• Aucun résultat trouvé

Tourmaline as a tracer of late-magmatic to hydrothermal fluid evolution: The world-class San Rafael tin (-copper) deposit, Peru

N/A
N/A
Protected

Academic year: 2022

Partager "Tourmaline as a tracer of late-magmatic to hydrothermal fluid evolution: The world-class San Rafael tin (-copper) deposit, Peru"

Copied!
69
0
0

Texte intégral

(1)

Article

Reference

Tourmaline as a tracer of late-magmatic to hydrothermal fluid evolution: The world-class San Rafael tin (-copper) deposit, Peru

HARLAUX, Matthieu, et al.

Abstract

The world-class San Rafael tin (-copper) deposit (central Andean tin belt, southeast Peru) is an exceptionally large and rich (>1 million metric tons Sn; grades typically >2% Sn) cassiterite-bearing hydrothermal vein system hosted by a late Oligocene (ca. 24 Ma) peraluminous K-feldspar-megacrystic granitic complex and surrounding Ordovician shales affected by deformation and low-grade metamorphism. The mineralization consists of NWtrending, quartz-cassiterite-sulfide veins and fault-controlled breccia bodies (>1.4 km in vertical and horizontal extension). They show volumetrically important tourmaline alteration that principally formed prior to the main ore stage, similar to other granite-related Sn deposits worldwide. We present here a detailed textural and geochemical study of tourmaline, aiming to trace fluid evolution of the San Rafael magmatic-hydrothermal system that led to the deposition of tin mineralization. Based on previous works and new petrographic observations, three main generations of tourmaline of both magmatic and hydrothermal origin were distinguished and were analyzed in situ for their major, minor, and [...]

HARLAUX, Matthieu, et al. Tourmaline as a tracer of late-magmatic to hydrothermal fluid evolution: The world-class San Rafael tin (-copper) deposit, Peru. Economic Geology, 2020, vol. 115, no. 8, p. 1665-1697

DOI : 10.5382/econgeo.4762

Available at:

http://archive-ouverte.unige.ch/unige:148392

Disclaimer: layout of this document may differ from the published version.

1 / 1

(2)

Tourmaline as a tracer of late-magmatic to hydrothermal fluid evolution:

1

The world-class San Rafael tin (-copper) deposit, Peru 2

3

Matthieu Harlaux1,2*, Kalin Kouzmanov1, Stefano Gialli1, Oscar Laurent3, Andrea Rielli4, 4

Andrea Dini4, Alain Chauvet5, Andrew Menzies6, Miroslav Kalinaj7, and Lluís Fontboté1 5

6 7

1 Department of Earth Sciences, University of Geneva, 1205 Geneva, Switzerland 8

2 Present address: Nevada Bureau of Mines and Geology, University of Nevada, Reno, NV 9

89557-0178, USA 10

3 Institute of Geochemistry and Petrology, ETH Zürich, 8092 Zürich, Switzerland 11

4 Istituto di Geoscienze e Georisorse, CNR, 56124 Pisa, Italy 12

5 Géosciences Montpellier, CNRS-UMR 5243, Université de Montpellier, 34095 Montpellier, 13

France 14

6 Bruker Nano GmbH, Am Studio 2D, 12489 Berlin, Germany 15

7 Minsur S.A., Jr. Lorenzo Bernini 149, San Borja, Lima 27, Peru 16

17

* Corresponding author: mharlaux@unr.edu 18

19

v. 115, no. 8, p. 1665-1697. (Changes introduced at the proof stage not included in this version)

(3)

Abstract 20

21

The world-class San Rafael tin (-copper) deposit (central Andean tin belt, southeast Peru) 22

is an exceptionally large and rich (>1 Mt Sn; grades typically >2% Sn) cassiterite-bearing 23

hydrothermal vein system hosted by a late Oligocene (ca. 24 Ma) peraluminous K-feldspar- 24

megacrystic granitic complex and surrounding Ordovician shales affected by deformation and 25

low-grade metamorphism. The mineralization consists of northwest-trending, quartz- 26

cassiterite-sulfide veins and fault-controlled breccia bodies (>1.4 km in vertical and horizontal 27

extension). They show volumetrically important tourmaline alteration that principally formed 28

prior to the main ore stage, similar to other granite-related Sn deposits worldwide. We present 29

here a detailed textural and geochemical study of tourmaline, aiming to trace fluid evolution of 30

the San Rafael magmatic-hydrothermal system that led to the deposition of tin mineralization.

31

Based on previous works and new petrographic observations, three main generations of 32

tourmaline of both magmatic and hydrothermal origin were distinguished and were analyzed in 33

situ for their major, minor, and trace element composition by electron microprobe analyzer 34

(EMPA) and laser ablation–inductively coupled plasma–mass spectrometry (LA-ICP-MS), as 35

well as for their bulk Sr, Nd, and Pb isotopic compositions by multi-collector–inductively 36

coupled plasma–mass spectrometry (MC-ICP-MS). A first late-magmatic tourmaline 37

generation (Tur 1) occurs in peraluminous granitic rocks as nodules and disseminations, which 38

do not show evidence of alteration. This early Tur 1 is texturally and compositionally 39

homogeneous, it has a dravitic composition, with Fe/(Fe+Mg) = 0.36-0.52, close to the schorl- 40

dravite limit, and relatively high contents (10s-100s ppm) of Li, K, Mn, LREE, and Zn. The 41

second generation (Tur 2), the most important volumetrically, is pre-ore, high-temperature 42

(>500°C), hydrothermal tourmaline occurring as phenocryst replacement (Tur 2a) and open- 43

space fillings in veins and breccias (Tur 2b), and microbreccias (Tur 2c) emplaced in the host 44

granites and shales. Pre-ore Tur 2 typically shows oscillatory zoning, possibly reflecting rapid 45

changes in the hydrothermal system, and has a large compositional range that spans the schorl 46

to dravite fields with Fe/(Fe+Mg) = 0.02-0.83. Trace element contents of Tur 2 are similar to 47

Tur 1. Compositional variations within Tur 2 may be explained by the different degree of 48

interaction of the magmatic-hydrothermal fluid with the host rocks (granites and shales), in part 49

due to the effect of replacement vs. open-space filling. The third generation is syn-ore 50

hydrothermal tourmaline (Tur 3). It forms microscopic veinlets and overgrowths, partly cutting 51

previous tourmaline generations, and is locally intergrown with cassiterite, chlorite, quartz, and 52

minor pyrrhotite and arsenopyrite from the main ore assemblage. Syn-ore Tur 3 has schorl- 53

(4)

foititic compositions with Fe/(Fe+Mg) = 0.48-0.94 that partly differ from those of late- 54

magmatic Tur 1 and pre-ore hydrothermal Tur 2. Relative to Tur 1 and Tur 2, syn-ore Tur 3 has 55

higher contents of Sr and HREE (10s-100s ppm), and unusually high contents of Sn (up to 56

>1000 ppm). Existence of these three main tourmaline generations, each having specific 57

textural and compositional characteristics, reflects a boron-rich protracted magmatic- 58

hydrothermal system with repeated episodes of hydrofracturing and fluid-assisted reopening 59

generating veins and breccias. Most trace elements in the San Rafael tourmaline do not correlate 60

with Fe/(Fe+Mg) ratios, suggesting that their incorporation was likely controlled by the 61

melt/fluid composition and local fluid-rock interactions. The initial radiogenic Sr and Nd 62

isotopic compositions of the three aforementioned tourmaline generations (0.7160-0.7276 for 63

87Sr/86Sr(i) and 0.5119-0.5124 for 143Nd/144Nd(i)) mostly overlap those of the San Rafael granites 64

(87Sr/86Sr(i) = 0.7131-0.7202 and 143Nd/144Nd(i) = 0.5121-0.5122) and support a dominantly 65

magmatic origin of the hydrothermal fluids. These compositions also overlap the initial Nd 66

isotope values of Bolivian tin porphyries. The initial Pb isotopic compositions of tourmaline 67

show larger variations, with 206Pb/204Pb(i), 207Pb/204Pb(i), and 208Pb/204Pb(i) ratios mostly falling 68

in the range of 18.6-19.3, 15.6-16.0, and 38.6-39.7, respectively. These compositions partly 69

overlap the initial Pb isotopic values of the San Rafael granites (206Pb/204Pb(i) = 18.6-18.8, 70

207Pb/204Pb(i) = 15.6-15.7, and 208Pb/204Pb(i) = 38.9-39.0) and are also similar to those of other 71

Oligocene to Miocene Sn-W ± Cu-Zn-Pb-Ag deposits in southeast Peru. REE patterns of 72

tourmaline are characterized, from Tur 1 to Tur 3, by decreasing (Eu/Eu*)N ratios (from 20 to 73

2) that correlate with increasing Sn contents (from 10s to >1000 ppm). These variations are 74

interpreted to reflect evolution of the hydrothermal system from reducing towards relatively 75

more oxidizing conditions, still in a low-sulfidation environment as indicated by the pyrrhotite- 76

arsenopyrite assemblage. The changing textural and compositional features of Tur 1 to Tur 3 77

reflect the evolution of the San Rafael magmatic-hydrothermal system and support the model 78

of fluid mixing between reduced, Sn-rich magmatic fluids and cooler, oxidizing meteoric 79

waters as the main process that caused cassiterite precipitation.

80 81

(5)

Introduction 82

83

Tourmaline is the main borosilicate mineral in granitic rocks and associated magmatic- 84

hydrothermal ore deposits (e.g., Sn-W veins and greisens, pegmatites, Cu-Au breccia pipes, 85

Cu-Mo-(Au) porphyries, IOCG deposits), where it is mostly found as late-magmatic 86

disseminations, hydrothermal veins and breccias, and metasomatic replacements (e.g., London 87

et al., 1996; Slack, 1996; Černý et al., 2005; Slack and Trumbull, 2011). Due to its stability 88

over a wide pressure-temperature range and negligible intra-crystalline diffusion, tourmaline is 89

an especially refractory mineral that preserves the initial signature of physicochemical 90

processes during its crystallization (e.g., Dutrow and Henry, 2011; Marschall and Jiang, 2011;

91

van Hinsberg et al., 2011). Tourmaline is common in granite-related Sn-W deposits and has 92

been used as a proxy for tracing ore-forming processes and fluid origins (e.g., Duchoslav et al., 93

2017; Codeço et al., 2017, 2019; Hong et al., 2017; Launay et al., 2018; Harlaux et al., 2019a).

94

Previous studies have shown that major element compositions of tourmaline reflects 95

dominantly those of the host rocks and mineralizing fluids (e.g., Henry and Guidotti, 1985;

96

Slack et al., 1993; Henry and Dutrow, 1996; van Hinsberg et al., 2011), whereas incorporation 97

of trace elements is controlled by melt/fluid composition, local fluid-rock interactions, and 98

crystal-chemical effects (e.g., van Hinsberg, 2011; Marks et al., 2013; Hazarika et al., 2015;

99

Yang et al., 2015; Kalliomäki et al., 2017).

100

In this study, we address the question of how textural and compositional variations of 101

tourmaline reflect the evolution of late-magmatic to hydrothermal processes leading to the 102

precipitation of ore minerals. Herein, tourmaline was analyzed from the world-class San Rafael 103

Sn (-Cu) deposit (central Andean tin belt, southeast Peru) that represents an exceptionally large 104

and rich, granite-related cassiterite-bearing vein system extensively studied by previous 105

workers (Arenas, 1980; Palma, 1981; Kontak and Clark, 2002; Mlynarczyk et al., 2003;

106

Mlynarczyk, 2005; Mlynarczyk and Williams-Jones, 2006; Wagner et al., 2009; Corthay, 2014;

107

Prado, 2015). The central Andean tin belt is a classic metallogenic province for Sn-W deposits 108

in which widespread tourmaline occurs in veins, stockworks, breccias, and greisens (e.g., 109

Sillitoe et al., 1975; Lehmann et al., 1990, 2000). Tourmalinization is related to the highly 110

evolved character of magmas of the central Andean tin belt as indicated by high boron contents 111

(avg = 225 ppm B) measured in melt inclusions from Bolivian tin porphyries (Dietrich et al., 112

2000; Lehmann et al., 2000; Wittenbrink et al., 2009). The volumetrically important occurrence 113

of tourmaline at San Rafael, principally formed prior to the main ore stage, can be compared to 114

other granite-related Sn-dominated deposits, such as those of the Cornwall province in 115

(6)

southwest England (e.g., Drivenes et al., 2015; Duchoslav et al., 2017), as well as to granite- 116

related W-dominated deposits, such as Panasqueira in Portugal (e.g., Codeço et al., 2017, 2019;

117

Launay et al., 2018).

118

Several generations of tourmaline, from late-magmatic to syn-ore hydrothermal, have been 119

previously documented in the San Rafael deposit (Table 1; Kontak and Clark, 2002;

120

Mlynarczyk et al., 2003; Mlynarczyk and Williams-Jones, 2006), making it a particularly 121

appropriate case study to test the use of tourmaline as a tracer of fluid evolution. Based on 122

previous studies on the San Rafael deposit and new petrographic observations, we distinguish 123

for the scope of the present work three main tourmaline generations: (i) late-magmatic Tur 1 124

occurring as nodules and disseminations in peraluminous granites, (ii) pre-ore hydrothermal 125

Tur 2 forming replacements and open-space fillings in veins, breccias, and microbreccias, and 126

(iii) syn-ore hydrothermal Tur 3 occurring as late microscopic veinlets and overgrowths, locally 127

intergrown with cassiterite, chlorite, and quartz from the main ore assemblage. Using a large 128

set of representative samples of these three tourmaline generations, we present here a 129

comprehensive dataset of in situ analysis of major, minor, and trace elements and bulk Sr, Nd, 130

and Pb isotopic compositions. We show that the changing textural and compositional features 131

of Tur 1 to Tur 3 reflect the evolution of the San Rafael magmatic-hydrothermal system, 132

including increasing Sn contents (from 10s to >1000 ppm) and changes from reducing toward 133

relatively more oxidizing conditions. These results are consistent with and support the model 134

of fluid mixing between reduced, Sn-rich magmatic fluids and cooler, oxidizing meteoric 135

waters, as the main process that caused cassiterite precipitation at San Rafael.

136 137

Regional geology of the central Andean tin belt 138

139

The San Rafael Sn (-Cu) deposit (latitude 14°13’58” S, longitude 70°19’18” W) is located 140

within the Cordillera de Carabaya in the Eastern Cordillera of southeast Peru and defines the 141

northern end of the central Andean tin belt (Fig. 1A). This metallogenic province extends ca.

142

1000 km to the southeast as a 30- to 130-km-wide belt from southern Peru through Bolivia to 143

northern Argentina, and hosts hundreds of Sn-W ± Cu-Zn-Pb-Ag vein deposits (e.g., Kelly and 144

Turneaure, 1970; Turneaure, 1971; Grant et al., 1979; Lehmann et al., 1990). These deposits 145

are spatially associated with peraluminous granitoid intrusions and subvolcanic stocks that were 146

emplaced during three major metallogenic periods: (i) Late Devonian - early Carboniferous;

147

(ii) Late Triassic - Early Jurassic, restricted to the northern part of the belt; and (iii) Late 148

Oligocene - Early Miocene, affecting the entire belt (e.g., Grant et al., 1979; Clark et al., 1983, 149

(7)

1990; McBride et al., 1983; Rice et al., 2005). The most important episode for deposition of 150

Sn-W ore was the last, between 25 and 12 Ma, including several world-class deposits such as 151

San Rafael, Llallagua, Cerro Rico de Potosi, and Chorolque (e.g., Kelly and Turneaure, 1970;

152

Turneaure, 1971; Sillitoe et al., 1975; Lehmann et al., 1990). The Cordillera de Carabaya in 153

southeast Peru consists of a more than 10-km-thick sequence of lower Paleozoic 154

metasedimentary rocks (San José, Sandia, and Ananea formations) overlying unexposed 155

Precambrian gneissic basement that were all deformed and metamorphosed during the Variscan 156

orogeny (Laubacher, 1974; Clark et al., 1990; Sandeman et al., 1996). The metasedimentary 157

rocks were intruded by Oligocene to Miocene S-type granitic bodies and overlain by volcanic 158

rocks that are part of the Crucero Supergroup (e.g., Kontak et al., 1987; Laubacher et al., 1988;

159

Clark et al., 1990; Kontak et al., 1990; Cheilletz et al., 1992; Sandeman et al., 1997).

160 161

Geology of the San Rafael Sn (-Cu) deposit 162

163

San Rafael currently is one of the largest and highest-grade primary Sn deposits in the 164

world, with total past production of >1 Mt of Sn, and remaining total reserves estimated at 8 165

Mt of ore at 1.74% Sn (Minsur S.A., 2018). According to internal Minsur Reports, between 166

1969, when the production started at around 4900 masl, and 2019, when the underground 167

workings have deepened down to 3600 masl, the San Rafael mine has produced 26,608,702 t 168

ore with 3.7% Sn on average. Production in 2019 amounted 1,159,299 t ore at 1.86% Sn. Until 169

1978, when only the upper parts of the deposit were mined, Cu grades were higher than those 170

of Sn (for example 29,100 t at 4.62% Cu and 1.14% Sn in 1969, and 111,926 t at 1.49% Cu and 171

1.31% Sn in 1978). Subsequently, Cu grades decreased until 1986 (229,784 t at 0.37% Cu and 172

2.84% Sn), last year when Cu production has been reported. Between 1987 and 2012, Sn grades 173

were >3% with a maximum of 6.76% Sn for 356,106 t ore in 1994. At that time, bonanza ore 174

with exceptional grades reaching as high as >20% Sn was mined. Since 2013, Sn grades have 175

decreased from 2.72% in 2013 to 1.86% in 2019, in line with the progressive deepening of the 176

mining operations and the increase of mining efficiency.

177

The mineralization consists of a northwest-trending, quartz-cassiterite-sulfide vein system 178

spatially associated with a late Oligocene granitic complex hosted in Ordovician shales of the 179

Sandia Formation that was affected by contact metamorphism during granite emplacement (Fig.

180

1B; Kontak and Clark, 2002; Mlynarczyk et al., 2003; Gialli et al., 2017). The intrusive complex 181

includes the San Rafael granite that crops out in the southwestern part of the district and the 182

Quenamari granite in the northeast, both being part of the same intrusion at depth, as indicated 183

(8)

by drilling and underground workings (Fig. 1C). The San Rafael and Quenamari granitic bodies 184

consist dominantly of peraluminous porphyritic biotite-cordierite-bearing monzogranite, 185

exhibiting several centimeter size phenocrysts of K-feldspar, and were extensively studied by 186

Kontak and Clark (2002). Minor enclaves of biotite microgranite and diorite, showing mingling 187

textures with the host K-feldspar-megacrystic granite, have also been described in the upper 188

part of the stock (Kontak and Clark, 2002; Mlynarczyk et al., 2003). The San Rafael and the 189

Quenamari granitic intrusions are surrounded by porphyritic ring dikes, which are 190

petrologically similar to the central granites and exhibit quenched textures (Fig. 1B; Kontak 191

and Clark, 2002). In the southwestern part of the San Rafael granite, a greisen (quartz + 192

muscovite ± tourmaline assemblage) crops out at an altitude of >4800 masl for a few hundred 193

meters, elongated in a NNW-SSE direction (Fig. 1B). This area of intense greisenization is 194

interpreted to represent the apical part of the granitic pluton, where acidic magmatic fluids 195

accumulated during cooling of the parental magma (Gialli et al., 2019). Clark et al. (2000) and 196

Kontak and Clark (2002) dated the San Rafael granite at 24.6 to 24.7 ± 0.2 Ma (U-Pb zircon 197

and monazite ages). Hydrothermal activity is dated between 24.1 ± 0.1 Ma and 22.0 ± 0.2 Ma 198

(40Ar/39Ar muscovite and adularia plateau ages; Clark et al., 2000; Kontak and Clark, 2002;

199

Gialli et al., 2017). Dikes of lamprophyres, 10 to 50 cm in thickness and having a phlogopite + 200

K-feldspar + plagioclase ± quartz dominated assemblage, cut the San Rafael intrusive complex 201

and partly crop out in the southwestern margin of the Quenamari granite (Fig. 1B-C).

202

Lamprophyre enclaves within the San Rafael K-feldspar-megacrystic granite have also been 203

reported (Kontak and Clark, 2002). Lamprophyric magmas mixing with peraluminous S-type 204

biotite-cordierite magmas (“hybridization”) represent a common feature of the Crucero 205

Supergroup in the Eastern Cordillera, where they emplaced between ca. 26 and 22 Ma (e.g., 206

Carlier et al., 1997; Sandeman et al., 1997; Sandeman and Clark, 2003).

207

The paragenetic sequence of the San Rafael deposit comprises four main stages (Palma, 208

1981; Kontak and Clark, 2002; Mlynarczyk et al., 2003): (i) Stage I corresponds to formation 209

of early barren quartz-tourmaline veins and breccia bodies, which cut the granitic intrusion and 210

surrounding host rocks. This stage was preceded by early episodes of hydrothermal alteration, 211

including incipient potassic and sericitic alteration of magmatic plagioclase and K-feldspar in 212

the granites, and pervasive sodic alteration forming hydrothermal albite haloes around the 213

quartz-tourmaline veins and affecting the granitic groundmass (Kontak and Clark, 2002; Gialli 214

et al., 2019). (ii) Stage II consists of the Sn ore-stage assemblage composed of quartz, 215

cassiterite, chlorite, and minor tourmaline, pyrrhotite, and arsenopyrite, together forming veins 216

and breccias that locally reopened pre-existing quartz-tourmaline veins-breccias. This stage is 217

(9)

associated with pervasive chloritization, mainly along veins and breccia bodies, that affected 218

both the granites and the metasedimentary host rocks. (iii) During stage III a sulfide-dominant 219

assemblage, comprising mostly pyrrhotite, chalcopyrite, arsenopyrite, galena, and sphalerite, 220

was deposited. This stage occurs as quartz-sulfide veins and as late infill that cut or reopened 221

pre-existing vein generations. (iv) Late barren quartz-carbonate (calcite, siderite) stage IV 222

veins, containing minor fluorite and adularia, cut all previous veins and breccias.

223

In the present investigation, we studied representative samples from the San Rafael deposit 224

covering a large spectrum of different generations of magmatic and hydrothermal tourmaline.

225

The majority of the studied samples come from the San Rafael mine and were collected from 226

underground workings between 3610 masl and 4475 masl.

227 228

Analytical methods 229

230

Scanning electron microscopy and automated mineralogy (QEMSCAN) 231

232

Mineralogical observations of tourmaline were carried out using transmitted-light 233

microscopy combined with scanning electron microscopy (SEM) using a JEOL JSM7001F 234

SEM equipped with an energy-dispersive X-ray spectrometer (EDS) at the University of 235

Geneva, Switzerland. Back-scattered electron (BSE) images were acquired on carbon-coated 236

polished thin sections using an acceleration voltage of 15 kV, adjusting the image contrast to 237

reveal internal zoning within the tourmaline grains. Scanning electron microscopy was used to 238

study internal textures of tourmaline and to select grains for in situ chemical analyses.

239

Automated mineral analysis and textural imaging were performed using an FEI QEMSCAN 240

Quanta 650F facility at the University of Geneva. The QEMSCAN system is equipped with 241

two Bruker QUANTAX light-element EDS detectors. Analyses were conducted at high 242

vacuum, an accelerating voltage of 25 kV, and with a beam current of 10 nA on carbon-coated 243

polished thin sections. The field image operating mode (Pirrie et al., 2004) was used for 244

analyses. In total, 221 individual fields were measured per sample, with a field size of 1500 x 245

1500 µm and a point spacing of 5 µm. The standard 1000 counts per point were acquired, 246

yielding a limit of detection of approximately 2 wt.% per element for mineral classification.

247

Measurements were performed using the iMeasure v5.3.2 software; the iDiscover v5.3.2 248

software package was used for data processing. Results consist of: (i) spatially resolved and 249

fully quantified mineralogical maps; (ii) BSE images with identical resolution as in the 250

mineralogical maps; and (iii) X-ray elemental distribution maps.

251

(10)

252

Electron microprobe analysis 253

254

Major and minor element compositions of tourmaline were determined by electron 255

microprobe analyses (EMPA) using a JEOL JXA-8200 Superprobe microanalyzer equipped 256

with five wavelength-dispersive X-ray spectrometers (WDS) at the University of Geneva.

257

Operating conditions were as follows: acceleration voltage of 15 kV, beam current of 20 nA, 258

and beam diameter of 5 μm. Counting times on element peaks and backgrounds were 16 sec 259

and 8 sec, respectively, for Si, Al, K, Ca, Fe, Mg, Mn, Ti, and 30 sec and 15 sec, respectively, 260

for Na, Cr, F, Cl. The following standards were used for calibration: albite (Si, Al, Na), olivine 261

(Mg), fayalite (Fe), synthetic MnTiO3 (Mn), rutile (Ti), orthoclase (K), wollastonite (Ca), Cr2O3

262

(Cr), topaz (F), and tugtupite (Cl). Limits of detection are approximately: 200 ppm for Fe and 263

F; 170 ppm for Si; 130 ppm for Al, Mg, Mn, and Cr; 100 ppm for Na, K, Ti, and Ca; and 40 264

ppm for Cl. Structural formulae of tourmaline were calculated using the WinTcac software of 265

Yavuz et al. (2014) by normalizing to 15 cations for the Y+Z+T sites and assuming a 266

stoichiometric 3 atoms for B and 4 atoms for OH+F, based on the general formula 267

XY3Z6(T6O18)(BO3)3V3W, where X = Na+, Ca2+, K+, and vacancy (X□); Y = Fe2+, Mg2+, Mn2+, 268

Al3+, Li+, Fe3+, and Cr3+; Z = Al3+, Fe3+, Mg2+, Ti4+, and Cr3+; T = Si4+ and Al3+; B = B3+; V = 269

OH-, O2-; and W = OH-, F-, and O2- (Henry et al., 2011). Chemical compositions of tourmaline 270

are reported in weight per cent (wt.%) oxides and structural formulae are expressed in atoms 271

per formula unit (apfu). EMPA X-ray elemental maps were acquired for F, Ca, Na, Ti, and Cl, 272

using an acceleration voltage of 15 kV, beam current of 20 nA, beam diameter of 5 μm, and 273

dwell times of 320 msec. The presented map is 1024 x 1024 pixels in size with a 5 µm per pixel 274

resolution, corresponding to an investigated area of 5120 x 5120 µm. Total acquisition time 275

was ca. 103 hrs. The program XMapTools 3.1.2 (Lanari et al., 2014) was used for processing 276

QEMSCAN and EMPA maps, including classification and analytical standardization based on 277

EMPA spot analyses along the investigated grains.

278 279

Laser ablation–inductively coupled plasma–mass spectrometry 280

281

Tourmaline trace element analyses were carried out at the ETH Zürich, by laser ablation–

282

inductively coupled plasma–mass spectrometry (LA-ICP-MS) using a RESOlution (Australian 283

Scientific Instruments) 193 nm ArF excimer laser system attached to an Element XR (Thermo 284

Scientific, Germany) sector-field mass spectrometer. Analyses were performed directly on 285

(11)

polished thin sections of ca. 30 µm thickness, loaded in a Laurin Technic S-155 dual-volume 286

ablation cell fluxed with carrier gas consisting of ca. 0.5 L·min−1 He (5.0 grade) and sample 287

gas from the ICP-MS consisting of ca. 1 L·min−1 Ar (6.0 grade). We used a laser repetition rate 288

of 5 Hz, spot diameters of 43 or 51 μm, and a laser output energy of ca. 40 to 45 mJ, 289

corresponding to an on-sample energy density of ca. 4 J·cm–2. Three pre-ablation pulses were 290

applied immediately before each analysis for surface cleaning. Signal homogenization was 291

performed using in-house Squid tubing. The ICP-MS was tuned for maximum sensitivity on 292

the high mass range while keeping the production of oxides low (248ThO+/232Th+ <0.25%).

293

Intensities for the 40 following isotopes were acquired using time resolved-peak jumping and 294

triple detector mode: 7Li, 9Be, 27Al, 29Si, 31P, 39K, 43Ca, 45Sc, 49Ti, 51V, 53Cr, 55Mn, 59Co, 60Ni, 295

65Cu, 66Zn, 85Rb, 88Sr, 89Y, 90Zr, 93Nb, 118Sn, 133Cs, 139La, 140Ce, 141Pr, 146Nd, 147Sm, 151Eu, 157Gd, 296

159Tb, 163Dy, 165Ho, 166Er, 169Tm, 172Yb, 175Lu, 181Ta, 208Pb, and 209Bi. Dwell times were set to 297

10 msec, except for 27Al, 29Si, and 49Ti (5 msec), 9Be and 118Sn (20 msec), and all REEs (30 298

msec). With these settings, the total sweep time was 914 msec. Each measurement consisted of 299

70 mass sweeps acquired over ca. 60 sec, split in 30 sec of gas blank measurement followed by 300

30 sec of sample ablation. The raw intensities were processed using the Matlab-based SILLS 301

software (Guillong et al., 2008). LA-ICP-MS spectra were individually checked for possible 302

presence of micro-inclusions whereas spikes unrelated to the analyzed sample were 303

systematically eliminated. Integration windows were defined only for plateau-like signals thus 304

avoiding parts of the sample signal contaminated by the matrix or micro-inclusion. NIST SRM 305

612 glass (Jochum et al., 2011) was used as a calibration reference material (analyzed with a 306

spot diameter of 43 µm), via conventional standard-sample bracketing to correct for sensitivity 307

drift throughout the analytical session. Matrix effects were corrected using as an internal 308

standard the wt.% SiO2 content determined by EMPA for each analytical spot. Repeated 309

analyses of glasses prepared from USGS natural reference materials (GSD-1G and GOR128- 310

G1; analyzed with spot diameter of 43 µm) were processed as unknowns to check accuracy and 311

reproducibility of the analyses. Results show that reproducibility ranges from 3 to 12% (2σ, 312

increasing with decreasing concentration) for all analyzed trace elements, and that the analyses 313

are accurate within this level of analytical uncertainty. Limits of detection (LOD) were 314

calculated using the equation of Pettke et al. (2012) and are reported in Supplementary Table 315

2. We also assessed trace element contamination of the LA-ICP-MS system, which can be 316

significant for Sn (Schlöglova et al., 2017), by repeated analysis of fused silica glass using a 317

laser repetition rate of 5 Hz, spot diameters of 43 and 51 μm, and laser output energy of ca. 168 318

mJ, corresponding to an on-sample energy density of ca. 10 J·cm–2. Results show that 319

(12)

contamination is negligible for most trace elements, i.e., <0.5 ppm and even <0.1 ppm for m/z 320

ratios >85. Notable exceptions are 39K (<2 ppm), 43Ca (<40 ppm), and 118Sn (<1 ppm).

321

Nevertheless, the corresponding levels of contamination are within analytical uncertainty (i.e., 322

<10% and mostly <3% relative) for these elements, and therefore can also be considered 323

negligible.

324 325

Multi-collector–inductively coupled plasma–mass spectrometry 326

327

Bulk Sr, Nd, and Pb isotopic analyses of tourmaline were performed at the IGG-CNR in 328

Pisa, Italy. Tourmaline samples were crushed and sieved to 75-350-µm mesh and then 329

processed through a multi-step heavy-liquid procedure using sodium heteropolytungstate 330

solution (LST Fastfloat; density 2.9 g·cm−3) followed by centrifugation at 1000-2000 rpm. The 331

recovered heavy fractions were cleaned with deionized water and then dried in a furnace at 332

40°C overnight. Pure tourmaline concentrates were handpicked (~99% purity) under a 333

binocular microscope and were ground with an agate mortar to <80 µm. Tourmaline separates 334

were powdered and ~100 mg aliquots were taken for Sr, Nd, and Pb isotopic analyses. The 335

aliquots were digested in perfluoroalkoxy alkane (PFA) vials with mixed HF + HNO3 on a hot 336

plate at ~100°C for 5 days until complete sample digestion. The solutions were evaporated to 337

dryness and residues were re-dissolved in HNO3 and HCl 6.6N at ~100°C for 1 day. Strontium, 338

Nd, and Pb were separated using conventional ion-exchange procedures. Isotopic analyses were 339

performed using a Thermo Fisher Neptune Plus multi-collector–inductively coupled plasma–

340

mass spectrometer (MC-ICP-MS) at the IGG-CNR in Pisa. The instrument was equipped with 341

a combined cyclonic and Scott-type quartz-spray chamber, Ni-cones, and a MicroFlow PFA 342

100 µL·min−1 self-aspiring nebulizer. Strontium isotopic analyses were corrected for mass bias 343

fractionation using the 88Sr/86Sr ratio (8.375209) and for mass interference using the ratios 344

83Kr/84 Kr (0.201750), 83Kr/86Kr (0.664740), and 85Rb/87Rb (2.592310). Analytical accuracy 345

and long-term external reproducibility of Sr isotopic analyses were assessed using the reference 346

material NIST SRM 987 that yielded 87Sr/86Sr ratios of 0.710257 ± 0.000019. Instrumental 347

mass fractionation during Nd analyses was corrected using the 146Nd/144Nd ratio (0.7219). Mass 348

interference correction was done using the ratios 147Sm/144Sm(4.838710) and 147Sm/148Sm 349

(1.327400). Analytical accuracy and long-term external reproducibility of Nd isotopic analyses 350

were assessed using the reference material J-Ndi-1 that yielded 143Nd/144Nd ratios of 0.512098 351

± 0.000005. During Pb analyses, mass interference was corrected using the ratio 202Hg/204Hg 352

(4.350370). Samples were spiked with an in-house Tl standard and the ratio 203Tl/205Tl 353

(13)

(0.418882) was used to correct for mass bias fractionation. Analytical accuracy and long-term 354

external reproducibility of the Pb isotopic analyses were assessed over the measurement period 355

by analyzing replicates of the international standard SRM 981. The 2σ uncertainties for the 356

206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios are 0.02%, 0.03%, and 0.04%, respectively.

357 358

Results 359

360

Tourmaline textures and paragenetic sequence 361

362

Characteristics of the three main generations of tourmaline of both magmatic (Tur 1) and 363

hydrothermal origin (Tur 2 and Tur 3) distinguished in the San Rafael deposit are summarized 364

in the following. This classification is based on new petrographic evidence (crosscutting 365

relationships, replacements, overgrowths) and incorporates the findings of Kontak and Clark 366

(2002) and Mlynarczyk and Williams-Jones (2006). Differences with the classification of these 367

authors (mainly regarding Tur 1 and Tur 2) are shown in Table 1 and are also indicated in the 368

following.

369

Late-magmatic tourmaline (Tur 1): is found in peraluminous granites as quartz-tourmaline 370

nodules (Tur 1a) and disseminated grains (Tur 1b) that represent less than 2% of the rock 371

volume. Tourmaline is intergrown with magmatic quartz and K-feldspar that are devoid of 372

hydrothermal alteration, suggesting that Tur 1 is late-magmatic. Tur 1a nodules are hosted in 373

both the K-feldspar-megacrystic biotite-cordierite monzogranite and the microgranite, and form 374

centimeter-wide, rounded aggregates of intergrown tourmaline and quartz rimmed by a 0.5- to 375

1-cm-thick leucocratic halo composed of quartz and K-feldspar (Fig. 2C). The morphology, 376

distribution, and size of the quartz-tourmaline nodules at the scale of the granitic stock, both 377

vertically and laterally, have not been investigated. At a microscopic scale, Tur 1a forms brown- 378

orange, anhedral to subhedral millimeter size grains having a skeletal radial texture with 379

interstitial 200-1000 µm quartz and K-feldspar grains (Fig. 3A-B). Accessory minerals include 380

fluorapatite, rutile, and ilmenite present as micro-inclusions (10-200 µm) in quartz and 381

tourmaline. Tourmaline Tur 1a shows discrete core-to-rim zonation but lacks oscillatory zoning 382

(Fig. 4A-B). Disseminated tourmaline Tur 1b, described previously only in minor leucogranite 383

plugs and dikes along the southwestern and northeastern margins of the San Rafael granite 384

(Kontak and Clark, 2002; Mlynarczyk and Williams-Jones, 2006), is also observed in 385

microgranite. This tourmaline occurs as millimeter size greenish to orange-brownish, euhedral 386

to sub-euhedral grains intergrown with magmatic quartz, K-feldspar, plagioclase, and accessory 387

(14)

fluorapatite (Figs. 2D and 3C), similar to the descriptions of Kontak and Clark (2002) and 388

Mlynarczyk and Williams-Jones (2006). Back-scattered electron images of Tur 1b show a 389

relatively homogeneous internal texture displaying locally a discrete core-to-rim zoning (Fig.

390

4C).

391

Pre-ore hydrothermal tourmaline (Tur 2): occurs as post-magmatic replacements and in 392

veins and breccias in the granites and enclosing shales. Based on new petrographic 393

observations, three sub-generations (Tur 2a, Tur 2b, and Tur 2c) are distinguished. Tur 2a, in 394

part equivalent to the “early post-magmatic tourmaline” of Mlynarczyk and Williams-Jones 395

(2006), locally replaces partly or wholly magmatic minerals (K-feldspar phenocrysts, 396

cordierite, biotite) in the megacrystic monzogranite and microgranite (Fig. 2F). Tur 2a locally 397

forms randomly distributed rosettes that replaced K-feldspar phenocrysts within metasomatized 398

granites, as previously documented by Kontak and Clark (2002). Under the petrographic 399

microscope, Tur 2a is orange-brownish to bluish and typically forms sub-euhedral crystals 400

ranging in size typically from 100 to 500 µm (Fig. 3D-E). In BSE images, tourmaline reveals a 401

complex internal texture with a 50-100 µm-thick core rimmed by thin bands (down to <10 µm) 402

showing oscillatory and sector zoning (Fig. 4D-E). Tur 2b occurs in quartz-bearing veins and 403

clast-supported breccias cemented by quartz. This tourmaline generation represents the most 404

abundant at San Rafael and was previously described as “early hydrothermal tourmaline” by 405

Mlynarczyk and Williams-Jones (2006). Tur 2b veins and breccias are observed at surface up 406

to more than 5000 masl and underground at least down to 3600 masl, thus having a continuous 407

vertical extent of >1.4 km. They cut both the granitic intrusion (Tur 2b(g)) and the shales in the 408

contact metamorphic aureole (Tur 2b(s)). Tourmaline-quartz veins are generally narrow (<3 cm) 409

and systematically surrounded by an alteration halo <1 to 10 cm-wide composed of albite and 410

sericite, with minor apatite and rutile (Fig. 2E). In thin section, tourmaline hosted in the granite 411

(Tur 2b(g)) is orange-brown to green and forms acicular, sub-euhedral crystals ranging from 100 412

µm up to several millimeters in size (Fig. 3F). The tourmaline grains show a typical open-space 413

filling texture with c-axes perpendicular to the host rock contact and growth directions towards 414

the center of the vein. Characteristic is intense oscillatory zoning at the µm-scale as evidenced 415

by BSE images (Fig. 4F) and X-ray elemental mapping (Fig. 5). Prismatic sectors are 416

characterized by higher concentrations of Ca, Ti, and F, and lower Al, whereas oscillatory 417

zoning is marked mainly by variations of the Fe/Mg ratio (Fig. 5). Tur 2b in the veins and 418

breccias is considered broadly synchronous owing to (i) similar color, texture, and zoning 419

patterns; (ii) subsequent re-openings and fillings commonly observed within a single sample;

420

and (iii) absence of systematic crosscutting evidence (Figs. 3G and 4G). Tourmaline from veins 421

(15)

hosted in the metamorphic shales (Tur 2b(s)) shows features similar to those in the granite- 422

hosted tourmaline but is typically finer grained (50-200 µm) and is green-blue under plane- 423

polarized transmitted light (Fig. 3H). Tur 2c in tourmaline-quartz microbreccias is 424

volumetrically the most abundant (Fig. 2G). The microbreccias are commonly associated with 425

reopening of pre-existing quartz-tourmaline veins and contain clasts of early tourmaline veins 426

and breccias. Under the microscope, Tur 2c occurs as fine-grained (<10-50 µm) and randomly- 427

oriented brown grains that form a dense aggregate cementing clasts of tourmalinized wall rocks, 428

tourmaline veins, and relict magmatic quartz (Figs. 3I and 4H). The acicular crystals of older 429

quartz-tourmaline veins are in places broken, partially corroded, and replaced by Tur 2c that 430

makes up the microbreccia (Fig. 4I).

431

Syn-ore hydrothermal tourmaline (Tur 3): occurs as microscopic veinlets and overgrowths 432

that are partly cutting previous tourmaline generations (Fig. 2H). Previously identified by 433

Mlynarczyk and Williams-Jones (2006) as “ore-stage tourmaline,” it is mostly blue (Tur 3a) or 434

in places green (Tur 3b). Under the microscope, Tur 3a is pale to dark blue and forms either 50 435

to 200 µm-thick veinlets cutting orange-brown acicular tourmaline grains (Tur 2) or 10 to 100 436

µm-thick overgrowths in optical continuity with the host crystal (Figs. 3J-K and 4H-K). Due to 437

their relatively small size, these fine-scale overprints of Tur 3 onto Tur 2 can be easily 438

overlooked when using transmitted light imagery only. Overprinting of pre-ore hydrothermal 439

quartz-tourmaline (Tur 2) by fine-grained Tur 3 can be recognized and quantified using 440

combined X-ray elemental and QEMSCAN mapping, as Tur 3 has higher Fe concentrations. In 441

the sample shown in Figure 6, more than 15% of the analyzed area corresponds to Tur 3a. Green 442

tourmaline (Tur 3b) and blue tourmaline (Tur 3a) veinlets locally show inter-crosscutting 443

relationships (Fig. 3K), suggesting penecontemporaneous formation. Green tourmaline (Tur 444

3b) forms khaki-green to dark-green acicular, 50 to 200 µm-long grains commonly intergrown 445

with quartz and chlorite in open-space fillings. Locally, green tourmaline occurs as small 446

needle-like crystals intergrown with cassiterite, chlorite, and quartz from the main ore 447

assemblage (Figs. 3L and 4L). This textural relationship indicates that tourmaline continued to 448

form during Sn ore deposition but in volumetrically minor amounts compared to the pre-ore 449

hydrothermal stage.

450 451

Major and trace element composition of tourmaline 452

453

The average major and minor element composition and structural formula of the three 454

tourmaline generations are reported in Table 2 and the full dataset in Supplementary Table 1.

455

(16)

Tourmaline from San Rafael shows major element variations, principally for Fe (0.05-2.21 456

apfu), Mg (0.13-2.90 apfu), and Al (5.41-7.20 apfu), as well as for minor elements such as Na 457

(0.30-0.86 apfu), Ca (0.01-0.34 apfu), and Ti (0.01-0.50 apfu). The different tourmaline 458

generations (Tur 1 to Tur 3) have compositions belonging to the schorl-dravite and foitite-Mg- 459

foitite solid solutions (Figs. 7 and 8), which is consistent with the EMPA data reported by 460

Kontak and Clark (2002) and Mlynarczyk and Williams-Jones (2006). Tur 1 belongs to the 461

alkali group (Fig. 7) and has dravitic compositions, close to the schorl-dravite limit, with 462

Fe/(Fe+Mg) = 0.36-0.52 and X□/(X□+Na+K) = 0.08-0.43. Tur 2 is also dominantly within the 463

alkali group and has intermediate compositions corresponding to the schorl-dravite solid 464

solution with Fe/(Fe+Mg) = 0.02-0.83 and X□/(X□+Na+K) = 0.02-0.62. Compared to Tur 2b(g), 465

shale-hosted Tur 2b(s) is distinguished by higher X□/(X□+Na+K) = 0.44-0.62 and Fe/(Fe+Mg) 466

= 0.61-0.80. In contrast, Tur 3 belongs to the alkali to X-vacant group and has schorl to foitite 467

compositions with Fe/(Fe+Mg) = 0.48-0.94 and X□/(X□+Na+K) = 0.11-0.70. The strongest 468

compositional variations relate to the Fe/(Fe+Mg) ratio, reflecting the substitution vector 469

Fe2+Mg2+-1 as indicated by a general linear trend in the Fe vs. Mg diagram (Fig. 8). All 470

tourmaline analyses show values of Fe+Mg <3 apfu and Al >6 apfu, suggesting the presence of 471

excess Al in the Y-site. The positive correlation between X-site vacancies and Al contents 472

indicates a combination of the three substitution vectors (☐Al3+)(Na+Mg2+)-1, (Al3+O2- 473

)(Mg2+OH-)-1, and (☐Al3+2)(Ca2+Mg2+2)-1 (Fig. 9), which also explains the variations in Ca and 474

Na.

475

Trace element contents of tourmaline from San Rafael are summarized in Table 3 and the 476

full dataset is reported in Supplementary Table 2. Tourmaline from San Rafael has generally 477

very low to low concentrations (<0.1 to 10 ppm) of Be, Co, Cu, Rb, Y, Zr, Nb, Ta, Cs, Bi, Pb, 478

and most REE; intermediate concentrations (10s to 100 ppm) of Li, P, Sc, V, Cr, Ni, Zn, Sr, 479

and partly Sn; and high to very high concentrations (100s to >1000 ppm) of K, Ca, Ti, and Mn, 480

along with Sc, V, Cr, and Sn in some cases. Variation diagrams of some trace elements as a 481

function of Fe/(Fe+Mg) ratio are shown in Figure 10, and percentile box and whisker plots are 482

shown in Supplementary Figure 1. The different tourmaline generations have overlapping and 483

highly variable concentrations for some trace elements such as Cr (1-700 ppm), Sc (1-1000 484

ppm), Nb (0.05-35 ppm), Ta (0.01-10 ppm), REE (0.05-50 ppm), and Pb (0.1-14 ppm). Other 485

trace elements show more systematic variations across the three main tourmaline generations.

486

Late-magmatic Tur 1 is characterized by high contents of Li (60-200 ppm), K (150-500 ppm), 487

Ti (1200-12000 ppm), Zn (83-360 ppm), and Mn (65-500 ppm). In contrast, syn-ore 488

(17)

hydrothermal Tur 3 is distinguished by higher contents of Be (2-50 ppm), Sr (17-530 ppm), and 489

Sn (105-2240 ppm). Intermediate trace element concentrations between those of Tur 1 and Tur 490

3 characterize pre-ore hydrothermal Tur 2, particularly for Li (3-250 ppm), Be (0.5-65 ppm), K 491

(32-400 ppm), Zn (1-300 ppm), and Sn (3-200 ppm). Compared to Tur 2b(g), shale-hosted Tur 492

2b(s) contains lower Li (3.5-15 ppm) and K (75-112 ppm) contents, and higher V (280-402 493

ppm), Mn (89-121 ppm), and Zn (235-290 ppm) contents (Fig. 10).

494

Positive linear correlations are observed for some trace elements such as Nb vs. Ta (R2 = 495

0.48), K vs. Pb (R2 = 0.52), Mn vs. Zn (R2 = 0.69), V vs. Cr (R2 = 0.58), Sc vs. V (R2 = 0.69), 496

Ti vs. Ca (R2 = 0.42), LREE vs. Ca (R2 = 0.79), and Ca vs. Sr (R2 = 0.52) (Supplementary 497

Figure 2). The observed co-variations likely indicate substitution-controlled mechanisms for 498

their incorporation in the tourmaline structure, principally the X-, Y-, and Z-sites. In multi- 499

element diagrams normalized to upper continental crust (UCC), tourmaline is characterized by 500

relative enrichments in Li, Be, Sn, and Zn between 2 and 1000 times higher than the UCC 501

values, and relative depletions in Sr, K, Ca, Cs, Rb, Y, P, Zr, Bi, Pb, Mn, Co, Ni, and Cu 502

between 2 and 1000 times lower than the UCC values (Fig. 11). The REE content of tourmaline 503

is uniformly low (<0.1 to 50 ppm), between 10 and 1000 times lower than the UCC values, and 504

is characterized by variable LREE/HREE ratios with systematic positive Eu anomalies (Fig.

505

11), which is quantified by the ratio (Eu/Eu*)N = EuN/(SmN x GdN)0.5 where N corresponds to 506

the chondrite-normalization. Tur 1 and Tur 2 are enriched in LREE over HREE, whereas Tur 3 507

has higher contents of HREE relative to LREE. Tourmaline Tur 1 to Tur 3 shows a progressive 508

decrease of the (La/Yb)N (from 1000 to 0.01) and (Eu/Eu*)N (from 20 to 2) ratios, correlating 509

with a progressive increase of the Sn content, from 10s to 100 ppm for Tur 1 and Tur 2, and as 510

high as >1000 ppm for Tur 3 (Fig. 12).

511 512

Multivariate statistical analysis 513

514

Two types of multivariate statistical techniques, principal component analysis (PCA) and 515

discriminant projection analysis (DPA), were applied to the major and trace element dataset of 516

tourmaline in order to understand and quantitatively classify the observed compositional 517

variations.

518

Principal component analysis is a classical multivariate statistical technique particularly 519

useful for treating large geochemical datasets, such as LA-ICP-MS trace element data (e.g., 520

Winderbaum et al., 2012; Belissont et al., 2014; Harlaux et al., 2018, 2019a). The aim of PCA 521

is to provide dimensionality reduction of correlated variables into a reduced set of orthogonal 522

(18)

linear combinations, so-called principal components, which maximize the variance and 523

minimize information loss (Izenman, 2008). A classical PCA has been applied to the major and 524

trace element dataset of San Rafael tourmaline (n = 258 spot analyses, separating the three main 525

generations Tur 1 to Tur 3). A total of 18 variables has been selected for the PCA including 526

four major elements (Al, Fe, Mg, Na) quantified by EMPA and 14 trace elements (Li, K, Ti, V, 527

Cr, Sc, Mn, Zn, Sr, Sn, Nb, Ca, LREE, HREE) determined by LA-ICP-MS. Other elements 528

present at very low concentrations (<1 ppm) or below the limits of detection were not included 529

in the analysis. Results of the PCA applied to the log-transformed dataset of major and trace 530

elements in tourmaline are shown in Figure 13A. Data are represented by a two-dimensional 531

projection of the two first principal components (PC1 vs. PC2), which describe the statistical 532

correlations between the investigated variables (n = 18). The elements projected on the PC1 vs.

533

PC2 plane account for 51.2% of element content variability. Four main groups of element 534

correlation clusters are discriminated by the PCA. A first group consisting of K, Nb, Mn, Zn, 535

and Fe, characterizes Tur 1. A second group composed of Li, Ca, Ti, LREE, Sr, and Na, and a 536

third group including Mg, Cr, Sc, and V, explain the variability of Tur 2. A fourth group 537

comprises Sn and HREE and represents mainly Tur 3. These statistical correlation clusters 538

reflect partly the crystal chemical control on the incorporation of major and trace elements in 539

tourmaline. Indeed, elements such as Fe, Mg, Zn, Mn, Cr, Sc, V, Li, and REE are incorporated 540

into the Y-site, whereas Na, Ca, and Sr enter the X-site (Henry et al., 2011; van Hinsberg, 541

2011).

542

Discriminant projection analysis is another multivariate statistical technique that aims to 543

classify a high-dimensional dataset into predefined groups known a priori by calculating a 544

linear discriminant function that maximizes ratios between the groups (Izenman, 2008). This 545

method has been used for geological or forensic investigations of trace element or isotopic 546

analyses of minerals (e.g., Dalpé et al., 2010). We have applied DPA to the same element 547

dataset for San Rafael tourmaline (n = 258 spot analyses, 18 variables) as done for PCA, and 548

defined three groups corresponding to the three generations of tourmaline (Tur 1 to Tur 3).

549

Results of the DPA applied to the log-transformed dataset of major and trace elements in 550

tourmaline are shown in Figure 13B. Data are represented by the two discriminant projections 551

(DP1 vs. DP2), which maximize the separation between the predefined groups. Tur 1 is mainly 552

projected on the DP2 axis and is defined by correlations of K, Mn, Zn, Ti, Ca, Li, Mg, LREE, 553

and V. Tur 2 lies at the center of the DP1 vs. DP2 projection, mainly reflecting correlations of 554

Mg, Al, Cr, Li, and LREE. Tur 3 is defined by the DP2 vector, reflecting correlations of Fe, Sn, 555

HREE, and Sr.

556

(19)

557

Sr, Nd, and Pb isotopic compositions of tourmaline 558

559

The Sr, Nd, and Pb radiogenic isotopic compositions of tourmaline are shown in Figure 14 560

and data are reported in Table 4. Results show variable initial ratios of 87Sr/86Sr(i) (0.7160- 561

0.7276) and 143Nd/144Nd(i) (0.5119-0.5124) that plot within the fields of Coastal and Western 562

Cordilleran intrusions and Arequipa-Antafolla metamorphic basement (see references in Fig.

563

14). Most samples of tourmaline have 87Sr/86Sr(i) and 143Nd/144Nd(i) values that overlap those of 564

the San Rafael granites (87Sr/86Sr(i) = 0.7131-0.7202 and 143Nd/144Nd(i) = 0.5121-0.5122;

565

Supplementary Table 3), thus supporting a dominantly magmatic origin for the Sr and Nd, and 566

by extension, for the tourmaline-precipitating hydrothermal fluids. Only one sample of 567

hydrothermal tourmaline from a vein hosted in shale has a relatively high Sr radiogenic 568

composition (sample SRG-21A, 87Sr/86Sr(i) = 0.7276), likely related to fluid-rock interaction 569

with the host rock. The initial Nd isotopic values of tourmaline fall in the compositional range 570

of Bolivian tin porphyries of Llallagua, Chorolque, and Cerro Rico de Potosi (143Nd/144Nd(i) = 571

0.5121-0.5124; Dietrich et al., 2000). The initial Pb isotopic compositions of tourmaline are 572

more variable, with 206Pb/204Pb(i), 207Pb/204Pb(i), and 208Pb/204Pb(i) ratios mostly falling in the 573

range of 18.6-19.3, 15.6-16.0, and 38.6-39.7, respectively. Three samples (PSR-24A and PSR- 574

24A-bis with 206Pb/204Pb(i) = 21.2-21.5, and PSR-12B with 206Pb/204Pb(i) = 22.8) have markedly 575

higher 206Pb/204Pb(i) values and are not shown in Figure 14. These highly radiogenic Pb 576

compositions likely reflect the precipitating fluid that had a high initial 238U/204Pb ratio. Most 577

Pb isotopic compositions of tourmaline plot close to the upper crust and orogen curves of 578

Zartman and Doe (1981) and fall within the compositional ranges of the Arequipa-Antafolla 579

metamorphic basement, the central volcanic zone, and the Coastal and Western Cordilleran 580

intrusions (see references in Fig. 14). These compositions partly overlap the initial Pb isotope 581

values of the San Rafael granites (206Pb/204Pb(i) = 18.6-18.8, 207Pb/204Pb(i) = 15.6-15.7, and 582

208Pb/204Pb(i) = 38.9-39.0; Supplementary Table 3) and are also similar to those of other 583

Oligocene to Miocene Sn-W ± Cu-Zn-Pb-Ag deposits in southeast Peru (206Pb/204Pb(i) = 18.5- 584

25.2, 207Pb/204Pb(i) = 15.6-16.0, and 208Pb/204Pb(i) = 38.6-40.1; Kontak et al., 1990).

585 586

Discussion 587

588

Tourmaline textures as indicator of formation conditions 589

590

(20)

The textural features of tourmaline may record particularities of the environment of 591

formation and provide valuable information on physicochemical conditions during its 592

crystallization (e.g., van Hinsberg et al., 2001; Dutrow and Henry, 2016, 2018). The textural 593

variations of tourmaline generations (from Tur 1 to Tur 3) reflect evolving physical and 594

chemical parameters of the San Rafael system during the magmatic-hydrothermal transition.

595

Texturally homogeneous nodules and disseminations of Tur 1 intergrown with the quartz- 596

feldspar granitic groundmass (Figs. 3 and 4) are typical of magmatic tourmaline in evolved 597

peraluminous granites (e.g., London and Manning, 1995; London et al., 1996; Balen and 598

Broska, 2011; Balen and Petrinec, 2011; Drivenes et al., 2015). The formation of tourmaline 599

nodules has been widely discussed in the literature and three main hypotheses have been 600

proposed for their origin: (i) post-magmatic hydrothermal alteration of granitic bodies by 601

externally derived boron-rich fluids (e.g., Rozendaal and Bruwer, 1995); (ii) crystallization 602

from immiscible, hydrous, boron-aluminosilicate melts or boron-rich aqueous fluids that 603

separated from coexisting silicate melt (e.g., Dini et al., 2007; Balen and Broska, 2011;

604

Drivenes et al., 2015; Burianek et al., 2016); and (iii) products of magmatic crystallization on 605

the liquid line of descent of boron-rich granitic melts (e.g., Perugini and Poli, 2007; Balen and 606

Petrinec, 2011; Valentini et al., 2015). The Tur 1 nodules observed in the San Rafael granites 607

are devoid of alteration features (i.e., absence of veins, dissolution-reprecipitation texture, and 608

pervasive alteration), thus arguing against a hydrothermal origin related to post-magmatic 609

alteration. Instead, the rounded shapes of the nodules, presence of leucocratic rims, and limited 610

intergrowths with the granitic groundmass better suggest their physical separation from the 611

silicate melt during crystallization. Experimental studies indicate that tourmaline saturation in 612

strongly peraluminous granitic melts (ASI >1.2) can be reached after extended crystal 613

fractionation when the boron content of the melt attained >2 wt.% B2O3 at 750°C and 2 kbar 614

(Wolf and London, 1997). Other works based on petrographic relations, geochemical data, and 615

compositional phase diagrams suggested that minimum boron content in the range of 0.05-0.3 616

wt.% B2O3 is sufficient to saturate a peraluminous granitic melt in tourmaline between 600 and 617

750ºC (Pesquera et al., 2013). The San Rafael megacrystic granite has a moderately fractionated 618

peraluminous S-type composition (ASI = 1.1-1.5), as indicated by whole-rock and mineral 619

chemistry data (Kontak and Clark, 2002; Mlynarczyk, 2005; Corthay, 2014; Prado, 2015). Its 620

whole-rock boron content (B = 60-160 ppm, avg = 115 ppm; Mlynarczyk, 2005) falls in the 621

same range of peraluminous granites hosting tourmaline nodules elsewhere (B = 10-500 ppm;

622

Dini et al., 2007; Balen and Broska, 2011; Pesquera et al., 2013; Drivenes et al., 2015; Burianek 623

et al., 2016) as well as quartz-hosted melt inclusions from Bolivian tin porphyries (B = 35-640 624

(21)

ppm; Dietrich et al., 2000; Lehmann et al., 2000; Wittenbrink et al., 2009). This suggests that 625

the boron content of the San Rafael granitic melt may have been initially much higher owing 626

to the preferential partitioning of boron into a fluid phase relative to the silicate melt during 627

water saturation (e.g., Pichavant, 1981; London et al., 1988; Hervig et al., 2002; Thomas et al., 628

2003; Schatz et al., 2004). In addition, the presence of 0.5- to 1-cm-thick quartz-K-feldspar 629

leucocratic rims surrounding the Tur 1 nodules suggest that tourmaline crystallization was 630

favored by decomposition reactions of biotite and cordierite, possibly at temperatures <750ºC 631

close to solidus (Wolf and London, 1997). Based on the arguments above, we propose that the 632

Tur 1 nodules crystallized from an immiscible, hydrous, boron-aluminosilicate melt that 633

separated from the San Rafael granitic melt during the late-magmatic stage prior or 634

concomitantly to the magmatic-hydrothermal transition.

635

Pre-ore hydrothermal Tur 2 is characterized by intense oscillatory zoning at a micrometer 636

scale (Figs. 3 to 4), which is characterized by fluctuating concentrations of major, minor and 637

trace elements and progressively increasing Fe/(Fe+Mg) ratios along the growth direction 638

(Supplementary Figure 3). Oscillatory zoning in hydrothermal minerals is generally interpreted 639

to reflect (i) periodic changes in the external environment due to varying fluid properties such 640

as chemical composition, pressure, temperature, or fO2 (Holten et al., 1997), (ii) intrinsic self- 641

organization processes during crystal growth controlled by absorption-diffusion reactions at the 642

crystal-fluid interface (L’Heureux and Jamtveit, 2002), or (iii) a combination of both external 643

and internal fluctuating factors (Shore and Fowler, 1996). Quartz- and tourmaline-hosted 644

primary fluid inclusions from hydrothermal veins cutting the San Rafael megacrystic granite 645

yielded trapping temperatures of >500ºC and high salinity (34-62 wt.% NaCl equiv) in 646

lithostatic pressure conditions of 0.8 kbar (Kontak and Clark, 2002; Mlynarczyk et al., 2003;

647

Wagner et al., 2009; Corthay, 2014; Prado, 2015). Under such pressure-temperature conditions, 648

the oscillatory zoning observed in Tur 2 crystals is attributed to rapid changes in the 649

hydrothermal system, possibly caused by evolving physicochemical fluid conditions and/or by 650

rapid crystal growth. This interpretation is supported by (i) the open-space filling textures 651

typical of the Tur 2b crystals that grew perpendicular to the vein selvages, (ii) the increasing 652

Fe/(Fe+Mg) ratios along growth directions, and (iii) evidence for multiple vein-opening events;

653

all together indicating a dynamic fluid environment. Experimental synthesis of tourmaline 654

single crystals from boron-rich hydrothermal solutions (400-750°C, 1 kbar) reported a growth 655

rate of 0.05 mm/day (Setkova et al., 2011), which possibly reproduces the fluid dynamics in 656

natural magmatic-hydrothermal systems. Typical textures of repeated hydrofracturing and 657

fluid-assisted reopening (e.g., Jébrak, 1997) observed in quartz-tourmaline veins and breccias 658

Références

Documents relatifs

Whole rock compositions of rocks from the SRIC are shown in Figs. Photographs of representative rock samples from the SRIC. A) Biotite-cordierite megacrystic

No remnants of calc-silicate alteration typical of the volcaniclastics or the Susana Breccia matrix have been observed in fragments of the Ore Breccia pipe, and early

Carbon and oxygen isotope records of fluid mixing and fluid-rock interaction in carbonates from the zinc-lead deposit of san vicente (central

We present a combined study of oxygen and boron isotopes (δ18O and δ11B) measured in situ in tourmaline from the giant San Rafael tin deposit (Central Andean Tin Belt, Southeast

At the giant tin deposit of San Rafael, Peru, the early hydrothermal stages, preceding the economic tin mineralization, include successively: 1) potassic alteration represented

The recently discovered world class San Gregorio Zn-Pb deposit in Central Peru, has formed by high level carbonate replacement through very acid fluids coming from a high

Ore-related structures at the Zn-Pb Missi ssippi Valley-type San Vicente deposit, Central Peru, display a system atic geom etry with respect to bedding and a regional

The country rock consists in a Lower Cretaceous shallow water volcano-sedimentary sequence, crosscut by Middle to Upper Cretaceous intrusions of the Peruvian Coastal