.
473
Further, the occurrence of larger breccias that are supported by smaller breccias and
474
alteration materials, probably indicating that the clasts moved over short distances,
475
creating open spaces with clasts of different lithologies characterized by rounded shapes
476
further supported a fluidized breccia origin (e.g., Zhang et al., 2007). It is speculated that
477
brecciation and mineralization/hydrothermal alteration likely coincided with fault development in
478
the Musina area such as the Messina and associated faults, probably during Karoo times.
479 480
AC C
EP
TE D
M AN U
SC
RI PT
460
The occurrence of hydrous minerals like amphibole, sericite and biotite in the Musina copper deposits samples investigated is consistent with a hydrothermal
ACCEPTED MANUSCRIPT 22 overprinting phase(s) which affected the breccias. Hydrothermal overprinting (often
482
multistage) is a common occurrence in breccia pipe deposits (e.g., Taylor, 2009). Such
483
hydrothermal overprinting and associated alteration can be very intense to the extent of
484
obliterating the original breccia textures (e.g., Taylor, 2009). The lack of clear breccia
485
textures in thin sections could be due to this hydrothermal overprinting. The mineral
486
assemblages at the two deposits of potassium feldspar, sericite, muscovite, chlorite,
487
calcite, and amphibole are consistent with sericitic or potassium silicate alteration (e.g.,
488
Meyer and Hemley, 1967; Rose and Burt, 1979; Pirajno, 1992; Reed, 1997). The
489
development of sericite in the deposits investigated suggests that the samples are from the
490
outer zones of a hydrothermally system (Jacobsen and McCarthy, 1976). The occurrence
491
of pyrite at Artonvilla Mine as in the Emery Lode also lends support to our interpretation
492
that samples from the Musina copper deposits investigated here are part of the outer
493
zones of a hydrothermal system. The association of chalcopyrite and chalcocite, which
494
are present in the internal parts of hydrothermal systems (Jacobsen and McCarthy, 1976;
495
Bahnemann, 1986), were not encountered in our suite of samples.
SC
M AN U
TE D
Chalcopyrite-bearing samples from Campbell Mine, the largest of the Musina
EP
496
RI PT
481
district (Bahnemann, 1986), are probably from the upper levels of the hydrothermal
498
system where this sulfide is predominant, and where chalcopyrite occurs together with
499
minor bornite (e.g., Bahnemann, 1986). This observation supports the interpretations of
500
Bahnemann (1986) who reported comparable hydrothermal alteration zones between
501
some ore shoots at both Campbell and Artonvilla Mines. This hydrothermal activity,
502
therefore, must have affected a large area of at least 20km by 1 km wide, from Campbell
503
Mine in the west to Artonvilla Mine in the east.
AC C
497
ACCEPTED MANUSCRIPT 23 504
The occurrence of felsic, porphyritic calc-alkaline rocks in the Central Zone of the Limpopo orogenic belt is consistent with dacites, quartz-feldspar-biotite porphyries,
506
quartz-latite porphyries, and rhyolitic porphyries encountered in porphyry copper
507
deposits such as at Escondida (e.g., Richards et al., 2001) and the Miduk porphyry Cu
508
deposit, Kerman, Iran (e.g., Boomeri et al., 2009). Some younger magmatic rocks
509
associated with, or close to, the Musina deposits are those that are thought to have been
510
emplaced during the break-up of Pangaea. The Limpopo-Lebombo-Sabi triple junction,
511
encompassing an aborted rift and associated sedimentation, formed a thick volcanic and
512
sedimentary succession formed during the Mesozoic (Truswell, 1977; Jourdan et al.,
513
2007). In the Musina area, copper deposits showing evidence of this age include basalt
514
intrusions of Karoo age (Jacobsen and McCarthy, 1975). Ore-related magmatism,
515
however, could have occurred much earlier in which case it was likely to be associated
516
with the intrusion of the porphyritic 2612 Ma Bulai calc-alkaline pluton (Zeh et al., 2007;
517
Millonig et al., 2008) in the Central Zone of the Limpopo orogenic belt. Thus, it is
518
possible that the Musina ores could have been mobilized at least twice after
519
emplacement: first at 1000 Ma based on a Rb-Sr age of Campbell Mine albites (Ryan et
520
al., 1983), and then during Karoo times (Jacobsen and McCarthy, 1975). Dating of
521
sericite, presumed here to be cogenetic with the mineralization phase (and also the
522
brecciation?), would help constrain the age of the mineralization. Occurrence of Cu
523
mineralization in an orogenic belt is not unique to the Musina area. In the Qiyugou
524
breccia pipe, Henan Province, China, brecciation and associated gold mineralization is
525
interpreted to have been related to granite porphyry emplaced at ~134 Ma which occurred
526
during continental collision (Li et al., 2012). The 347 Ma Santa Olalla granodiorite
AC C
EP
TE D
M AN U
SC
RI PT
505
ACCEPTED MANUSCRIPT 24 complex Cu–Ni deposit, for example, is a breccia deposit which is hosted by quartz
528
diorite to granodiorite-monzogranite rocks (e.g., Ordoñez-Casado et al., 2008). This
529
breccia deposit occurs in calc-alkaline rocks of the Iberian Variscan orogenic belt in
530
Spain (e.g., Casquet et al., 2001; Piña et al., 2006; Tornos et al., 2006).
531
RI PT
527
Sulfide compositions from the Artonvilla and Campbell Mines are not unusual
when compared to those from various settings (Fig. 11). Sulfide compositions from our
533
Musina samples are compared to sulfides from gabbros from the South West Indian
534
Ridge spreading center (Miller and Cervantes, 2002); an ophiolite complex, the south
535
west Oregon ophiolite (Foose, 1986); a layered intrusion, the Duluth Complex (Pasteris,
536
1984); and the metamorphosed Keivitsansarvi Ni-Cu-PGE deposit in the Keivitsa
537
intrusion, Finland (Gervilla and Kojonen, 2002) in Fig 11.
M AN U
SC
532
For Artonvilla Mine and gneiss pyrite samples, Ni contents are comparable to
539
those from various tectonic settings (except the Keivitsansarvi deposit) but Fe is only
540
comparable to concentrations from the SW Indian Ridge gabbros (Fig. 11a). Fe, however,
541
is higher than that from the Keivitsansarvi deposit but lower than Fe from the SW Oregon
542
ophiolite (Fig. 11a). This probably reflects less substitution of Fe by both Ni and Co,
543
common substituents for Fe in pyrite (e.g., Vaughan, 2011) in the Keivitsansarvi deposit
544
than the rest of the samples. The Keivitsansarvi deposit, a Ni-Cu-PGE rich deposit,
545
clearly shows this to be the case as it has both elevated Ni (Fig. 11a) and Co (Fig. 11b)
546
concentrations compared to the rest of the samples plotted.
EP
AC C
547
TE D
538
A plot of Fe versus Cu for Campbell Mine samples shows a positive correlation
548
between the two and that Campbell Mine samples have both lower Fe and Cu
549
concentrations compared to other samples plotted (Fig. 11c). The observation that Cu
ACCEPTED MANUSCRIPT 25 concentration from Campbell Mine are lower than other samples may suggest that at
551
Campbell Mine, the Cu mineralization is not hosted by chalcopyrite, but rather either by
552
other sulfides not in the present suite of samples such as bornite, chalcocite, or as native
553
copper (e.g., Bahnemann, 1986). Another possible explanation is that remobilization of
554
the copper (originally hosted in chalcopyrite and bornite) occurred during the
555
hydrothermal process, as in other breccia deposits such the copper-rich Donoso breccia
556
pipe of the Río Blanco-Los Bronces deposit in central Chile, the copper is mostly
557
contained in chalcopyrite and/or bornite (e.g., Skewes et al., 2003).
SC
Another plausible explanation is that the samples studied are not from the ore
M AN U
558
RI PT
550
zone. This observation also supports our earlier interpretation, based on mineral
560
assemblages from this study, that our samples are most likely from the outer zone of
561
hydrothermal alteration. Further, this observation is also consistent with previous studies
562
that have documented hydrothermal alteration zones whereby native copper occupies the
563
central zone, followed by a zone of chalcocite, followed by a zone of chalcopyrite as one
564
moves from the center of the alteration zone outwards (Jacobsen and McCarthy, 1976).
565
The lack of spread in Cu and Fe concentrations in chalcopyrite is consistent with the lack
566
of significant substitution in chalcopyrite by other cations (Vaughan, 2011). The lower
567
Cu concentrations in Campbell Mine samples, however, may also be due to the lower
568
abundances of sulfides in the crustal rocks (i.e., quartz veins and amphibolites) compared
569
to mantle-derived mafic rocks with higher concentrations of these elements than our
570
samples (Fig. 11). Campbell Mine quartz vein and amphibolite samples are both
571
characterized by chalcopyrite samples with Pb concentrations which show much wider
AC C
EP
TE D
559
ACCEPTED MANUSCRIPT 26 572
ranges over smaller Cu ranges with no correlation between Cu and Pb observed (Fig.
573
11d). Pyrite compositions from Artonvilla Mine are plotted on a S-Fe-Ni ternary
575
diagram in Fig. 12a and it can be observed that low Ni+Co wt. % values are characteristic
576
of these samples (Kullerud et al., 1969). The low wt. % Ni+Co concentrations from
577
Artonvilla Mine pyrites plot at and near the pyrite point which is consistent with pyrite
578
crystallization well before both the onset of chalcopyrite crystallization and expansion of
579
the monosulfide solid solution with further cooling (Fig 12a; e.g., Kullerud et al., 1969;
580
Guo et al., 1999). A simplified phase relations diagram in the S-Fe-Cu system at 300oC
581
(Barton and Skinner, 1979) is plotted in Fig. 12b. Campbell Mine chalcopyrite
582
compositions plot close to the ideal chalcopyrite but define a trend from high S and low
583
Fe to low S and high Fe at near constant copper concentrations (~33 wt. % Cu),
584
especially in Campbell Mine quartz vein samples (Fig. 12b). In amphibolites, however,
585
Fe and Cu concentrations define a trend from high Fe and low Cu concentrations of ~31
586
wt. % Cu to low Fe and high Cu concentrations of ~36 wt. % Cu (Fig. 12b). From phase
587
relationships, such a trend shown by amphibolites (Fig. 12b) is consistent with
588
immiscibility of both S and Fe, leading to increased concentrations of Cu in the sulfide;
589
which then likely settled out later, leading to the formation of the Musina copper
590
deposits.
SC
M AN U
TE D
EP
AC C
591
RI PT
574
A probable interpretation for such differing trends is that some of the Cu was
592
probably sourced from minerals such as chalcopyrite and bornite in amphibolites having
593
widely varying Cu concentrations and was then concentrated by hydrothermal fluids
594
which are characterized by more uniform Cu concentrations. This interpretation is
ACCEPTED MANUSCRIPT 27 595
supported by BEI images in which sulfide inclusions and sulfides enclosed in silicates
596
were only observed in amphibolites but not in quartz veins and other crustal rocks such as
597
gneisses (Fig. 8). A plot of the frequency versus δ34S values for samples from both the Artonvilla
599
and Campbell Mines, combined with results from Sawkins and Rye (1979), is shown in
600
Fig. 13a. Data from the present investigation compare very well with data from Sawkins
601
and Rye (1979) (Fig. 13a), and data from the Mantos Blancos breccia copper deposit in
602
Chile are broadly comparable to data from Campbell Mine (Fig. 13a). Of significance is
603
that samples from Campbell Mine are consistent with a magmatic source for the sulfides,
604
even though the sulfides now occur in amphibolites and quartz veins, and these data lend
605
support to the hypothesis that the mineralization in the Musina copper deposits was
606
sourced from hidden magmatic rocks of alkaline affinity most likely related to the Karoo
607
igneous phase (Söhnge, 1945; Jacobsen et al., 1976). If this is the case, hydrothermal
608
fluids could have been triggered by intrusion of these alkaline intrusive complexes,
609
resulting in the Cu being carried upwards towards the surface by heated meteoric waters
610
during the breakup of Pangaea. Sawkins and Rye (1979, 1980) suggested that the source
611
of the magmatic sulfides may well have been Karoo lavas that have since been eroded
612
away, an interpretation not ruled out by the findings from this investigation. However, it
613
would be difficult to explain the origin of the sulfide inclusions in amphibolites (which
614
most likely were formed contemporaneously with their enclosing rocks) if the sulfides
615
were concentrated by meteoric fluids only as suggested by Sawkins and Rye (1979,
616
1980).
AC C
EP
TE D
M AN U
SC
RI PT
598
ACCEPTED MANUSCRIPT 28 Artonvilla Mine samples have δ34S values of ~4 ‰, whereas a country rock
618
granitic gneiss to this deposit has a δ34S value of 8.2 ‰ (Fig. 10). Contamination of
619
magmatic sulfur by sulfur from Artonvilla Mine granitic gneisses can produce the
620
measured Artonvilla Mine δ34S values in Fig. 10 if the Campbell Mine sulfides are
621
representative of the sulfur source. The sulfides in Artonvilla Mine could also have been
622
mantle-derived, and the sulfides were then subsequently affected by hydrothermal fluids
623
that had interacted with sulfides from Artonvilla Mine country rocks. The slightly 34S-
624
enriched values in Artonvilla Mine samples allow some interpretations to be drawn based
625
on the timing of mineralization at Artonvilla Mine. If the mineralization predated
626
metamorphism, Artonvilla Mine samples and its country rocks would likely be
627
characterized by similar δ34S ratios, which is not the case here, suggesting that the
628
mineralization most likely occurred after regional metamorphism.
SC
M AN U
The mantle is characterized by δ 34S values of 0 ± 5‰ (Ohmoto and Rye, 1979;
TE D
629
RI PT
617
Ohmoto and Goldhaber, 1997). Sulfur isotope thermometry carried out for the Campbell
631
Mine quartz vein using a ∆34Schalcopyrite-bornite = 0.3 ‰ value (sample CpQv01) yielded a
632
temperature of 359oC (Li and Liu, 2006) consistent with the stability of this mineral pair
633
(Barton and Skinner, 1979). Such a temperature is broadly comparable with a
634
hydrothermal imprint on the Musina copper deposits where replacement minerals were
635
also inferred to have been formed at temperatures of ~375oC (e.g., Jacobsen et al., 1976).
636
Other copper deposits such as Mines Gaspé in Quebec, Canada (e.g. Shelton and Rye,
637
1982; Yang and Bodnar, 2004), the Ilkwang Mine breccia pipe deposit (δ34S values of
638
between -1.6 to 0.9 ‰), Republic of Korea (So and Shelton, 1983), and Mantos Blancos
639
breccia copper deposit (e.g. Ramírez et al., 2006), record comparable temperatures of
AC C
EP
630
ACCEPTED MANUSCRIPT 29 formation. This temperature is consistent with the interpretation that the mantle-derived
641
(due to the cluster of Musina δ34S values around 0 ± 2 ‰ in Fig. 13a, e.g. Sakai et al.,
642
1984; Chaussidon et al., 1991) mineralization is related to hydrothermal circulation
643
driven by magmatic intrusions, based on other evidence presented in this work such as
644
petrography and sulfide mineral compositions. Samples from Artonvilla Mine, however,
645
did not yield any temperatures, probably due to resetting of the δ34S ratios at this deposit
646
by hydrothermal fluids which resulted in sulfur isotope disequilibrium. δ34S results for
647
samples from Campbell Mine are within the range of δ34S values expected from mantle
648
reservoirs. It is, thus, concluded that δ34S results for Campbell Mine samples are within
649
the range of typical mantle-derived S, whereas δ34S values of the country rock granitic
650
gneiss at Artonvilla Mine is within the range of metamorphic rocks (e.g., Seal, 2006;
651
Hoefs, 2015). Samples from Artonvilla Mine plot at the high end of samples
652
characteristic of mantle-derived S, and also within the metamorphic δ34S range,
653
suggesting that the Artonvilla Mine samples may well have been formed during
654
metamorphism. The implication of this observation is that some of the mineralization at
655
Musina may have occurred very early, as metamorphism in the Central Zone occurred
656
between 2.7 and 2.04 Ga, during individual terranes accretion under high-grade granulite
657
facies metamorphism (e.g., Barton et al., 2006). The source of the mineralization could
658
also have been associated with the generation of voluminous granitic bodies during
659
Neoarchaean granulite metamorphism at ca 2.6 Ga (Holzer et al., 1999). Later
660
hydrothermal alteration, therefore, would have had little or no effect on the δ34S isotope
661
signatures. Interaction of mantle-derived S, such as encountered at Campbell Mine, with
AC C
EP
TE D
M AN U
SC
RI PT
640
ACCEPTED MANUSCRIPT 30 662
fluids that had previously interacted with country rock gneisses could also have produced
663
the enriched 34S values at Artonvilla Mine.
664
Assimilation of country-rock S is considered as the most reasonable mechanism for producing enriched S isotope signatures such as those at Artonvilla Mine (e.g.,
666
Ripley, 1999; Ripley and Li, 2003, 2013). Determination of the extent of assimilation of
667
crustal S by mafic magmas can be modeled using a simple two-component mixing model
668
(e.g., Ripley, 1999). If we assume an isotope composition of a mantle-derived magma of
669
δ34S = 0.0‰, a S concentration of 800 ppm in primitive magma, an Artonvilla Mine
670
country rock granitic gneiss δ34S value of 8.2‰ (Table 4), and a concentration of 1 wt. %
671
S in fresh biotite-garnet-cordierite gneisses (Jacobsen and McCarthy, 1976),
672
approximately 38% to 48% assimilation of fresh biotite-garnet-cordierite gneisses would
673
be required to produce δ34S values of between 3 ‰ and 4 ‰. A curved trajectory for this
674
two-component mixing equation is shown in Figure 13b. However, despite the
675
occurrence of sulfide-bearing country rocks to the Musina copper deposits, assimilation
676
of country rock sulfur was not the only cause of mineralization. At Campbell Mine, δ34S
677
results are indistinguishable from mantle values, indicating no contribution of crustal
678
sulfur or very little to negligible degrees of contamination.
SC
M AN U
TE D
EP
Due to the hydrothermal alteration that affected the Musina deposits, assessing the
AC C
679
RI PT
665
680
assimilation of crustally derived sulfur by the use of two-component mixing models (e.g.,
681
Lesher and Burnham, 1999; Ripley, 1999; Faure, 2001; Ripley and Li, 2003), however,
682
must be interpreted with caution. Physical addition of a contaminant or processes that
683
affect sulfide masses such as the hydrothermal alteration documented for the deposits
684
under study may render such mixing models inapplicable (e.g., Ripley and Li, 2003).
ACCEPTED MANUSCRIPT 31 Sulfur isotope studies in the Limpopo orogenic belt are lacking: however, such
686
data exists for adjacent Archean cratons and intrusions of the Zimbabwe and Kaapvaal
687
cratons. Mineralized samples from the Phoenix and Selkirk Ni-Cu-(PGE) deposits in the
688
Tati greenstone belt in eastern Botswana display slightly positive δ34S isotope values,
689
which range from 0.2 to 0.8‰ (Fiorentini et al., 2012). These S isotope values were
690
interpreted by Fiorentini et al. (2012) to be consistent with a dominantly mantle source
691
for the sulfur. In the Selebi-Phikwe belt, however, a granite-gneiss terrane with abundant
692
amphibolite lenses of either volcanic and/or intrusive nature, mineralized lower grade
693
samples from the Phikwe, Phokoje, and Dikoloti Ni-Cu-(PGE) deposits are characterized
694
by more variable δ34S values ranging from −3.1 to +0.3‰, suggesting that barren sulfides
695
associated with distal or low temperature sea-floor hydrothermal activity contributed
696
sulfur to these deposits (Fiorentini et al., 2012).
SC
M AN U
In the Damara orogenic belt located between the Kaapvaal and the Congo cratons
TE D
697
RI PT
685
southwest of the study area, Pirajno et al. (1992/93) reported possible granite related
699
mineral deposits (Group 2) composed of a wide range of sulfide minerals characterized
700
by low δ34S values (-1.2 ± 1.9‰). Mineral deposits in this orogenic belt which include
701
the Onguati and Brown Mountain marble-hosted Au-bearing hydrothermal quartz veins
702
and stockworks were interpreted by Pirajno et al. (1992/93) to be orogenic magmatic-
703
related. The formation of hydrothermal mineral deposits in the Damara orogenic belt was
704
related to the northwest subduction of the Kalahari plate below the Congo craton
705
accompanied by the emplacement of more than 200 plutons in the Central Zone (Pirajno
706
et al., 1992/93). The emplacement mechanism proposed for the Damara orogenic belt
707
deposits may also have been broadly comparable to what may have lead to the formation
AC C
EP
698
ACCEPTED MANUSCRIPT 32 708
of the Musina copper deposits, with crustal contamination having been more pronounced
709
at some of the Musina copper deposits such as at Artonvilla Mine being the exception. Crustal contamination of mantle-derived magmas is widespread even in the
711
Kaapvaal craton. In the northern part of the Kaapvaal craton, crustal sulfur contamination
712
signatures have been documented within all Platreef rock types of the mafic-ultramafic
713
layered intrusion of the Bushveld Complex from sulfur isotopes (e.g., Holwell et al.,
714
2007; Sharman et al., 2013). Further, from sulfur isotopes, Li et al. (2007) document
715
magmatic hydrothermal alteration helped concentrate platinum-group elements in the
716
main mineralized zone of the Great Dyke, a major Neoarchean mafic-ultramafic complex
717
intruding the Zimbabwe craton. Varying degrees of crustal contamination of mantle
718
derived magmas in the Kalahari craton have been described for various mineralized
719
bodies using other stable (e.g., Schiffries and Rye, 1989, Chaumba and Wilson, 1997)
720
and radiogenic (e.g., Kruger, 1994; McCandless et al., 1999; Schoenberg et al., 1999,
721
2003) isotopes. Thus, both hydrothermal alteration and crustal contamination of mantle-
722
derived magmas was widespread in the Kalahari craton and it was therefore not unique to
723
the Musina copper deposits.
SC
M AN U
TE D
EP
724
RI PT
710
Although the Musina breccia pipes are well known and have been described previously (e.g., Bahnemann, 1986, and references therein), a breccia pipe origin has not
726
yet been linked to the origin of the Musina ore bodies and it is hypothesized in this work
727
that these deposits originated as breccia pipes. In the Artonvilla and Campbell deposits,
728
in general >50 % of the volume of breccias we investigated are composed of clasts that
729
are at least 2 mm in diameter, consistent with an origin as fault breccias (e.g., Woodstock
730
and Mort, 2008). The documented hydrothermal event later affected the rocks under
AC C
725
ACCEPTED MANUSCRIPT 33 study thus overprinting the fault breccias, although it is possible that some breccias may
732
have been formed during the hydrothermal alteration stage (e.g., Jébrak, 1997). A breccia
733
pipe origin may better explain some of the features of the Musina copper deposits.
734
Alteration which has been documented in all previous studies at Musina copper deposits
735
(e.g., Mihalik et al., 1974; Jacobsen and McCarthy, 1976; McCarthy and Jacobsen, 1976;
736
Jacobsen et al., 1976; Bahnemann, 1986) and it is also common in some breccia pipes
737
where it has also been interpreted to be a result of both orthomagmatic and hydrothermal
738
events (e.g., Williams et al., 1999). Although brecciation may redistribute and overprint
739
earlier formed orebodies, breccia pipes also can provide pathways for fluids to permeate
740
through rocks resulting in high-grade deposits (e.g., Taylor and Pollard, 1983; Taylor,
741
2009), such as the high-grade deposits at Musina.
SC
M AN U
742
RI PT
731
Hydrothermal breccias develop early during the formation of veins, probably in response to the process of fracture propagation (e.g., Scholz, 1990). Hydrothermal
744
brecciation is one of the most common mechanisms of brecciation (e.g., Jebrak, 1997).
745
Formation of the Musina copper deposits at collisional tectonic environments that we
746
have suggested in this study is consistent with breccia pipe deposits formed in other
747
collisional tectonic regimes such as the Belif breccias in Tunisia (e.g., Decrée et al.,
748
2013) and the Qiyugou breccia pipe in China (e.g., Zhang et al., 2007; Li et al., 2012).
749
Evidence exists for a breccia pipe origin for the Musina copper deposits.
AC C
EP
TE D
743
750
Numerous ore concentrations in the Musina copper deposits are associated with breccias
751
(e.g., Söhnge, 1945; Van Graan, 1964; Bahnemann, 1986). Breccia pipes are also a
752
common occurrence in porphyry systems (e.g., Skewes et al., 2003), resulting in
753
extensive hydrothermal alteration in both the breccia pipes and porphyry systems. It is
ACCEPTED MANUSCRIPT 34 also possible that the Musina deposits may have originated as a porphyry system as most
755
breccia pipes are associated with porphyry deposits (e.g., Sillitoe, 2010) but the evidence
756
linking the deposit to that origin may have been destroyed during subsequent processes.
757
Although classic porphyry-style mineralization characterized by different alteration zones
758
which are often a guide to ore from series of mineral assemblages which extend into
759
larger adjacent rocks has not been reported in all breccia pipes, a link between the two
760
can often be ubiquitous (e.g., Sillitoe, 1985; Seedorff et al., 2005). Oxygen isotope values
761
of quartz veins of 9.8 ‰ reported by Sawkins and Rye (1979) are also consistent with an
762
igneous origin, although no hydrogen isotopes have yet been performed to help constrain
763
the source of the fluids.
SC
M AN U
764
RI PT
754
On discussing a model for the origin of mineralization at Musina, Bahnemann (1986) concluded that “…it must be able to explain the unusual purity of the ore at
766
Musina that allows the production of fire-refined copper ingots of almost wire-bar
767
quality”. A breccia origin for the Musina copper deposits can help answer this as high-
768
grade copper deposits occur in breccias such as the Donoso breccia deposit at Rio
769
Blanco-Los Bronces (e.g., Skewes et al., 2003) where mineral deposition in the breccia
770
matrix cooled rapidly, resulting in metal deposition due to expansion of magmatic fluids
771
that are released from cooling stocks (e.g., Sillitoe, 1985). Indeed, the El Teniente, one of
772
largest and high-grade Cu-Mo deposits in central Chile (a deposit largely classified as a
773
porphyry Cu-Mo deposit, e.g., Klemm et al., 2007), is now also being re-interpreted as a
774
breccia deposit by Skewes et al. (2002) and here we also suggest a breccia origin for the
775
Musina copper deposits.
776
AC C
EP
TE D
765
ACCEPTED MANUSCRIPT 35 777
6. Conclusions
778 The Musina copper deposits display evidence of hydrothermal activity based on their
780
mineralogy. Hydrothermal alteration products include potassium feldspar, sericite,
781
muscovite, and pyrite. Sulfide mineralogy and sulfide compositions at Artonvilla and
782
Campbell Mines are also consistent with a hydrothermal overprint on earlier formed
783
sulfide mineralization. δ34S values of the Musina copper deposits are consistent with a
784
hydrothermal imprint of mantle-derived sulfur at Campbell Mine. Mantle-derived sulfur
785
affected by either contamination by country-rock sulfides or affected by hydrothermal
786
activity that had interacted with country rocks, or both, may explain the elevated δ34S
787
isotope compositions at Artonvilla Mine. Sulfur isotope thermometry carried out for the
788
Campbell Mine quartz vein using a ∆34Schalcopyrite-bornite = 0.3 ‰ value yield a temperature
789
of 359oC that is consistent with the expected stability of these mineral phases. Features
790
such as the occurrence of faults, breccias, high-grade copper ores, and hydrothermal
791
alteration displayed by the Musina copper deposits are consistent with a breccia origin.
TE D
M AN U
SC
RI PT
779
EP
792 Acknowledgements
794
JBC gratefully acknowledges funding from a UNC Pembroke Summer Research
795
Fellowship which covered the costs of S isotope work at The University of Georgia, and
796
a Research Fellowship from UNC Pembroke Office of Graduate Studies and Research
797
which helped cover the costs of microprobe analyses at Fayetteville State University.
798
Mike Roden is gratefully acknowledged for critically reading an earlier version of the
799
manuscript. Doug Crowe commented on an earlier draft of the manuscript and facilitated
AC C
793
ACCEPTED MANUSCRIPT 36 800
access to the UGA Stable Isotope Lab. Thoughtful reviews by Joyashish Thakurta and an
801
anonymous reviewer helped improve the manuscript and are acknowledged.
802 References
804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840
Aldiss, D. T., 1991. The Motloutse Complex and the Zimbabwe Craton/Limpopo Belt transition in Botswana. Precambrian Research, 50: 89-109. Armstrong, J.T., 1988. Quantitative analysis of silicate and oxide minerals: comparison of the Monte Carlo, ZAF and Φ(ρΖ) procedures: in D.E. Newbury (ed.), San Francisco Press, San Francisco, USA. Microbeam Analysis - 1988: 239-246. Ault, W.U. and Jensen, M.L., 1963. Summary of sulfur isotope standards: In M.L. Jensen (Ed.), Biogeochemistry of Sulfur Isotopes, National Science Foundation Symposium Proceedings, Yale University. Bahnemann, K.P., 1986. A review of the geology of the Messina copper deposits, Northern Transvaal: in C.R. Anhaeusser and S. Maske (eds.), Mineral deposits of Southern Africa, v. 2, Geological Society of South Africa, p. 1671-1688. Barton, J., Klemd, R., Zeh, A., 2006. The Limpopo belt: a result of Archean to Proterozoic, Turkic-type orogenesis?: In W.U. Reimold and R.L. Gibson (Eds.), Processes on Early Earth: Geological Society of America Special Paper 405, 405. Geological Society of America, p. 315-332. Barton, P.B. Jr. and Skinner, B.J., 1979. Sulfide mineral stabilities, in H.L. Barnes (ed.), Geochemistry of Hydrothermal Ore Deposits, 2nd Ed., J. Wiley and Sons, p. 278403. Bertelli, M. and Baker, T., 2010. A fluid inclusion study of the Suicide Ridge breccia pipe, Cloncurry district, Australia: implication for breccia genesis and IOCG mineralization, Precambrian Research, 179: 69-87. Boomeri, M., Nakashima, K. and Lentz, D.R., 2009, The Miduk porphyry Cu deposit, Kerman, Iran: A geochemical analysis of the potassic zone including halogen element systematics related to Cu mineralization processes, Journal of Geochemical Exploration, 103: 17-29. Bushnell, S.E., 1988, Mineralization at Cananea, Sonora, and the paragenesis and zoning of breccia pipes in quartzofeldspathic rock: Economic Geology, 83: 1760-1781. Cas, R., Giordano, G., Balsamo, F., Esposito, A and Lo Mastro, S., 2011. Hydrothermal breccia textures and processes: Lisca Bianca Inlet, Panarea Volcano, Italy, Economic Geology, 106: 437-450 Casquet, C., Galindo, C., Tornos, F., Velasco, F., and Canales, A. 2001. The Aguablanca Cu–Ni ore deposit (Extremadura, Spain), a case of synorogenic orthomagmatic mineralization: age and isotope composition of magmas (Sr, Nd) and ore (S). Ore Geology Reviews, 18: 237-250. Chaumba, J.B. and Wilson, A.H., 1997. An Oxygen isotope study of the Lower Mafic Succession of the Great Dyke, Zimbabwe. Chemical Geology, 135: 293-305. Chaumba, J.B., Reid, J.C. and Parker, J.C., 2015. Hamme (Tungsten Queen) mine
AC C
EP
TE D
M AN U
SC
RI PT
803
ACCEPTED MANUSCRIPT 37
EP
TE D
M AN U
SC
RI PT
tailings: Vance County, North Carolina: Mineral chemistry of potentially recoverable huebnerite-wolframite and sulfide minerals, Southeastern Geology, 51: 33-49. Chaussidon, M., Albarede, F., and Michard, A., 1991. Hydrogen, sulphur, and neodymium isotope variations in the mantle beneath the EPR at 12°50' N: Geochemical Society Special Publication 3: 325–337. Chavagnac, V., Kramers, J.D., Naegler,T.F. & Holzer, L., 2001. The behaviour of Nd and Pb isotopes during 2.0 Ga migmatization in paragneisses of the Central Zone of the Limpopo Belt (South Africa and Botswana). Precambrian Research, 112: 51-86. Cox, K.G., 1988. The Karoo Province: in D. Macdougall, (ed.), Continental Flood Basalts, J Kluwer Acad., Norwell, Massachusetts, p. 239-271. Cox, K.G., 1992. Karoo igneous activity, and the early stages of the breakup of Gondwana: in B.C. Storey et al. (ed.), Magmatism and the Causes of Continental Breakup, Geological Society Special Publication no. 68: 137-148. Decrée, S., Marignac, Ch., De Putter, Th., Yans, J., Clauer, N., Dermech, M., Aloui, K., and Baele, J.M., 2013, The Oued Belif hematite-rich breccias: a Miocene iron oxide Cu-Au (U-REE) deposit in the Nefza Mining District, Tunisia, Economic Geology, 108: 1379-1396. Deyell, C.L., 2005. Sulfur isotope zonation at the Mt Polley alkalic porphyry Cu-Au deposit, British Columbia, Canada: in J. Mao and F.P. Bierlein (eds.), Mineral Deposit Research - Meeting the Global Challenge, Springer, Heidelberg, p 373376. Dill, H.G., Garrido, M.M., Melcher, F., Gomez, M.C. and Luna, L.I., 2012. Depth-related variation of tourmaline in the breccia pipe of the San Jorge porphyry copper deposit, Mendoza, Argentina, Ore Geology Reviews, 48: 271-77. Duncan, R.A, Hooper, P.R., Rehacek, J., Marsh, J.S., and Duncan, A.R., 1997. The timing and duration of the Karoo igneous event, southern Gondwana: Journal of Geophysical Research, 102: 18,127-18,138A. Faure, G., 2001. Origin of igneous rocks, the isotopic evidence, Springer, New York, 496p. Field, C.W. and Gustafson, L.B., 1976. sulfur isotopes in the porphyry copper deposit at el Salvador, Chile, Economic Geology, 71: 1533-1548. Fiorentini, M.L., Bekker, A., Rouxel, O., Wing, B., Maier, W., and Rumble, D., 2012. Multiple sulfur and iron isotope composition of magmatic Ni-Cu-(PGE) sulfide mineralization from Eastern Botswana: Economic Geology, 107: 105-116. Foose, M.P., 1986. Setting of a magmatic sulfide occurrence in a dismembered ophiolite, southwestern Oregon: USGS Bulletin 1626A, p. A1-A23. Gervilla, F. and Kojonen, K., 2002. The platinum-group minerals in the upper section of the Keivitsansarvi Ni-C-PGE deposit, Northern Finland: Canadian Mineralogist, 40: 377-394. Guilbert, J.M. and Park, C.F., 1986. The geology of ore deposits: Freeman, San Francisco, 985 p. Guo, J., Griffin, W.L. and O’Reilley, S.Y., 1999. Geochemistry and origin of sulphide minerals in mantle xenoliths: Qilin, Southeastern China: Journal of Petrology, 40: 1125-1149.
AC C
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
ACCEPTED MANUSCRIPT 38
EP
TE D
M AN U
SC
RI PT
Hoefs, J., 2015. Stable isotope geochemistry, 7th Edition, Springer, New York, 389 p. Holwell, D.A., Boyce, A.J., McDonald, I., 2007. Sulfur isotope variations within the Platreef Ni-Cu-PGE deposit: genetic implications for the origin of sulfide mineralization. Economic Geology, 102: 1091-1110. Holzer, L., Barton, J.M., Paya, B.K. and Kramers, J.D., 1999. Tectonothermal history of the western part of the Limpopo Belt: tectonic models and new perspectives, Journal of African Earth Sciences, 28: 383-402. Jacobsen, J.B.E. and McCarthy, T.S., 1975. Possible late Karroo carbonatite and basalt intrusions at Messina, Transvaal, Transaction of the Geological Society of South Africa, 78: 117-130. Jacobsen, J.B.E. and McCarthy, T.S., 1976. An unusual hydrothermal copper deposit at Messina, South Africa: Economic Geology, 78: 153-159. Jacobsen, J.B.E., McCarthy, T.S. and Laing, G.J., 1976 The copper-bearing breccia pipes of the Messina District, South Africa: Mineralium Deposita, 11: 33-45. Jébrak, M., 1997. Hydrothermal breccias in vein-type ore deposits: a review of mechanisms, morphology and size distribution. Ore Geology Reviews, 12: 111134. Jourdan, F., Bertrand, H., Scharer, U., Blichert-Toft, J, Feraud, G., and Kampunzu, A.B., 2007. Major and Trace Element and Sr, Nd, Hf, and Pb isotope compositions of the Karoo large igneous province, Botswana-Zimbabwe: lithosphere vs mantle plume contribution: Journal of Petrology, 48: 1043-1077. Khoza, D., Jones, A.G., Muller, M.R., Evans, R.L., Webb, S.J., Miensopust, M. and The SAMTEX team, 2013, Tectonic model of the Limpopo belt: Constraints from magnetotelluric data: Precambrian Research, 226: 143-156. Klemm, L.M., Pettke, T., Heinrich, C.A., and Campos, E., 2007. Hydrothermal evolution of the El Teniente deposit, Chile: Porphyry Cu-Mo ore deposition from lowsalinity magmatic fluids: Economic Geology, 102: 1021-1046. Kruger, F.J., 1994. The Sr-isotopic stratigraphy of the Western Bushveld Complex, South African Journal of Geology, 97: 393-398. Kullerud, G., Yund, R.A. and Moh, G., 1969. Phase relations in the Cu-Fe-S, Cu-N-S and Fe-Ni-S systems: Economic Geology Monograph, 4: 323-343. Landtwing, M.R., Dillenbeck, E.D., Leake, M.H., and Heinrich, C.A., 2002. Evolution of the breccia-hosted porphyry Cu–Mo–Au deposit at Agua Rica, Argentina: progressive unroofing of a magmatic hydrothermal system: Economic Geology, 97: 1273-1292. Lesher, C.M., and Burnham, O.M., 1999. Mass balance and mixing in dynamic oreforming magmatic system, in Keays, R.R., Lesher, C.M., Lightfoot, P.C., and Farrow, C.E.G., eds., Dynamic processes in magmatic ore deposits and their application in mineral exploration: Geological Association of Canada Short Course Notes, 13, p. 413-450. Li, C. Ripley, E.M., Oberthür, T., Miller, J.D. Jr., & Joslin, G.D., 2007. Textural, mineralogical and stable isotope studies of hydrothermal alteration in the main sulfide zone of the Great Dyke, Zimbabwe and the precious metals zone of the Sonju Lake Intrusion, Minnesota, USA, Mineralium Deposita, 43: 97-110. Li, N., Carranza, E.J.M., Ni, Z., and Guo, D., 2012. The CO2-rich magmatic-
AC C
887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931
ACCEPTED MANUSCRIPT 39
EP
TE D
M AN U
SC
RI PT
hydrothermal fluid of the Qiyugou breccia pipe, Henan Province, China: implication for breccia genesis and gold mineralization, Geochemistry: Exploration, Environment, Analysis, 12: 147-160. Li, Y.B. and Liu, J.M., 2006. Calculation of sulfur isotope fractionation in sulfides: Geochimica et Cosmochimica Acta , 70: 1789-1795. Locmelis, M., Melcher, F. and Oberthür, T., 2010. Platinum-group element distribution in the oxidized Main Sulfide Zone, Great Dyke, Zimbabwe, Mineralium Deposita, 45: 93-109. Maydagán, L., Franchini, M., Lentz, D. Pons, J. & McFarlane, C., 2013. Sulfide composition and isotopic signature of the Alter Cu-Au deposit, Argentina: constrains on the evolution of the porphyry-epithermal system, Canadian Mineralogist, 51: 813-840. McCandless, T.E., Ruiz, J., Adair, B.I., Freydier, C., 1999. Re–Os isotope and Pd/Ru variations in chromitites from the Critical Zone, Bushveld Complex, South Africa. Geochimica. Cosmochimica. Acta, 63: 911-923. McCarthy, T.S. and Jacobsen, J.B.E., 1976. The mineralizing fluids at the Artonvilla copper deposit: an example of a silica-deficient, alkaline hydrothermal system: Economic Geology, 71: 131-138. McCarthy, T.S. and Jacobsen, J.B.E., 1980. Additional geochemical data on the Messina copper deposits, Transvaal, South Africa - A discussion: Economic Geology, 75: 478-483. McCourt, S., Kampunzu, A.B., Bagai, Z. and Armstrong, R.A., 2004. The crustal architecture of Archaean terranes in Northeastern Botswana, South African Journal of Geology, 107: 146-158. Meyer, C. and Hemley, J.J., 1967. Wall rock alteration, in H.L. Barnes (ed.), Geochemistry of Hydrothermal Ore Deposits, Holt, Rinehart and Winston, New York, p. 166-235. Mihalik, P., Jacobsen, J. B. E., and Hiemstra, S. A., 1974. Platinum group minerals from a hydrothermal environment, Economic Geology, 69: 257-262. Miller, D.J. and Cervantes, P., 2002. Sulfide mineral chemistry and petrography and platinum group element composition in gabbroic rocks from the Southwest Indian Ridge: In J.H. Natland, H.J.B. Dick, D.J. Miller and R.P. Von Herzen (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, 176: 1-29. Millonig, L., Zeh, A., Gerdes, A. and Klemd, R., 2008. Neoarchaean high-grade metamorphism in the Central Zone of the Limpopo Belt (South Africa): Combined petrological and geochronological evidence from the Bulai pluton: Lithos, 103: 333-351. Millonig, L., Zeh, A., Gerdes, A. Klemd, R. and Jackson, J.M. Jr., 2010. Decompressional heating of the Mahalapye Complex (Limpopo Belt, Botswana): a response to Palaeoproterozoic magmatic underplating? Journal of Petrology, 51: 703-729. Ordoñez-Casado, B., Martin-Izard, A. and García-Nieto, J., 2008. SHRIMP-zircon U–Pb dating of the Ni–Cu–PGE mineralized Aguablanca gabbro and Santa Olalla granodiorite: Confirmation of an Early Carboniferous metallogenic epoch in the Variscan Massif of the Iberian Peninsula, Ore Geology Reviews, 34:343-353. Ohmoto, H. and Rye, R.O., 1979. Sulfur and carbon isotopes: in H.L. Barnes (ed.),
AC C
932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977
ACCEPTED MANUSCRIPT 40
EP
TE D
M AN U
SC
RI PT
Geochemistry of Hydrothermal Ore Deposits, 2nd Ed., J. Wiley and Sons, p. 509567. Ohmoto, H. and Goldhaber, M.B., 1997. Sulfur and carbon isotopes: in H. L. Barnes (ed.), Geochemistry of Hydrothermal Ore Deposits, 3rd Ed., J. Wiley and Sons, p. 517-611. Pasteris, J. D., 1984. Further interpretation of the Cu-Fe-Ni sulfide mineralization in the Duluth Complex, northeastern Minnesota, Canadian Mineralogist, 22: 39-53. Pelletier, R.A., 1964. Mineral Resources of South Central Africa, Oxford University Press, London. 37 p. Penniston-Dorland, S.C., Mathez, E.M., Wing, B.A., Farquhar, J. and Kinnaird, J.A., 2012. Multiple sulfur isotope evidence for surface-derived sulfur in the Bushveld Complex. Earth Planetary Science Letters, 337/338: 236-242. Perelló, J., Cox, D., Garamjav, D., Sanjdorj, S., Diakov, S., Schissel, D., Munkhbat, T.O., and Oyun, G., 2001. Oyu Tolgoi, Mongolia: Siluro-Devonian porphyry Cu-Au(Mo) and high sulfidation Cu mineralization with a cretaceous chalcocite blanket: Economic Geology, 96: 1407-1428. Piña, R., Lunar, R., Ortega, L., Gervilla, F., Alapieti, T., Martínez, C., 2006. Petrology and geochemistry of mafic–ultramafic fragments from the Aguablanca (SW Spain) Ni-Cu ore breccia: implications for the genesis of the deposit. Economic Geology 101, 865–881. Pirajno, F., 1992. Hydrothermal mineral deposits, principles and fundamental concepts for the exploration geologist: Springer-Verlag, Berlin, 709 p. Pirajno, F., Kinnaird, J.A., Fallick, A.E., Boyce, A.J., and Petzel, V.W.F., 1992/93. Communications of the Geol. Survey Namibia, 8: 87-104. Que, M. and Allen, A.R., 1996. Sericitization of plagioclase in the Rosses Granite Complex, Co. Donegal, Ireland, Mineralogical Magazine, 60: 927-936. Ramírez, L.E., Palacios, C., Townley, B., Parada, M.A., Sial, A.N., Fernandez-Turiel, J.L., Gimeno, D., Garcia-Valles, M. and Lehmann, B., 2006. The Mantos Blancos copper deposit: an upper Jurassic breccia-style hydrothermal system in the Coastal Range of Northern Chile, Mineralium Deposita, 41: 246-258. Reed, M.H., 1997. Hydrothermal alteration and its relationship to ore fluid composition: in H. L. Barnes (ed.), Geochemistry of Hydrothermal Ore Deposits, 3rd Ed., J. Wiley and Sons, p. 303-365. Richards, J.P., Boyce, A.J., Pringle, M.S., 2001. Geologic evolution of the Escondida area, Northern Chile: a model for spatial and temporal localization of porphyry Cu mineralization, Economic Geology, 96: 271-305. Ripley, E.M., 1999. Systematics of sulphur and oxygen isotopes in mafic igneous rocks and related Cu–Ni–PGE mineralization. In: Keays, R.R., Lesher, C.M., Lightfoot, P.C., Farrow, C.E.G. (Eds.). Dynamic Processes in Magmatic Ore Deposits and their Application to Mineral Exploration: Geological Association of Canada. Short Course Notes, 13: 133-158. Ripley, E.M., and Li, C., 2003. Sulfur isotope exchange and metal enrichment in the formation of magmatic Cu-Ni-(PGE) deposits. Economic Geology, 98: 635-641. Ripley, E.M., Li, C., 2013. Sulfide saturation in mafic magmas: is external sulfur required for magmatic Ni-Cu-(PGE) ore genesis? Economic Geology, 108: 45-58. Rose, A.W. and Burt, D.M., 1979. Hydrothermal alteration: in H.L. Barnes (ed.),
AC C
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 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023
ACCEPTED MANUSCRIPT 41
EP
TE D
M AN U
SC
RI PT
Geochemistry of Hydrothermal Ore Deposits, 2nd Ed., J. Wiley and Sons, p. 173235. Ryan, B.D., Kramers, J.D., Stacey, J.S., Deleveaux, M., Barton, J.M. and Fripp, R.E.P., 1983. Strontium and lead isotope studies and K/Rb ration measurements relating to the origin and emplacement of the copper deposits near Messina, South Africa: in W.J. van Biljon and J.H. Legg (eds.) Geological Society of South Africa, Special Publication no. 8: 47-53. Sakai, H., DeMarais, D., Veda, A., and Moore, J.G., 1984. Concentrations and isotope ratios of carbon, nitrogen, and sulfur in ocean-floor basalts: Geochimica et Cosmochimica Acta, 48: 2433-2441. Sawkins, F. J., 1977. Fluid inclusion studies of the Messina copper deposits, Transvaal, South Africa, Economic Geology, 72: 619-631. Sawkins, F. J. and Rye, R. O., 1979. Additional geochemical data on the Messina copper deposits, Transvaal, South Africa: Economic Geology, 74: 684-689. Sawkins, F. J. and Rye, R. O., 1980. Additional geochemical data on the Messina copper deposits, Transvaal, South Africa-A reply: Economic Geology, 75: 481-482. Schiffries, C.M. and Rye, D.M., 1989. Oxygen isotopic systematics of the Bushveld Complex: I. Constrains on magmatic processes in layered intrusions, American Journal of Science, 289: 841-873. Schoenberg, R., Kruger F.J., Nagler T.F., Meisel T., and Kramers J.D., 1999. PGE enrichment in chromitite layers and the Merensky Reef of the Western Bushveld Complex; a Re–Os and Rb–Sr isotope study. Earth Planet Science Letters 172: 49-64. Schoenberg, R., Nagler, T.F., Gnos, E., Kramers, J.D., and Kamber, B.S., 2003. The source of the Great Dyke, Zimbabwe, and its tectonic significance: evidence from Re–Os isotopes, Journal of Geology, 111: 565- 578. Seal, R.R. II, 2006. Sulfur isotope geochemistry of sulfide minerals: Reviews in Mineralogy and Geochemistry, 61: 633-677. Seedorff, E., Dilles, J.H., Proffett, J.M. Jr., Einaudi, M.T., Zurcher, L., Stavast, W.J.A., Johnson, D.A., and Barton, M.D., 2005, Porphyry deposits: Characteristics and origin of hypogene features, in Hedenquist, J. W., Thompson, J. F. H., Goldfarb, R. J., and Richards, J. P., eds.: Economic Geology 100th Anniversary Volume, p.251-298 Sharman, E.R., Penniston-Dorland, S.C., Kinnaird, J.A., Nex, P.A.M., Brown, M., Wing, B.A., 2013. Primary origin of marginal Ni-Cu-(PGE) mineralization in layered intrusions: ∆33 S, evidence from the Platreef, Bushveld, South Africa. Economic Geology, 108: 365-377. Shelton, K.L. and Rye, D.M., 1982. Sulfur isotopic compositions of ores from Mines Gaspé, Quebec: An example of sulfate-sulfide isotopic disequilibria in oreforming fluids with applications to other porphyry-type deposits, Economic Geology, 77: 1688-1709. Sillitoe, R.H., 1985. Ore-related breccias in volcanoplutonic arcs: Economic Geology, 80: 1467-1514. Sillitoe, R.H., 2010. Porphyry copper systems, Economic Geology, 105: 3-41. Skewes, M.A., Arévalo, A., Floody, R., Zuñiga, P.H., and Stern, C.R., 2002. The giant El
AC C
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
ACCEPTED MANUSCRIPT 42
EP
TE D
M AN U
SC
RI PT
Teniente breccia deposit: Hypogene copper distribution and emplacement: Society of Economic Geologists Special Publication 9: 299–332. Skewes, M.A., Holmgren, C., Stern, C.R., 2003. The Donoso copper-rich, tourmalinebearing breccia pipe in central Chile: petrologic, fluid inclusion and stable isotope evidence for an origin from magmatic fluids, Mineralium Deposita, 38: 2–21. So, C.-S. and Shelton, K.L., 1983. A sulfur isotopic and fluid inclusion study of the CuW-bearing tourmaline breccia pipe, Ilkwang Mine, Republic of Korea, Economic Geology, 78: 329-333. Söhnge, P.G., 1945. The geology of the Messina copper mines and surrounding country: Geological Survey of South Africa, Memoir 40: 272 p. Taylor, R., 2009. Ore textures - recognition and interpretation, Springer, Dordrecht, 288p. Taylor, R.G., and Pollard, P.J., 1993. Mineralized breccia systems, methods of recognition and interpretation: James Cook University, Geology Department, Economic Geology Research Unit Contribution 46: 31 p. Tornos, F., Galindo, C., Casquet, C., Pevida, L.R., Martínez, C., Martínez, E., Velasco, F., and Iriondo, A., 2006. The Aguablanca Ni-(Cu) sulfide deposit, SW Spain: geologic and geochemical controls and the relationship with a midcrustal layered mafic complex. Mineralium Deposita, 41: 737-769. Truswell, J.F., 1977. The geological evolution of South Africa, Purnell, Cape Town, 218 p. Van Breemen, O. and Dodson, M. H., 1972. Metamorphic chronology of the Limpopo Belt, southern Africa, Geological Society of America Bulletin, 83: 2005-2018. Van Graan, S.J., 1964. The geology of the Messina copper deposits: in S.H. Haughton (ed.), The geology of ore deposits in southern Africa, v. II, Geological Society of South Africa, p. 303-314. Vargas, R., Gustafson, L.B., Vucasovic, M., Tidy, E., and Skewes, M.A., 1999. Ore breccias in the Rio Blanco-Los Bronces porphyry copper deposit, Chile: Society of Economic Geologists Special Publication 7: 281-297. Vaughan, D.J., 2011. Sulphides: in J.F.W. Bowles, R.A. Howie, D.J. Vaughan and J. Zussman (eds.), Rock-forming minerals, 2nd Edition, Non silicates: Oxides, hydroxides and Sulfides, volume 5A, The Geological Society, London, p. 627892. Warnaars, F.W., Holmgren , C., and Barassi, S., 1985. Porphyry copper and tourmaline breccias at Los Bronces-Río Blanco, Chile: Economic Geology, 80: 1544-1565. Wenk, H.-R. and Bulakh, A., 2004. Minerals: their constitution and origin, Cambridge University Press, Cambridge, UK, 646 p. Williams, W.C., Meissl, E., Madrid, J., and de Machuca, B.C., 1999. The San Jorge porphyry copper deposit, Mendoza, Argentina: a combination of orthomagmatic and hydrothermal mineralization, Ore Geology Reviews, 14: 185-201. Wilson, M.G.C., 1998. Copper, in Wilson, M.G.C. and Anhaeusser, C.R., (eds.), Mineral Resources of South Africa, Council for Geoscience, Pretoria, 6th Edition, pp. 209-227. Woodstock, N.H. and Mort, K., 2008. Classification of fault breccias and related fault rocks, Geological Magazine, 145: 435-440. Yanagisawa, F. and Sakai, H., 1983. Thermal decomposition of barium sulfate-vanadium
AC C
1069 1070 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 1113 1114
ACCEPTED MANUSCRIPT 43
EP
TE D
M AN U
SC
RI PT
pentoxide-silica glass mixtures for preparation of sulfur dioxide in sulfur isotope ratio measurements: Analytical Chemistry, 55: 985-987. Yang, K. and Bodnar, R.J., 2004. Orthomagmatic deposit for the Ilkwang Cu-W brecciapipe deposit, southeastern Kyongsang Basin, South Korea, Journal of Asian Earth Sciences, 24: 259-270. Zeh, A., Gerdes, A., Klemd, R. and Barton, J. M., Jr., 2007. Archaean to Proterozoic crustal evolution in the Central Zone of the Limpopo Belt (South AfricaBotswana): Constraints from combined U-Pb and Lu-Hf isotope analyses of zircon. Journal of Petrology, 48: 1605-1639. Zhang, Y.H., Zhang, S.H. and Pirajno, F., 2007. Fluidization: an important process in the formation of the Qiyugou Au-bearing breccia pipe in central China. Acta Geologica Sinica, 81: 226-238.
AC C
1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 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
ACCEPTED MANUSCRIPT 44 Figure captions
RI PT
Fig. 1. Simplified geological map of part of Southern Africa showing the location of Musina in the Central Zone of the Limpopo orogenic belt. Abbreviations: GGT – granitic greenstone terranes of the cratons, NMZ – North Marginal Zone, CZ – Central Zone, SMZ – South Marginal Zone. Proterozoic and younger formations are unornamented. The location of Fig. 2a is also shown.
SC
Fig. 2. (a) Simplified geologic map of part of the Limpopo orogeny showing location of the two groups of samples used in this study. (b) shows detailed geology of an area between Messina Mine and Artonvilla Mine (both maps modified from Jacobsen and McCarthy, 1976). Abbreviations: ZW – Zimbabwe, L.R. – Limpopo River, Undifferenti. – undifferentiated.
M AN U
Fig. 3. Generalized geology in the vicinity of the Musina copper deposits showing extent of underground workings (Modified from Bahnemann, 1986). Fig. 4. (a) Vertical projection of various lodes in the Campbell Mine area transverse to the Musina Fault in a SW view, and (b) longitudinal projection of the Harper mining area parallel to the Musina Fault in a NW view (After Bahnemann, 1986). Note the association of sulfide mineralization with breccias in both deposits.
EP
TE D
Fig. 5. Photographs showing textures of brecciated quartz and amphibole macroscopic crystals occurring in association with copper sulfide mineralization from the Musina copper deposits. Pencil shown for scale is ~18 cm long. (a) Samples from Artonvilla Mine showing breccias that are composed of mostly amphibole, potassium feldspar, and angular quartz. (b) Campbell Mine sample showing very coarse-grained and sometimes rounded angular quartz, amphibole, and potassium feldspar clasts. (c) Brecciated sample from Campbell Mine of mostly pyrite hosted in quartz veins. (d) Brecciated sample showing rounded and angular quartz, chalcopyrite, pyrite, and amphibole. Width of photograph is ~25 cm. Specimen from Campbell Mine area. Fig. 6. Photomicrographs for samples from Artonvilla Mine. Scale bar shown in (a) is applicable to all photomicrographs. (a) and (b): The two fields of view are occupied by hydrothermally altered clasts (altered to micas) and amphibole crystals, plane plolarized light (PPL) for both images, sample AtBr02. In (b) angular points are consistent with a breccia texture. (c): photomicrograph showing angular points at the contact of amphibole and mica clasts which is consistent with broken rock, sample ATBr02, PPL. (d) shows rounded (milled) mica clasts in amphibole which are consistent with rotated, intrusive breccias, sample ATBr02, PPL. (e): Field of view dominated by disseminated sulfides (Sulf), calcite (Cal), and potassium feldspar (Kfs) in PPL (e) and crossed polars (XPL) (f), sample AtCal01. Upper left view shows bifurcating sulfides and muscovite. (g) View dominated by altered plagioclase (Pl) and altered alkali feldspar (Kfs) in granitic gneiss, XPL, sample ATGN04. (h): Granitic gneiss showing sillimanite (Sil) together with biotite (Bt) which occupies smaller portions to the right (h) of center and altered Pl and
AC C
1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205
ACCEPTED MANUSCRIPT 45 Kfs occupying most of the field of view in the photomicrograph, sample AtGn03. The feldspars in (h) are less altered compared to those in previous photomicrographs (a)-(g).
SC
RI PT
Fig. 7. Photomicrographs for samples from Campbell Mine. Scale bar shown in (a) is same for all photomicrographs. (a): Sulfides (Sulf), sericite (Ser), and potassium feldspar (Kfs) in quartz vein in crossed polars (XPL), sample CpQV01. (b) shows an orthoclase (Or) crystal displaying typical Carlsbad twinning together with other alkali feldspar crystals (Kfs), XPL. (c) shows hydrothermally altered Kfs crystals with sulfides associated with the veinlets together with altered Kfs being replaced by sericite (Ser), sample CpQv01, XPL. (d): view dominated by Ser with remnant Kfs at center of photomicrograph, view of a breccia dominated by Ser and Kfs in XPL, sample CpQv01. (e) Highly altered amphibolite showing field of view dominated by amphibole and minor feldspar (Fsp), XPL, sample CpAM04. (f): muscovite (Ms) and chlorite (Chl) at center of photomicrograph in breccia, together with Kfs in XPL, sample CpQv02.
TE D
M AN U
Fig. 8. Backscattered electron images (BEI) from Artonvilla and Campbell Mines. (a) and (b) are BSE images showing coarse-grained disseminated pyrite occurring together with Ser, Ms, and Kfs and from Artonvilla Mine, sample AtCal01. Upper left view of (a) shows bifurcating sulfides and muscovite. (c): BEI showing very fine-grained chalcopyrite (Ccp) crystals completely enclosed in an amphibole crystal, sample CpQv02. (d): BSE image showing fine-grained chalcopyrite crystals completely enclosed in feldspar crystals (Kfs), sample CpAm04. (e): BSE image showing fine-grained chalcopyrite crystals occurring in association with amphibole (amp) and plagioclase feldspar crystals (Pl), sample CpAm04. Note one chalcopyrite crystal in center right “enclosing” a very fine-grained Pl crystal. (f): BSE image showing fine-grained chalcopyrite crystals completely enclosed in plagioclase (Pl) crystals, sample CpAm04.
EP
Fig. 9. Plots of pyrite and chalcopyrite compositions from Artonvilla and Campbell Mines. Plot of wt. % Fe versus wt. % S (a), wt. % Ni versus wt. % Fe (b), and wt. % Co versus wt. % Fe (c) from Artonvilla Mine. Plot of wt. % Fe versus wt. % S (d), wt. % Cu versus wt. % S (e), and wt. % Fe versus wt. % Cu (f) from Campbell Mine. Fig. 10. (a) Frequency plot of pyrite or chalcopyrite δ34S values for samples from both Artonvilla and Campbell Mines. (b) Frequency plot of pyrite or chalcopyrite δ34S values for samples from both the Artonvilla and Campbell Mines.
AC C
1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246 1247 1248 1249 1250 1251
Fig, 11. Plots of pyrite and chalcopyrite compositions from Artonvilla and Campbell Mines compared to compositions from Southwest Indian Ridge (Miller and Cervantes, 2002), Southwest Oregon ophiolite (Foose, 1986), the Duluth Complex (Pasteris, 1984), and the Keivitsansarvi Ni-Cu-PGE deposit, Finland (Gervilla and Kojonen, 2002). Plot of wt. % Ni versus wt. % Fe (a), wt. % Co versus wt. % Fe (b), and wt. % Fe versus wt. % Cu (c) from Artonvilla Mine. Plot of wt. % Cu versus wt. % Pb (d) from Campbell Mine. Fig. 12. (a) Pyrite compositions, in wt. %, from Artonvilla Mine plotted on S-Fe-Ni ternary diagram. The labeled dashed fields and the shaded field are high-temperature monosulfide solid solution (mss) fields of Kullerud et al. (1969), Py – pyrite. (b)
ACCEPTED MANUSCRIPT 46 Simplified phase relations for Campbell Mine chalcopyrite samples in the S-Fe-Cu system at 300oC (Barton and Skinner, 1979). Abbreviations for phases are bn – bornite, iss – intermediate solid solution, py – pyrite, po – pyrrhotite, cc – chalcocite, ccp – chalcopyrite, and cv – covellite, The composition of an ideal ccp is also shown.
EP
TE D
M AN U
SC
RI PT
Fig. 13. (a) Frequency plot of δ34S values for samples from both the Artonvilla and Campbell Mines together with results from Sawkins and Rye (1979) for pyrite, chalcopyrite, bornite and chalcocite. Mantos Blancos breccia pyrite δ34S data from Ramírez et al. (2006) and Ilkwang breccia pipe, South Korea, data from So and Shelton (1983) and Yang and Bodnar (2004). (b) δ34S mixing diagram for the Artonvilla and Campbell Mines using δ34S value of 0 ‰ and a S concentration of 800 ppm; a fresh biotite-garnet-cordierite gneisses contaminant with a S concentration of 10,000 ppm (Jacobsen and McCarthy, 1976) and a δ34S value of 8.2 ‰ (Table 4).
AC C
1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264
ACCEPTED MANUSCRIPT
Table 1. Description of samples from Campbell and Artonvilla Mines for sulfur isotopes
Mafic
Medium
Mafic
Low
Coarse grained
Medium to coarse grained, gneissic foliation Coarse grained, foliated
Atgn05
Quartz-feldsparthic gneiss
Low
Felsic to mafic felsic
CpQv01 CpQv02
Quartz vein Amphibolite with mineralization within quartz veins Amphibolite: minor quartz veins hosting sulfides
Medium Low
Felsic Mafic
Coarse grained Coarse grained
Low
Mafic
Medium to coarse grained
CpAm04
Mineral composition
RI PT
Low
Mafic
Texture/ structure Medium to coarse grained, veinlets of quartz/calcite Medium to coarse grained
TE D
Atgn04
Color
Amphiboles, mafic minerals, calcite, quartz, disseminated sulfides within amphibole Amphiboles, quartz, minor micas, low content of sulfides
SC
Atgn03
Amphibole with elongated calcareous and quartz crystals forming veins Brecciated amphibole with angular and veined quartz grains Biotite garnet gneiss with disseminated sulfide mineralization Biotite gneiss with disseminated sulfides
Degree of alteration Low
EP
Campbell
Artonvilla
Atbr02
Rock type
M AN U
Sample ID Atcal01
AC C
Deposit
1
Amount sulfide mineralization Medium
Low
Layers of mafic and felsic minerals, micas and spotted garnets
Medium
Alternating layers of mafic and felsic minerals, minor spots of garnets, sulfides disseminated No fully developed interlayers of quartz feldsparthic and minor mafic minerals Quartz , chalcopyrite, pyrite Quartz , amphibole, chalcopyrite, pyrite, bornite,
Low
Amphiboles, Chalcopyrite- pyrite, quartz, chalcopyrite partly oxidized
Very low
Medium-high Low
Low
ACCEPTED MANUSCRIPT
Table 2. Representative pyrite chemical analyses b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. 0.0053
Pb (wt.%)
b.d.l. b.d.l. b.d.l. b.d.l. 0.025 0.009 0.014 b.d.l. b.d.l. 0.01 b.d.l. 0.012 0.0044
0.056 0.068 0.149 0.155 0.158 0.064 0.148 0.189 b.d.l. 0.142 0.097 0.068 0.0273
Fe (wt.%)
1
Co (wt.%)
RI PT
Cu (wt.%)
46.037 46.073 45.332 45.005 46.058 45.93 46.255 46.104 44.72 45.194 44.549 45.434 0.0027
SC
b.d.l. b.d.l. b.d.l. 0.438 b.d.l. b.d.l. b.d.l. 0.008 0.005 0.074 0.062 0.07 0.0043
Zn (wt.%)
M AN U
53.784 53.086 52.815 53.193 53.268 52.697 53.176 53.285 51.436 53.934 55.959 52.238 0.003
Ni (wt.%)
TE D
ATCA1-P76 ATCA1-P77 ATCA1-P78 ATCA1-P79 ATCA1-P81 ATCA1-P82 ATCA1-P83 ATCA1-P84 ATCA1-P85 ATGN04-P1 ATGN04-P2 ATGN04-P3 M.D.L.
S (wt.%)
EP
76 77 78 79 81 82 83 84 85 1 2 3
Comment
AC C
Point
0.078 0.071 0.088 0.228 0.094 0.1 0.086 0.072 0.074 0.113 0.215 0.145 0.0033
Mn (wt.%)
b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. 0.0032
Total
99.958 99.298 98.39 99.028 99.613 98.805 99.681 99.663 96.235 99.467 100.882 97.967
ACCEPTED MANUSCRIPT
CPQV01-P17 CPQV01-P18 M.D.L.
0.007 b.d.l. 0.006 b.d.l. b.d.l. b.d.l. b.d.l. b.d.l. 0.006 b.d.l. b.d.l. b.d.l. 0.0043
SC
33.706 33.088 33.813 33.794 34.115 34.034 33.589 33.697 34.101 33.673 35.706 36.34 0.003
Zn Cu (wt.%) Pb (wt.%) Fe (wt.%) Co Mn Total (wt.%) (wt.%) (wt.%) b.d.l. 33.474 b.d.l. 29.432 0.057 b.d.l. 96.68 b.d.l. 33.591 0.214 29.051 0.074 b.d.l. 96.018 b.d.l. 33.405 0.038 29.489 0.053 b.d.l. 96.804 b.d.l. 33.722 0.1 29.742 0.069 b.d.l. 97.427 b.d.l. 33.452 b.d.l. 29.477 0.067 b.d.l. 97.128 b.d.l. 33.473 0.079 29.475 0.058 b.d.l. 97.126 b.d.l. 33.501 0.147 29.488 0.046 b.d.l. 96.771 b.d.l. 33.702 0.108 29.63 0.074 b.d.l. 97.23 b.d.l. 33.626 0.048 29.584 0.055 b.d.l. 97.42 b.d.l. 33.438 0.03 29.336 0.059 b.d.l. 96.557 b.d.l. 33.798 b.d.l. 29.117 0.053 b.d.l. 98.718 b.d.l. 32.861 0.108 28.096 0.066 0.004 97.479 0.0053 0.0044 0.0273 0.0027 0.0033 0.0032
M AN U
CPAM04-P1 CPAM04-P2 CPAM04-P3 CPAM04-P4 CPAM04-P5 CPAM04-P13 CPAM04-P14 CPAM04-P15 CPAM04-P16 CPAM04-P17
Ni (wt.%)
TE D
1 2 3 4 5 13 14 15 16 17 17 18
S (wt.%)
EP
Comment
AC C
Point
RI PT
Table 3. Representative chalcopyrite chemical analyses
1
AC C
EP
TE D
SC
M AN U
Table 4. δ34S(CDT) data for Artonvilla Mine and Campbell Mine samples Mine Sample Phase δ34S (‰, VCDT) Artonvilla AtCal01py pyrite 3.1 Artonvilla AtCal01py2 pyrite 3.6 Artonvilla AtCal01Ccp chalcopyrite 3.9 Artonvilla AtGn04 wr whole-rock 8.2 Campbell CpQv01py pyrite 0.5 Campbell CpQv01Ccp2 chalcopyrite 0.7 Campbell CpQv01Ccp chalcopyrite 0.4 Campbell CpQv02Ccp chalcopyrite -0.3 Campbell CpAm04Ccp chalcopyrite 0.3 Campbell CpAm04Ccp2 chalcopyrite -1.1 Campbell CpQv01bn bornite 0.4
RI PT
ACCEPTED MANUSCRIPT
1
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT