Extension and gold mineralisation in the hanging walls of active convergent continental shear zones

Extension and gold mineralisation in the hanging walls of active convergent continental shear zones

Accepted Manuscript Extension and gold mineralisation in the hanging walls of active convergent continental shear zones Phaedra Upton, Dave Craw PII: ...

3MB Sizes 0 Downloads 27 Views

Accepted Manuscript Extension and gold mineralisation in the hanging walls of active convergent continental shear zones Phaedra Upton, Dave Craw PII:

S0191-8141(13)00146-6

DOI:

10.1016/j.jsg.2013.08.004

Reference:

SG 2948

To appear in:

Journal of Structural Geology

Received Date: 14 February 2013 Revised Date:

7 August 2013

Accepted Date: 8 August 2013

Please cite this article as: Upton, P., Craw, D., Extension and gold mineralisation in the hanging walls of active convergent continental shear zones, Journal of Structural Geology (2013), doi: 10.1016/ j.jsg.2013.08.004. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

1

Extension and gold mineralisation in the hanging walls of active convergent

2

continental shear zones

3 Phaedra Uptona,* and Dave Crawb

RI PT

4 5 6

a

7

5704198

8

b

9

[email protected]

SC

GNS Science, PO Box 30368, Lower Hutt, New Zealand, [email protected], +64 4

M AN U

Geology Department, University of Otago, PO Box 56, Dunedin, New Zealand,

10

Abstract: Orogenic gold-bearing quartz veins form in mountain belts adjacent to

12

convergent tectonic boundaries. The vein systems, hosted in extensional structures

13

within compressively deformed rocks, are a widespread feature of these orogens. In

14

many cases the extensional structures that host gold-bearing veins have been

15

superimposed on, and locally controlled by, compressional structures formed within

16

the convergent orogen. Exploring these observations within the context of a three-

17

dimensional mechanical model allows prediction of mechanisms and locations of

18

extensional zones within convergent orogens.

19

convergence angle and mid-crustal strength on stress states and compare them to the

20

Southern Alps and Taiwan. The dilatation zones coincide with the highest mountains,

21

in the hanging walls of major plate boundary faults, and can extend as deep as the

22

brittle-ductile transition. Extensional deformation is favoured in the topographic

23

divide region of oblique orogens with mid-lower crustal rheology that promotes

24

localisation rather than diffuse deformation.

25

influences the stress state to a depth approximately equal to the topographic relief,

Our models explore the effect of

AC C

EP

TE D

11

1

In the near surface, topography

ACCEPTED MANUSCRIPT

26

bringing the rock closer to failure and rotating

to near vertical. The distribution of

27

gold-bearing extensional veins may indicate the general position of the topographic

28

divide within exhumed ancient orogens.

30

RI PT

29 Keywords: gold; structure; extension; tectonics; stress field; convergent orogen

31 1. Introduction

SC

32

Orogenic gold deposits have formed in convergent orogens throughout geological

34

time, and are still forming in modern orogens (Bierlein and Crowe, 2000; Craw et al.,

35

2002; Goldfarb et al., 2005; Groves et al., 1998; Groves et al., 2003). These gold

36

deposits are hosted in a wide range of lithologies in primarily greenschist facies

37

metamorphic rocks that have undergone compressional deformation (Bierlein and

38

Crowe, 2000; Goldfarb et al., 2005; Groves et al., 1998; Groves et al., 2003). Most of

39

these orogenic gold deposits are dominated by quartz veins that have filled dilational

40

structural sites in compressively-deformed host rocks (Allibone et al., 2002; Cox et

41

al., 1995; Kontak and Horne, 2010; MacKenzie et al., 2008; Sibson et al., 1988; Witt

42

and Vanderhor, 1998). Many of these extensional structures formed locally during

43

compressional deformation (Cox et al., 1995; Goldfarb et al., 2005; Witt and

44

Vanderhor, 1998). In addition, structures that form in a regional extensional stress

45

regime, within regionally compressively-deformed rocks, are widespread features in

46

orogenic gold deposits (Bierlein and Crowe, 2000; Goldfarb et al., 2005; Vielreicher

47

et al., 2010; Witt and Vanderhor, 1998). These structures include normal faults and

48

steeply dipping extensional vein arrays, with dilational openings on the centimetre to

49

metre scale, and the structures commonly post-date, and overprint, compressional

50

structures.

AC C

EP

TE D

M AN U

33

2

ACCEPTED MANUSCRIPT

This common occurrence of regional-scale superimposition of structures

52

formed in an extensional stress regime on compressionally-deformed rocks in

53

orogenic gold systems does not require changes in tectonic plate vectors, although

54

plate tectonic vectors have apparently changed in some ancient settings (Allibone et

55

al., 2002; Goldfarb et al., 1991). Instead, on-going transfer of rocks from a

56

compressional regime to an extensional regime within the same orogen (Craw et al.,

57

2002; Koons, 1995; Koons et al., 1998; Phillips and Powell, 2009; Upton et al.,

58

2009a) can achieve the same end result. This process is different from the more

59

localised changes in stress orientation at the local scale (metres to hundreds of metres)

60

including variations in the interactions between different lithologies and structures

61

(Cox et al., 1995; Upton et al., 2008; Witt and Vanderhor, 1998), or from variations in

62

fluid pressures (Ridley, 1993; Sibson et al., 1988).

M AN U

SC

RI PT

51

Transfer of rocks from a compressional stress regime to an extensional stress

64

regime without a change in far-field plate vector direction can be difficult to infer

65

from ancient orogens, as evidence for the plate boundary geometry and motions are

66

typically poorly recorded. However, direct observation of active orogens can provide

67

information on the structural evolution of mineralised rocks in the context of directly-

68

observable plate vectors and stress regimes. In this paper, we present observations

69

from two active convergent orogens, Taiwan and New Zealand (Fig. 1), where gold

70

deposits are forming in extensional sites in the hanging walls of major compressional

71

plate boundary structures. We present these observations in the context of a numerical

72

model that reflects the same types of deformation that are occurring in our examples,

73

to show where and how extensional zones form to facilitate fluid flow and gold

74

mineralisation, and the geometry of rock motion that results in transfer of rocks from

75

one stress regime to another.

AC C

EP

TE D

63

3

ACCEPTED MANUSCRIPT

76

2. General settings Both New Zealand and Taiwan (Fig. 1) are oblique convergent tectonic zones on the

78

margin of the Pacific Ocean (Wu et al., 2007). The Southern Alps of southern New

79

Zealand are forming in the hanging wall of the Alpine Fault, a major transcurrent

80

structure that forms the plate boundary through pre-existing Paleozoic-Mesozoic

81

continental crust (Fig. 1A,B,C) (Norris et al., 1990; Walcott, 1986). Taiwan is a

82

young orogen constructed from Cenozoic marine sediments that lay on the Chinese

83

continental margin before the current collision with the Luzon volcanic arc, with a

84

component of pre-existing metamorphic crust at depth (Byrne et al., 2013; Fisher et

85

al., 2002; Ho, 1986; Suppe, 1981). The main mountain ranges of Taiwan are in the

86

hanging wall of the Longitudinal Valley Fault that forms the on-land plate boundary

87

(Fig. 1D,E,F).

M AN U

SC

RI PT

77

Rapid uplift of mountains in both New Zealand and Taiwan has resulted in

89

regional-scale conductive thermal anomalies (Koons, 1989) that are driving meteoric

90

hot-spring systems in regions of elevated thermal gradients (Allis, 1981; Upton et al.,

91

2011). Deep-sourced fluids, derived from progressive dehydration of metamorphic

92

rocks beneath the orogens, rise beneath the mountains to contribute to the tectonic-

93

hydrothermal systems, and mix with circulating meteoric waters (Bertrand et al.,

94

2009; Upton et al., 2003; Wannamaker et al., 2002). The tectonic-hydrothermal

95

system in Taiwan is distributed right across the island (Chen, 1985; Lee and Cheng,

96

1986), whereas that in the Southern Alps occurs primarily beneath, and to the west of,

97

the topographic divide (Allis, 1981; Barnes et al., 1978; Upton et al., 2011).

98

AC C

EP

TE D

88

2.1.

Extensional veins and gold deposits

99

Gold deposits associated with active transpressional tectonism are small in both New

100

Zealand and Taiwan, and most deposits have no known economic significance. Minor

4

ACCEPTED MANUSCRIPT

historic gold mining has occurred in one of these deposits in Taiwan (Tan et al., 1991)

102

and in several in New Zealand (Becker et al., 2000; Craw et al., 2010; Craw et al.,

103

2006b). Placer gold derived by erosion of these small but numerous vein deposits

104

forms the most significant resource, and beach placers in particular have been mined

105

extensively in New Zealand and locally in Taiwan (Craw et al., 1999; Tan et al.,

106

1991; Williams, 1974). The large volumes of exhumed and eroded rock containing

107

gold-bearing veins in the hanging wall of the plate boundary fault zones continue to

108

contribute to those placer deposits (Craw et al., 1999).

SC

RI PT

101

The hydrothermal vein systems that host gold deposits in the Southern Alps of

110

New Zealand fill extensional fractures and faults that cut across variably folded

111

metamorphic fabrics (Fig. 2A,B,C). The metamorphic fabrics are mostly Mesozoic in

112

origin, with some Cenozoic overprint, and these metamorphic fabrics have been

113

folded during Cenozoic deformation related to the rise of the Southern Alps (Craw et

114

al., 2002). Folds are typically upright and shallow-plunging, with rounded hinges.

115

Folds at the outcrop scale occur in swarms that are part of larger scale (1-3 km) fold

116

zones that trend parallel or subparallel to the mountain chain axis. Extensional

117

fractures, many of which host veins, occur in swarms with a range of orientations but

118

fractures parallel to Cenozoic fold axial surfaces are particularly common (Fig. 2A).

119

Most gold-bearing veins in extensional fractures strike north to northeast (Craw,

120

2006). Normal faults that cut and deform the metamorphic fabric also host gold

121

deposits, in well-defined veins (cm to m scale; Fig. 2C). Veins form at a range of

122

depths, from near-surface to >6 km (Craw et al., 2002; Craw et al., 2009). Most of the

123

Southern Alps gold deposits are characterised by distinctive ankeritic alteration of

124

metamorphic chlorite in host rocks, and coarse grained ankerite in veins, with stable

AC C

EP

TE D

M AN U

109

5

ACCEPTED MANUSCRIPT

125

isotopic signatures implying variable mixtures of metamorphic and meteoric fluids

126

(Fig. 2B,C; Craw et al., 2009). Gold-bearing veins in Taiwan occur immediately east of the topographic

128

divide in Miocene sediments that have been metamorphosed and recrystallised during

129

the collisional deformation that characterises the modern orogen (Craw et al., 2010;

130

Upton et al., 2011). These auriferous veins were emplaced between 4 and 10 km

131

depth, with the deeper veins forming under late metamorphic (greenschist facies)

132

conditions (Craw et al., 2010). Extensional veins in and near gold-bearing systems in

133

Taiwan are dominated by ankeritic quartz veins that cut greenschist facies

134

metamorphic foliation and post-metamorphic folds of that foliation (Fig. 2D,E). Gold-

135

bearing quartz-ankerite veins fill fold axial surface fractures of upright NE and ESE

136

trending post-metamorphic kink folds, and shallow-dipping fractures that crosscut the

137

metamorphic foliation. Historic mining focussed on these extensional features,

138

especially where they intersected each other (Craw et al., 2010). Earlier generations of

139

veins, emplaced during compressional deformation in the latter stages of

140

metamorphism, are subparallel to foliation, and some of these veins contain gold as

141

well (Craw et al., 2010).

EP

TE D

M AN U

SC

RI PT

127

2.2.

143

2.2.1. Tectonic extension zones

144

Extension in the collisional orogens

AC C

142

An examination of the 3D stress tensor and strain rates in Taiwan has previously

145

been outlined in the context of the permeability structure at depth (Upton et al., 2011).

146

In summary, Mouthereau et al. (2009) use earthquakes combined with geodetic

147

measurements to invert for the 3D strain rate field and find that surface strain in the

148

topographic divide region is extensional. They determine a maximum extension

6

ACCEPTED MANUSCRIPT

149

direction in central Taiwan ranging from NNE-SSW to NE-SW. Extension beneath

150

the Central Range persists to depths of more than 15 km. Evidence for tectonic extension in the central Southern Alps is largely geological

152

as a high resolution inversion of earthquake events similar to that for Taiwan does not

153

exist at present. A small number of earthquakes with a normal fault solution are

154

recorded in the central Southern Alps (Anderson et al., 1993; Leitner et al., 2001).

155

Models of elastic deformation in the central Southern Alps based on geodetic

156

measurements imply a NW-SE extension direction in the topographic divide region

157

north of Lakes Pukaki and Tekapo (Beavan et al., 1999). The principal geological

158

evidence for extension in the Southern Alps is the widespread occurrence of small-

159

displacement normal faults and extensional quartz and carbonate filled extensional

160

veins, in addition to the extensional gold-bearing veins, that cut the basement rocks

161

throughout the mountains (Fig. 2A-C) (Cox et al., 1997; Craw et al., 2002; Craw et

162

al., 2009).

SC

M AN U

TE D

163

RI PT

151

2.2.2. Topographic extension zones

In addition to the tectonically driven extensional processes described in the previous

165

section, there is commonly an overprint of shallow extensional fractures associated

166

with the steep topography of active mountains. The steep slopes facilitate mass

167

movement of surficial zones, in variably-coherent landslides that are typically tens or

168

even hundreds of metres thick (Beck, 1968). The headward parts of these landslides

169

are commonly marked by abundant ridge rents (sackungen), or normal faults, with

170

scarps 1-10 m high. The landslides can extend for kilometres downslope to adjacent

171

axial valleys, with total vertical extent of ~2000 m. These landslides dominate most

172

mountain hillsides in Taiwan, up to 3000 m above sea level (Upton et al., 2011).

173

Landslides are also common on New Zealand mountains, especially those that were

AC C

EP

164

7

ACCEPTED MANUSCRIPT

174

deglaciated in the early Holocene. The surficial landslides reflect overall extensional

175

strain and topographic collapse of the steep mountain terrain (Koons and Kirby,

176

2007). 3. Numerical Methods

RI PT

177

We use three-dimensional numerical models to explore the stress state and

179

deformation within generic oblique orogens. Our approach is to use simple models to

180

test the influence of various parameters on the dynamics of the model. We then

181

compare our model results to structural data, including vein orientations, stress

182

inversions and gold deposits, from the Southern Alps and Taiwan. Our models are a

183

generic version of a previously published geometry and rheological description of the

184

Southern Alps (Koons et al., 1998; Koons et al., 2012; Upton and Koons, 2007; Upton

185

et al., 2009a) (Fig. 3A). Although the model geometry is based on the Southern Alps,

186

the models are generic enough, representing the collision of a rigid block with a

187

readily deformed crustal wedge, that we can also apply it them to Taiwan by changing

188

the convergence direction (Suppe, 1980, 1981).

TE D

M AN U

SC

178

Models were developed using the numerical code FLAC3D, a three-dimensional

190

finite difference code (Version 4.0, Itasca, 2009) which we have modified to

191

accommodate local erosion.

192

geological problems ranging from plate-scale tectonics to consideration of material

193

dynamics at the thin-section scale (Johnson et al., 2004; Koons et al., 1998; Oliver et

194

al., 2006; Upton et al., 2009b). Our large scale model domain extends 500 km normal

195

to the plate boundary (x) by 1000 km parallel to the plate boundary (y) by 50 km

196

depth (z). We focus our analysis on a region in the centre of the model, at a distance

197

from the boundaries, to ensure that the magnitudes and orientations we observe are

198

not influenced by these boundary conditions. We use a two-layered crust overlying

FLAC3D has been used to model a wide range of

AC C

EP

189

8

ACCEPTED MANUSCRIPT

an elastic mantle (Fig. 3). The western edge of the model consists of an elastic block

200

which deforms minimally and represents a strong indentor (corresponding to the

201

Australian plate in the case of the Southern Alps or the Luzon Arc in the case of

202

Taiwan). The erosional regime of an orographic mountain range, similar to the

203

Southern Alps, is simulated by maintaining the western slope at a constant elevation

204

(Koons et al., 1998; Koons et al., 2012; Upton and Koons, 2007; Upton et al., 2009a).

205

The models are run to represent snapshots in time.

SC

RI PT

199

We vary the obliquity of collision and the mid-crustal strength to explore how

207

these parameters influence the degree of extension predicted to occur during

208

convergence. All model runs are listed in Table 2. Three model velocity conditions

209

are used: (1) orthogonal collision, (2) oblique collision where vx = vy and (3) highly

210

oblique collision where vy = 4vx. As the Mohr-Coulomb rheology is rate independent,

211

the absolute magnitudes of the velocity boundary conditions do not influence the

212

results. The boundary conditions are imposed on the edges of the model and the base

213

(Fig. 3A). The velocity of the mantle simulates a subduction style boundary with

214

crustal material being dragged down by the crust-mantle coupling at the base of the

215

orogeny (Ellis et al., 2006; Upton and Koons, 2007). We use material properties

216

identical to previous modelling efforts where we compared the different inherited

217

rheological properties of the Canterbury and Otago regions of the Southern Alps

218

regions (Table 2) (Koons et al., 2012; Upton and Koons, 2007; Upton et al., 2009a).

219

Here we generalise those models to a strong or weak thermally activated mid-lower

220

crust beneath a frictional upper crust (Fig. 3). The strength profiles are static initial

221

conditions and are based on thermal equilibration of Mesozoic thinning (Canterbury)

222

and thickening (Otago) of quartzofeldspathic crust (Upton et al., 2009a).

AC C

EP

TE D

M AN U

206

9

ACCEPTED MANUSCRIPT

In addition to generic models of oblique orogens, we build a topographic model of

224

the central Southern Alps by placing a DEM smoothed to 500 m onto a numerical

225

elasto/Mohr-Coulomb model crust. The material is modelled using a friction angle of

226

30° and a cohesion of 2e7 Pa (Clark, 1966). Using this model, we are able to explore

227

the influence of fine-scale (ridges and valleys) topography on the stress state.

228

4. Model results

229

4.1.

RI PT

223

SC

Orthogonal convergence (Model 1)

The strong two-dimensional case, where

231

conjugate shears uplifting a block of material between them (Koons, 1990). The

232

horizontal component of velocity decreases across the shear zones which are regions

233

of contractional deformation as shown by negative

234

crust, between the conjugate shear zones, a small zone of extensional (positive)

235

237

(Fig. 4A). In the upper

A weaker mid-crust produces more diffuse deformation and

negligible extension in the topographic divide region.

TE D

236

occurs.

, produces a two-sided wedge with

M AN U

230

The stress state through most of the model domain is very simple with

parallel

238

to the x-axis, with

239

small region of extension does exist close to the surface the stress tensor will have

240

rotated (Fig. 4A). The relative magnitudes of

241

sectional area of the model domain, producing a relatively low differential stress for

242

the strong model (Fig. 6A).

EP

AC C

243

parallel to the y-axis, with

4.2.

near vertical (Fig. 5A). Where the

and

are similar through the cross-

Oblique (1:1) convergence (Model 2)

244

In this model convergence is oblique such that

245

wedge is similar to the 2D model except that a higher velocity zone has developed

246

beneath the topographic divide at depths of ~5-20 km (Fig. 4B), leaving a rock packet

247

at the topographic divide that is moving to the left more slowly than the packets either

10

. The resulting two-sided

ACCEPTED MANUSCRIPT

248

side of it. A weaker mid-crust again produces more diffuse deformation, no localised

249

high velocity zone, and negligible topographic divide extension.

250

The stress tensor has rotated relative to the 2D model and varies through the model domain.

252

horizontal and

253

deforming zone),

254

However, beneath the developing topography, where the obliquity has increase the

255

horizontal shear stress

,

256

relative magnitude of

is similar to the 2D case but

257

shear stress producing higher differential stress in the mid-crust (Fig. 6B). 4.3.

is sub-

is subvertical. In the outboard region (distal from the boundary and

is near vertical (crosses on Fig 5B).

becomes steeper as

SC

is near horizontal and

rotates away from vertical. The increases with increasing

M AN U

258

is still horizontal but has rotated ~25° north of east.

RI PT

251

Highly oblique (4:1) convergence (Model 3)

259

In this model

260

beneath the topographic divide region is more pronounced in the strong model and

261

there is more significant extension developed beneath the topographic divide (Fig.

262

4C). Extension in the weaker case is not as pronounced but it is more significant than

263

it was in the corresponding weak orthogonal and oblique models (Fig. 4D,E,F).

264

The stress tensor varies significantly though the model domain. In the outboard

265

region,

267

TE D

EP

has rotated ~40° north of east from the 2D model,

is near horizontal and

is near vertical. Beneath the topographic divide, the stress tensor rotates such that

AC C

266

, producing a highly oblique system. The higher velocity zone

is now sub-vertical and

is sub-horizontal (Fig. 5C). Differential stress in this

268

model is much higher than in the previous models, a combination of

269

magnitude and

270

4.4.

reducing in

increasing (Fig. 6C).

Model of finer-scale topographic variations (Model 4)

271

Koons and Kirby (2007) show that the surface state (topography) can influence crustal

272

deformation, especially in regions of high relief. Topographic stresses, arising from

11

ACCEPTED MANUSCRIPT

273

changes in vertical loading (σzz) and changes in both vertical (τxz and τyz) and

274

horizontal (τxy) shear stresses, are greatest near the free surface for the Southern Alps

275

example (Koons and Kirby, 2007). We calculate the contribution of topographic stress to the present stress state.

277

The topographic stress state is represented by a ratio of strength to stress which is an

278

indication of the proximity of the rock mass to failure. The ratio varies from >10, far

279

from failure, to 1 where the stresses match the strength and the rock mass is at failure.

280

If failure occurs, this reduces the topographic stresses and the strength/stress ratio

281

increases toward 10. Topography influences the stress state in the model to a depth

282

extent approximately equal to the relief (Fig. 7A).

283

through the topographic divide of the Southern Alps, material down to ~3 km b.s.l.

284

has been brought closer to failure by the topographic stresses. The stress tensor is

285

also rotated by the topography. In the near surface

286

horizontal and their orientation varies significantly (Fig 7A insert).

M AN U

SC

RI PT

276

5. Discussion

288

5.1.

is vertical,

and

are

TE D

287

On the cross-section shown

Development of extensional zones

That extensional deformation is occurring in both Taiwan and the central Southern

290

Alps is evident from geodetic measurements (Beavan et al., 1999; Hsu et al., 2009;

291

Lin et al., 2010; Mouthereau et al., 2009; Wu et al., 2010), a small number of

292

earthquakes with normal focal mechanisms (Anderson et al., 1993; Crespi et al., 1996;

293

Leitner et al., 2001; Upton et al., 2011) and the presence of extensional fractures and

294

dilational veins (Craw et al., 2010). Our models suggest that extension beneath the

295

topographic divide, to depths below the brittle-ductile transition in some cases, is a

296

predictable result of oblique convergence, especially when the mid-crust is strong

297

enough to focus deformation into discrete shear zones rather than distributing strain

AC C

EP

289

12

ACCEPTED MANUSCRIPT

298

throughout the deforming region (Fig. 4). A weaker mid-crust results in more diffuse

299

deformation and negligible extension in the topographic divide region. Other factors may enhance the degree of extensional deformation that occurs.

301

Upton and Koons (2007) show in models based specifically on the Southern Alps, that

302

upper crustal NW-SE oriented dilatation is enhanced by the presence of elevated pore

303

pressure at depth beneath the topographic divide (Wannamaker et al., 2002). The

304

isotopic signature of Plio-Pleistocene mineralised veins in the topographic divide of

305

the Southern Alps is dominated by a rock-exchanged signature, reflecting rock-

306

buffered mineralising fluids (Becker et al., 2000; Cox et al., 1997; Koons et al., 1998).

307

In those models, elevated pore pressure in the mid-lower crust acts to weaken the

308

mid-crust and enhances the development of a higher velocity zone leading to material

309

in the mid-crust moving toward the Australian plate more rapidly than the upper

310

crustal material overlying it (Upton and Koons, 2007). The net result is extension in

311

the x-direction between the nearly stationary material and the material to the west

312

(Upton and Koons, 2007).

SC

M AN U

5.2.

TE D

313

RI PT

300

Variation of the stress tensor through the oblique orogen

Our models allow us to track the orientation of the stress tensor through the orogeny

315

as a function of obliquity (Fig. 8). As a rock moves through an orogen, the rock will

316

move through one or more stress regimes depending on the trajectory of the rock and

317

the degree of stress variability within the orogeny. Thus it should be expected to see

318

exhumed rocks with evidence for superimposition of structures from different stress

319

regimes through which they had passed.

320

represent changing stress states they do not necessarily represent a change in the far

321

field tectonic plate vectors (Platt and Rubie, 1987).

AC C

EP

314

13

Whilst these overprinting relationships

ACCEPTED MANUSCRIPT

In the simplest case, an orthogonal collision zone, where the transport direction is

323

perpendicular to the boundary, the stress tensor is invariant (Fig. 5A) and exhumed

324

rocks would be dominated by compressional structures with little evidence of

325

extensional structures overprinting that compression. With increasing obliquity, the

326

stress tensor rotates.

327

deforming zones the two minor principal stresses rotate away from horizontal and

328

vertical (Fig. 5B, E). In the highly oblique case, similar to that of the Southern Alps, and

rotates away from the transport direction and in the

SC

329

RI PT

322

have rotated so much that they have flipped with

now associated with the

330

horizontal axis and

331

mechanics for a strike slip fault (Anderson, 1905). Comparison of these results to

332

stresss inversions (Boese et al., 2012) and vein orientations from the Southern Alps

333

(highly oblique) shows good agreement with the models (Fig. 8A, C). Comparison of

334

model results to stress inversions (Hsu et al., 2009; Mouthereau et al., 2009; Wu et al.,

335

2010) and vein orientations (Craw et al., 2010) from Taiwan (oblique) is more

336

variable (Fig. 8B, D). Hsu et al. (2009) determine that σ1 has a plunge close to

337

horizontal and trend which is consistent with the compressive stress regime of the

338

collision zone. They find that σ2 and σ3 vary a lot across most of Taiwan (Fig. 8B)

339

and that the stress regime of the Central Range is transitional to either thrust or

340

normal faulting (Hsu et al., 2009; Fig. 7). Wu et al. (2010)also predict variable σ1 and

341

σ2 in the Central Range. The complexity in Taiwan is likely due to two factors.

342

Firstly, as an oblique orogen, the regionally controlled stress state is transitional

343

between convergent and strike slip, and is thus more likely to be influenced by local

344

variability. This means it is difficult to distinguish between region or first order

345

controls verses topographic effects which will extend to ~3 km b.s.l. as discussed

346

below.

AC C

EP

TE D

M AN U

sub-vertical (Fig. 5C, F), consistent with Andersonian

Secondly, the stress state in the Central Ranges experienced significant

14

ACCEPTED MANUSCRIPT

perturbation as a result of the 1999 Chichi earthquake (Hsu et al., 2009; Lin et al.,

348

2010). Temporal variability in stress state is beyond the scope of the study presented

349

here but it likely to have contributed to the complexity of observations (Hsu et al.,

350

2009; Lin et al., 2010).

351

5.3.

RI PT

347

Superposition of topographic effects

In detail, vein orientations in both orogens suggest more complex stress orientations.

353

In Taiwan, many of the veins emplaced at shallow levels have steep dips and most

354

strike approximately NW-SE (Fig. 8D), implying that

355

(Upton et al., 2011). In the Southern Alps, gold bearing veins and partly-coeval

356

lamprophyre dykes show a range of orientations, but three stand out: EW, NW-SE

357

and NE-SW (Fig. 8D) (Cooper et al., 1987; Craw et al., 2006a). The regional stress

358

tensor, as measured by Boese et al. (2012) and calculated from our models predicts

359

relatively steep extension fractures and extension shears striking NW-SE,

360

perpendicular to

M AN U

is horizontal not vertical

TE D

(Fig. 8B).

SC

352

In both cases, we suggest that the variability in the stress state we see is the near

362

surface influence of the considerable topography in these regions. Our model of the

363

topographic stresses in the central Southern Alps shows that topography can rotate the

364

stress tensor so that

365

determined by the regional topographic stresses. In general, given a vertical

366

orientation of

367

perpendicular to the strike of the mountain ranges.

368

topographic stresses produce steep extension fractures with a strike of NE-SW,

369

parallel to the main mountain chain. This extension orientation is consistent with the

370

orientations of some of the Haast Pass lamprophyre dikes and dipping sills and many

371

extensional veins in the central Southern Alps (Fig. 8D) that formed as the mountains

EP

361

AC C

becomes vertical and the orientation of the minor stresses is

will tend to be along the strike of the mountains and

15

, the will be

In the Southern Alps case,

ACCEPTED MANUSCRIPT

developed progressively towards the NE since the Miocene (Cooper et al., 1987;

373

Craw and Campbell, 2004). Finer scale ridge and valley topography will be more

374

variable resulting in a complex array of orientations of near surface fractures and

375

shears. Near surface fractures and shears in Taiwan strike approximately parallel to

376

the relative plate-vector, implying

377

(Craw et al., 2010; Upton et al., 2011).

RI PT

372

oriented along the strike of mountain ranges

This topographically controlled extension controls the shallowest mineralised

379

zones. The rock mass close to the surface hosts hydrothermal fluid flow in fractures

380

with numerous open cavities (Fig. 7B,C). Large (cm scale) prismatic crystals of

381

quartz and adularia, and bladed calcite crystals, are common in these cavities (Fig.

382

7B,C), some of which have grown on pre-existing hydrothermal deposits (Fig. 7C).

383

Vein formation typically occurred between 200 and 300°C from meteoric water that

384

has had variable amounts of isotopic exchange with host rocks (Craw et al., 2010;

385

Jenkin et al., 1994). Co-existing liquid and vapour fluid inclusions in many of the

386

New Zealand examples attest to boiling during entrapment (Craw, 1997; Craw et al.,

387

2009). Likewise, the mineral assemblage quartz-bladed calcite-adularia is typical of

388

geothermal boiling zones (Simmons and Christenson, 1994). Boiling of these fluids

389

occurs at depths of <2 km, and possibly only a few hundred metres, below the

390

hydrological surface, where fluid temperature exceeded rock temperature (Craw,

391

1997). This hydrological surface may be near the top of a mountain ridge, but could

392

also be at or near valley level (Fig. 7D). Hence, the location of the boiling zone is

393

poorly defined but typically lies beneath the highest mountains (Craw, 1997; Craw et

394

al., 2009). These shallow extensional vein systems generally do not contain gold, but

395

some very shallow level (near-surface) gold deposition has occurred locally (Craw et

396

al., 2002).

AC C

EP

TE D

M AN U

SC

378

16

ACCEPTED MANUSCRIPT

397

5.4.

Significance for ancient orogens

The extensional zones described in the Taiwan and Southern Alps convergent orogens

399

are being actively uplifted and eroded, and the gold deposited in them contributes to

400

nearby placers (Craw et al., 2010; Craw et al., 1999). Hence, individual veins have a

401

limited life within the basement rocks of the orogen, and are replaced by more of the

402

same as the convergence continues. When convergence ceases, and the locus of

403

tectonism changes because of plate boundary evolution, the last-formed vein systems

404

are preserved in the quiescent orogen until slower uplift and erosion over longer time

405

periods exposes veins in the deeper parts of the orogens (Craw et al., 2013). At this

406

stage, almost all evidence of the topography and tectonic processes of the active

407

orogens have been removed. Nevertheless, the distribution of zones of any remaining

408

gold-bearing extensional veins may give some clues to the general position of the

409

topographic divide within the orogen, as seen in the active orogens described above.

M AN U

SC

RI PT

398

At the outcrop or mine scale, extensional veins that fill fractures in

411

compressional structural elements, particularly fold axial surface fractures, are

412

widespread in metamorphic belts within ancient collisional orogens, and many of

413

these veins are gold-bearing (Bierlein and Crowe, 2000; Cox et al., 1995; Goldfarb et

414

al., 2005; Witt and Vanderhor, 1998) For example, numerous swarms of small

415

auriferous extensional veins hosted in Jurassic fold axial surfaces have provided

416

sources for the rich Yukon placers of Canada (MacKenzie et al., 2008) (Fig. 2F), and

417

gold-bearing veins were emplacement in fold axial surfaces in the late Mesozoic

418

collisional Zagros Orogen in Iran (Niroomand et al., 2011).

AC C

EP

TE D

410

419

On a larger scale, there is a strong association between gold mineralisation and

420

regional faults or shear zones, with the gold emplaced in minor structures adjacent to,

421

but not in, the regional structures (Goldfarb et al., 2005). Many of these gold-bearing

17

ACCEPTED MANUSCRIPT

structures are extensional, and have been superimposed on compressionally-deformed

423

host rocks (Bierlein and Crowe, 2000; Cox et al., 1995; Goldfarb et al., 2005; Witt

424

and Vanderhor, 1998). Rather than reflecting regional changes in tectonic regimes,

425

this structural superimposition may reflect ancient translation of host rocks into the

426

root zones of the extensional portions of the collisional orogens, similar to the

427

processes described in the active orogens in this study. However, this late extension-

428

hosted mineralisation is distinctly different from mineralisation that occurs in local

429

extensional sites during compressional deformation (Cox et al., 1995) or localised

430

extensional mineralisation associated with earthquake events (Cox and Ruming, 2004;

431

Micklethwaite et al., 2010).

432

Conclusions

433

The common occurrence of regional-scale superimposition of extensional structures

434

on compressionally-deformed rocks in ancient orogenic gold systems is often cited as

435

requiring a change in the tectonic plate vectors. While this has apparently occurred in

436

some ancient settings (Allibone et al., 2002; Goldfarb et al., 1991), it is not required.

437

We use two active oblique orogens, the Southern Alps of New Zealand, and Taiwan,

438

as examples of the on-going transfer of rocks from a compressional regime to an

439

extensional regime within the same orogeny.

EP

TE D

M AN U

SC

RI PT

422

Extensional deformation within a convergent orogen develops beneath the

441

region of the topographic divide. Extension is favoured by a high degree of obliquity

442

and by a mid-lower crustal rheology that promotes localisation of shearing rather than

443

diffuse deformation.

444

observed in both Taiwan and the Southern Alps, the stress tensor must rotate from the

445

standard convergent situation where σ3 is vertical and σ3 must be small or

446

extensional.

AC C

440

In order to produce the sub-vertical extensional structures

Increasing obliquity rotates the stress tensor and, by increasing the

18

ACCEPTED MANUSCRIPT

447

horizontal shear stress τxz, reduces the magnitude of σ3, thereby increasing the

448

differential stress. In the near surface, topographic stresses are superimposed onto the tectonic

450

stresses and in regions of high relief such as the Southern Alps and Taiwan, can have

451

a significant effect. In the central Southern Alps, material to depths of ~3 km b.s.l. is

452

brought closer to failure by the topographic stresses. The stress tensor is rotated so

453

that σ1 is vertical and the orientation of the vertical extensional shears that develop is

454

largely controlled by the local topography.

SC

RI PT

449

With cessation of the convergence, the last-formed vein systems are preserved

456

in the quiescent orogeny while longer term erosion exposes veins from deeper in the

457

orogeny. By then almost all evidence of topography and tectonic processes will have

458

gone. Nevertheless, the distribution of any remaining extensional vein swarms may

459

point to the general position of the topographic divide. Similarly, igneous rocks may

460

be preferentially emplaced in the extensional zone, and their eroded roots form linear

461

intrusive belts in ancient orogens.

462

Acknowledgements

463

This research was funded by NZ Ministry for Business, Innovation and Employment,

464

and University of Otago. The authors contributions are as follows: PU and DC:

465

conceptual development of the study; PU: did the numerical modelling; PU and DC

466

wrote the manuscript. Discussions with Peter O Koons, Doug MacKenzie and Grant

467

Caldwell and reviews by Shaun Barker and one anonymous referee have helped to

468

develop the ideas expressed here.

AC C

EP

TE D

M AN U

455

469

19

ACCEPTED MANUSCRIPT

470 References

472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517

Allibone, A., Teasdale, J., Cameron, G., Etheridge, M.A., Uttley, P., Soboh, A., Appiah-Kubi, J., Adanu, A., Arthur, R., Mamphey, J., Odoom, B., Zuta, J., Tsikata, A., Pataye, F., Famiyeh, S., Lamb, E., 2002. Timing and structural control on gold mineralisation at the Bogoso Gold Mine, Ghana, West Africa. Economic Geology 97, 949-969. Allis, R.G., 1981. Continental underthrusting beneath the Southern Alps of New Zealand. Geology 9, 303-307. Anderson, E.M., 1905. The dynamics of faulting. Transactions of the Edinburgh Geological Society 8, 387-402. Anderson, H., Webb, T., Jackson, J., 1993. Focal mechanisms of large earthquakes in the South Island of New Zealand: implications for the accommodation of PacificAustralia plate motion. Geophysical Journal International 115, 1032-1054. Barnes, I., Downes, C.J., Hulston, J.A., 1978. Warm springs, South Island, New Zealand, and their potential to yield laumontite. American Journal of Science 278, 1412-1427. Beavan, J., Moore, M., Pearson, C., Henderson, M., Parsons, B., Bourne, S., England, P., Walcott, R.I., Blick, G., Darby, D., Hodgkinson, K., 1999. Crustal deformation during 1994-98 due to oblique continental collision in the central Southern Alps, New Zealand, and implications for seismic potential of the Alpine fault. Journal of Geophysical Research 104, 25233-25255. Beck, A.C., 1968. Gravity faulting as a mechanism of topographic adjustment. New Zealand Journal of Geology and Geophysics 11, 191-199. Becker, J.A., Craw, D., Horton, T., Chamberlain, C.P., 2000. Gold mineralisation near the Topographic divide, upper Wilberforce valley, Southern Alps, New Zealand. New Zealand Journal of Geology and Geophysics 43, 199-215. Bertrand, E., Unsworth, M., Chiang, C.W., Chen, C.S., Chen, C.C., Wu, F., Turkoglu, E., Hsu, H.L., Hill, G., 2009. Magnetotelluric evidence for thick-skinned tectonics in central Taiwan. Geology 37, 711-714. Bierlein, F.P., Crowe, D.E., 2000. Phanerozoic orogenic lode gold deposits, in: Hagemann, S.G., Brown, P.E. (Eds.), Gold in 2000. Society of Economic Geologists, pp. 103-139. Boese, C.M., Townend, J., Smith, E.G.C., Stern, T., 2012. Microseismicity and stress in the vicinity of the Alpine Fault, central Southern Alps, New Zealand. Journal of Geophysical Research 117. Byrne, T., Chan, Y.-C., Rau, R.-J., Lu, C.-Y., Lee, Y.-H., Wang, Y.-J., 2013. The Arc-Continent Collision in Taiwan, in: Brown, D., Ryan, P.D. (Eds.), Arc-Continent Collision, Frontiers in Earth Sciences. Chen, C.-H., 1985. Chemical characteristics of thermal waters in the Central Range of Taiwan, R.O.C. Chemical Geology 49, 303-317. Clark, S.P.J., 1966. Handbook of Physical Constants. The Geological Society of America, Inc., New York. Cooper, A.F., Barreiro, B.A., Kimbrough, D.L., Mattinson, J.M., 1987. Lamprophyre dike intrusion and the age of the Alpine fault, New Zealand. Geology 15, 941-944. Cox, S.C., Craw, D., Chamberlain, C.P., 1997. Structure and fluid migration in a late Cenozoic duplex system forming the Topographic divide in the central Southern Alps, New Zealand. New Zealand Journal of Geology and Geophysics 40, 359-373.

AC C

EP

TE D

M AN U

SC

RI PT

471

20

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Cox, S.F., Ruming, K., 2004. The St Ives mesothermal gold system, Western Australia - A case of golden aftershocks? Journal of Structural Geology 26, 11091125. Cox, S.F., Sun, S.S., Etheridge, M.A., Wall, V.J., Potter, T.F., 1995. Structural and geochemical controls on the development of turbidite- hosted gold quartz vein deposits, Wattle Gully mine, central Victoria, Australia. Economic Geology 90, 17221746. Craw, D., 1997. Fluid inclusion evidence for geothermal structure beneath the Southern Alps, New Zealand. New Zealand Journal of Geology and Geophysics 40, 43-52. Craw, D., 2006. Gold-bearing veins in the Southern Alps, in: Christie, A.B., Brathwaite, R.L. (Eds.), Geology and Exploration of New Zealand Mineral Deposits, pp. 285-288. Craw, D., Begbie, M., MacKenzie, D., 2006a. Structural controls on tertiary orogenic gold mineralization during initiation of a mountain belt, New Zealand. Mineralium Deposita 41, 645-659. Craw, D., Campbell, J.R., 2004. Tectonic and structural setting for active mesothermal gold vein systems, Southern Alps, New Zealand. Journal of Structural Geology 26, 995-1005. Craw, D., Koons, P.O., Horton, T., Chamberlain, C.P., 2002. Tectonically driven fluid flow and gold mineralisation in active collisional orogenic belts: Comparison between New Zealand and western Himalaya. Tectonophysics 348, 135-153. Craw, D., Upton, P., Horton, T., Williams, J., 2013. Migration of hydrothermal systems in an evolving collisional orogen, New Zealand. Mineralium Deposita 48, 233-248. Craw, D., Upton, P., Mackenzie, D.J., 2009. Hydrothermal alteration styles in ancient and modern orogenic gold deposits, New Zealand. New Zealand Journal of Geology and Geophysics 52, 11-26. Craw, D., Upton, P., Yu, B.-S., Horton, T., Chen, Y.-G., 2010. Young orogenic gold mineralisation in active collisional mountains, Taiwan. Mineralium Deposita 45, 631646. Craw, D., Youngson, J.H., Koons, P.O., 1999. Gold dispersal and placer formation in an active oblique collisional mountain belt, Southern Alps, New Zealand. Economic Geology 94, 605-614. Craw, D., Youngson, J.H., Leckie, D.A., 2006b. Transport and concentration of detrital gold in foreland basins. Ore Geology Reviews 28, 417-430. Crespi, J.M., Chan, Y.C., Swaim, M.S., 1996. Synorogenic extension and exhumation of the Taiwan hinterland. Geology 24, 247-250. DeMets, C., Gordon, R.G., Argus, D., Stein, S., 1994. Effect of recent revisions to the geomagnetic reversal time scale on estimates of current plate motion. Geophyiscal Research Letters 21, 2191-2194. Ellis, S., Beavan, J., Eberhart-Phillips, D., 2006. Bounds on the width of mantle lithosphere flow derived from surface geodetic measurements: Application to the central Southern Alps, New Zealand. Geophysical Journal International 166, 403-417. Fisher, D.M., Lu, C., Chu, H.T., 2002. Taiwan Slate Belt: Insights into the ductile interior of an arc-continent collision, in: Byrne, T., Liu, C.S. (Eds.), Geology and Geophysics of an Arc-Continent Collision, Taiwan. Geological Society of America, pp. 93-106. Goldfarb, R.J., Baker, T., Dube, B., Groves, D.I., Hart, C.J., Gosselin, P., 2005. Distribution, character and genesis of gold deposits in metamorphic terranes, in:

AC C

518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567

21

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), Economic Geology 100th Anniversary Volume, pp. 407-450. Goldfarb, R.J., Newberry, R.J., Pickthorn, W.J., Gent, C.A., 1991. Oxygen, hydrogen, and sulfur isotope studies in the Juneau gold belt, southeastern Alaska: constraints on the origin of hydrothermal fluids. Economic Geology 86, 66-80. Groves, D.I., Goldfarb, R.J., Gebre-Mariam, M., Hagemann, S.G., Robert, F., 1998. Orogenic gold deposits: A proposed classification in the context of their crustal distribution and relationship to other gold deposit types. Ore Geology Reviews 13, 227. Groves, D.I., Goldfarb, R.J., Robert, F., Hart, C.J.R., 2003. Gold deposits in metamorphic belts: overview of current understanding, outstanding problems, future research, and exploration significance. Economic Geology 98, 1-29. Ho, C.S., 1986. A synthesis of the geological evolution of Taiwan. Tectonophysics 125, 1-16. Hsu, Y.-J., Yu, B.-S., Simons, M., Kuo, L.-C., Chen, H.-Y., 2009. Interseismic crustal deformation in the Taiwan plate boundary zone revealed by GPS observations, seismicity and earthquake focal mechanisms. Tectonophysics 479, 4-18. Itasca, 2009. FLAC3D (Fast Lagrangian Analysis of Continua in 3 Dimensions) User's Guide, 4.0 ed. Itasca Consulting Group Inc., Minneapolis, Minnesota. Jenkin, G., Craw, D., Fallick, A., 1994. Stable isotopic and fluid inclusion evidence for meteoric fluid penetration into an active mountain belt; Alpine Schist, New Zealand. Journal of Metamorphic Geology 12, 429-444. Johnson, S.E., Vernon, R.H., Upton, P., 2004. Foliation development and progressive strain-rate partitioning in the crystallizing carapace of a tonalite pluton: Microstructural evidence and numerical modeling. Journal of Structural Geology 26, 1845-1865. Kontak, D.J., Horne, R.J., 2010. A Multi-stage Origin for the Meguma Lode Gold Deposits, Nova Scotia, Canada, in: Deb, M. (Ed.), Gold Metallogeny. Alpha Science, Narosa, p. 320 Koons, P.O., 1989. The topographic evolution of collisional mountain belts: a numerical look at the Southern Alps, New Zealand. American Journal of Science 289, 1041-1069. Koons, P.O., 1990. Two-sided orogen: collision and erosion from the sandbox to the Southern Alps, New Zealand. Geology 18, 679-682. Koons, P.O., 1995. Modeling the topographic evolution of collisional belts. Annual Review of Earth and Planetary Sciences 23, 375-408. Koons, P.O., Craw, D., Cox, S.C., Upton, P., Templeton, A.S., Chamberlain, C.P., 1998. Fluid flow during active oblique convergence: A Southern Alps model from mechanical and geochemical observations. Geology 26, 159-162. Koons, P.O., Kirby, E., 2007. Topography, Denudation, and Deformation. The Role of Surface Processes in Fault Evolution, in: Handy, M.R., Hirth, G., Hovius, N. (Eds.), Tectonic Faults; Agents of change on a dynamic Earth. MIT Press, pp. 205230. Koons, P.O., Upton, P., Barker, A.D., 2012. The influence of mechanical properties on the link between tectonic and topographic evolution. Geomorphology 29. Lee, C.R., Cheng, W.T., 1986. Preliminary heat flow measurements in Taiwan, Fourth Circum-Pacific Energy and Mineral Resources Conference, Singapore.

AC C

568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615

22

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Leitner, B., Eberhart-Phillips, D., Anderson, H., Nabelek, J.L., 2001. A focused look at the Alpine fault, New Zealand: Seismicity, focal mechanisms, and stress observations. Journal of Geophysical Research B: Solid Earth 106, 2193-2220. Lin, K.-C., Hu, J.C., Ching, K.-E., Angelier, J., Rau, R.-J., Yu, S.-B., Tsai, C.-H., Shin, T.-C., Huang, M.-H., 2010. GPS crustal deformation, strain rate, and seismic activity after the 1999 Chi-Chi earthquake in Taiwan. Journal of Geophysical Research 115. MacKenzie, D.J., Craw, D., Mortensen, J., 2008. Structural controls on orogenic gold mineralisation in the Klondike goldfield, Canada. Mineralium Deposita 43, 435-448. Micklethwaite, S., Sheldon, H.A., Baker, T., 2010. Active fault and shear processes and their implications for mineral deposit formation and discovery. Journal of Structural Geology 32, 151-165. Mouthereau, F., Fillon, C., Ma, K.F., 2009. Distribution of strain rates in the Taiwan orogenic wedge. Earth and Planetary Science Letters 284, 361-385. Niroomand, S., Goldfarb, R.J., Moore, F., Mohajjel, M., Marsh, E.E., 2011. The Kharapeh orogenic gold deposit: geological, structural, and geochemical controls on epizonal ore formation in West Azerbaijan Province, Northwestern Iran. Mineralium Deposita 46, 409-428. Norris, R.J., Koons, P.O., Cooper, A.F., 1990. The obliquely-convergent plate boundary in the South Island of New Zealand: implications for ancient collision zones. Journal of Structural Geology 12, 715-725. Oliver, N.H.S., McLellan, J.G., Hobbs, B.E., Cleverley, J.S., Ord, A., Feltrin, L., 2006. Numerical models of extensional deformation, heat transfer, and fluid flow across basement-cover interfaces during basin-related mineralisation. Economic Geology 101, 1-31. Phillips, G.N., Powell, R., 2009. Formation of gold deposits: review and evaluation of the continuum model. Earth Science Reviews 94, 1-21. Platt, J.P., Rubie, D.C., 1987. The uplift of high-pressure - low-temperature metamorphic rocks. Philosophical Transactions of the Royal Society, London 321, 87-103. . Ridley, J., 1993. The relations between mean rock stress and fluid flow in the crust, with reference to vein- and lode-style gold deposits. Ore Geology Reviews 8, 23-37. Sibson, R.H., Robert, F., Poulsen, K.H., 1988. High-angle reverse faults, fluidpressure cycling, and mesothermal gold-quartz deposits. Geology 16, 551-555. Simmons, S.F., Christenson, B.W., 1994. Origins of calcite in a boiling geothermal system. American Journal of Science 294, 361-400. Suppe, J., 1980. A retrodeformable cross section of northern Taiwan. Proceedings Geological Society of China 23, 46-55. Suppe, J., 1981. Mechanics of mountain building and metamorphism in Taiwan. Memoir of the Geological Society of China (Taiwan) 4, 67-89. Tan, L.P., Chen, C.H., Yeh, T.K., A, T., 1991. Tectonic and geochemical characteristics of Pleistocene gold deposits in Taiwan, in: Jones, M., Cosgrove, J. (Eds.), Neotectonics and Resources. Bellhaven Press, London, pp. 290-305. Upton, P., Begbie, M., Craw, D., 2008. Numerical modelling of mechanical controls on coeval steep and shallow dipping auriferous quartz vein formation in a thrust zone, Macraes mine, New Zealand. Mineralium Deposita 43, 23-35. Upton, P., Craw, D., Caldwell, T.G., Koons, P.O., James, Z., Wannamaker, P.E., Jiracek, G.J., Chamberlain, C.P., 2003. Upper crustal fluid flow in the outboard region of the Southern Alps, New Zealand. Geofluids 3, 1-12.

AC C

616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664

23

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

Upton, P., Craw, D., Yu, B., Chen, Y.-G., 2011. A comparison of orogenic fluid flow: Southern Alps, New Zealand and Taiwan, in: Fagereng, A., Toy, V., Rowland, J.V. (Eds.), Geology of the earthquake source: a volume in honour of Rick Sibson. London: Geological Society, London, pp. 249-265. Upton, P., Koons, P.O., 2007. Three-dimensional geodynamic framework for the Central Southern Alps, New Zealand: Integrating geology, geophysics and mechanical observations, in: Okaya, D., Stern, T., Davey, F. (Eds.), A Continental Plate Boundary: Tectonics at South Island, New Zealand. American Geophysical Union, Washington, DC. Upton, P., Koons, P.O., Craw, D., Henderson, C.M., Enlow, R., 2009a. Along-strike differences in the Southern Alps of New Zealand: Consequences of inherited variation in rheology. Tectonics 28. Upton, P., Mueller, K., Chen, Y.G., 2009b. Three-dimensional numerical models with varied material properties and erosion rates: Implications for the mechanics and kinematics of compressive wedges. Journal of Geophysical Research B: Solid Earth 114. Vielreicher, N.M., Groves, D.I., Snee, L.W., Fletcher, I.R., McNaughton, N.J., 2010. Broad synchroneity of three gold mineralization styles in the Kalgoorlie Gold Field: SHRIMP, U-Pb, and 40Ar/39Ar geochronological evidence. Economic Geology 105, 187-227. Walcott, R.I., 1986. The kinematics of the plate boundary zone through New Zealand: a comparison of short- and long-term deformations. Geophysical Journal of the Royal Astronomical Society 79, 613-633. Wannamaker, P.E., Jiracek, G.R., Stodt, J.A., Caldwell, T.G., Gonzalez, V.M., McKnight, J.D., Porter, A.D., 2002. Fluid generation and pathways beneath an active compressional orogen, the New Zealand Southern Alps, inferred from magnetotelluric data. Journal of Geophysical Research B: Solid Earth 107, 6-1. Williams, G.J., 1974. Economic geology of New Zealand. . Austral Inst. Mining Metall Monog 4. Witt, W.K., Vanderhor, F., 1998. Diversity within a unified model for Archean gold mineralization in the Yilgarn Craton of Western Australia: an overview of the lateorogenic structurally controlled gold deposits. Ore Geology Reviews 13, 29-64. Wu, F.T., Davey, F.J., Okaya, D.A., 2007. Taiwan and South Island, New Zealand: A Comparison of Continental Collisional Orogenies, in: Okaya, D., Stern, T., Davey, F. (Eds.), A Continental Plate Boundary: Tectonics at South Island, New Zealand. American Geophysical Union, Washington, DC, pp. 329-346. Wu, W.-N., Kao, H., Hsu, S.-K., Lo, C.-L., Chen, H.-W., 2010. Spatial variation of the crustal stress field along the Ryukyu-Taiwan-Luzon convergent boundary. Journal of Geophysical Research 115.

AC C

665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704

24

ACCEPTED MANUSCRIPT

705 Figure captions

707

Figure 1: A: Plate tectonic setting of New Zealand. B: Hillshade map of the South

708

Island of New Zealand showing the Southern Alps. Elevations range from 0 to

709

3754 m. C: Map of the South Island showing major active faults, relative plate

710

motion from De Mets et al. (1994), Mesozoic and late Cenozoic gold bearing

711

veins, hot springs, and the region of the Southern Alps where extensional

712

structures are found. D: Hillshade map of Taiwan. Elevations range from 0 to

713

3952 m. E: Map of Taiwan showing late Cenozoic gold bearing veins, hot

714

springs, and the region of the Central Ranges where extensional structures are

715

found. F: Plate tectonic setting of Taiwan.

M AN U

SC

RI PT

706

Figure 2: Photographs of extensional veins (Ex. vein), with summary sketches, in

717

folded rocks. FAS = fold axial surface, commonly with parallel joints or

718

fractures that host veins. A-C: Southern Alps, New Zealand. The FAS fractures

719

in the cliff face in A host extensional veins. D-E: Slate Range, Taiwan; veins in

720

E are gold-bearing and are exposed in an underground mine face cut

721

perpendicular to fold axial surface fractures. F: Jurassic extensional gold-

722

bearing veins formed in fold axial surface fractures, Klondike goldfield, Yukon,

723

Canada.

AC C

EP

TE D

716

724

Figure 3: A: Geometry and boundary conditions for the three-dimensional

725

mechanical model. Model is of a two layered crust which is subjected to three

726

different velocity boundary conditions as shown by the three arrows. Material

727

is pushed from the right and dragged along the base.

728

approximately horizontal except for an elevated brittle-ductile transition close to

729

the model plate boundary; the upper crust is represented by a pressure-

25

Crustal rheology is

ACCEPTED MANUSCRIPT

dependent rheology and the lower crust and mantle by a temperature-dependent

731

rheology. Approximately steady state elevations are maintained adjacent to the

732

plate boundary. Axes define reference frame used where y is parallel to the

733

plate boundary and x is normal to the plate boundary. B: Cross-section through

734

the centre of the model showing locations of stress tensor plots in figures 5 and

735

8.

Figure 4: Horizontal velocity component ( ) and its spatial derivative (

) for

SC

736

RI PT

730

the series of models described in Table 2. A: Strong crust with an orthogonal

738

boundary condition. B: Strong crust with an oblique boundary condition. C:

739

Strong crust with a highly oblique boundary condition. D: Weak crust with an

740

orthogonal boundary condition.

741

condition. F: Weak crust with a highly oblique boundary condition. See text

742

for more details.

E: Weak crust with an oblique boundary

Figure 5: A-C: Stereonets showing principal stress orientations at 14 points within

TE D

743

M AN U

737

the three models, all with strong rheology.

745

outboard region and circles are values from beneath the deforming region (Fig.

746

3B). See text for more details. D-F: Block diagrams showing model geometry,

747

boundary conditions and resultant rock trajectories.

748

principal stress directions. On E and F the plane containing

AC C

749

sectional view orientated perpendicular to

750

Figure 6: Stress magnitudes,

751

Orthogonal convergence.

752

Crosses are values from the

EP

744

,

and

Also shown are the and

is a cross-

. for the three strong models.

B: Oblique convergence.

A:

C: Highly oblique

convergence.

753

Figure 7: A: Model of topographic stresses in the central Southern Alps. Map view

754

contours are elevation in metres. Cross-section contours are the strength/stress

26

ACCEPTED MANUSCRIPT

755

ratio resulting from topography which is a measure of how close the rock mass

756

is to failure. Areas in blue are close to failure, areas in red are far from failure.

757

Purple band along the top shows regions that have failed in tension. Inset shows

758

sample model stress tensors close to the surface.

759

orientations of σ2 and σ3 depend on the nearly topography.

760

adularia (Ad) and quartz (Q) from a meteoric water boiling zone in the Southern

761

Alps. C: Prismatic quartz (Q) and bladed calcite (Ca) from a meteoric water

762

boiling zone in Taiwan Central Range. D: Sketch cross section through a

763

typical mountain range in an active collisional orogen (based on Southern Alps

764

and Taiwan Central Range). Meteoric and deep-sourced fluids mix beneath the

765

mountains via an extensional fracture network that is partly formed by

766

topographic collapse. Rising hot fluids boil in these fractures at depths (fluid

767

pressures) controlled by fracture connections to the surface at various

768

topographic levels, depositing mineral assemblages such as those shown in B

769

and C.

RI PT

σ1 is vertical while the

TE D

M AN U

SC

B: Euhedral

Figure 8: A: Model stress tensors for the highly oblique model rotated into a

771

Southern Alps reference frame for comparison with structural data from that

772

orogeny. Large open circles are stress tensors from the topographic model. The

773

maximum horizontal stress (MHS) determined by Boese et al. (2012) is also

775 776

AC C

774

EP

770

shown. B: Model stress tensors for the oblique model rotated into a Taiwan reference frame for comparison with structural data from that orogeny. C: Example stress tensor from Boese et al. (2012) from the central Southern Alps,

777

and rose diagrams showing vein and dyke orientations from the length of the

778

Southern Alps (after Cooper et al., 1987; Craw et al., 2006a).D: Extent of

779

regional extension in Taiwan as determined from geodetic measurements and

27

ACCEPTED MANUSCRIPT

780

strain directions from earthquakes (Mouthereau et al 2009).

781

tensors from Hsu et al. 2009.

Sample stress

782

AC C

EP

TE D

M AN U

SC

RI PT

783

28

ACCEPTED MANUSCRIPT

Table 1. Tectonic and geological settings of extensional zones and gold-bearing vein systems in the hanging wall of major crustal structures, Taiwan and South Island, New Zealand.

Movement sense Lithologies

Principal mountains Maximum altitude Exhumation rate Extensional zone

Location

Extension direction Gold-bearing veins

Distribution

TE D

Late metamorphic veins Extensional ankeritic veins

Fluid composition

Distribution Geological control Temperature Water source Heat source

AC C

Warm/hot springs

EP

Fluid source

New Zealand SI Pacific-Australia Oblique collision 70° from perpendicular 37 mm/year Miocene Alpine Fault

RI PT

Geology

Taiwan Eurasia-Phillipine Oblique collision 20° from perpendicular 88 mm/year Pliocene Longitudinal Valley Fault Sinistral reverse Cenozoic sediments, minor crystalline basement; outboard Pliocene- Recent molasse Central Range, including Slate Belt 3952 m >5 mm/year in mountains Immediately east (inboard) of topographic divide NE, high angle to convergence vector Focussed zones near topographic divide Shallow dip, subparallel to folded foliation Cut steeply across folded foliation, fill fold axial surface fractures

Dextral reverse Mesozoic basement; outboard Cenozoic cover & PlioceneRecent molasse

SC

Principal on-land structure

Plate boundary Motion Obliquity Rate Initiation Surface feature

Southern Alps

3754 m Up to 8 mm/year at plate boundary Centred on present topographic divide and paleo-divide NW, high angle to convergence vector Widely dispersed near topographic divide Irregular veins, abundant host rock alteration Cut steeply across folded foliation, fill fractures & normal faults Low-salinity rockexchanged fluid Metamorphic devolatilisation beneath mountains Near plate boundary Extensional fractures Max. 56°C Meteoric Rapid rock uplift

M AN U

Tectonics

Brines; rock-exchanged fluid Metamorphic devolatilisation beneath mountains Throughout mountains Extensional fractures Up to 99°C Meteoric Rapid rock uplift

ACCEPTED MANUSCRIPT

Table 2: Parameters, and material properties used in the numerical models Model name Orthogonal

Boundary condition vy = 0

Rheology

Results: extension

10

K = 1x10 Pa 9 G = 3x10 Pa ρ = 2800 kg m-3 φ = 30° coh = 2e7 Pa 1

6

1a

kσ = 1 x 10 – 1 x 10

1b 2

kσ = 1 x 10 – 6 x 10

5

6

M AN U 5

7

kσ = 1 x 10 – 6 x 10

yv = 4×vx

3a

K = 1x10 Pa 9 G = 3x10 Pa ρ = 2800 kg m-3 φ = 30° coh = 2e7 Pa 6

8

5

7

TE D

kσ = 1 x 10 – 6 x 10

10

Topographic stresses

K = 1x10 Pa 9 G = 3x10 Pa ρ = 2800 kg m-3 φ = 30° coh = 2e7 Pa

AC C

EP

Southern Alps topography

Minor extension to ~10km depth beneath topographic divide Negligible extension

10

kσ = 1 x 10 – 1 x 10

3b

4

8

kσ = 1 x 10 – 1 x 10

Highly oblique

Minor extension to ~9km depth beneath topographic divide. Negligible extension

SC

K = 1x10 Pa G = 3x109 Pa -3 ρ = 2800 kg m φ = 30° coh = 2e7 Pa

2a

2b 3

7

10

vx = vy

Oblique

8

RI PT

Model number 1

Significant extension to ~19 km depth beneath the topographic divide Minor extension to ~10 km depth beneath topographic divide Depends on local topography. σ1 near vertical, σ2 subparallel to mountain range, σ3 perpendicular to mountain range

K = bulk modulus, G = shear modulus, ρ = density, φ = friction angle, coh = cohesion.

Creep flow law is εɺ = Aσ e A is a pre-exponential factor, σ is the differential stress (MPa), n is the power-law exponent, Q is the activation energy, R the universal gas constant and T is temperature (K). A, Q and n are determined from extrapolation of laboratory creep experiments on wet synthetic quartzite (Paterson and Luan, 1990). A = 6.5x10-8 MPa-ns-1; Q = 135 KJ mol-1; n=3.1. In the lower crust the rheology is that of diabase, A = 2 x 10-4 MPa-ns-1; Q = 260 KJ mol-1; n = 3.4 (Shelton and Tullis, 1981). kσ is determined from the calculated differential stress, kσ = 0.5σ. 1

n

( − Q / RT )

1

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

A B

ACCEPTED MANUSCRIPT 20 m Ex. vein

D

FAS, veins

Foliation

Foliation

Ex. veins

Reverse faulted quartz veins

M AN U

10 cm

SC

C

Normal fault zone

TE D

Ex. veins

EP

E

AC C

Ex. vein

Foliation

Foliation

FAS

FAS

10 cm

FAS

Veins

n

tio

lia Fo

Foliation

Altered rock

Ex. vein

RI PT

Folded metamorphic veins

F

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Topographic divide extension is a predicable feature of oblique convergence



Folded rocks are laterally translated into the extensional regime



Extensional structures are overprinted on compressional structures



Gold is emplaced in extensional sites beneath the mountains



Vein orientations in the upper 3 km reflect both tectonic and topographic stresses

AC C

EP

TE D

M AN U

SC

RI PT