Journal Pre-proofs Fractionation of Cadmium isotope caused by vapor-liquid partitioning in hydrothermal ore-forming system: A case study of the Zhaxikang Sb–Pb–Zn–Ag deposit in Southern Tibet Da Wang, Youye Zheng, Ryan Mathur, Miao Yu PII: DOI: Reference:
S0169-1368(19)30139-8 https://doi.org/10.1016/j.oregeorev.2020.103400 OREGEO 103400
To appear in:
Ore Geology Reviews
Received Date: Revised Date: Accepted Date:
16 February 2019 26 January 2020 9 February 2020
Please cite this article as: D. Wang, Y. Zheng, R. Mathur, M. Yu, Fractionation of Cadmium isotope caused by vapor-liquid partitioning in hydrothermal ore-forming system: A case study of the Zhaxikang Sb–Pb–Zn–Ag deposit in Southern Tibet, Ore Geology Reviews (2020), doi: https://doi.org/10.1016/j.oregeorev.2020.103400
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. 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.
© 2020 Elsevier B.V. All rights reserved.
1
Fractionation of Cadmium isotope caused by vapor-liquid partitioning in
2
hydrothermal ore-forming system: A case study of the Zhaxikang Sb–Pb–Zn–
3
Ag deposit in Southern Tibet
4
Da Wang1,2, Youye Zheng1,3*, Ryan Mathur2*, Miao Yu4
5
1State
6
and Resources, China University of Geosciences, Beijing 100083, China
7
2Department
8
3State
9
Resources, China University of Geosciences, Wuhan 430074, China
Key Laboratory of Geological Processes and Mineral Resources, School of Earth Sciences
of Geology, Juniata College, Huntingdon, Pennsylvania 16652, USA
Key Laboratory of Geological Processes and Mineral Resources, and Faculty of Earth
10
4Beijing
11
*Corresponding author: Youye Zheng:
[email protected]
12
Institute of Geology, Beijing, 100120, China
Ryan Mathur:
[email protected]
1
13
Abstract
14
Here, we conducted a case study of the Zhaxikang Sb–Pb–Zn–Ag deposit to explore Cd
15
isotopic fractionation mechanisms in hydrothermal ore-forming system. The δ114/110Cd values of
16
sphalerite systematically decrease from ore mineralization stage 1 (–0.30‰ to 1.01‰; average
17
value = 0.17‰; n = 4) through stage 2 (–0.51‰ to –0.09‰; average value = –0.23‰; n = 3) to
18
stage 3 (–0.34‰ to –0.23‰; average value = –0.285‰; n = 2). A simple Rayleigh distillation
19
models this temporally decreasing trend resulted from Cd isotopic fractionation that is most likely
20
related to vapor-liquid dynamics of ore-forming fluid. This mechanism for Cd isotopic
21
fractionation is further augmented by the general geochemical characteristics and fluid inclusion
22
data of sulfides and cogenetic gangue minerals. Firstly, the sphalerite has low Cd concentrations
23
(1183 – 2199 ppm), correspondingly high Zn/Cd ratios (248 – 421), and large Cd isotopic variation
24
range with relatively lower δ114/110Cd values (–0.51‰ to 1.01‰). Commonly, sphalerite from
25
sedimentary exhalative systems possess these characteristics caused by vapor-liquid interactions.
26
Secondly, fluid inclusion data from the Pb–Zn sulfides and cogenetic carbonate alteration minerals
27
indicates that vapor-liquid two-phase inclusions are in dominance (more than 90%). Meanwhile,
28
the vapor-liquid ratios (20% – 50%) of these two-phase inclusions are consistent with those of the
29
established Cd (25.48% – 55.07%) and Zn (26.70% – 48.70%) isotopic Rayleigh distillation model.
30
To conclude, vapor-liquid partitioning is main cause for observed Cd isotopic variations in
31
Zhaxikang deposit, and is demonstrated as an important Cd isotopic fractionation mechanism in
32
hydrothermal ore-forming system.
33
Keywords: Cd isotope; Rayleigh distillation model; vapor-liquid partitioning; hydrothermal ore-
2
34
forming system; Zhaxikang deposit
35 36
1. Introduction
37
One of the fundamental questions associated with ore deposits is to understand where metals
38
originated and how these systems became endowed with the metal budget they possess. Relatively
39
new techniques of transition metal and metal isotope have begun to further advance our
40
understanding of metal source, migration and concentration in the crust (Duan et al., 2016;
41
Saunders et al., 2015; Lv et al., 2016; Mathur et al., 2018). The advantage of using these metal
42
isotopic systems is that we can derive information directly from the ore minerals themselves, which
43
has led to the measurement of significant Fe-Cu-Zn-Mo-Sn isotopic variations in sulfides from
44
different types of ore deposits (e.g., Wilkinson et al., 2005; Markl et al., 2006; Mathur et al., 2009,
45
2012; Yao et al., 2016, 2018; Wang et al., 2017; Wu et al., 2017; Gao et al., 2018).
46
From a geochemical perspective, Cadmium (Cd) is isotopically unique. The reason is that the
47
Cd has a large number (8) of naturally formed stable isotopes (Cd106: 1.25‰, Cd108: 0.89‰,
48
Cd110:12.5‰, Cd111: 12.8‰, Cd112: 24.1‰, Cd113: 12.2‰, Cd114: 28.7‰ and Cd116: 7.49‰) and
49
exhibits a great mass range (108 to 116 amu) on the periodic table (Cloquet et al., 2005; Zhu et al.,
50
2015a, b). The relative mass difference of the isotopes and the larger mass range of Cd may allow
51
for monitoring the fractionation of a heavy metal within hydrothermal systems (Tu et al., 2004;
52
Yao et al., 2018). With the regards to ore deposit investigation, Cd is an economically significant
53
metal that may exhibit sulphophile, volatile and lithophile behavior. Because of the difficulty to
54
form independent minerals, Cd is found enriched by isomorphic replacement with other elements
3
55
(e.g., replace Zn in sphalerite) within the mineral structure. This is the case in Pb–Zn deposits
56
where the element is mined economically. Accordingly, the related research of Cd isotopes has the
57
potential to trace metal sources and provide insights into ore-forming processes within Pb–Zn
58
deposits.
59
The attempts of Cd isotopic measurement can date back to 1970s (Rosman and De Laeter,
60
1975, 1976, 1978), the analytical precision was limited at that time yet. In recent years, the
61
optimized TIMS (thermal ionization mass spectrometer) double spiked method (e.g., Sands et al.,
62
2001; Schmitt et al., 2009) and newly-developed MC-ICP-MS (Multicollector-Inductively
63
Coupled Plasma Mass Spectrometer) measuring technique (e.g., Lacan et al., 2006; Gao et al., 2008)
64
have much improved the analytical precision (δ114/110Cd: ± 0.10‰; Ripperger et al., 2007).
65
Currently, measurable Cd isotopic fractionation has been demonstrated to occur in nature and
66
significant Cd isotopic variations have been reported in ores (Wen et al., 2015, 2016; Yang et al.,
67
2015; Zhu et al., 2016a, 2017), igneous rocks (Wombacher et al., 2003; Schmitt et al., 2009),
68
carbonate and sedimentary rocks (Wombacher et al., 2003; Horner et al., 2011), chondrite and lunar
69
samples (Rosman and De Laeter., 1975; Sands et al., 2001; Schediwy et al., 2006; Wombacher et
70
al., 2004, 2008), sea and river water (Lacan et al., 2006; Ripperger et al., 2007; Abouchami et al.,
71
2011; Gault-Ringold et al., 2012; Yang et al., 2012; Lambelet et al., 2013), soil and sediments (Gao
72
et al., 2008; Schmitt et al., 2009; Zhang et al., 2016). Cd isotopic fractionation reported in these
73
contributions has been attributed to evaporation and condensation, biological and inorganic
74
processes (Wang et al., 2013; Zhu et al., 2013).
75
The research of these works focused on factors for Cd isotopic variation related to
4
76
environmental science, marine science and cosmochemistry. Mechanisms for Cd isotopic
77
fractionation in hydrothermal ore-forming system are nascently studied, with only the kinetic
78
Rayleigh fractionation mechanism related to mineral precipitation (solid-liquid partitioning; Wen
79
et al., 2016; Zhu et al., 2017). In hydrothermal ore-forming system, the kinetic Rayleigh
80
fractionation related to vapor-liquid partitioning is also considered as an important transition metal
81
isotopic fractionation mechanism (e.g., Cu: Graham et al., 2004; Cu and Mo: Yao et al., 2016; Fe
82
and Zn: Wang et al., 2018a). Therefore, we carried out a case study of the super-large complex
83
polymetallic Zhaxikang deposit in Southern Tibet, which is representative and ideal for Cd isotope
84
related study in hydrothermal ore-forming system. The investigation centers on δ114/110Cd values
85
of sulfides and establishment of Cd isotopic Rayleigh distillation model to demonstrate the
86
influence of vapor-liquid partitioning on observed Cd isotopic variations.
87 88
2. Geological Background
89
The North Himalayan Metallogenic Belt (NHMB) is an important component of the Tethys-
90
Himalaya Metallogenic Domain (Fig. 1A). There are three main regional mineralization events
91
within NHMB that have generated a series of Sb, Sb–Au, Au, W–Sn(Be), Pb–Zn(Ag) and Sb–Pb–
92
Zn–Ag deposits (Fig. 1B; Yang et al., 2009; Zheng et al., 2012, 2014; Li et al., 2017). The first
93
regional mineralization event is related to multiple seafloor volcanic events during synsedimentary
94
period (220 to 130 Ma); the second regional mineralization event is associated with the
95
metamorphic fluid system during syn-collision period (60 to 42 Ma); and the third regional
96
mineralization event relates to the magmatic-hydrothermal activity during post-collision period (25
5
97
Ma to now; Wang et al., 2018b). As the regional geology of NHMB has been described in detail
98
by several previous studies (e.g., Wang et al., 2017; Sun et al., 2018), it is not necessary to be
99
repeated here.
100
The Zhaxikang deposit, in the location of ~48 km west from Longzi County Town, is the only
101
super-large deposit identified within NHMB (Fig. 1B). According to the latest exploration results
102
of Huayu Mining Company, the Zhaxikang deposit contains Zn+Pb 2.066 Mt at average grade of
103
6.38%, Sb 0.235 Mt at an average grade of 1.14%, Ag 2660.6 t at an average grade of 101.64g/t,
104
Au 10.4 t at an average grade of 2.9 g/t, associated Ga 361 t, and 20 Mt Mn–Fe carbonate ores at
105
an average grade of 42% for Fe+Mn. The Lower Jurassic Ridang formation consists of epi-
106
metamorphic marine clastic rocks and hosts majority of the mineralization within the orefield (Fig.
107
1C). Some Upper Jurassic Weimei formation, composed of fine-grained metamorphic quartzose
108
sandstone, silty slate, and calcarenite, also outcrops within the orefield. Meanwhile, there is a small
109
amount of Quaternary sediments accumulated within valleys (Fig. 1C; Zheng et al., 2012). The
110
engineering and geological mapping projects have identified sixteen faults (F1 – F16) that host
111
nine orebodies (I–IX; Fig. 1C and D) within the orefield. Among these orebodies, orebody V is the
112
largest and richest that hosts more than 80% of the reserves. The orefield records magmatism that
113
have formed diabase, porphyritic rhyolite, basalt, leucogranite units and some granite porphyry
114
dykes (Fig. 1C). Various types of alteration associated with mineralization have occurred within
115
the orefield, such as the silicification associated with Sb mineralization, carbonatization that is
116
related to Pb–Zn mineralization in the form of Mn–Fe carbonate veins, the chloritization, weak
117
sericitization and clayization. In addition, ore-forming elements exhibit a vertical sequence that is
6
118
zoned from a lowermost Zn (Pb + Ag) zone through a central Zn + Pb + Ag–(Sb) zone to an
119
uppermost Pb + Zn + Sb + Ag zone, whereas there is no horizontal zoning (Wang et al., 2018a).
120
The ore paragenetic sequence in Zhaxikang deposit has been divided into six stages of ore
121
formation based on hand specimen and microscopic observations (Fig. 2). These six stages are
122
assigned to two clear pulses. The first pulse (Pb–Zn mineralization) consists of stages 1 and 2 that
123
is primarily dominated by Mn–Fe carbonates and sulfides (Figs. 2 and 3A-D). The second pulse
124
(Sb mineralization) includes stages 3 to 6 that is principally characterized by quartz, calcite,
125
sulfosalt minerals, and sulfides (Figs. 2 and 3E-I). As the transitional stage between the two pulses
126
of mineralization, the sulfides in stage 3 mainly form by replacement of earlier sulfides (Wang et
127
al., 2017, 2018a). Additionally, a supergene stage is distinguished by the formation of ferrohydrite,
128
smithsonite, sardinianite, valentinite, travertine, malachite and siliceous sinter in the shallowest
129
elevation (Fig. 2).
130 131
3. Sampling and Analytical methods
132
The most common isomorphic replacement for Cd is to replace Zn and then enter into the
133
crystal lattices of sphalerite, hence sphalerite is the most ideal mineral for Cd isotopic research in
134
hydrothermal ore-forming system. Due to the absence of Zn-bearing minerals from stages 4 to 6,
135
nine sphalerite and four galena samples from stages 1 to 3 are chosen for relevant analyses in
136
Zhaxikang deposit. In these samples, there are four sphalerite-galena coexisting mineral pairs.
137
These ore samples were crushed to around 100 – 200 mesh for separation of sphalerite and 200 –
138
300 mesh for separation of galena, then the separations of sulfide crystals were prepared with
7
139
careful handpicking under a binocular microscope on the basis of size, clarity, color, and
140
morphology to achieve a purity of 99.99%.
141
The measurement of Zn and Cd concentrations were conducted on an ICP–OES (Inductively
142
Coupled Plasma Optical Emission Spectrometer) at Pennsylvania State University. Approximately
143
20 to 70 mg of sulfide powders were dissolved in 4 ml of heated (120℃) ultrapure aquaregia for
144
12 hours. Then the aliquots from each solution were acidified and diluted in 2% nitric acid for
145
chemical analysis. Zn and Cd concentrations were determined with standard calibration curves that
146
ranged from 0.6 to 10 ppm and Indium was used as an internal standard for analysis. The errors of
147
the concentration data are usually within 5%, which were used to determine the amount of Cd
148
needed for isotope analysis.
149
The Cd isotopic compositions were measured on the Neptune MC–ICP–MS at Pennsylvania
150
State University. Cd was purified using the anion exchange chromatography (Cloquet et al., 2005)
151
with volumetric yields for the samples greater than 94% after two rounds of column
152
chromatography. In this procedure, 2ml of wet BioRad AG MP-1 resin chloride form (100 – 200
153
mesh) was added to a 10ml BioRad chromatography column. The resin was sequentially cleaned
154
with 10ml of 2% HNO3, 10ml of MQ water (18.2), 5ml and of 1.2 molar HCl. The sample was
155
loaded onto the resin with 1ml of 1.2 molar HCl and the unwanted ions were sequentially eluted
156
with another 4ml of 1.2 molar HCl, 15ml of 0.3 molar HCl and 16ml of 0.012 molar HCl. The Cd
157
was collected in 17ml of 0.0012 molar HCl in the final elution. This process was repeated with the
158
use of new resin for the second column to eliminate Sn. The 116Sn mass was monitored in H4 cup,
159
with 107Ag in L4 cup, 109Ag in L2 cup, 110Cd in L1 cup, 111Cd in Ax cup, 112Cd in H1 cup, 113Cd in
8
160
H2 cup, 114Cd in H3 cup (Table 1). Instrumentation setup and introduction was similar to Wasylenki
161
et al. (2014). All samples were doped with 100 ppb NIST 987 Ag isotope standard which was used
162
to correct for mass bias using the exponential fractionation correction (Marechal et al., 1999). The
163
107Ag/109Ag
164
at 100 ppb Cd, with on peak blank subtraction in 2 blocks of 15 ratios. The reported values are an
165
average of 2 separate measurements, and the data are presented relative to the NIST SRM 3108 Cd
166
of the NIST 987 Ag isotope standard is reported at 1.07638. Solutions were measured
standard in per mil notation defined as: δ114/110Cd (‰) =
(
)
( )sample ―1 ( ) NIST SRM 3108 114Cd 110Cd
114Cd 110Cd
× 1000
167
(Abouchami et al., 2013). All the cited data from previous literatures in this paper are converted
168
relative to NIST SRM 3108 Cd standard according to Cloquet et al. (2005); Abouchami et al. (2013)
169
and Zhu et al. (2017). The variation of NIST SRM 3108 throughout the measuring session was 0.04
170
‰ (2 n=14) and is considered the error of measurements.
171 172
4. Results
173
All the analytical results are listed in Table 2. The galena has Zn concentrations of 0.030% –
174
0.153% (average value: 0.084%, n = 4) and Cd concentrations of 14 – 26 ppm (average value: 20
175
ppm, n = 4), respectively. The Zn/Cd ratios of galena range from 14 to 82 (average value: 41, n =
176
4). Besides, stage 2 galena shows δ114/110Cd values of –2.29‰ and 2.24‰, and stage 3 galena
177
exhibits δ114/110Cd values of –0.21‰ and 0.01‰. By contrast, the sphalerite has Zn and Cd
178
concentrations of 40.456% – 58.440% (average value: 51.502%, n = 9) and 1183 – 2199 ppm
179
(average value: 1596 ppm, n = 9), respectively. Meanwhile, the Zn/Cd ratios of sphalerite are 248
9
180
– 421 (average value: 332, n = 9). The overall δ114/110Cd values of sphalerite range from –0.51‰
181
to 1.01‰ with an average value of –0.08‰ ± 0.83‰ (2SD, n = 9). Separately, δ114/110Cd values of
182
sphalerite in different mineralization stages are as follow: (1) stage 1: –0.30‰ to 1.01‰ (average
183
value: 0.17‰ ± 1.01‰, 2SD, n = 4); (2) stage 2: –0.51‰ to –0.09‰ (average value: –0.23‰ ±
184
0.39‰, 2SD, n = 3); (3) stage 3: –0.34‰ to –0.23‰ (average value: –0.285‰ ± 0.11‰, 2SD, n =
185
2).
186 187
5. Discussion
188
5.1 The occurrence of Cd in sulfides
189
The occurrence of Zn and Cd in galena have been studied and used to relate to specific ore
190
genetic models (Palero-Fernández and MartínIzard, 2005; Zhou et al., 2011). In this instance, their
191
data has been used to understand the Zhaxikang and Fule MVT deposits. Specifically, both the
192
Zn2+ and Cd2+ should enter into the galena by directly isomorphic replacement with Pb2+ in
193
Zhaxikang deposit, which differs from the situation in Fule MVT deposit that both Zn and Cd
194
present as sphalerite micro-inclusions within galena (Zhu et al., 2017). This inference is
195
demonstrated by the following comparisons: (1) Zhaxikang deposit: ①No correlation between Zn
196
and Cd concentrations for galena samples (Fig. 4A); ②Cd concentrations (14 – 26 ppm; average
197
value = 20 ppm, n = 4) of galena samples fall into the range of hydrothermal deposits (10 – 500
198
ppm; Schwartz, 2000), but much lower than those of Fule MVT deposit (48 – 1163 ppm; average
199
value = 367.3 ppm; n = 6; Fig. 4C) (T test p=0.0548); ③Zn/Cd ratios (14 – 82) of galena samples
200
are not within the range of sphalerite (248 – 421); ④The BSE (Back Scattered Electron) image
10
201
(Fig. 5A; Wang et al. 2018a) and photomicrograph (Fig. 5B) show that there is no sphalerite micro-
202
inclusions within galena; (2) Fule MVT deposit: ①Cd/Zn ratios of bulk galena samples (159 –
203
512) are within the range of sphalerite samples (23 – 588); ②There is a positive correlation (R2 =
204
0.94) between Cd and Zn concentrations for galena samples; ③BSE images confirm that sphalerite
205
micro-inclusions exist within galena (Zhu et al., 2017). These Cd characteristics suggest that the
206
occurrence of Zn–Cd within galena in Zhaxikang and Fule deposits are different.
207
On the other hand, two different kinds of isomorphic replacement between Zn2+ and Cd2+
208
within sphalerite have been confirmed by LA–ICP–MS (Laser Ablation Inductively Coupled
209
Plasma Mass Spectrometer) measurements: (1) Cook et al. (2009) suggested that the Zn2+ is directly
210
substituted by Cd2+ during crystallization; whereas (2) Belissont et al. (2014) considered that both
211
Cd2+ and Fe2+ enter into sphalerite by direct substitution with Zn2+. In the study of Fule MVT
212
deposit, the sphalerite has high Cd concentrations (5240 – 35000 ppm) but low Fe concentrations
213
(365 – 970 ppm), thus Zhu et al. (2017) thought that the Cd2+ enter into sphalerite by direct
214
substitution with Zn2+. By comparison, the sphalerite from Zhaxikang deposit has lower Cd
215
concentrations of 1183 – 2199 ppm with no correlation between Cd and Zn concentrations (Fig.
216
4B), but much higher Fe concentrations (4.05% – 9.44%, Sun et al., 2018), which indicates that the
217
Zn2+ is directly substituted by both of Cd2+ and Fe2+ during crystallization of sphalerite in
218
Zhaxikang deposit.
219 220 221
5.2 Comparison with previous Cd isotopic genetic and evolution models By investigating the Cd characteristics of sphalerite that include Cd concentrations, Zn/Cd
11
222
ratios and Cd isotopic compositions, Wen et al. (2016) divided nine Pb–Zn deposits with different
223
geneses in China into three categories: (1) The Low-T (temperature) systems (MVT Pb–Zn deposits)
224
have high Cd concentrations (2415 – 34981 ppm), relatively higher δ114/110Cd values (–0.10‰ to
225
0.59‰) and low Zn/Cd ratios (17 – 201); (2) The High-T systems (porphyry, magmatic
226
hydrothermal, skarn, and VMS Pb–Zn deposits) show moderate Cd concentrations (2410 – 4126
227
ppm), tight Cd isotopic variation range (–0.25‰ to 0.05‰) and moderate Zn/Cd ratios (155 – 223);
228
(3) The Exhalative systems exhibit low Cd concentrations, correspondingly high Zn/Cd ratios and
229
large Cd isotopic variation range with relatively lower δ144/110Cd values, which include the SEDEX
230
(sedimentary exhalative) Pb–Zn deposits (595 – 996 ppm, 316 – 368, –0.29‰ to 0.22‰) and
231
seafloor hydrothermal sulfides (295 – 1174 ppm, 211 – 510, –0.49‰ to 0.35‰). Comparing with
232
these three aforesaid categories, the Cd characteristics of sphalerite from Zhaxikang deposit (low
233
Cd concentrations: 1183 – 2199 ppm, correspondingly high Zn/Cd ratios: 248 – 421, large Cd
234
isotopic variation range with relatively lower δ144/110Cd values: –0.51‰ to 1.01‰) conform to
235
Exhalative system, yet are obviously distinct from Low-T and High-T systems (Fig. 6). Therefore,
236
these Cd characteristics of sphalerite support a SEDEX genetic model for the Pb–Zn mineralization
237
in Zhaxikang deposit.
238
This interpretation is consistent with previous geology, mineralogy, elements, isotopes and
239
geochronology evidence: (1) the ores from the first pulse of mineralization exhibit lamellar, banded,
240
massive, globular, concentric annular, disseminated, brecciated, fine-grained layered and colloform,
241
net-veined and Dal Matianite ore textures, as well as syndepositional structure and even ancient
242
hydrothermal vent (Zheng et al., 2012, 2014), which are similar to those of the ores from Red Dog
12
243
SEDEX-type ore district in Alaska with typical submarine hydrothermal sedimentation
244
(metasomatism) origin (Moore, et al., 1987; Leach et al., 2005); (2) the Fe–Zn isotopic
245
compositions of the first pulse of ore-forming fluid (δ56Fe: –1‰ to –0.5‰, δ66Zn: –0.28‰ to 0‰)
246
overlaps those of seafloor hydrothermal fluids, which are calculated from Fe–Zn isotopic Rayleigh
247
fractionation models (Wang et al., 2017, 2018a); (3) the Sm–Nd isochron age of Mn–Fe carbonate
248
(173.7 ± 7.4 Ma) and Rb–Sr isochron age (147 ± 3.2 Ma) of stage 2 sphalerite are in keeping with
249
the age of marine volcanic rocks (220 – 130 Ma) within regional strata, which are dominated by a
250
set of Late Triassic-Early Cretaceous flysch formations formed by turbidity sediment and
251
carbonaceous-siliceous-argillaceous rock series related to hydrothermal sedimentation (Wang et al.,
252
2019).
253
On the basis of unique characteristics for different ore-forming systems, Wen et al. (2016)
254
selected different initial δ114/110Cd values and fractionation factors (α) to establish the Cd isotopic
255
evolution models for Low-T, High-T and Exhalative systems (Fig. 7). These evolution models can
256
illustrate the Cd isotopic variations during the deposition of aqueous Cd in different hydrothermal
257
fluids. The aforementioned comparisons about genetic models have proved that SEDEX genesis is
258
most plausible for the Pb–Zn mineralization in Zhaxikang deposit, which can well explain why
259
neither of the evolution models for Low-T and High-T systems is suitable for Cd isotopic
260
compositions of sphalerite in Zhaxikang deposit (Fig. 7). However, the δ114/110Cd values of
261
sphalerite in Zhaxikang deposit are also still not in exact consistency with the evolution model for
262
Exhalative system (Fig. 7). This phenomenon reveals that the metallogenic model in Zhaxikang
263
deposit is not simple deposition of aqueous Cd, and the Cd isotopic variations in Zhaxikang deposit
13
264
haven’t absolutely induced by the kinetic Rayleigh fractionation related to mineral precipitation
265
(Wen et al., 2016; Zhu et al., 2017). In the following section, we will discuss about the Cd isotopic
266
fractionation mechanisms in Zhaxikang deposit in detail.
267 268
5.3 Background on Cd isotopic fractionation and possible causes of Cd isotopic variations
269
At present, the mechanisms for Cd isotopic fractionation in hydrothermal ore-forming
270
system is still not clear. Nevertheless, a diverse set of hypotheses have been presented to explain
271
isotopic values in ores derived from geochemical reactions during ore formation. In general, two
272
hypotheses regarding the interpretation of the transition metal isotopic data from hydrothermal
273
mineralization exist: (1) the physiochemical nature of the fluids evolve during the mineralization
274
process which cause isotopic changes in the minerals of the system; (2) the addition/subtraction of
275
different metal sources leads to mixing before which causes isotopic changes in the minerals of the
276
system (Yao et al., 2016). The processes that could operate during these two scenarios include
277
disequilibrium chemical diffusion or changes in temperature, salinity and pH (e.g., Fe: Huang et
278
al., 2010; Zn: Toutain et al., 2008; Ag: Mathur et al., 2018; Cu: Maher et al., 2011), electron transfer
279
induced by redox reactions (e.g., Cu: Mathur et al., 2005; Fe: Kavner et al., 2005; Sn: Yao et al.,
280
2018; Ag: Fujii and Albarede, 2018), equilibrium fractionation (e.g., Fe: Polyakov et al., 2000,
281
2007, 2011), kinetic Rayleigh fractionation related to solid-liquid partitioning (mineral
282
precipitation; e.g., Cd: Zhu et al., 2017; Zn: Wilkinson et al., 2005; Kelley et al., 2009) or vapor-
283
liquid partitioning (e.g., Cu: Maher et al., 2011; Sn: Mathur et al., 2017; Cu and Mo: Yao et al.,
284
2016), heterogeneity metal source and mixing of isotopic reservoirs (e.g., Zn: Wilkinson et al.,
14
285
2005; Fe: Hou et al., 2014). Nevertheless, if all these factors are appropriate for Cd isotope and
286
which one or several are the main causes of Cd isotopic variations in Zhaxikang deposit? Next, we
287
will discuss one by one.
288
Firstly, the disequilibrium chemical diffusion or changes in temperature, salinity and pH can’t
289
lead to the significant Cd isotopic variations in Zhaxikang deposit. The Cd isotope should show a
290
smaller fractionation at high temperature, thus the temperature decreasing of ore-forming fluids
291
could lead to heavier Cd isotopes enriched in earlier precipitated sphalerite (Yang et al., 2015).
292
However, both the high- and low-temperature sphalerite has heavier δ114/110Cd values, although the
293
fluids temperature decreased from 300℃ (bottom) to 150℃ (top) in the Tianbaoshan deposit (Zhu
294
et al., 2016a). Meanwhile, based on the study of Cd isotopic fractionation during inorganic calcite
295
growth, Horner et al. (2011) discovered that the fractionation factor (αCalcite–Cd(aq)) is insensitive to
296
temperature, ambient [Mg2+] and precipitation rate, and the Cd isotopic fractionation related to
297
kinetic isotope effects during largely unidirectional incorporation of Cd at the mineral surface will
298
become more pronounced at high salinity. In Zhaxikang deposit, Cd is mainly enriched by
299
isomorphic replacement with Zn in sphalerite. According to the previous fluid inclusion data
300
trapped on sphalerite and cogenetic Mn–Fe carbonate (Yu, 2015), from stage 1 (203 – 273℃, 9.0
301
– 14.5 wt% NaCl equiv.) through stage 2 (215 – 265℃, 8.9 – 16.7 wt% NaCl equiv.) to stage 3
302
(237 – 259℃, 12.0 – 16.1wt% NaCl equiv.), the temperature and salinity show little changes.
303
Notably, the disequilibrium chemical diffusion or changes in temperature and salinity of ore-
304
forming fluid can’t result in the significant Cd isotopic variations of sulfides (sphalerite: –0.51‰
305
to 1.01‰; galena: –2.29‰ to 2.24‰) in Zhaxikang deposit. Furthermore, significant Cu (Maher et
15
306
al., 2011; Mathur et al., 2013) and Ag (Mathur et al., 2018) isotopic fractionation have been linked
307
to the changes in pH. Whereas, currently, no data exists to support the pH induced Cd isotopic
308
fractionation.
309
Secondly, there is no electron transfer caused Cd isotopic fractionation during ore-forming
310
process with respects to Zhaxikang deposit. In Pb–Zn deposits, Cd is mainly involved in sphalerite
311
by directly substitutions with Zn, the equation is ZnSsolid + Cd2+ Liquid = CdSSolid + Zn2+ Liquid
312
(Mookherjee, 1962; Tsusue and Holland, 1996; Sverjensky et al., 1997; Wen et al., 2016).
313
Meanwhile, Zn2+ and Cd2+ should enter into the galena by directly isomorphic replacement with
314
Pb2+ in Zhaxikang deposit as discussed in Section 5.1, too. Specifically, most of the Zn2+ and Cd2+
315
are subordinated and form the aqueous complexes with ligands of Cl−, HS− and OH− in
316
hydrothermal fluids (Bazarkina et al., 2010; Tagirov and Seward, 2010). Over a wide range of
317
temperatures (20℃ – 450℃), acidities (pH: 1 – 8) and chloride concentrations (mCl: 0.04 – 18
318
mol/kg H2O), aqueous Cd speciation is dominated by chloride species [CdClm(H2O)2−m n]
319
(Giordano, 2002; Mei et al., 2015; Wen et al., 2016). Accordingly, Cd ions exist in the valence
320
state of +2 both in the ore-forming fluids and sulfides, there is no redox reactions occurred during
321
ore formation.
322
Thirdly, heterogeneity metal source and mixing of isotopic reservoirs are common in
323
hydrothermal ore-forming system. Zhu et al. (2016a) thought that the ore-forming fluids are
324
heterogeneous and heavy Cd isotopes were enriched in original ore-forming fluids, which result in
325
heavier Cd isotopic enrichments in early precipitated sphalerite in Tianbaoshan deposit. The
326
Zhaxikang deposit is a complex polymetallic deposit that has experienced two pulses of
16
327
mineralization. Therefore, heterogeneity metal source and mixing of isotopic reservoirs are more
328
likely to occur. However, these two factors still not the dominated causes induced the Cd isotopic
329
variations in Zhaxikang deposit. The evidences are as follow: (1) There is no correlation between
330
Cd concentrations and δ114/110Cd values of sulfides (Figs. 8A and B), and the previous H–O isotopic
331
data (Xie et al., 2017) of cogenetic Mn–Fe carbonate are almost the same and all fall into the
332
formation water area (Fig. 9), which indicates that there is no heterogeneity in the metal source for
333
Pb–Zn mineralization (stages 1 and 2). (2) According to the commonsense metallogenic element
334
associations characteristics in ore deposits, the Cd should mainly source from the first pulse of
335
mineralization (Pb–Zn), rather than the latter second pulse of mineralization that is dominated by
336
Sb mineralization, which implies that the Cd is single-source and there is no mixing of Cd isotopic
337
reservoirs.
338
Fourthly, the transition metal isotopic fractionation between coexisting mineral pairs have
339
attracted extensive attentions. For instance, Fe isotopic fractionation between magnetite and pyrite,
340
magnetite and pyrrhotite, pyrrhotite and pyrite (Zhu et al., 2016b), Mn–Fe carbonate and pyrite
341
(Wang et al., 2017), as well as Zn isotopic fractionation between Mn–Fe carbonate and sphalerite
342
(Wang et al., 2017). In this paper, we compare Cd isotopic fractionation within four sphalerite-
343
galena coexisting mineral pairs. The results show there is Cd isotopic fractionation between
344
sphalerite and galena (Fig. 10A), yet these mineral pairs have not attained Cd isotopic equilibrium
345
(Fig. 10B). The galena has very large Cd isotopic variation range (–2.19‰ to 2.24‰) with low Cd
346
concentrations (14 – 26 ppm), which can be well explained by Cd isotopic fractionation between
347
sphalerite and galena. According to the mass balance theory, as Cd concentrations of sphalerite
17
348
(1183 – 2199 ppm) are dozens higher than galena in Zhaxikang deposit, little Cd isotopic variation
349
in sphalerite will cause large Cd isotopic variation in galena within same mineral pair. From the
350
above, the Cd isotopic fractionation between sphalerite-galena mineral pairs must have some
351
contribution to the Cd isotopic variations in Zhaxikang deposit, yet still can’t well decipher the
352
significant Cd isotopic variations of sphalerite (–0.51‰ to 1.01‰).
353
Lastly, the kinetic Rayleigh fractionation mechanism is most commonly used to explain the
354
transition metal isotopic variations in hydrothermal deposits. The research of Schmitt et al. (2009)
355
and Horner et al. (2011) indicate that the minerals preferentially enrich lighter Cd isotopes than
356
related fluids. Similarly, based on the density functional theory, Yang et al. (2015) calculated the
357
fractionation properties of Cd species in hydrothermal fluids, and then suggested that the light Cd
358
isotopes preferentially fractionate into solid phase relative to relevant fluids. The Zn and Cd have
359
shown similar geochemical behavior and experience similar hydrothermal evolution histories in
360
Pb–Zn deposits (Schwartz, 2000; Schmitt et al., 2009). Consequently, this kinetic Rayleigh
361
fractionation mechanism is suitable for Zn isotope, too (Gagnevin et al., 2012). This equilibrium
362
induced fractionation because of partitioning into solid-liquid phases has been modeled with
363
Rayleigh distillation to decipher increasing δ114/110Cd (e.g. MVT: Zhu et al., 2017) and δ66Zn (e.g.
364
Irish-type: Wilkinson et al., 2005; SEDEX: Kelley et al., 2009) values with time in Pb–Zn deposits.
365
With the regards to Zhaxikang deposit, it is an entirely different situation. The ore samples for Zn
366
and Cd analyses are both collected from the same location in orebody V, hence the Cd and Zn
367
isotopic values still have good correlation although they are not from the same suite of samples.
368
Both the δ114/110Cd and δ66Zn values of sphalerite display a gradually decreasing trend from stage
18
369
1 (δ114/110Cd: –0.30‰ to 1.01‰, average value = 0.17‰, n =4; δ66ZnAA-ETH: –0.12‰ to 0.07‰,
370
average value = 0‰, n = 4) through stage 2 (δ114/110Cd: –0.51‰ to –0.09‰, average value = –
371
0.23‰, n = 3; δ66ZnAA-ETH: –0.38‰ to –0.05‰, average value = –0.189‰, n = 14) to stage 3
372
(δ114/110Cd: –0.34‰ to –0.23‰, average value = –0.285‰, n = 2; δ66ZnAA-ETH: –0.32‰ to –0.06‰,
373
average value = –0.193‰, n =7) (Fig. 11A and B; Wang et al., 2017, 2018a). The number of
374
measurements for Cd isotope are too limited to do statistically analysis (T Test ①stages 1 and 2:
375
p = 0.138475, ②stages 2 and 3: p = 0.377581). Yet, the number of measurements for Zn isotope
376
are enough to do the T Test and there are statistically differences in Zn isotopic compositions for
377
each stage (①stages 1 and 2: p = 0.001814, ②stages 2 and 3: p = 0.472327). Meanwhile, the
378
decreasing trend in δ66Zn and δ56Fe values have also been found within concentric ore samples
379
(Fig. 12; Wang et al., 2018a), which enhances the authenticity for the temporally decreasing trend
380
in transition metal isotopic compositions. In brief, this temporally decreasing trend indeed exists
381
both macroscopically and microscopically in Zhaxikang deposit. This is in accord with the
382
inference in Section 5.2 that the metallogenic model in Zhaxikang deposit is not simple deposition
383
of aqueous Cd, and the Cd isotopic variations in Zhaxikang deposit haven’t absolutely induced by
384
the kinetic Rayleigh fractionation related to solid-liquid partitioning (mineral precipitation).
385
This different situation in Zhaxikang deposit has also been found in other ore deposits (e.g.,
386
δ114/110Cd values of sphalerite in MVT deposit: Zhu et al., 2016a; δ65Cu values of chalcopyrite in
387
porphyry deposit: Graham et al., 2004, Yao et al., 2016). Zhu et al. (2016a) have proposed that
388
there must be a potential mechanism resulting in early precipitated minerals enriched in heavy
389
isotopes. We argue that the potential mechanism in Zhaxikang deposit should be the kinetic
19
390
Rayleigh fractionation related to vapor-liquid partitioning. This inference is supported by the
391
following evidence: (1) The Pb–Zn mineralization is demonstrated to have the SEDEX genesis,
392
hence the degasification is ubiquitous in this kind of ore-forming system. (2) The fluid inclusion
393
data from the Pb–Zn sulfides and cogenetic carbonate alteration minerals indicates that vapor-
394
liquid two-phase inclusions are in dominance (more than 90%) with the vapor-liquid ratios of 20%
395
– 50% (Yu, 2015). (3) The evaporation (vapor-liquid partitioning) has been used to explain the
396
measurable Cd isotopic fractionation occurs in nature in previous studies (Wang et al., 2013; Zhu
397
et al., 2013). (4) Transition metal isotopic fractionation initiated by vapor-liquid partitioning could
398
be the cause for fractionation factors used in Rayleigh distillation models, which can result in the
399
temporally decreasing trend of transition metal isotopic values (e.g., Cu: Graham et al., 2004; Yao
400
et al., 2016). (5) We have confirmed that the metallogenic model in Zhaxikang deposit is not simple
401
deposition of aqueous Cd in Section 5.2. Meanwhile, the Fe–Zn–Cd should experience similar
402
evolutionary process during Pb–Zn mineralization, and the kinetic Rayleigh fractionation related
403
to vapor-liquid partitioning have been confirmed to cause the decreasing trend of Fe–Zn isotopic
404
compositions with time in Zhaxikang deposit (Wang et al., 2018a). This Rayleigh distillation
405
mechanism models the temporally decreasing trend as follow: During the ore formation, there is
406
vapor-liquid partitioning within ore-forming system and the ratios change with the temperature
407
decreasing. Initially, the increment of vapor phase preferentially enriched in light isotopes relative
408
to ore-forming fluid, which lead to the early minerals precipitated from the ore-forming fluid
409
enriched heavy isotopes. Then with the removing of precipitated mineral from the ore-forming
410
system and condensation of vapor phase back to ore-forming fluid, the isotopic values of residual
20
411
ore-forming fluid and subsequent minerals have lower values (Fig. 13; Graham et al., 2004; Yao et
412
al., 2016; Wang et al., 2018a).
413
On the other hand, Wang et al. (2017) have found that the higher δ56Fe values and lower δ66Zn
414
values coincide with an increase in alteration for stage 3 minerals in Zhaxikang deposit. Meanwhile,
415
there is obvious temporally decreasing trend in δ114/110Cd and δ66Zn values from stages 1 to 2, yet
416
the δ114/110Cd and δ66Zn values are almost same in stages 2 and 3 (Fig. 11A and B). In view of the
417
fact that stage 3 is a transitional stage between the two pulses of mineralization, the overprint by
418
the second pulse of mineralization have also changed the Cd isotopic compositions of sulfides in
419
stage 3 to some extent although there is no mixing of Cd sources.
420
On the whole, we propose that the kinetic Rayleigh fractionation related to vapor-liquid
421
partitioning is the main cause for Cd isotopic variations in Zhaxikang deposit, overprint by the
422
latter second pulse of mineralization and Cd isotopic fractionation between sphalerite-galena
423
mineral pairs also have some contribution.
424 425
5.4 Cd isotopic Rayleigh distillation model for vapor-liquid partitioning mechanism
426
To further verify the correctness of kinetic Rayleigh fractionation mechanism related to vapor-
427
liquid partitioning in section 5.3 and constrain some aspects associated with this cause for observed
428
Cd isotopic fractionations, we established a simple open system Cd isotopic Rayleigh distillation
429
model for Zhaxikang deposit. The objective of this model is to demonstrate that vapor-liquid
430
partitioning induced Cd isotopic fractionation can explain the measured δ114/110Cd values and the
431
temporally decreasing trend. For this discussion, we assume that the simplest explanation for Cd
432
isotopic fractionation that would lead to the temporally decreasing trend in δ144/110Cd values would 21
433
be a physiochemical change in hydrothermal fluid. The rationale for Cd isotopic fractionation
434
model is described below. The Rayleigh distillation equations are found in Sharp. (2007) and Yao
435
et al. (2016):
436
―1 – 103 δXvap= (δX0vap+ 103) × Fαmin
437
δXaq = (δXvap + 103) × α – 103
438
where X = Cd isotope ratio of interest, vap = vapor, aq = aqueous, 0 = original/starting Cd isotopic
439
composition, F = fraction of X/X0, α = fractionation factor defined as Xp/Xs (p = product of reaction,
440
s = substrate of reaction).
441
These two equations are used to predict Cd isotopic compositions given different fractions of
442
vapor and liquid phases. The following geological facts and rationale assumptions were applied in
443
construction of the Cd isotopic Rayleigh distillation model:
444
(1) In Zhaxikang deposit, the sphalerite is the dominated carrier for Cd, the Cd concentrations of
445
sphalerite are dozens higher than galena (Table 2). Accordingly, this Cd isotopic fractionation
446
model just compares with δ114/110Cd values of sphalerite;
447
(2) The above discussion has demonstrated that SEDEX genesis is most plausible for the Pb–Zn
448
mineralization in Zhaxikang deposit, consequently the initial δ144/110Cd value (–0.11%; Fig. 7) of
449
exhalative system (Wen et al., 2016) should be regarded as the δX0vap;
450
(3) Fractionation factors must be chosen accurately to fit the decreasing trend in δ144/110Cd values
451
so that the Rayleigh distillation model would generate values seen in the natural samples. Horner
452
et al. (2011) have studied the Cd isotopic fractionation for calcite in ocean environment, and found
453
that the αCalcite–Cd(aq) (0.99955 ± 0.00012) is insensitive to temperature, ambient [Mg2+] and
22
454
precipitation rate (across the range studied). In light of that the Pb–Zn mineralization has marine
455
genesis, the sphalerite exhibits a great Cd isotopic variation range (–0.51‰ to 1.01‰) and there
456
are large amounts of cogenetic carbonatite (Mn–Fe carbonate) in Zhaxikang deposit, the lowest
457
fractionation factor (0.99943; Horner et al., 2011) is selected in the construction of this model.
458
Fig. 14 is the established Cd isotopic Rayleigh distillation model that illustrates how δ114/110Cd
459
values of sphalerite change due to different proportions of vapor-liquid partitioning in the ore-
460
forming system (variable defined by F in equations above), which is suitable for the δ144/110Cd
461
values of sphalerite in Zhaxikang deposit (Fig. 14A). The total δ144/110Cd values (–0.51‰ to 1.01‰)
462
cover the F (vapor-liquid ratios) ranging from 5.23% to 74.23% (Fig. 14A), and the concentrated
463
δ144/110Cd values (–0.34‰ to 0.10‰) define the F range of 25.48% – 55.07% (Fig. 14A).
464
Importantly, these F ranges calculated from Cd isotopic Rayleigh distillation model are in keeping
465
with previous Zn isotopic Rayleigh distillation model for vapor-liquid partitioning (Fig. 14B; Wang
466
et al., 2018a) and fluid inclusions data of Zhaxikang deposit (Fig. 14C; Yu, 2015). When the mean
467
δ66Zn value of bulk earth (0‰; Chen et al., 2013) is considered as the initial value (δ66Zni), the
468
concentrated δ66ZnAA-ETH values (–0.38‰ to 0.07‰) of sphalerite cover the F range of 26.70% –
469
48.70% in the Zn isotopic Rayleigh distillation model (Fig. 14B; Wang et al., 2018a). Furthermore,
470
the vapor-liquid two-phase fluid inclusions are in dominance (more than 90%) with the vapor-
471
liquid ratios ranging from 10% to 70% and concentrating on 20% – 50% (Yu, 2015), which are
472
also in line with those of Cd isotopic Rayleigh distillation model (Fig. 14C). The aforesaid
473
comparisons further evidence the reasonability of vapor-liquid partitioning mechanism for Cd
474
isotopic fractionation in Zhaxikang deposit.
23
475 476
6. Conclusions
477
In the case study of Zhaxikang deposit, the kinetic Rayleigh fractionation related to vapor-
478
liquid partitioning has been proved as an important Cd isotopic fractionation mechanism in
479
hydrothermal ore-forming system, which is the main cause for observed Cd isotopic variations in
480
Zhaxikang deposit. The liquid-vapor transitions are processes associated directly with formation of
481
ore deposits, genesis of igneous rocks, and meteorite condensation processes. Consequently, this
482
paper is beneficial to further identify the Cd isotopic fractionation mechanism in ore, rock,
483
sediments and solar system forming processes, which will contribute to future research efforts.
484 485 486 487
Conflict of interest We declare that we do not have any commercial or associative interest that represents a conflict of interest in connection with the submitted work.
488 489
Acknowledgments
490
We would like to express our gratitude to three anonymous reviewers, for their constructive
491
comments and suggestions that have largely improved this manuscript, and Associated Editor Dr.
492
Yanbo Cheng and Editor-in-Chief Prof. Huayong Chen for the efficient editorial handling. Mr.
493
Matthew Gonzalez (Pennsylvania State University) are also thanked for aid in measuring and
494
access to the Neptune instruments. We acknowledge the financially support by the Deep Resources
495
Exploration and Mining, National Key R&D Program of China (2018YFC0604104,
24
496
2017YFC0601506), the China Postdoctoral Science Foundation Funded Project (2019M650785)
497
and the Open Research Project from the State Key Laboratory of Geological Processes and Mineral
498
Resources, China University of Geosciences (GPMR201811).
25
499
References
500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539
Abouchami, W., Galer, S.J.G., De Baar, H.J.W., Alderkamp, A.C., Middag, R., Laan, P., Feldmann, H., & Andreae, M.O., 2011. Modulation of the Southern Ocean cadmium isotope signature by ocean circulation and primary productivity. Earth and Planetary Science Letters 305(1), 83–91. Abouchami, W., Galer, S.J., Horner, T.J., Rehkämper, M., Wombacher, F., Xue, Z., Myriam, L., Melanie, G.R., Claudine, H.S., Maria, S., Alyssa, E.S., Dominique, W., & Philip, F.H., 2013. A common reference material for cadmium isotope studies–NIST SRM 3108. Geostandards and Geoanalytical Research 37, 5–17. Archer, C., & Vance, D., 2004. Mass discrimination correction in multiple-collector plasma source mass spectrometry: an example using Cu and Zn isotopes. Journal of Analytical Atomic Spectrometry 19, 656– 665. Bazarkina, E.F., Pokrovski, G.S., Zotov, A.V., & Hazemann, J.L., 2010. Structure and stability of cadmium chloride complexes in hydrothermal fluids. Chemical Geology 276(1–2), 1–17. Belissont, R., Boiron, M.C., Luais, B., & Cathelineau, M., 2014. LA-ICP-MS analyses of minor and trace elements and bulk Ge isotopes in zoned Ge-rich sphalerites from the Noailhac-Saint-Salvy deposit (France): Insights into incorporation mechanisms and ore deposition processes. Geochimica et Cosmochimica Acta 126, 518–540. Chen, H., Savage, P.S., Teng, F.Z., Helz, R.T., and Moynier, F., 2013. Zinc isotope fractionation during magmatic differentiation and the isotopic composition of the bulk Earth. Earth and Planetary Science Letters 369, 34– 42. Cloquet, C., Rouxel, O., Carignan, J., & Libourel, G., 2005. Natural Cadmium Isotopic Variations in Eight Geological Reference Materials (NIST SRM 2711, BCR 176, GSS-1, GXR-1, GXR‐2, GSD-12, Nod-P1, Nod-A-1) and Anthropogenic Samples, Measured by MC-ICP-MS. Geostandards and Geoanalytical Research 29(1), 95–106. Cook, N.J., Ciobanu, C.L., Pring, A., Skinner, W., Shimizu, M., Danyushevsky, L., Bernhardt, S.E., & Melcher, F., 2009. Trace and minor elements in sphalerite: A LA-ICPMS study. Geochimica et Cosmochimica Acta 73, 4761–4791. Duan, J., Tang, J., Li, Y., Liu, S.-A., Wang, Q., Yang, C., and Wang, Y., 2016. Copper isotopic signature of the Tiegelongnan high-sulfidation copper deposit, Tibet: implications for its origin and mineral exploration. Mineralium Deposita 51(5), 591–602. Fujii, T., & Albarede, F., 2018. 109 Ag–107 Ag fractionation in fluids with applications to ore deposits, archeometry, and cosmochemistry. Geochimica et Cosmochimica Acta 234, 37–49. Gao, B., Liu, Y., Sun, K., Liang, X., Peng, P. A., Sheng, G., & Fu, J., 2008. Precise determination of cadmium and lead isotopic compositions in river sediments. Analytica Chimica Acta 612(1), 114–120. Gao, Z., Zhu, X., Sun, J., Luo, Z., Bao, C., Tang, C., & Ma, J., 2018. Spatial evolution of Zn-Fe-Pb isotopes of sphalerite within a single ore body: A case study from the Dongshengmiao ore deposit, Inner Mongolia, China. Mineralium Deposita 53(1), 55–65. Gault-Ringold, M., Adu, T., Stirling, C.H., Frew, R.D., & Hunter, K.A., 2012. Anomalous biogeochemical behavior of cadmium in subantarctic surface waters: mechanistic constraints from cadmium isotopes. Earth and Planetary Science Letters 341, 94–103. Gagnevin, D., Boyce, A.J., Barrie, C.D., Menuge, J.F., Blakeman, R.J., 2012. Zn, Fe and S isotope fractionation in a large hydrothermal system. Geochimica et Cosmochimica Acta 88, 183–198.
26
540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581
Giordano, T.H., 2002. Transport of Pb and Zn by carboxylate complexes in basinal ore fluids and related petroleum-field brines at 100℃: the influence of pH and oxygen fugacity. Geochemical Transactions 3(8), 56–72. Graham, S., Pearson, N., Jackson, S., Griffin, W., & O'reilly, S.Y., 2004. Tracing Cu and Fe from source to porphyry: in situ determination of Cu and Fe isotope ratios in sulfides from the Grasberg Cu–Au deposit. Chemical Geology 207(3-4), 147–169. Horner, T.J., Rickaby, R.E., & Henderson, G.M., 2011. Isotopic fractionation of cadmium into calcite. Earth and Planetary Science Letters 312(1-2), 243–253. Hou, K., Li, Y., Gao, J., Liu, F., Qin, Y., 2014. Geochemistry and Si–O–Fe isotope constraints on the origin of banded iron formations of the Yuanjiacun Formation, Lvliang Group, Shanxi, China. Ore Geology Reviews 57, 288–298. Huang, F., Chakraborty, P., Lundstrom, C.C., Holmden, C., Glessner, J.J.G., Kieffer, S.W., Lesher, C.E., 2010. Isotope fractionation in silicate melts by thermal diffusion. Nature 464, 396–400. Kavner, A., Bonet, F., Shahar, A., Simon, J., & Young, E., 2005. The isotopic effects of electron transfer: An explanation for Fe isotope fractionation in nature. Geochimica et Cosmochimica Acta 69(12), 2971–2979. Kelley, K.D., Wilkinson, J.J., Chapman, J.B., Crowther, H.L., & Weiss, D.J., 2009. Zinc isotopes in sphalerite from base metal deposits in the Red Dog district, northern Alaska. Economic Geology 104(6), 767-773. Lacan, F., Francois, R., Ji, Y., & Sherrell, R.M., 2006. Cadmium isotopic composition in the ocean. Geochimica et Cosmochimica Acta 70(20), 5104–5118. Lambelet, M., Rehkämper, M., van de Flierdt, T., Xue, Z., Kreissig, K., Coles, B., Porcelli, D., & Andersson, P., 2013. Isotopic analysis of Cd in the mixing zone of Siberian rivers with the Arctic Ocean—New constraints on marine Cd cycling and the isotope composition of riverine Cd. Earth and Planetary Science Letters 361, 64-73. Leach, D.L., Bradley, D.C., Huston, D., Pisarevsky, S.A., Taylor, R.D., Gardoll, S.J., 2010. Sediment-hosted lead-zinc deposits in earth history. Economic Geology 105, 593–625. Li, G.M., Zhang, L.K., Jiao, Y.J., Xia, X.B., Dong, S.L., Fu, J.G., Liang, W., Zhang, Z., Wu, J.Y., Dong, L., and Huang, Y., 2017. First discovery and implications of Cuonadong superlarge Be-W-Sn polymetallic deposit in Himalayan metallogenic belt, southern Tibet. Mineral Deposits 36, 1003–1008 (In Chinese with English abstract). Lv, Y.W., Liu, S.A., Zhu, J.M., Li, S.G., 2015. Copper and zinc isotope fractionation during deposition and weathering of highly metalliferous black shales in central China. Chemical Geology 422, 82–93. Maher, K.C., Jackson, S., & Mountain, B., 2011. Experimental evaluation of the fluid–mineral fractionation of Cu isotopes at 250° C and 300° C. Chemical Geology 286(3-4), 229–239. Marechal, C.N., Telouk, P., Albarede, F., 1999. Precise analysis of copper and zinc isotopic compositions by plasma-source mass spectrometry. Chemical Geology 156, 251–273. Markl, G., Wagner, T., Blanckenburg, F.V., 2006. Iron isotope fractionation during hydrothermal ore deposition and alteration. Geochimica et Cosmochimica Acta 70, 3011–3030. Mathur, R., Titley, S., Barra, F., Brantley, S., Wilson, M., Phillips, A., Munizaga, F., Maksaev, V., Vervoort, J., & Hart, G., 2009. Exploration potential of Cu isotope fractionation in porphyry copper deposits. Journal of Geochemical Exploration 102(1), 1–6. Mathur, R., Ruiz, J., Titley, S., Liermann, L., Buss, H., & Brantley, S., 2005. Cu isotopic fractionation in the supergene environment with and without bacteria. Geochimica et Cosmochimica Acta 69(22), 5233–5246.
27
582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623
Mathur, R., Munk, L., Nguyen, M., Gregory, M., Annell, H., & Lang, J., 2013. Modern and paleofluid pathways revealed by Cu isotope compositions in surface waters and ores of the Pebble porphyry Cu-Au-Mo deposit, Alaska. Economic Geology 108(3), 529-541. Mathur, R., Powell, W., Mason, A., Godfrey, L., Yao, J., & Baker, M.E., 2017. Preparation and measurement of cassiterite for Sn isotope analysis. Geostandards and Geoanalytical Research 41(4), 701–707. Mathur, R., Ruiz, J., Casselman, M. J., Megaw, P., & van Egmond, R., 2012. Use of Cu isotopes to distinguish primary and secondary Cu mineralization in the Cañariaco Norte porphyry copper deposit, Northern Peru. Mineralium Deposita 47(7), 755–762. Mathur, R., Arribas, A., Megaw, P., Wilson, M., Stroup, S., Meyer-Arrivillaga, D., and Arribas, I., 2018. Fractionation of silver isotopes in native silver explained by redox reactions. Geochimica et Cosmochimica Acta 224, 313–326. Mei, Y., Sherman, D.M., Liu, W., Etschmann, B., Testemale, D., & Brugger, J., 2015. Zinc complexation in chloride-rich hydrothermal fluids (25–600℃): A thermodynamic model derived from ab initio molecular dynamics. Geochimica et Cosmochimica Acta 150, 265–284. Mookherjee, A., 1962. Certain aspects of the geochemistry of cadmium. Geochimica et Cosmochimica Acta 26(2), 351–360. Moore, D.W., Young, L.E., Modene, J.S., Plahuta, J.T., 1987. Geologic setting and genesis of the Red Dog zinclead-silver deposit, western Brooks Range, Alaska. Economic Geology 81, 1696–1727. Palero-Fernández, F.J., & Martín-Izard, A., 2005. Trace element contents in galena and sphalerite from ore deposits of the Alcudia Valley mineral field (Eastern Sierra Morena, Spain). Journal of geochemical exploration 86, 1–25. Polyakov, V.B., Mineev, S.D., 2000. The use of Mössbauer spectroscopy in stable isotope geochemistry. Geochimica et Cosmochimica Acta 64, 849–865. Polyakov, V.B., Clayton, R.N., Horita, J., Mineev, S.D., 2007. Equilibrium iron isotope fractionation factors of minerals: reevaluation from the data of nuclear inelastic resonant X-ray scattering and Mössbauer spectroscopy. Geochimica et Cosmochimica Acta 71, 3833–3846. Polyakov, V.B., Soultanov, D.M., 2011. New data on equilibrium iron isotope fractionation among sulfides: Constraints on mechanisms of sulfide formation in hydrothermal and igneous systems. Geochimica et Cosmochimica Acta 75, 1957–1974. Ripperger, S., Rehkämper, M., Porcelli, D., & Halliday, A.N., 2007. Cadmium isotope fractionation in seawater-A signature of biological activity. Earth and Planetary Science Letters 261(3), 670–684. Rosman, K.J.R., & De Laeter, J.R., 1975. The isotopic composition of cadmium in terrestrial minerals. Journal of Mass Spectrometry and Ion Physics 16(4), 385–394. Rosman, K.J.R., & De Laeter, J.R., 1976. Isotopic fractionation in meteoritic cadmium. Nature 261(5557), 216– 218. Rosman, K.J.R., & De Laeter, J.R., 1978. A survey of cadmium isotopic abundances. Journal of Geophysical Research: Solid Earth 83(B3), 1279–1287. Sands, D.G., Rosman, K.J.R., & De Laeter, J.R., 2001. A preliminary study of cadmium mass fractionation in lunar soils. Earth and Planetary Science Letters 186(1), 103–111. Saunders, J. A., Mathur, R., Kamenov, G. D., Shimizu, T., and Brueseke, M. E., 2015. New isotopic evidence bearing on bonanza (Au-Ag) epithermal ore-forming processes. Mineralium Deposita 51(1), 1–11. Schediwy, S., Rosman, K.J.R., & De Laeter, J.R., 2006. Isotope fractionation of cadmium in lunar material. Earth
28
624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665
and Planetary Science Letters 243(3), 326–335. Schmitt, A.D., Galer, S.J., & Abouchami, W., 2009. Mass-dependent cadmium isotopic variations in nature with emphasis on the marine environment. Earth and Planetary Science Letters 277(1), 262–272. Schwartz, M.O., 2000. Cadmium in zinc deposits: Economic geology of a polluting element. International Geology Review 42(5), 445–469. Sharp, Z., 2007. Principles of stable isotope geochemistry. Upper Saddle River, NJ: Pearson education, 78–80. Sun, X., Zheng, Y.Y., Pirajno, F., McCuaig, T.C., Yu, M., Xia, S.L., Song, Q.J., & Chang, H.F., 2018. Geology, S–Pb isotopes, and 40Ar/39Ar geochronology of the Zhaxikang Sb–Pb–Zn–Ag deposit in Southern Tibet: implications for multiple mineralization events at Zhaxikang. Mineralium Deposita 53(3), 435–458. Sverjensky, D.A., Shock, E.L., & Helgeson, H.C., 1997. Prediction of the thermodynamic properties of aqueous metal complexes to 1000℃ and 5 kb. Geochimica et Cosmochimica Acta 61(7), 1359–1412. Tagirov, B.R., & Seward, T.M., 2010. Hydrosulfide/sulfide complexes of zinc to 250 C and the thermodynamic properties of sphalerite. Chemical Geology 269(3–4), 301–311. Toutain, J.P., Sonke, J., Munoz, M., Nonell, A., Polvé, M., Viers, J., Freydier, R., Sortino, F., Joron, J.L., Sumarti, S., 2008. Evidence for Zn isotopic fractionation at Merapi volcano. Chemical Geology 253(1–2), 74–82. Tsusue, A., & Holland, H.D., 1966. The coprecipitation of cations with CaCO3-III. The coprecipitation of Zn2+ with calcite between 50 and 250℃. Geochimica et Cosmochimica Acta 30(4), 439–453. Tu, G.C., Gao, Z.M., Hu, R.Z., Zhang, Q., Li, C.Y., Zhao, Z.H and Zhang, B.G., 2004. The Geochemistry and Deposit-Forming Mechanism of Disperse Element. Geological Publishing House, Beijing. Chapter 1, 423 (In Chinese). Wang, D., Sun, X., Zheng, Y.Y., Wu, S., Xia, S.L, Chang, H.F., & Yu, M., 2017. Two pulses of mineralization and genesis of the Zhaxikang Sb–Pb–Zn–Ag deposit in southern Tibet: Constraints from Fe–Zn isotopes. Ore Geology Reviews 84, 347–363. Wang, D., Zheng, Y.Y., Mathur, R., & Wu, S., 2018a. The Fe-Zn Isotopic Characteristics and Fractionation Models: Implications for the Genesis of the Zhaxikang Sb-Pb-Zn-Ag Deposit in Southern Tibet. Geofluids 2018. Wang, D., Zheng, Y.Y., Yang, W.T., & Gyatso, N., 2018b. Geology, Mineralogy, Fluid Inclusion, and H–O–S– Pb Isotope Constraints on Ore Genesis of the Keyue Sb–Pb–Zn–Ag Deposit in Southern Tibet. Geofluids 2018. Wang, D., Zheng, Y.Y., Mathur, R., Jiang, J.S., Zhang, S.K., Zhang, J.F., & Yu, M., 2019. Multiple mineralization events in the Zhaxikang Sb–Pb–Zn–Ag deposit and their relationship with the geodynamic evolution in the North Himalayan Metallogenic Belt, South Tibet. Ore Geology Reviews 105, 201–215. Wang, D.N., Jin, L.L., Chen, B., Xie, X.J., & Hu, S.H., 2013. A review of the isotope system of cadmium and its applications in geosciences and environmental sciences. Rock & Mineral Analysis 30(2), 356–368 (In Chinese with English abstract). Wasylenki, L.E., Swihart, J.W., Romaniello, S.J., 2014. Cadmium isotope fractionation during adsorption to Mn oxyhydroxide at low and high ionic strength. Geochimica et Cosmochimica Acta 140, 212–226. Wen, H.J, Zhang, Y.X, Cloquet, C., Zhu, C.W, Fan, H.F, & Luo, C.G., 2015. Tracing sources of pollution in soils from the Jinding Pb–Zn mining district in China using cadmium and lead isotopes. Applied Geochemistry 52, 147–154. Wen, H.J, Zhu, C.W, Zhang, Y.X, Cloquet, C., Fan, H.F, & Fu, S.H., 2016. Zn/Cd ratios and cadmium isotope evidence for the classification of lead-zinc deposits. Scientific Reports 6, 25273.
29
666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707
Wilkinson, J.J., Weiss, D.J., Mason, T.F.D., & Coles, B.J., 2005. Zinc isotope variation in hydrothermal systems: preliminary evidence from the Irish Midlands ore field. Economic Geology 100(3), 583-590. Wombacher, F., Rehkämper, M., Mezger, K., & Münker, C., 2003. Stable isotope compositions of cadmium in geological materials and meteorites determined by multiple-collector ICPMS. Geochimica et Cosmochimica Acta 67(23), 4639–4654. Wombacher, F., Rehkämper, M., & Mezger, K., 2004. Determination of the mass-dependence of cadmium isotope fractionation during evaporation. Geochimica et Cosmochimica Acta 68(10), 2349–2357. Wombacher, F., Rehkämper, M., Mezger, K., Bischoff, A., & Münker, C., 2008. Cadmium stable isotope cosmochemistry. Geochimica et Cosmochimica Acta 72(2), 646–667. Wu, S., Zheng, Y., Wang, D., Chang, H., & Tan, M., 2017. Variation of copper isotopes in chalcopyrite from Dabu porphyry Cu-Mo deposit in Tibet and implications for mineral exploration. Ore Geology Reviews, 90, 14–24. Xie, Y., Li, L., Wang, B., Li, G., Liu, H., Li, Y., Dong, S., & Zhou, J., 2017. Genesis of the Zhaxikang epithermal Pb-Zn-Sb deposit in southern Tibet, China: Evidence for a magmatic link. Ore Geology Reviews, 80, 891– 909. Yang, J.L, Li, Y.B, Liu, S.Q, Tian, H.Q, Chen, C.Y, Liu, J.M., & Shi, Y.L, 2015. Theoretical calculations of Cd isotope fractionation in hydrothermal fluids. Chemical Geology 391, 74–82. Yang, S.C., Lee, D.C., & Ho, T.Y., 2012. The isotopic composition of cadmium in the water column of the South China Sea. Geochimica et Cosmochimica Acta 98, 66–77. Yang, Z.S, Hou, Z.Q, White, N.C., Chang, Z.S, Li, Z.Q, & Song, Y. C., 2009. Geology of the post-collisional porphyry copper–molybdenum deposit at Qulong, Tibet. Ore Geology Reviews 36(1-3), 133-159. Yao, J.M, Mathur, R., Sun, W.D, Song, W.L, Chen, H.Y, Mutti, L., Xiang, X.K., & Luo, X.H., 2016. Fractionation of Cu and Mo isotopes caused by vapor‐liquid partitioning, evidence from the Dahutang W ‐Cu‐Mo ore field. Geochemistry, Geophysics, Geosystems 17(5), 1725–1739. Yao, J.M., Mathur, R., Powell, W., Lehmann, B., Tornos, F., Wilson, M., & Ruiz, J., 2018. Sn-isotope fractionation as a record of hydrothermal redox reactions. American Mineralogist, 103(10), 1591–1598. Yin, A., 2006. Cenozoic tectonic evolution of the Himalayan orogen as constrained by along-strike variation of structural geometry, exhumation history, and foreland sedimentation. Earth Science Review 76, 1–131. Yu, M., 2015. Characteristics of ore geology and ore-forming fluid in the Zhaxikang Sb-Pb-Zn-Ag deposit, Southern Tibet, China. Master's Thesis, China University of Geoscience, Beijing (in Chinese). Zhang, Y.X, Wen, H.J, Zhu, C.W, Fan, H.F, Luo, C.G, Liu, J., & Cloquet, C., 2016. Cd isotope fractionation during simulated and natural weathering. Environmental pollution 216, 9–17. Zheng, Y.Y., Liu, M.Y., Sun, X., Yuan, E.H., Tian, L.M., Zheng, H.T., Zhang, G.Y., Zhang, L.H., 2012. Type, discovery process and significance of Zhaxikang antimony polymetallic ore deposit, Tibet. Earth Science 37, 1004–1014 (In Chinese with English abstract). Zheng, Y.Y., Sun, X., Tian, L.M., Zheng, H.T., Miao, Y., Yang, W.T, Zhou, T.C., Geng, X.B., 2014. Mineralization, deposit type and metallogenic age of the gold antimony polymetallic belt in the eastern part of north Himalayan. Geotectonica Et Metallogenia 38, 108–118 (In Chinese with English abstract). Zhou, J., Huang, Z., Zhou, G., Li, X., Ding, W., & Bao, G., 2011. Trace Elements and Rare Earth Elements of Sulfide Minerals in the Tianqiao Pb-Zn Ore Deposit, Guizhou Province. China. Acta Geologica Sinica (English Edition) 85, 189–199. Zhu, B., Zhang, H.F., Zhao, X.M., & He, Y.S., 2016b. Iron isotope fractionation during skarn-type alteration:
30
708 709 710 711 712 713 714 715 716 717 718 719 720
implications for metal source in the Han-Xing iron skarn deposit. Ore Geology Review 74, 139–150. Zhu, C.W, Wen, H.J, Zhang, Y.X, Fan, H.F, Fu, S.H, Xu, J., & Qin, T.R., 2013. Characteristics of Cd isotopic compositions and their genetic significance in the lead-zinc deposits of SW China. Science China Earth Science 56(12), 2056–2065. Zhu, C.W., Wen, H.J., Zhang, Y.X., Fan, H.F., Liu, J., Zhou, Z.B., 2015a. Analytical technique for cadmium stable isotopes and its applications. Earth Science Frontiers 22(5), 115–123 (In Chinese with English abstract). Zhu, C.W., Wen, H.J., Zhang, Y.X., Liu, Y.Z., & Wei, R.F., 2015b. Isotopic geochemistry of cadmium: a review. Acta Geologica Sinica (English Edition) 89(6), 2048–2057. Zhu, C.W, Wen, H.J, Zhang, Y.X, & Fan, H.F., 2016a. Cadmium and sulfur isotopic compositions of the Tianbaoshan Zn–Pb–Cd deposit, Sichuan Province, China. Ore Geology Reviews 76, 152-162. Zhu, C.W, Wen, H.J, Zhang, Y.X, Fu, S.H, Fan, H.F., & Cloquet, C., 2017. Cadmium isotope fractionation in the Fule Mississippi Valley-type deposit, Southwest China. Mineralium Deposita 52(5), 675–686.
31
721
Figure captions
722
Fig. 1. (A) Tectonic framework of the Himalayan Terrane (Yin, 2006); (B) Regional geological map of the North
723
Himalayan Metallogenic Belt (modified from Zheng et al., 2014; Wang et al., 2018a); (C) Geological map of the
724
Zhaxikang Sb–Pb–Zn–Ag deposit (modified from Zheng et al., 2012; Wang et al., 2018a); (D) Cross-section
725
along the Exploration Line 8 (modified from Wang et al., 2019).
726
IYZS: the Indus-Yarlung Zangbo Suture Zone; STDS: the South Tibet Detachment System; MCT: the Main
727
Central Thrust Fault; MBT: the Main Boundary Thrust Fault; MFT: the Main Frontal Thrust; NH: the North
728
Himalayan Tethys sedimentary fold belt; HH: the High Himalayan crystalline rock belt; LH: the Low Himalayan
729
fold belt; SH: the Sub-Himalayan tectonic belt.
730 731
Fig. 2. Ore paragenetic sequence within the Zhaxikang deposit (modified from Wang et al., 2018a).
732 733
Fig. 3. Hand specimen photographs of representative samples from the Zhaxikang deposit. (A) Stage 1 lamellar
734
sphalerite-pyrite-arsenopyrite and stage 2 massive and banded sphalerite-pyrite hosted by fine-grained Mn–Fe
735
carbonate. The mineral assemblage is in turn cross-cut by stage 4 quartz-boulangerite veins (cited from Wang et
736
al., 2018a). (B) Stage 1 lamellar sphalerite-pyrite-arsenopyrite and stage 2 massive sphalerite-pyrite hosted within
737
fine-grained Mn–Fe carbonate (cited from Wang et al., 2018a). (C) Stage 2 massive sphalerite-galena-pyrite
738
hosted by coarse-grained Mn–Fe carbonate. (D) Stage 2 massive, globular and concentric annular sphalerite-
739
pyrite hosted by coarse-grained Mn–Fe carbonate (cited from Wang et al., 2018a). (E) Stage 3 quartz veins cross-
740
cut the pyrite-sphalerite to form the banded and brecciated texture. The sample also contains minor amounts of
741
early Mn–Fe carbonate. (F) Stage 3 brecciated sphalerite within stage 3 quartz-calcite (cited from Wang et al.,
32
742
2018a). (G) Stage 3 sphalerite-galena veins cross-cut by stage 6 quartz-calcite veins (cited from Wang et al.,
743
2018a). (H) Stage 4 boulangerite-quartz (cited from Wang et al., 2018a). (I) Elongate stage 5 stibnite hosted by
744
stage 5 quartz.
745
Abbreviations are as follows: Mcar1 = stage 1 fine-grained Mn–Fe carbonate; Apy1 = stage 1 lamellar
746
arsenopyrite; Py1 = stage 1 lamellar pyrite; Sp1 = stage 1 lamellar sphalerite; Mcar2 = stage 2 coarse-grained
747
Mn–Fe carbonate; Py2 = stage 2 pyrite; Sp2 = stage 2 sphalerite; Gn2 = stage 2 galena; Py3 = stage 3 pyrite; Sp3
748
= stage 3 sphalerite; Gn3 = stage 3 galena; Ccp3 = stage 3 chalcopyrite; Qtz3 = stage 3 quartz; Cal3 = stage 3
749
calcite; Blr4 = stage 4 boulangerite; Qtz4 = stage 4 quartz; Stb5 = stage 5 stibnite; Qtz5 = stage 5 quartz; Cal6 =
750
stage 6 calcite; Qtz6 = stage 6 quartz.
751 752
Fig. 4. (A) Plot of Cd vs Zn concentrations for galena from Zhaxikang deposit; (B) Plot of Cd vs Zn
753
concentrations for sphalerite from Zhaxikang deposit; (C) The comparison in Cd concentrations of galena
754
between Zhaxikang and Fule MVT deposits (the data of Fule MVT deposit is cited from Zhu et al., 2017).
755 756
Fig. 5. The (A) BSE image (cited from Wang et al. 2018a) and (B) photomicrograph for ore samples from
757
Zhaxikang deposit. Abbreviations are as Fig. 3.
758 759
Fig. 6. Distribution of Zn/Cd ratios versus δ144/110Cd values for sphalerite in different mineralization systems,
760
compared with Zhaxikang deposit (modified from Wen et al., 2016).
761 762
Fig. 7. Evolution of δ144/110Cd values during the deposition of aqueous Cd in different hydrothermal fluids,
763
compared with Zhaxikang deposit. Dashed lines represent the evolution of the deposited minerals, solid lines 33
764
represent the evolution of residual aqueous Cd, grey fields represent the observed fractionation range of the
765
sphalerite samples (modified from Wen et al., 2016).
766 767
Fig. 8. (A) Diagram of δ114/110Cd values versus Cd concentrations of sphalerite in Zhaxikang deposit; (B) Diagram
768
of δ114/110Cd values versus Cd concentrations of galena in Zhaxikang deposit.
769 770
Fig. 9. The δD–18O H2O diagram of Mn–Fe carbonate in Zhaxikang deposit (the H–O isotopic data are cited
771
from Xie et al., 2017; and the base map are modified from Yang et al., 2009).
772 773
Fig. 10. (A) The Cd isotopic fractionation within galena-sphalerite coexisting mineral pairs in Zhaxikang deposit;
774
(B) The Cd isotopic equilibrium fractionation diagram of galena-sphalerite coexisting mineral pairs in Zhaxikang
775
deposit.
776 777
Fig. 11. (A) The temporally decreasing δ114/110Cd values from stages 1 to 3 for sphalerite in Zhaxikang deposit;
778
(B) The temporally decreasing δ66Zn values from stages 1 to 3 for sphalerite in Zhaxikang deposit.
779 780
Fig. 12. The temporally decreasing trend in δ66Zn and δ56Fe values within concentric ore samples from Zhaxikang
781
deposit (modified from Wang et al., 2018a).
782 783
Fig.13. Conceptual model illustrating how Cd partition among mineralizing vapor and liquid phases. Box
784
indicates the minerals formed, pentagon represents the vapor and ellipse indicates the starting hydrothermal fluid
34
785
(modified from Yao et al., 2016).
786 787
Fig.14. (A) The Cd isotopic Rayleigh distillation model for vapor-liquid partitioning, comparing with the total
788
and concentrated δ144/110Cd values of sphalerite in Zhaxikang deposit; (B) The Zn isotopic Rayleigh distillation
789
model for vapor-liquid partitioning, comparing with the concentrated δ66Zn values of sphalerite in Zhaxikang
790
deposit (cited from Wang et al., 2018a); (C) Comparison between the F ranges of sphalerite in Cd isotopic
791
Rayleigh distillation model for vapor-liquid partitioning and the vapor-liquid ratios of vapor-liquid two-phase
792
fluid inclusions in Zhaxikang deposit (Yu, 2015).
793 794 795 796
Conflict of interest We declare that we do not have any commercial or associative interest that represents a conflict of interest in connection with the submitted work.
797 798 799
Highlights
800
(1) Vapor-liquid partitioning is main cause for observed Cd isotopic variations in Zhaxikang
801
deposit.
802
(2) The SEDEX genesis is most plausible for the Pb–Zn mineralization in Zhaxikang deposit.
803
(3) Kinetic Rayleigh fractionation related to vapor-liquid partitioning is demonstrated as an
804
important Cd isotopic fractionation mechanism in hydrothermal ore-forming system.
805
(4) Cd isotopic fractionation induced by vapor-liquid partitioning can be used to identify ore, rock,
806
sediments and solar system forming processes. 35
807 808 809
810 811 812
813 814
Table 1 The cup configuration of the MC-ICP-MS for Cd isotopic analyses. Cup H4 L4 L2 L1 Ax H1 H2 Number 116Sn 107Ag 109Ag 110Cd 111Cd 112Cd 113Cd Isotope
H3 114Cd
Table 2 The Zn and Cd concentrations and δ114/110Cd values of sulfide samples from the Zhaxikang deposit. Sample Number Mineral Zn (%) Cd (ppm) Zn/Cd δ114/110CdNIST SRM 3108 2σ (‰) ZXK-PD9-B2-1 Sp1 40.456 1183 342 1.01 0.04 Stage 1 lamellar spha sphalerite-pyrite hos D52-3 Sp1 58.440 1606 364 0.10 0.04 Stage 1 lamellar spha and banded sphale carbonate. The mine quartz-boulangerite v 14-6 Sp1 55.308 1543 358 –0.13 0.04 Stage 1 lamellar spha sphalerite hosted by 14-9 Sp1 50.618 1202 421 –0.30 0.04 Stage 1 lamellar spha sphalerite hosted by also contains a Mn–F columnar quartz and ZXK-PD9-B2-2 Sp2 49.599 1581 314 –0.10 0.04 The same as ZXK-PD 9-4 Sp2 53.821 1437 375 –0.09 0.04 Stage 2 globular and 9-4 Gn2 0.030 14 22 –2.19 0.04 coarse-grained Mn–F ZXK-12-B134 Sp2 50.325 1961 257 –0.51 0.04 Massive ore contain ZXK-12-B134 Gn2 0.030 22 14 2.24 0.04 stage 2 coarse-graine of quartz grains. ZK007-B4 Sp3 54.433 2199 248 –0.34 0.04 Stage 3 brecciated an ZK007-B4 Gn3 0.121 26 46 0.01 0.04 calcite, and the sam hosted by sphalerite. ZXK-XC-3 Sp3 50.521 1655 305 –0.23 0.04 Stage 3 quartz vein ZXK-XC-3 Gn3 0.153 19 82 –0.21 0.04 banded and brecciate Abbreviations: Sp1= stage 1 sphalerite, Sp2 = stage 2 sphalerite, Sp3 = stage 3 sphalerite, Gn2 = stage 2 galena, Gn3 = stage 3 galena.
815
36