processes controlling vertical coupling and …kklepeis/publications/klepeis et al....

27
UNCORRECTED PROOF Processes controlling vertical coupling and decoupling between the upper and lower crust of orogens: results from Fiordland, New Zealand Keith A. Klepeis a, * , Geoffrey L. Clarke b , George Gehrels c , Jeff Vervoort c a Department of Geology, University of Vermont, Burlington, VT, 05405-0122, USA b School of Geosciences, Division of Geology and Geophysics, University of Sydney, NSW 2006, Australia c Department of Geosciences, University of Arizona, Tucson, AZ 85721, USA Received 27 January 2002; received in revised form 15 July 2003; accepted 25 August 2003 Abstract The pre-Cenozoic configuration of western New Zealand allows determination of the effects of magmatism and a changing lower crustal rheology on the evolution of a Cretaceous orogen from upper to lower crustal levels (10 – 50 km). Beginning at , 126 Ma, a composite batholith dominated by diorite was emplaced into the lower crust. During emplacement, deformation was partitioned into zones weakened by magma and heat, leading to the development of two layer-parallel shear zones at the upper and lower contacts of the batholith. Transient vertical decoupling of the crust above and below the batholith occurred from , 126 Ma until , 120 Ma as magma was emplaced into and moved through a weak, thick lower crust. By , 116 Ma, however, much of the batholith had crystallized and the lowermost crust had cooled from 750 8C , T , 850 8C to T ¼ 650–700 8C. Cooling was aided by the juxtaposition of pre-existing crust against hot new crust and by the efficient extraction of partial melts out of the lower crust. Cooling together with dehydration of the lower crust and mafic compositions led to the development of a strong, dry, lower crustal root by , 116 Ma. A strong lower crust resulted in high degrees of vertical coupling between the upper and lower crust during contraction from , 116 to , 105 Ma even as magma continued to be emplaced into the mid-upper crust. A narrow, focused orogenic style in the upper crust at this time reflected a highly viscous lower crust through which compressional stresses were transferred vertically. The results imply that changes in plate boundary dynamics rather than the thermal weakening of thick lower crust during convergence controlled the onset of regional extension at , 108 – 105 Ma. q 2003 Published by Elsevier Ltd. Keywords: Vertical coupling and decoupling; Magmatism; Orogen 1. Introduction Studies of convergent margins worldwide have shown that deformation patterns and the mechanical behavior of continental crust vary according to crustal level and tectonic setting (e.g. Sisson and Pavlis, 1993; Axen et al., 1998; Klepeis and Crawford, 1999; Miller and Paterson, 2001; Karlstrom and Williams, 2002; Teyssier et al., 2002). Experimental data (Wilks and Carter, 1990; Rushmer, 1995; Rutter and Neumann, 1995), numerical simulations (Harry et al., 1995; Ellis et al., 1998; McKenzie et al., 2000), and analytical models (Royden, 1996) indicate that lower crustal strength and rheology especially affect how deformation is partitioned vertically through the lithosphere during con- vergence. These studies emphasize the critical role the lower crust plays in linking the upper mantle with the upper crust of orogens. Despite this work, however, we still lack direct information on the mechanisms by which deformation is relayed vertically between different sections of the litho- sphere, especially as physical and chemical conditions in the lower crust change. Large, dipping shear zones that divide the crust and upper mantle into different structural domains have been observed or postulated in many orogenic belts (Oldow et al., 1990; Harry et al., 1995; Willett, 1998; McKenzie et al., 2000; Teyssier et al., 2002) but we do not fully understand how deformation above, below, and within these potentially transient features relate to one another or affect orogenic evolution. This gap in knowledge arises partly because orogens that allow direct observation of processes at lower crustal levels and their relationship with the upper crust are rare. In addition, the age and kinematic 0191-8141/$ - see front matter q 2003 Published by Elsevier Ltd. doi:10.1016/j.jsg.2003.08.012 Journal of Structural Geology xx (0000) xxx–xxx www.elsevier.com/locate/jsg * Corresponding author. Tel.: þ 1-802-656-0246; fax: þ1-802-656-0045. E-mail address: [email protected] (K.A. Klepeis). SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor ARTICLE IN PRESS 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112

Upload: others

Post on 26-Jun-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Processes controlling vertical coupling and decoupling between the upper

and lower crust of orogens: results from Fiordland, New Zealand

Keith A. Klepeisa,*, Geoffrey L. Clarkeb, George Gehrelsc, Jeff Vervoortc

aDepartment of Geology, University of Vermont, Burlington, VT, 05405-0122, USAbSchool of Geosciences, Division of Geology and Geophysics, University of Sydney, NSW 2006, Australia

cDepartment of Geosciences, University of Arizona, Tucson, AZ 85721, USA

Received 27 January 2002; received in revised form 15 July 2003; accepted 25 August 2003

Abstract

The pre-Cenozoic configuration of western New Zealand allows determination of the effects of magmatism and a changing lower crustal

rheology on the evolution of a Cretaceous orogen from upper to lower crustal levels (10–50 km). Beginning at ,126 Ma, a composite

batholith dominated by diorite was emplaced into the lower crust. During emplacement, deformation was partitioned into zones weakened by

magma and heat, leading to the development of two layer-parallel shear zones at the upper and lower contacts of the batholith. Transient

vertical decoupling of the crust above and below the batholith occurred from ,126 Ma until ,120 Ma as magma was emplaced into and

moved through a weak, thick lower crust. By ,116 Ma, however, much of the batholith had crystallized and the lowermost crust had cooled

from 750 8C , T , 850 8C to T ¼ 650–700 8C. Cooling was aided by the juxtaposition of pre-existing crust against hot new crust and by the

efficient extraction of partial melts out of the lower crust. Cooling together with dehydration of the lower crust and mafic compositions led to

the development of a strong, dry, lower crustal root by ,116 Ma. A strong lower crust resulted in high degrees of vertical coupling between

the upper and lower crust during contraction from ,116 to ,105 Ma even as magma continued to be emplaced into the mid-upper crust. A

narrow, focused orogenic style in the upper crust at this time reflected a highly viscous lower crust through which compressional stresses

were transferred vertically. The results imply that changes in plate boundary dynamics rather than the thermal weakening of thick lower crust

during convergence controlled the onset of regional extension at ,108–105 Ma.

q 2003 Published by Elsevier Ltd.

Keywords: Vertical coupling and decoupling; Magmatism; Orogen

1. Introduction

Studies of convergent margins worldwide have shown

that deformation patterns and the mechanical behavior of

continental crust vary according to crustal level and tectonic

setting (e.g. Sisson and Pavlis, 1993; Axen et al., 1998;

Klepeis and Crawford, 1999; Miller and Paterson, 2001;

Karlstrom and Williams, 2002; Teyssier et al., 2002).

Experimental data (Wilks and Carter, 1990; Rushmer, 1995;

Rutter and Neumann, 1995), numerical simulations (Harry

et al., 1995; Ellis et al., 1998; McKenzie et al., 2000), and

analytical models (Royden, 1996) indicate that lower crustal

strength and rheology especially affect how deformation is

partitioned vertically through the lithosphere during con-

vergence. These studies emphasize the critical role the

lower crust plays in linking the upper mantle with the upper

crust of orogens.

Despite this work, however, we still lack direct

information on the mechanisms by which deformation is

relayed vertically between different sections of the litho-

sphere, especially as physical and chemical conditions in the

lower crust change. Large, dipping shear zones that divide

the crust and upper mantle into different structural domains

have been observed or postulated in many orogenic belts

(Oldow et al., 1990; Harry et al., 1995; Willett, 1998;

McKenzie et al., 2000; Teyssier et al., 2002) but we do not

fully understand how deformation above, below, and within

these potentially transient features relate to one another or

affect orogenic evolution. This gap in knowledge arises

partly because orogens that allow direct observation of

processes at lower crustal levels and their relationship with

the upper crust are rare. In addition, the age and kinematic

0191-8141/$ - see front matter q 2003 Published by Elsevier Ltd.

doi:10.1016/j.jsg.2003.08.012

Journal of Structural Geology xx (0000) xxx–xxx

www.elsevier.com/locate/jsg

* Corresponding author. Tel.: þ1-802-656-0246; fax: þ1-802-656-0045.

E-mail address: [email protected] (K.A. Klepeis).

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

ARTICLE IN PRESS

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

112

Page 2: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

significance of lower crustal fabrics identified in geophy-

sical studies (e.g. Warner, 1990; Mayer et al., 1997; Nemes

et al., 1997) commonly are difficult to confirm.

In this paper, we show how displacements were

transferred vertically from lower to upper crustal levels of

an ancient orogen by reconstructing pieces of a composite

crustal column now exposed in Fiordland and Westland

(Fig. 1). This approach is possible because of the exposure

of an Early Cretaceous mid–lower crustal section in

Fiordland (Fig. 1; 25–50 km paleodepths) and its originally

contiguous mid–upper crust in Westland (Fig. 1; 8–27 km

paleodepths). The Alpine Fault now separates rocks of the

Fiordland belt from those of similar Early Cretaceous and

older affinity in Westland. Excellent pre-Cenozoic markers,

including the western margin of the Median Batholith (Fig.

1, inset), indicate that ,460 km of offset have accumulated

along the Alpine Fault (Wellman, 1953; Molnar et al., 1999;

Sutherland et al., 2000). Once restored to their pre-Cenozoic

configuration, the Fiordland and Westland regions form

parts of the same orogenic belt (Fig. 1; Oliver, 1990;

Tulloch and Challis, 2000). Mid–late Cretaceous extension

exhumed much of the lower crustal parts of the belt in

Fiordland as parts of the upper plate (including Westland)

slid off to the SW and NE (Gibson et al., 1988; Tulloch and

Kimbrough, 1989; Gibson, 1990; Oliver, 1990). By

,90 Ma, the Fiordland rocks had cooled to ,400 8C and

were in the upper 10 km of the crust (Mattinson et al., 1986;

Nathan et al., 2000; Claypool et al., 2002). The results of

this differential exhumation and offset allowed us to

compare processes and events from ,126–90 Ma in

Fiordland with those that occurred during the same time

interval at upper crustal levels in Westland.

We present structural, metamorphic and geochronologic

data that reveal the evolution of shear zones that separate the

middle and lower crustal section into distinctive structural

domains. We compare these features to structural patterns

and events preserved in the mid–upper crust and describe

how strain was partitioned within the orogen during a

transition from lithospheric contraction to extension. The

data indicate that strong physical and kinematic links were

established between the different layers of the lithosphere

only a few (,3–4 Ma) million years after emplacement of a

major batholith in the lower crust. The results provide direct

physical evidence of transient vertical decoupling followed

by coupling between the upper and lower crust during the

period ,126–105 Ma. We discuss the controls on coupling

and decoupling processes and explain why the mechanical

behavior of the Fiordland–Westland orogen may differ

from other orogens that experienced larger degrees of partial

melting and pluton emplacement in the deep crust.

2. The Fiordland–Westland orogen

The Fiordland–Westland orogen (Fig. 1) records a

history of magmatism, metamorphism and deformation

that accompanied the development of an early Mesozoic arc

along the margin of Gondwana. A Western Belt (Fig. 1),

representing the ancient continental margin, contains

Paleozoic terranes that preserve a record of mostly

,380–300 Ma pluton emplacement, low- to high-grade

metamorphism, and convergence (Landis and Coombs,

1967; Bishop et al., 1985; Cooper and Tulloch, 1992; Muir

et al., 1996; Ireland and Gibson, 1998). An Eastern Belt

(Fig. 1) contains plutons and volcano-sedimentary terranes

that originally formed outboard of the margin during the

early Mesozoic (Mattinson et al., 1986; McCulloch et al.,

1987; Tulloch and Kimbrough, 2003). Between these two

provinces (Fig. 1) is a linear, N- and NE-trending belt of

early Mesozoic plutonic, volcanic and sedimentary rock

called the Median Tectonic Zone (Kimbrough et al., 1994;

Muir et al., 1994) or the Median Batholith (Mortimer,

1999a,b).

The Median Batholith contains several compositionally

distinctive plutonic suites. On the outboard (east) side of the

Gondwana margin, the Median Suite of Mortimer and

Tulloch (1996) and the Darran Suite (Fig. 2) of Muir et al.

(1998) were emplaced into a Permo–Triassic accretionary

complex (Brook Street terrane, Fig. 1) mostly during the

interval 170–128 Ma (Mortimer, 1999). These suites are

dominated by diorite although gabbro and smaller granite

plutons also are common. On the continent side of the

Median Batholith is a younger belt of ,126–105 Ma

plutonic rock that includes the Separation Point and Rahu

suites (Fig. 1; Bradshaw, 1990; Kimbrough et al., 1994;

Muir et al., 1994; Mortimer et al., 1999a; Tulloch and

Kimbrough, 2003). At mid–upper crustal levels, now

exposed in Westland and easternmost Fiordland, rocks of

these latter two suites are dominated by tonalitic, grano-

dioritic and granitic compositions. The lower crustal levels

of this belt, exposed in Fiordland, are represented by the

dioritic–monzodioritic Western Fiordland Orthogneiss

(WFO; Figs. 1 and 2). Gabbro also is common in the WFO.

The emplacement of plutons of the Separation Point

Suite into both Eastern and Western belts at ,126 Ma

indicate that these two provinces were together at that time

Fig. 1. Present configuration (top of inset) and Cretaceous reconstruction (bottom of inset and main diagram) of western New Zealand after Tulloch and Challis

(2000). Geologic relationships are from Wood (1972), Oliver and Coggon (1979), Bradshaw (1989), Daczko et al. (2002a) and Klepeis and Clarke (2003).

Abbreviations show key locations or features: MB—Median Batholith, SP—Separation Point, L—Largs Terrane; MS—Milford Sound; GS—George Sound;

CS—Caswell Sound; CHS—Charles Sound; DS—Doubtful Sound; LTA—Lake Te Anau; LM—Lake Manapouri. Metamorphic pressures from Fiordland

represent the peak of Early Cretaceous metamorphism at ,120 Ma and are from Bradshaw (1985, 1989a,b), Brown (1996), Klepeis et al. (1999), Clarke et al.

(2000) and Daczko et al. (2001a,b, 2002a,b). See text for discussion. Pressures from Westland show shallower early–mid-Cretaceous (125–105 Ma) pluton

emplacement depths (after Tulloch and Challis, 2000). Metamorphic and structural data from Fiordland show a south-tilted lower crustal section (b).

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx2

ARTICLE IN PRESS

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

220

221

222

223

224

Page 3: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 3

ARTICLE IN PRESS

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

Page 4: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

(Williams and Harper, 1978; Mortimer et al., 1999a,b).

Hollis et al. (2003) obtained ages that suggest this

amalgamation occurred as early as ,136 Ma and certainly

by ,129 Ma. Tulloch and Kimbrough (2003) determined

that differences in composition and age reflect a configur-

ation where mantle-derived plutons of the outboard belt

were underthrust beneath Gondwana where they partially

melted at high pressures producing magma of the inboard

belt. Daczko et al. (2001a, 2002a) and Klepeis and Clarke

(2003) describe the contraction that accompanied this

amalgamation.

By ,108–105 Ma, regional extension affected parts of

Fiordland and Westland (Bradshaw, 1989; Tulloch and

Kimbrough, 1989; Gibson and Ireland, 1995; Spell et al.,

2000). Extensional metamorphic core complexes in the

Paparoa and Victoria ranges (Fig. 1) formed in the mid–

upper crust beginning at this time. Emplacement of the

,110 Ma Hohonu granitoids (Fig. 1; Waight et al., 1998)

Fig. 2. Structural map of Fiordland. Only main lithologic divisions are shown (see Turnbull (2000) and Klepeis and Clarke (2003) for details). Bold black lines

show boundaries of major shear zones. Structural measurements are from Bradshaw (1985, 1990), Blattner (1991), Klepeis et al. (1999), Daczko et al. (2002a),

Claypool et al. (2002), Turnbull (2000) and Klepeis and Clarke (2003). Foliation trajectories (thin black lines) show interpolation of structural trends. Plotted

U–Pb dates are from Hollis et al. (2003; white boxes) and this study (black boxes). Abbreviations show site localities: Mt. Daniel (MD), Mt. Edgar (ME),

Camp Oven Creek (CO), the Pembroke Valley (P), Mt. Ada (MA), Selwyn Creek (SC), Mt. Kepka (MK).

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx4

ARTICLE IN PRESS

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

Page 5: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

and the youngest plutons of the Separation Point Suite at

,105 Ma (Tulloch and Kimbrough, 2003) may have

overlapped with the transition to extension. The Doubtful

Sound Shear Zone (Gibson et al., 1988; Gibson and Ireland,

1995) and the Anita Shear Zone (Hill, 1995; Klepeis et al.,

1999) record decompression and exhumation after ,108–

105 Ma (Fig. 1).

During the Cenozoic changes in relative motions among

the Pacific, Australian, and Antarctic plates led to the

development of the modern Pacific–Australian plate

boundary by ,25 Ma (Cooper et al., 1987; Sutherland,

1995; Lamarche et al., 1997). Approximately 70–75% of

current motion arising from the oblique convergence

between the Australian and Pacific plates is accommodated

by slip along the Alpine Fault (Sutherland et al., 2000;

Norris and Cooper, 2001). Claypool et al. (2002) review the

effects of late Cenozoic faulting and exhumation (#6 km)

on the structure of northern Fiordland.

2.1. Early Cretaceous crustal thickening and magmatism in

the mid–upper crust

In Westland, plutons of the ,126–105 Ma Separation

Point Suite (Fig. 1) record Early Cretaceous emplacement

depths of 8–27 km, with the greatest depths (17–27 km)

occurring on the western side of the Median Batholith

(Tulloch and Challis, 2000). The youngest plutons also

occur on this western side (Tulloch, 1979; Harrison and

McDougall, 1980; Kimbrough et al., 1994; Muir et al.,

1994). Early Cretaceous deformation was distributed across

a 50–75-km-wide zone in Westland. Plutons localized some

of this deformation. Ductile shear zones, including the

Wainui Shear Zone (Fig. 1), formed on the western side of

the Separation Point Suite, including in its amphibolite

facies contact aureole (Grindley, 1980; Bradshaw, 1993;

Tulloch and Challis, 2000). Paleozoic fault zones also

influenced Early Cretaceous deformation. A record of

multiple displacements in many fault zones (Cooper,

1979) probably reflects some Mesozoic reactivation.

Crustal thickening by magma addition in the Eastern Belt

is indicated by a peak in subduction-related magmatism at

,140 Ma (Tulloch and Kimbrough, 2003). A major thrust

fault on Stewart Island (Fig. 1) formed between the outboard

and inboard belts during ,125–105 Ma magmatism

(Allibone and Tulloch, 1997; Tulloch and Kimbrough,

2003). Southeast of Fiordland, the Early Cretaceous Largs

Terrane (L, Fig. 1; Williams, 1978; Mortimer et al., 1999b)

is folded and thrust over the Brook Street Terrane

(Bradshaw, 1989). In the Victoria Range (Fig. 1), argon

thermochronology (Spell et al., 2000) and metamorphic data

(Tulloch, 1979) suggest that thickening and burial of the

Western Belt to mid-crustal levels initiated by ,130–

120 Ma and occurred prior to ,105 Ma (Tulloch and

Challis, 2000; Tulloch and Kimbrough, 2003).

2.2. Early Cretaceous crustal thickening, magmatism and

partial melting in the mid–lower crust

In Fiordland, the WFO was emplaced into a mid–lower

crust composed of Paleozoic metasediment and orthogneiss,

and early Mesozoic intrusive rock (Hollis et al., 2003).

Published ages from the WFO range from ,126 to

,116 Ma (Mattinson et al., 1986; Muir et al., 1998). New

high-precision data reported by Tulloch and Kimbrough

(2003) suggest that the main dioritic phase of the WFO was

emplaced during the interval 126–124 ^ 2 Ma. Hollis et al.

(2004) reports zircon ion probe data from Mt. Daniel and

George Sound that suggest that the WFO crystallized at

121.8 ^ 1.7 and 120 ^ 2.6 Ma, respectively, at these

localities (Hollis et al., 2004). Part of the WFO near

Doubtful Sound may be as young as ,116 Ma (Hollis et al.,

2004). Tulloch and Kimbrough (2003) also report a

116.6 ^ 1.2 Ma age from Wet Jacket Arm south of

Doubtful Sound that supports this interpretation. Alterna-

tively, the young zircon may reflect recrystallization.

Bradshaw (1989a, 1990) and Bradshaw and Kimbrough

(1989) used metamorphic P–T paths to infer an up-pressure

metamorphic history for the WFO following its emplace-

ment. Clarke et al. (2000) confirmed an up-pressure loading

history from P , 8 to 14–16 kbar for rocks of the Arthur

River Complex on the basis of metamorphic mineral

assemblages. The Arthur River Complex (Fig. 2) is

composed of Paleozoic paragneiss and both Paleozoic and

Mesozoic orthogneiss that lie below the WFO (Tulloch et al.,

2000; Hollis et al., 2003).

Oliver (1990) and Brown (1996) inferred that the cause

of burial involved magma loading during and following

emplacement of the WFO. Muir et al. (1995, 1998)

suggested that part of the Median Batholith was thrust to

lower crustal depths beneath western Fiordland and melted

to produce the WFO. This interpretation is supported by

geochemical data reported by Tulloch and Kimbrough

(2003). Daczko et al. (2002a) showed that some tectonic

burial was accomplished by the stacking of thrust sheets

above the WFO. Klepeis and Clarke (2003) showed that

lower crustal thickening involved displacements on both

steep and gently dipping shear zones.

Granulite facies metamorphism in the lower crust

accompanied and followed WFO emplacement (Oliver,

1977; Bradshaw, 1985, 1989a,b; Gibson and Ireland, 1995;

Clarke et al., 2000; Daczko et al., 2001b). Thin meta-

morphic overgrowths on Paleozoic and Mesozoic zircon

yield an average age of ,120 Ma in northernmost Fiordland

(Tulloch et al., 2000; Hollis et al., 2003). This age is

interpreted to represent the peak of metamorphism at

temperatures of 750 8C , T , 850 8C. Pressures represent-

ing the peak of this metamorphism range from P ¼ 7–

9 kbar in the shallowest part of the section to P ¼ 11–

16 kbar at the deepest levels (data plotted in Fig. 1). Recent

work suggests that granulite facies metamorphism

accompanied the partial melting of metadiorite during and

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 5

ARTICLE IN PRESS

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

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

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

Page 6: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

after WFO emplacement and was promoted by the

mobilization of water-poor melt in fractures (Clarke et al.,

2000; Daczko et al., 2001b). Partial melting was controlled

by the decomposition of hornblende ^ clinozoisite (Anti-

gnano, 2002; Klepeis et al., 2003). Antignano (2002) used

experiments to show that positive volume changes associ-

ated with the reaction of hornblende ^ clinozoisite to

produce melt were sufficient to fracture matrix feldspar

and quartz. Klepeis and Clarke (2003) outline relationships

that suggest the fractures formed by high fluid pore pressure

in melt pockets and propagation during the fluid-absent

melting of mafic lower crust.

Following peak metamorphism at the granulite facies,

the batholith and its lower crustal host rock hydrated and

cooled (Daczko et al., 2002b). Quantitative cation mapping

and thermodynamic modeling of kyanite- and paragonite-

bearing assemblage by Daczko et al. (2002c) suggested

isobaric cooling of the Arthur River Complex to conditions

of T ¼ 650–700 8C and P ¼ 11–15 kbar.

3. Structure of the middle and lower crustal sections

Continuous exposure along Fiordland’s waterways

combined with ridge sections above the fjords were used

to construct serial cross-sections from Milford Sound to

Caswell Sound (Figs. 1b, 3 and 4). This approach allowed

determination of the three-dimensional structure of northern

Fiordland (Figs. 2 and 5). The results show a tilted middle–

lower crustal section (,80 km horizontal distance) on the

western side of Fiordland that includes both the upper and

lower contacts of the WFO. The uppermost contact of this

batholith is well exposed at Caswell Sound; the lowermost

contact is well exposed at Mt. Daniel (Figs. 1b and 2). These

contacts dip variably to the S and W and are deformed by

shear zones of variable thickness and geometry. The shear

zones at these two localities formed at different paleodepths

and divide the crustal section into distinctive structural

domains.

The geometry of the titled section, with Caswell and

George sounds exposing the middle crust (25–30 km depth)

and areas north of Mt. Daniel exposing the lower crust (45–

50 km depth) is consistent with variations in Early

Cretaceous paleodepths determined using published ther-

mobarometric data (Fig. 1). Pressures reflecting the peak of

Early Cretaceous (,120 Ma) metamorphism at Caswell and

George sounds have a range of P ¼ 7–9 kbar and occur

within the WFO and its contact aureole (Daczko et al.,

2002a). Pressures reflecting peak metamorphic conditions

during and following emplacement of the WFO at Mt.

Daniel and Milford Sound are P ¼ 10–13 kbar (Bradshaw,

1985, 1989a,b) and P ¼ 12–16 kbar (Clarke et al., 2000),

respectively. South of Charles Sound the dip of the WFO-

country rock contact changes to the NE, subparallel to the

Doubtful Sound Shear Zone (Fig. 1b). Peak pressures in the

WFO at Charles Sound are recorded at P ¼ 8–10 kbar

(Bradshaw, 1985; Brown, 1996). South of Charles Sound

paleodepths increase. In the footwall of the Doubtful Sound

Shear Zone, pressures reflecting the peak of Early Cretac-

eous metamorphism are P ¼ 11–12.5 kbar (Gibson and

Ireland, 1995). The tilted depth section between Caswell

and Milford sounds thus lies in the hanging wall of the

Doubtful Sound Shear Zone (Fig. 1b).

The western boundary of the lower crustal section in

northern Fiordland coincides with the steep, upper amphi-

bolite facies Anita Shear Zone (Fig. 2). This shear zone cuts

all Early Cretaceous fabrics within the Arthur River

Complex and the WFO (Figs. 2 and 5) and separates the

high-grade rocks from Paleozoic rocks (including the

Greenland Group and Saint Anne Gneiss) to the west. The

western boundary of the lower crustal section coincides with

the eastern margin of a 10–15-km-wide shear zone named

the Indecision Creek Shear Zone by Klepeis and Clarke

(2003). This shear zone separates the granulites from

weakly metamorphosed ,150–130 Ma plutonic rock of

the Darran and Roxburgh suites to the east (Fig. 2). Most of

the Darran Suite is only weakly deformed except on its

western sides (see also Blattner and Graham, 2000). Highly

deformed rocks on this western side include the Selwyn

Creek Gneiss (SC, Fig. 2). Brittle faults deform the steep

margins of the Anita Shear Zone and the Darran Suite.

East of the Anita Shear Zone the lower crustal section

displays two structural domains (Figs. 2 and 5) that are

separated by a 4–5-km-wide transitional zone (Fig. 4a). The

western domain is composed of Paleozoic metasediment,

mafic dikes, and layered intrusions, including the ,129 Ma

Mt. Edgar Diorite (Fig. 2; Hollis et al., 2003). Igneous

layering and gneissic foliations in this domain dip

moderately to the S, W and SW (Fig. 2). The eastern

domain, collectively termed the Indecision Creek Complex

by Bradshaw (1990), is dominated by steep foliations of the

Indecision Creek Shear Zone (Figs. 2 and 4). This latter

domain contains gabbroic gneiss, dioritic gneiss and

deformed felsic dikes. In the transitional zone, the foliations

of the western domain are folded and transposed parallel to

the margins of the Indecision Creek Shear Zone. At George

Sound, a 4–10-km-wide zone of high strain, named the

George Sound Shear Zone by Klepeis and Clarke (2003),

displays foliations that are similar in geometry to those of

the Indecision Creek Shear Zone (Fig. 2). This shear zone

lies structurally below a mid-crustal fold–thrust belt at the

Caswell Sound (Fig. 2) that was first identified by Daczko

et al. (2002a).

4. Shear zone evolution in the middle and lower crust

4.1. The Caswell Sound fold–thrust belt

At Caswell Sound, garnet granulite and upper amphibo-

lite facies thrusts sole into a subhorizontal shear zone

located at and below the contact between the WFO and its

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx6

ARTICLE IN PRESS

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

Page 7: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

host rock (Fig. 3). Inside the WFO, the shear zone is defined

by a ,1-km-thick section of subhorizontal to gently dipping

upper amphibolite facies foliations (Fig. 3d and e). Country

rock is composed of calcsilicate gneiss, marble, and

metapsammitic schist that probably form part of the

Paleozoic–Triassic sequences of Gondwana (Bradshaw

and Kimbrough, 1991; Hollis et al., 2004). At the east end

of the sound, imbricated, W-dipping thrust splays formed

within the contact aureole of the WFO and locally cut across

it (Fig. 3). Tight to isoclinal, S-plunging folds between

thrusts deform a penetrative gneissic foliation (S1) in

metasedimentary country rock. To the west of the imbricate

series the fold geometry changes from tight and overturned

to open and upright, reflecting a decrease in strain intensity

above and away from the WFO margin and the basal shear

zone. A steep upper amphibolite facies foliation parallels

the axial planes of the folds (Fig. 3b). Farther west

(,12 km), the folds gradually tighten and overturn to the

east above an E-dipping thrust fault. This E-dipping thrust

separates the high-grade rocks of the thrust belt to the east

from the weakly deformed McKerr monzodiorite to the west

(Fig. 3).

Flattened clusters of coarse hornblende, clinozoisite and

garnet in a feldspar matrix define thrust plane foliations

inside the WFO. Inside the WFO and within 500 m of its

uppermost contact, feldspar in the thrust zones was

dynamically recrystallized along grain boundaries. In

contrast, greater than 500 m above the contact, feldspar

behaved in a brittle manner during deformation in thrust

zones although mylonitic textures also are common. Thrust

faults greater than 500 m and up to 2.5 km from the contact

with the WFO are defined by aligned chlorite, muscovite,

quartz, feldspar, clinozoisite and amphibole. Daczko et al.

(2002a) described these variations in mineral assemblage in

detail and showed that they reflect a temperature gradient of

T ¼ 700–800 8C within the (500 m thick) contact aureole

and T ¼ 550–600 8C outside of it. These strong links

among increasing metamorphic grade, the recrystallization

of feldspar, and proximity to the WFO suggests that the

thrusts were preferentially partitioned into crust that was

thermally softened by the emplacement of the batholith. The

subsolidus character of the thrust fabrics also indicates that

contraction outlasted emplacement and crystallization of the

WFO.

Structurally below the Caswell thrusts, the WFO is

composed of layered dioritic intrusions and folded rafts of

upper amphibolite facies country rock (Fig. 3). At George

Sound (Fig. 2) the rafts contain a folded layer-parallel

foliation (S1) that is cut by the dioritic intrusions and

metasedimentary rock is migmatitic within 500 m of the

Fig. 3. Composite NW–SE profile of region between Charles Sound and George Sound. See Fig. 2 for section locations. Light shaded regions in B–B0 are

metasedimentary (Paleozoic–Mesozoic) country rock. Shaded and black units in D–D0 are felsic and mafic dikes, respectively. Thin, solid, black lines are

lithologic layering, dashed lines are foliation trends, and bold, black lines are shear zones. Equal area stereoplots (a–h) show poles to foliations, lineations and

fold axes. All data were measured by the authors and Daczko et al. (2002a) except those in plot e, which are from Bradshaw (1985). Locations of samples

(995a, 9928) for U–Pb analyses are shown. Sample 995a is from the Mckerr monzodiorite.

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 7

ARTICLE IN PRESS

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

707

708

709

710

711

712

713

714

715

716

717

718

719

720

721

722

723

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

746

747

748

749

750

751

752

753

754

755

756

757

758

759

760

761

762

763

764

765

766

767

768

769

770

771

772

773

774

775

776

777

778

779

780

781

782

783

784

Page 8: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx8

ARTICLE IN PRESS

785

786

787

788

789

790

791

792

793

794

795

796

797

798

799

800

801

802

803

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

822

823

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

849

850

851

852

853

854

855

856

857

858

859

860

861

862

863

864

865

866

867

868

869

870

871

872

873

874

875

876

877

878

879

880

881

882

883

884

885

886

887

888

889

890

891

892

893

894

895

896

Page 9: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

WFO margin (Fig. 3). The preservation of this migmatite

supports our interpretation that thrust zones were preferen-

tially partitioned into a similar zone that was weakened by

melt and heat at Caswell Sound. Foliations inside the diorite

form two dominant structural trends. The first trend includes

highly variable magmatic flow foliations (SWFO) defined by

the planar alignment of clinopyroxene, hornblende, tabular

plagioclase and other minerals. The second trend includes

steeply to moderately dipping subsolidus foliations (e.g.

Fig. 3f and g) that are heterogeneously developed within the

batholith. These latter foliations are best developed in the

George Sound Shear Zone (Figs. 2 and 3f) where they cut

the older synmagmatic SWFO foliations. Tight, upright, S-

plunging folds of dikes with steep axial planar foliations

display geometries that are similar to the upright folds and

steep foliations of the Caswell fold–thrust belt (compare

Fig. 3b and f). The George Sound Shear Zone flattens up

section and merges with the subhorizontal shear zone

beneath the Caswell fold–thrust belt (Fig. 3).

4.2. The Mount Daniel Shear Zone

A few hundred meters above the lowermost contact of

the WFO exposed at Mt. Daniel, the main phase of the WFO

is a coarse-grained diorite that contains pods of gabbro. A

hornblende cumulate layer occurs at the base of the diorite

(Fig. 6). These rocks display primary igneous layering and,

locally, a coarse-grained magmatic flow foliation (SWFO)

defined by aligned hornblende, clinozoisite, and plagioclase

Fig. 4. Cross-sections across the Mt. Daniel region (a) and north of Milford Sound (b). See Fig. 2 for locations. Equal area stereoplots show poles to foliations,

lineations and fold axes. Symbols the same as in Fig. 3. Plots i, ii, and iv in (a) include data from Bradshaw (1985). P and M are samples of orthogneiss dated by

Tulloch et al. (2000). PB3 in (b) is post-tectonic dike discussed in the text. # represent granulite facies fracture arrays.

Fig. 5. Composite block diagram of region between Caswell Sound (south end) and Milford Sound (north end) constructed using structural relationships shown

in Figs. 1a, 2, 3 and 4. Equal area stereoplots show poles to foliations, mineral lineations and fold axes. Plots (a)–(d) represent the western domain; plots (e)–

(h) represent the eastern domain. Symbols the same as in Fig. 3. Plot (c) is from Bradshaw (1985). WFO is the Western Fiordland Orthogneiss.

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 9

ARTICLE IN PRESS

897

898

899

900

901

902

903

904

905

906

907

908

909

910

911

912

913

914

915

916

917

918

919

920

921

922

923

924

925

926

927

928

929

930

931

932

933

934

935

936

937

938

939

940

941

942

943

944

945

946

947

948

949

950

951

952

953

954

955

956

957

958

959

960

961

962

963

964

965

966

967

968

969

970

971

972

973

974

975

976

977

978

979

980

981

982

983

984

985

986

987

988

989

990

991

992

993

994

995

996

997

998

999

1000

1001

1002

1003

1004

1005

1006

1007

1008

Page 10: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

that parallel a moderately (40–508) SW-dipping basal

contact (Fig. 4a, part iii).

Below the cumulate layer is a banded igneous complex

(Fig. 6) that preserves evidence of both suprasolidus and

subsolidus deformation. Sheeted tonalite intrusions in this

zone display undulate, diffuse contacts with slightly more

mafic tonalitic bodies that reflect injection into an

incompletely crystallized host (see also Fig. 5 of Klepeis

and Clarke, 2003). Discordant mafic dikes with sharp,

straight contacts cut some tonalite sheets and, in turn, are cut

by veins that originate from the surrounding tonalite host.

These mutually crosscutting relationships indicate the

simultaneous emplacement of tonalitic and more mafic

sheets.

The central and lower parts of the banded igneous

complex preserve relationships indicating the accumulation

of high strains while the rocks were partially molten.

Migmatitic tonalite and trondhjemite sheets are complexly

interfolded and stretched. Recumbent folds of dikes and

igneous layering are common. Interfolded, transposed

sheets are cut by less deformed sheets, indicating that

deformation coincided with the periodic emplacement of the

sheeted intrusions (see also Fig. 5d of Klepeis and Clarke,

2003). Tightly folded pegmatites display axial planes that

parallel the margins of tonalite layers. However, despite the

evidence of the high strains required to produce these tight

folds, many of the pegmatite dikes are not foliated. Coarse

biotite in the dikes forms radial or misaligned patterns and

plagioclase exhibits a clean, interlocking igneous texture

with little evidence of subsolidus recrystallization. These

features define a thin melt-enhanced shear zone at the base

of the WFO (Fig. 6).

Beneath the basal shear zone, metagabbroic–metadioritic

orthogneiss forms part of the Arthur River Complex (Fig. 6).

Greater than 100 m below the contact zone at Mt. Daniel,

the dominant rock type is a biotite-rich dioritic orthogneiss.

This latter unit is compositionally similar to the ,129 Ma

Mt. Edgar Diorite (Fig. 2). Between the biotite orthogneiss

and the basal shear zone is a 30–50-m-thick zone of

granulite facies, garnet-rich metagabbro. The metagabbro

contains little plagioclase and no biotite. These observations

suggest that the metagabbro represents a depleted part of the

Arthur River Complex that resulted from partial melting of a

biotite-rich dioritic host similar to that which occurs below

the metagabbro (see also Daczko et al., 2002b).

In addition to evidence of suprasolidus deformation, a

heterogeneous subsolidus overprint also occurs within the

basal shear zone. Thin (,10 m wide) upper amphibolite

facies shear zones locally cut the margins of some folded

intrusions and preserve evidence that plagioclase grain sizes

were reduced during dynamic recrystallization. The shear

zones parallel the axial planes of recumbent, SE-plunging

folds that are geometrically similar to the melt-enhanced

folds we described earlier. However, the former contain a

weak axial planar foliation (S2) defined by flattened

plagioclase and aligned biotite and hornblende (Fig. 4a,

parts ii and iv). The heterogeneous development of this

subsolidus foliation and a lack of transposition during

folding resulted in the preservation of the migmatitic

features at the base of the WFO. In addition, the subsolidus

folds deform the entire lower contact of the basal shear zone

and form part of a set that also occurs at the base of the Mt.

Edgar Diorite (Fig. 4a). We refer to these folds as F2

structures because they deform primary igneous layering

and older gneissic foliations (S1). The presence of S2 and the

subsolidus shear zones indicate that F2 folding either post-

dated or outlasted crystallization of the melt-enhanced shear

zone. We suggest that they are related to crustal thickening

Fig. 6. Vertical profile of the Mt. Daniel Shear Zone. Bold lines are upper amphibolite facies shear zones discussed in the text. WFO is Western Fiordland

Orthogneiss; ARC is Arthur River Complex. # represent granulite facies fracture arrays.

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx10

ARTICLE IN PRESS

1009

1010

1011

1012

1013

1014

1015

1016

1017

1018

1019

1020

1021

1022

1023

1024

1025

1026

1027

1028

1029

1030

1031

1032

1033

1034

1035

1036

1037

1038

1039

1040

1041

1042

1043

1044

1045

1046

1047

1048

1049

1050

1051

1052

1053

1054

1055

1056

1057

1058

1059

1060

1061

1062

1063

1064

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

1076

1077

1078

1079

1080

1081

1082

1083

1084

1085

1086

1087

1088

1089

1090

1091

1092

1093

1094

1095

1096

1097

1098

1099

1100

1101

1102

1103

1104

1105

1106

1107

1108

1109

1110

1111

1112

1113

1114

1115

1116

1117

1118

1119

1120

Page 11: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

during and slightly after emplacement and burial of the

WFO.

Another set of folds occurs in a narrow zone (,20 m

thick) between parallel minor shear zones in the Arthur

River Complex (Fig. 6). These folds are disharmonic,

mostly upright and plunge gently to moderately to the S

(Fig. 4a, part v). A spaced crenulation cleavage approxi-

mately parallels the axial planes of these folds. This same

style of folding occurs in the transitional zone between the

western and eastern domains (Figs. 2 and 4a). On the basis

of crosscutting relationships, we refer to these upright folds

as F3 structures (Fig. 4a, part v).

Finally, above and below the basal shear zone are arrays

of discordant veins and fractures filled with leucosome that

cut all ductile fabrics in the basal shear zone (# symbols in

Figs. 4 and 6). Narrow (6–7 cm) dehydration zones

containing garnet- and clinopyroxene-bearing assemblages

surround garnet-bearing leucosome in some fractures.

Inside these zones, hornblende–clinozoisite-bearing assem-

blages that define foliation in these rocks are replaced by

garnet, clinopyroxene, and rutile. Daczko et al. (2002b)

reported symplectic intergrowths of clinopyroxene and

kyanite and also of clinozoisite, quartz, kyanite and plagio-

clase that partially replace the older hornblende and

clinozoisite assemblage. These assemblages and reaction

textures record dehydration of the WFO and its host rocks at

the garnet granulite facies following cooling of the WFO

and its basal shear zone at Mt. Daniel (Daczko et al., 2002b).

None of the fractures or leucosome appears folded.

4.3. The Indecision Creek and George Sound Shear Zones

The Indecision Creek and George Sound Shear Zones

(Figs. 2–4) display vertical to steeply dipping, upper

amphibolite facies foliations (SSZ) that strike to the N, NE

and NNE and dip variably to the NW and SE. These steep

foliations parallel the axial planes of tight, S-plunging folds.

Steeply to gently plunging hornblende and plagioclase

mineral lineations occur on foliation planes (Figs. 2–4).

Rock fabrics are locally mylonitic and the folds and

intrusive contacts are transposed in the shear zone.

South of Milford Sound, the transitional domain (Fig. 4a)

preserves crosscutting relationships among shear zone

fabrics and other fabrics and fold sets of northern Fiordland.

In this zone, all igneous layering and gneissic foliations of

the western domain (S1, SWFO, S2) and F2 folds are tightly

folded into south-plunging F3 folds (Figs. 2 and 4a). From

W to E across this zone, F3 fold axial planes steepen to near

vertical, interlimb angles decrease, and the F3 folds

gradually are transposed parallel to the steep SSZ foliations

(Fig. 4a, parts v, vi and vii). The dominant SSZ foliation

parallels the axial planes of these tight F3 folds. These

changes define a 3–4-km-wide positive strain gradient that

increases from W to E into the central part of the shear zone.

North of Milford Sound, in the Pembroke Valley (P, Fig.

2), steep fabrics of the Indecision Creek Shear Zone envelop

a large lens of dioritic and gabbroic gneiss (Fig. 4b). This

locality preserves features that record the progressive

development of mineral assemblages and fabrics in the

Indecision Creek Shear Zone. One of the most spectacular

features is a lattice-like array of 3–5-cm-wide dehydration

zones that surround steep, orthogonal sets of leucosome-

filled fractures in dioritic and gabbroic orthogneiss. The

dehydration zones contain coronas of garnet and clinopyr-

oxene mantling hornblende. These reactions zones, like

those found at Mt. Daniel and elsewhere, record dehydration

at the garnet granulite facies. Descriptions of the petrology,

P–Tconditions (P ¼ 13–16 kbar, T . 750 8C), and origin

of these features are provided by Blattner (1976), Oliver

(1977), Bradshaw (1989a,b), Clarke et al. (2000) and

Daczko et al. (2001b).

The dehydration zones and vein sets form markers (Fig.

7a) that record the evolution of two sets of superposed shear

zones (Daczko et al., 2001a). The first set includes pairs of

thin, 1–3-m-wide sinistral and dextral shear zones (Fig. 7b).

The sinistral set is dominant and displays a steep mylonitic

foliation that strikes to the E and NE. The dextral set is

subordinate in size and abundance to the sinistral set. This

latter set dips gently to moderately to the SW. Both shear

zone sets contain gently plunging hornblende and clinozoi-

site mineral lineations. Superimposed on the sinistral and

dextral shear zone pairs is a younger set of vertically

stacked, layer-parallel shear zones that dip gently to the SE

(Fig. 4b, part i and Fig. 7c). Each of these shear zones

contains a 7–10-m-thick central zone where asymmetric

pods of coarse-grained gneiss is surrounded by thin (,1 m

thick) mylonitic to ultramylonitic shear bands. The asym-

metric pods form imbricated, antiformal stacks. The thin

shear bands locally dip steeply to the NW and SE and swing

into parallelism with the gently SE-dipping shear zones

located above and below them. Hornblende mineral

lineations on foliation planes plunge to the SE (Fig. 4b,

part i). The vertical spacing between parallel shear zones is

,50–100 m. The exact thickness of the stack is unknown.

East and west of the Pembroke Valley, the delicate

dehydration zones and the layer-parallel thrusts are mostly

transposed and recrystallized by the Indecision Creek Shear

Zone (Fig. 7d). This shear zones cuts the lower (eastern

side) contact of the WFO southeast of Mount Daniel (Figs.

2b, 4a and 5) and records retrogression of granulite facies

mineral assemblages to the upper amphibolite facies.

Southeast of the Indecision Creek Shear Zone, the George

Sound Shear Zone cuts across the central part of the WFO

(Figs. 2 and 3).

5. Kinematic relationships

The kinematics of deformation that occurred while the

WFO batholith was partially molten are recorded best in the

melt-enhanced basal shear zone exposed at Mt. Daniel.

Oblique foliations and the truncation of sheeted intrusions

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 11

ARTICLE IN PRESS

1121

1122

1123

1124

1125

1126

1127

1128

1129

1130

1131

1132

1133

1134

1135

1136

1137

1138

1139

1140

1141

1142

1143

1144

1145

1146

1147

1148

1149

1150

1151

1152

1153

1154

1155

1156

1157

1158

1159

1160

1161

1162

1163

1164

1165

1166

1167

1168

1169

1170

1171

1172

1173

1174

1175

1176

1177

1178

1179

1180

1181

1182

1183

1184

1185

1186

1187

1188

1189

1190

1191

1192

1193

1194

1195

1196

1197

1198

1199

1200

1201

1202

1203

1204

1205

1206

1207

1208

1209

1210

1211

1212

1213

1214

1215

1216

1217

1218

1219

1220

1221

1222

1223

1224

1225

1226

1227

1228

1229

1230

1231

1232

Page 12: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

by successive sheets indicate top-to-the-E and -NE, thrust-

style displacements parallel to W- and SW-plunging

mineral lineations (Fig. 4a, part iii). The plunges of the

lineations indicate a sinistral component to the deformation.

The subsolidus shear zones that parallel the axial planes of

F2 folds give identical oblique-thrust senses of shear.

Below the WFO, kinematic indicators in the steeply

dipping shear zone pairs of the Pembroke Valley (Fig. 7b)

include oblique foliations, micro-faulted garnet, and asym-

metric tails on feldspar porphyroclasts. These shear zones

record mostly NE–SW stretching parallel to the arc with a

component of sinistral displacement. Daczko et al. (2001a)

showed that they also record subhorizontal shortening at

high angles to the trend of the Median Batholith. The

Pembroke Thrust Zone (Fig. 7c) contains asymmetric

hornblende and clinozoisite fish, minor shear bands, and

asymmetric tails on feldspar clasts that record a top-to-the-

NW sense of shear. The style of displaced, asymmetric pods

that are stacked on top of one another also reflects a

component of layer-perpendicular thickening. Together, the

steep shear zone pairs and Pembroke Thrust Zone record

subhorizontal (layer-parallel) shortening normal to the

batholith, sinistral arc-parallel displacements, and vertical

(layer-perpendicular) thickening. This result is consistent

with the oblique-thrust style displacements recorded in the

melt-enhanced shear zone at Mt. Daniel, suggesting that this

style of deformation began during emplacement of the

WFO.

Above the WFO, the Caswell fold–thrust belt also

records arc-normal contraction and crustal thickening

following crystallization of the WFO. Lineation trends are

similar to those that characterized those in the basal shear

zone at Mt. Daniel (Fig. 5a and d). The spread of lineation

plunges on foliation planes (Fig. 3c) also suggests that the

imbricated thrusts record a component of sinistral displace-

ment. The conjugate style of W-dipping thrusts with an

E-dipping back thrust (Fig. 3) indicates compression

directions at high angles to the trend of the arc.

Below the Caswell Sound fold–thrust belt, the Indeci-

sion Creek Shear Zone records shortening at high angles to

the arc leading to the development of steep foliation planes.

This shortening is best illustrated by the progressive change

in fold geometry, including tightness and the steepening of

fold axial surfaces, from W to E across the transition zone.

Outcrop-scale sense of shear indicators, including hornble-

nde and clinozoisite fish, asymmetric tails of biotite and

hornblende around garnet porphyroblasts, asymmetric

boudinage, and minor shear zones, mostly occur in areas

of low–intermediate strain at the eastern and western edges

of the shear zone. The sense of shear in these areas is

dominantly sinistral parallel to gently and moderately

plunging mineral lineations.

In the central part of the Indecision Creek Shear Zone,

changes in orientation of hornblende lineations with

increasing strain provide additional kinematic information.

From W to E across the transition zone, the lineations

change from gently and moderately S-plunging to near

vertical and steeply plunging (compare Fig. 4a, parts vi and

vii). The migration of these mineral lineations toward the

dip line of the steep shear zone indicates stretching parallel

Fig. 7. Block diagrams showing the sequence of deformation recorded in structures at Pembroke Valley (modified from Fig. 3 of Clarke et al. (2000) and Fig. 13

of Daczko et al. (2001a)). (a) Garnet granulite facies fracture arrays cut S1. (b) Steep, sinistral–dextral shear zone pairs deform fracture arrays. (c) Gently SE-

dipping thrust zones (see also Fig. 4b) cut sinistral shear zones. (d) Steep upper amphibolite facies foliation (SSZ) envelops shear zones at Pembroke Valley.

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx12

ARTICLE IN PRESS

1233

1234

1235

1236

1237

1238

1239

1240

1241

1242

1243

1244

1245

1246

1247

1248

1249

1250

1251

1252

1253

1254

1255

1256

1257

1258

1259

1260

1261

1262

1263

1264

1265

1266

1267

1268

1269

1270

1271

1272

1273

1274

1275

1276

1277

1278

1279

1280

1281

1282

1283

1284

1285

1286

1287

1288

1289

1290

1291

1292

1293

1294

1295

1296

1297

1298

1299

1300

1301

1302

1303

1304

1305

1306

1307

1308

1309

1310

1311

1312

1313

1314

1315

1316

1317

1318

1319

1320

1321

1322

1323

1324

1325

1326

1327

1328

1329

1330

1331

1332

1333

1334

1335

1336

1337

1338

1339

1340

1341

1342

1343

1344

Page 13: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

to this direction (Lin et al., 1998; Jiang and Williams, 1998).

The reference frame provided by the gneissic layering of the

western domain and well-defined boundaries indicate that

the shear zone was thickening vertically during contraction.

The kinematic evolution of the George Sound Shear

Zone has not been studied in detail due to its remote locality.

However, the increase in fold tightness and rotation of

hornblende lineations to down dip with increasing strain

suggest that it also records arc-normal contraction, near

vertical stretching. Minor shear zones show both dextral and

sinistral displacements. Given that these styles are similar to

those of the Indecision Creek Shear Zone (Fig. 5) we

suggest that the kinematic evolution of the two shear zones

is similar also.

6. The ages of lower crustal deformation, magmatism,

and metamorphism

Published dates and four new age determinations were

used to estimate the age of high-grade fabrics and intrusions.

At Caswell Sound, U–Pb spot analyses on single zircons

from the McKerr monzodiorite (sample 995a, Figs. 2 and 3)

and from a dioritic dike within the zone of imbricated

thrusts (sample 9928, Figs. 2 and 3) allowed us to place a

lower limit on the age of deformation within eastern and

western parts of the fold–thrust belt. Twenty analyses were

conducted on the cores of zircon from sample 995a using a

beam diameter of 50 microns (analytical procedures are

described in Appendix A). Nineteen grains yielded analyses

that are apparently of the same 206Pb/238U age, and one

additional grain is discordant due to inheritance (Fig. 8a).

The weighted mean of these analyses yields an interpreted

crystallization age of 116.8 ^ 3.7 Ma at the 2s level (Fig.

8a and b). Twenty-two analyses were conducted on the

cores of zircon grains from sample 9928 using a slot

diameter of 50 microns. Twenty grains yield analyses that

are apparently of the same age, and two additional analyses

are apparently discordant due to a slight amount of

inheritance. The final 206Pb/238U age is 118.7 ^ 3.8 Ma

(2s level; Fig. 8c and d). These data suggest that the

Caswell Sound fold–thrust belt evolved during and after the

interval 122.5–113 Ma.

Two samples of syntectonic dikes from within the

Indecision Creek Shear Zone (Ada2 and 0221K, Fig. 2)

provided an approximate lower age limit of deformation in

the shear zone. Forty-five analyses were conducted on

zircon cores from sample Ada2 using a laser beam diameter

of 25 microns. This dike from Mt. Ada (MA, Fig. 2) cuts the

steep foliation (SSZ) of the shear zone and also is folded

within it. These analyses yield two clusters of ages (Fig. 8e

and f). The older age of 204.0 ^ 6.1 Ma from zircon cores is

interpreted to record igneous crystallization. The rim ages of

115.7 ^ 3.8 Ma are interpreted to record the growth of

metamorphic zircon. The spread of the rim ages indicate

that shear zone deformation continued through the interval

119.5–112 Ma.

Fifty analyses were conducted on zircon grains from

sample 0221K using a beam diameter of 25 microns. This

dike (from near Mt. Kepka, Fig. 2) also cut steep foliation

planes (SSZ) in the Indecision Creek Shear Zone and is

folded within it. Most analyses were conducted on core

areas of the zircon grains, with a smaller number of analyses

on the rims (tips) of the grains. The rim analyses generally

yield ages that are younger than the core ages. The

occurrence of two distinct clusters of ages (Fig. 8g and h)

suggests that the grains record two phases of zircon growth:

an older phase at 365.3 ^ 11.4 Ma that reflects crystal-

lization of the dike, and a younger phase at 129.5 ^ 4.2 Ma

that reflects the growth of metamorphic zircon. The young

rim ages obtained from the 0221K and Ada2 samples are in

agreement with crosscutting relationships indicating that

deformation in the Indecision Creek Shear Zone outlasted

emplacement of the WFO.

These new ages are compatible with other published ages

(Fig. 8i). Tulloch et al. (2000) identified Paleozoic

(355 ^ 10 Ma) oscillatory-zoned cores and Early Cretac-

eous (134 ^ 2 Ma) sector-zoned cores from the Arthur

River Complex. Both these core types displayed thin low-U

rims that yield an average age of ,120 Ma but some with

ages as young as ,105 Ma (Tulloch et al., 2000). Similar

rim ages (Fig. 2) have been obtained from the deformed

western margin of the Darran Suite (Selwyn Gneiss of

Hollis et al., 2003). The Jurassic core age of sample Ada2 is

compatible with similar ages obtained from the Darran Suite

(Muir et al., 1998; Blattner and Graham, 2000). The rim age

of sample 0221K is similar to ,136–129 Ma crystallization

ages from intrusive rocks in the Arthur River Complex (Fig.

2; Hollis et al., 2003). Metamorphism leading to zircon

growth also could have accompanied emplacement of

intrusive rocks prior to the WFO (see also Tulloch et al.,

2000).

Zircon ages of ,82 Ma from a post-tectonic dike (PB3;

Fig 4b) indicate that ductile deformation in northernmost

Fiordland terminated in the Late Cretaceous (Hollis et al.,

2003). This age of a few million years younger than K–Ar

ages on hornblende (Nathan et al., 2000) and U–Pb dates on

apatite (Mattinson et al., 1986) indicate that the Arthur

River Complex had cooled to T ¼ 300–400 8C by ,90 Ma.

Near Doubtful Sound, Gibson and Ireland (1995) dated

thermal conditions of T . 800 8C at 107.5 ^ 2.8 Ma from a

sample of the WFO deformed by the Doubtful Sound Shear

Zone (sample D in Fig. 8i). This age could reflect the

recrystallization of zircon in the shear zone. However, the

chemistry and age of the zircon suggested to Gibson and

Ireland (1995) that it represents a new generation of zircon

growth during metamorphism. The age is consistent with

regional geologic relationships indicating that extension

began by ,108–105 Ma (Tulloch and Kimbrough, 1989,

2003; Gibson and Ireland, 1995; Spell et al., 2000). K–Ar

cooling ages of ,93 and ,77 Ma on amphibole and biotite,

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 13

ARTICLE IN PRESS

1345

1346

1347

1348

1349

1350

1351

1352

1353

1354

1355

1356

1357

1358

1359

1360

1361

1362

1363

1364

1365

1366

1367

1368

1369

1370

1371

1372

1373

1374

1375

1376

1377

1378

1379

1380

1381

1382

1383

1384

1385

1386

1387

1388

1389

1390

1391

1392

1393

1394

1395

1396

1397

1398

1399

1400

1401

1402

1403

1404

1405

1406

1407

1408

1409

1410

1411

1412

1413

1414

1415

1416

1417

1418

1419

1420

1421

1422

1423

1424

1425

1426

1427

1428

1429

1430

1431

1432

1433

1434

1435

1436

1437

1438

1439

1440

1441

1442

1443

1444

1445

1446

1447

1448

1449

1450

1451

1452

1453

1454

1455

1456

Page 14: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Fig. 8. U–Pb isotopic data from zircon collected using an inductively coupled plasma mass spectrometer (ICPMS). Plots (a), (c), (e) and (g) are concordia plots (ellipses shown with dashed lines were not used to

calculate mean ages). An explanation is provided in Appendix A. Plots (b), (d), (f) and (h) show error analyses and distribution of analyses (solid lines) used to calculate mean ages and errors (2s level). Locations

of samples are shown in Fig. 2. Part (i) shows a comparison with published ages from Mattinson et al. (1986), Muir et al. (1998), Tulloch et al. (2000) and Hollis et al. (2003). Ages representing early phases of the

Median Batholith are from both the Darran Suite and the Arthur River Complex (A). Samples P and M are from Tulloch et al. (2000) located in Fig. 4b. Post-tectonic dike is sample PB 3 reported by Hollis et al.

(2003) and shown in Fig. 4b. Sample D is a zircon age from the extensional Doubtful Sound Shear Zone (Gibson and Ireland, 1995).

SG

14

56

—3

0/9

/20

03

—1

4:5

7—

SF

OR

ST

ER

—8

24

96

—M

OD

EL

5–

br,ed

,cor

K.A

.K

lepeis

eta

l./

Jou

rna

lo

fS

tructu

ral

Geo

log

yxx

(00

00

)xxx–

xxx1

4

AR

TIC

LE

INP

RE

SS

14

57

14

58

14

59

14

60

14

61

14

62

14

63

14

64

14

65

14

66

14

67

14

68

14

69

14

70

14

71

14

72

14

73

14

74

14

75

14

76

14

77

14

78

14

79

14

80

14

81

14

82

14

83

14

84

14

85

14

86

14

87

14

88

14

89

14

90

14

91

14

92

14

93

14

94

14

95

14

96

14

97

14

98

14

99

15

00

15

01

15

02

15

03

15

04

15

05

15

06

15

07

15

08

15

09

15

10

15

11

15

12

15

13

15

14

15

15

15

16

15

17

15

18

15

19

15

20

15

21

15

22

15

23

15

24

15

25

15

26

15

27

15

28

15

29

15

30

15

31

15

32

15

33

15

34

15

35

15

36

15

37

15

38

15

39

15

40

15

41

15

42

15

43

15

44

15

45

15

46

15

47

15

48

15

49

15

50

15

51

15

52

15

53

15

54

15

55

15

56

15

57

15

58

15

59

15

60

15

61

15

62

15

63

15

64

15

65

15

66

15

67

15

68

Page 15: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

respectively, also support a late Cretaceous age for the

Doubtful Sound Shear Zone (Gibson et al., 1988).

7. Correlation of structures within Northern Fiordland

Crosscutting relationships, geochronology, similarities in

style and metamorphic grade, and the results of physically

tracing structures above and below the WFO allowed us to

correlate fabrics between Milford and Caswell Sounds. We

use these correlations to reconstruct the sequential evolution

of the section (Table 1, Figs. 9 and 10).

One especially useful marker unit is the WFO. The

regional extent of this batholith and its ,126–120 Ma age

allowed us to divide structures into groups that predated,

accompanied, and post-dated its emplacement. Structures

that predate emplacement occur in Paleozoic and early

Mesozoic host gneiss located above and below the WFO or

as xenoliths within it. Primary igneous layering and the

gneissic foliations (S1) in the Arthur River Complex and

Darran Suite are included in this group (Table 1). These

structures are locally cut by the WFO and mostly occur in

the western domain (Figs. 2 and 4a).

The second group of structures includes all magmatic

foliations (SWFO) that formed within the WFO during its

emplacement, including the Mt. Daniel Shear Zone (Table

1). These structures all exhibit evidence of deformation

while the batholith was still partially molten. The links

among increasing metamorphic grade, feldspar recrystalli-

zation, and proximity to the WFO also suggest that

deformation in the Caswell Sound fold–thrust belt began

during this stage. We include in this group the recumbent F2

folds and axial planar foliations (S2) that formed at the lower

contact of the WFO.

In northernmost Fiordland the melt-induced fracture

arrays and dehydration zones record the peak of granulite

facies metamorphism in the lower crust during or immedi-

ately after WFO emplacement. The development of these

arrays across large areas of the section (Pembroke Valley,

Mt. Daniel, George Sound, Doubtful Sound) suggests that

similar processes controlled their development. However,

these features may exhibit slightly different ages across the

section. Metamorphic rims on zircon suggest that in the

Milford Sound region this metamorphism mostly occurred

between ,123 and ,116 Ma with a clustering at ,120 Ma

(Tulloch et al., 2000; Hollis et al., 2003; this study). Near

Doubtful Sound a zircon age of 107.5 ^ 2.8 Ma may

indicate that granulite facies metamorphism there is

younger (Gibson and Ireland, 1995). Despite this age

range, we correlate these distinctive garnet granulite

reaction zones (GRZ, Table 1). The recrystallization of

these features in the Indecision Creek and George Sound

shear zones indicates that they formed prior to the shear

zones north of Caswell Sound.

The third group of fabrics includes the sinistral–dextral

shear zone pairs (Fig. 7b), the Pembroke thrust zone (Fig.

7c), and the steep fabrics of the Indecision and George

Sound Shear Zones (Fig. 7d). On the basis of structural

relationships exposed in the Pembroke Valley, Clarke et al.

(2000) and Daczko et al. (2001a) referred to the shear zone

Table 1

Time–space correlation of structures (dashed lines) from SW (left) to NE (right) within northern Fiordland. GRZ is garnet granulite reactions zones, S is

foliation where subscript refers to rock unit and generation: WFO is Western Fiordland orthogneiss

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 15

ARTICLE IN PRESS

1569

1570

1571

1572

1573

1574

1575

1576

1577

1578

1579

1580

1581

1582

1583

1584

1585

1586

1587

1588

1589

1590

1591

1592

1593

1594

1595

1596

1597

1598

1599

1600

1601

1602

1603

1604

1605

1606

1607

1608

1609

1610

1611

1612

1613

1614

1615

1616

1617

1618

1619

1620

1621

1622

1623

1624

1625

1626

1627

1628

1629

1630

1631

1632

1633

1634

1635

1636

1637

1638

1639

1640

1641

1642

1643

1644

1645

1646

1647

1648

1649

1650

1651

1652

1653

1654

1655

1656

1657

1658

1659

1660

1661

1662

1663

1664

1665

1666

1667

1668

1669

1670

1671

1672

1673

1674

1675

1676

1677

1678

1679

1680

Page 16: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

pairs and the ductile thrust fabrics as D3 and D4 structures,

respectively. However, the occurrence of these features in

the transitional zone of the Indecision Creek Shear Zone and

evidence of arc-normal contraction during each phase of

deformation suggest that they all reflect slightly different

stages of the same contractional event. We also include in

this group the tight F3 folds. This correlation is consistent

with crosscutting relationships indicating that all minor

shear zones and the F3 folds in the transitional zone deform

S1, S2, F2, SWFO and GRZ structures. These relationships

and the rim ages obtained from sample Ada2 (Figs. 6g and

8e) indicate that this deformation occurred after ,120 Ma

and outlasted the emplacement of a ,116 Ma dike.

South of George Sound, the George Sound Shear Zone

merges with the subhorizontal shear zone exposed at

Caswell and Charles Sounds (Figs. 3 and 5). On the basis

of crosscutting relationships with respect to the WFO both

the Caswell fold–thrust belt and the George Sound Shear

Zone developed following crystallization of the WFO and

before the onset of regional extension (Table 1). Hollis et al.

(2004) reports metamorphic rims on zircon from George

Sound (an average age of ,120 Ma with ages spread across

the interval ,138–106 Ma) that support this interpretation.

Together, the combined isotopic data from Tulloch et al.

(2000), Hollis et al. (2004) and from samples 995a, 9928

and Ada2 indicate that the Caswell fold–thrust belt and the

Fig. 9. (a) Interpretive cross-section showing the setting of the Fiordland–Westland orogen during the period ,126–105 Ma (modified from Klepeis et al.,

2003). (b) Block diagram showing correlations of structures in Fiordland and Westland. Diagram was constructed using pre-Cenozoic configuration of the

orogen and the relative structural position of the Fiordland cross-sections shown in Figs. 1b, 3, 4 and 5 (localities listed on right side of diagram). Section

representative of Westland is from the north shore of Westland (separation point). Paleodepths were calculated from data shown in Fig. 1. Symbols (#)

represent garnet granulite facies fracture arrays.

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx16

ARTICLE IN PRESS

1681

1682

1683

1684

1685

1686

1687

1688

1689

1690

1691

1692

1693

1694

1695

1696

1697

1698

1699

1700

1701

1702

1703

1704

1705

1706

1707

1708

1709

1710

1711

1712

1713

1714

1715

1716

1717

1718

1719

1720

1721

1722

1723

1724

1725

1726

1727

1728

1729

1730

1731

1732

1733

1734

1735

1736

1737

1738

1739

1740

1741

1742

1743

1744

1745

1746

1747

1748

1749

1750

1751

1752

1753

1754

1755

1756

1757

1758

1759

1760

1761

1762

1763

1764

1765

1766

1767

1768

1769

1770

1771

1772

1773

1774

1775

1776

1777

1778

1779

1780

1781

1782

1783

1784

1785

1786

1787

1788

1789

1790

1791

1792

Page 17: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Indecision Creek and George Sound Shear Zones all

evolved during the interval ,122.5–105 Ma.

The last group of structures includes the Anita Shear

Zone (ASZ, Figs. 5 and 9b; Table 2) and the Doubtful Sound

Shear Zone (Table 1). Both shear zones record the

exhumation (Fig. 9b) of the lower crust after ,108–

105 Ma (Mattinson et al., 1986; Gibson et al., 1988; Nathan

et al., 2000; Claypool et al., 2002). The Anita Shear Zone

also preserves fabrics (including ASZ1 at P ¼ 12 kbar and

ASZ2 at P ¼ 8 kbar, Fig. 10) that record Late Cretaceous–

Cenozoic decompression (Klepeis et al., 1999).

8. Interpretation of vertical coupling and decoupling

within the crustal column

The results of our analyses show that deformation

accompanying the ,126–120 Ma emplacement of the

WFO was localized within and at its upper and lower

contacts (Figs. 1b, 9b and 10c). In contrast, deformation a

few kilometers below the WFO during this period was weak

to nonexistent (Table 1). The two layer-parallel shear zones

at Mt. Daniel and Caswell Sound both separate areas of

melt-enhanced deformation inside the WFO (SWFO) from

areas outside it where older structures (S1 and igneous

layering) and mineral assemblages are well preserved. In

addition, structural relationships above and below the Mt.

Daniel Shear Zone at the base of the WFO are discordant.

This discordance and evidence that the structures above and

below the basal shear zone formed at different times and

under different physical conditions suggest that the crust

above and below the lower contact of the WFO was

decoupled during emplacement of the batholith.

As the batholith cooled and crystallized, the role of the

layer-parallel shear zones began to change. The develop-

ment of melt-induced fracture arrays and dehydration zones

(Fig. 10c) that cut the lower contact of the WFO mark the

abandonment of the melt-enhanced shear zone at Mt.

Fig. 10. Pressure–temperature–time path (a) for the lower crust exposed between Caswell and Milford sounds (modified from Daczko et al., 2002c). P–T data

are compiled from Bradshaw (1985, 1989a,b), Gibson and Ireland (1995), Klepeis et al. (1999), Clarke et al. (2000) and Daczko et al. (2001a,b, 2002a,b,c).

Shaded regions incorporate errors for specific metamorphic mineral assemblages discussed by Daczko et al. (2002c). Block diagrams ((b)–(e)) illustrate four

stages in the tectonic evolution of the Fiordland–Westland belt. Patterns are the same as those in Fig. 1a. Parts (c)–(e) are modified from Klepeis et al. (2003).

Abbreviations are: ASZ—Anita Shear Zone, ARC—Arthur River Complex, WFO—Western Fiordland Orthogneiss, ME—Mt. Edgar, SC—Selwyn Creek,

C—Caswell Sound, G—George Sound, MD—Mt. Daniel, DSSZ—Doubtful Sound Shear Zone, P—Paparoa Range, V—Victoria Range.

SG 1456—30/9/2003—14:57—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 17

ARTICLE IN PRESS

1793

1794

1795

1796

1797

1798

1799

1800

1801

1802

1803

1804

1805

1806

1807

1808

1809

1810

1811

1812

1813

1814

1815

1816

1817

1818

1819

1820

1821

1822

1823

1824

1825

1826

1827

1828

1829

1830

1831

1832

1833

1834

1835

1836

1837

1838

1839

1840

1841

1842

1843

1844

1845

1846

1847

1848

1849

1850

1851

1852

1853

1854

1855

1856

1857

1858

1859

1860

1861

1862

1863

1864

1865

1866

1867

1868

1869

1870

1871

1872

1873

1874

1875

1876

1877

1878

1879

1880

1881

1882

1883

1884

1885

1886

1887

1888

1889

1890

1891

1892

1893

1894

1895

1896

1897

1898

1899

1900

1901

1902

1903

1904

Page 18: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Table 2

U–Pb isotopic data

Sample U

(ppm)

206Pb/204Pb

ratio

206Pb/238U ^ ratio

(%)

207Pb/235U ^ ratio

(%)

206Pb/207P ^ ratio

(%)

Error

correction

206Pb/238U ^ age

(Ma)

207Pb/235U ^ age

(Ma)

206Pb/207Pb ^ age

(Ma)

207Pb correction ^ age

(Ma)

28-1c 40 NA 0.0205 ^ 0.5 0.1308 ^ 5.8 21.579 ^ 5.8 0.09 130.6 ^ 0.7 124.8 ^ 6.8 15 ^ 70 131.0 ^ 0.8

28-2c 33 NA 0.0195 ^ 0.7 0.1491 ^ 4.7 18.022 ^ 4.6 0.16 124.4 ^ 0.9 141.1 ^ 6.1 432 ^ 51 123.3 ^ 1.0

28-3c 42 NA 0.0188 ^ 0.6 0.1113 ^ 6.1 23.238 ^ 6.1 0.09 119.8 ^ 0.7 107.1 ^ 6.2 2 166 ^ 76 120.6 ^ 0.8

28-4c 39 NA 0.0186 ^ 0.8 0.1282 ^ 4.5 19.962 ^ 4.4 0.17 118.6 ^ 0.9 122.5 ^ 5.2 199 ^ 51 118.3 ^ 1.0

28-5c 61 NA 0.0193 ^ 1.1 0.1293 ^ 3.9 20.555 ^ 3.7 0.29 123.1 ^ 1.3 123.4 ^ 4.5 131 ^ 43 123.0 ^ 1.4

28-6c 65 NA 0.0180 ^ 0.7 0.1324 ^ 5.0 18.723 ^ 5.0 0.14 114.9 ^ 0.8 126.3 ^ 5.9 346 ^ 56 114.2 ^ 0.9

28-7c 88 NA 0.0194 ^ 0.5 0.1615 ^ 1.6 16.558 ^ 1.5 0.30 123.8 ^ 0.6 152.0 ^ 2.2 618 ^ 16 122.0 ^ 0.6

28-8c 99 NA 0.0190 ^ 0.4 0.1491 ^ 1.6 17.582 ^ 1.5 0.26 121.4 ^ 0.5 141.1 ^ 2.1 487 ^ 17 120.1 ^ 0.5

28-9c 134 NA 0.0188 ^ 0.4 0.1415 ^ 1.4 18.321 ^ 1.3 0.31 120.1 ^ 0.5 134.4 ^ 1.8 395 ^ 15 119.1 ^ 0.5

28-10c 141 NA 0.0184 ^ 0.6 0.1387 ^ 1.3 18.293 ^ 1.2 0.44 117.6 ^ 0.7 131.9 ^ 1.6 399 ^ 13 116.7 ^ 0.7

28-11c 207 NA 0.0180 ^ 0.7 0.1252 ^ 1.6 19.831 ^ 1.4 0.44 115.1 ^ 0.8 119.8 ^ 1.8 215 ^ 16 114.8 ^ 0.8

28-12c 127 NA 0.0191 ^ 0.4 0.1546 ^ 1.7 17.014 ^ 1.6 0.26 121.9 ^ 0.5 146.0 ^ 2.3 559 ^ 18 120.3 ^ 0.5

28-13c 144 NA 0.0188 ^ 0.4 0.1343 ^ 1.6 19.262 ^ 1.6 0.25 119.8 ^ 0.5 127.9 ^ 1.9 282 ^ 18 119.3 ^ 0.5

28-14c 77 NA 0.0209 ^ 1.2 0.1565 ^ 3.0 18.402 ^ 2.8 0.40 133.2 ^ 1.6 147.6 ^ 4.1 385 ^ 31 132.3 ^ 1.6

28-15c 155 NA 0.0180 ^ 0.4 0.1224 ^ 2.4 20.320 ^ 2.4 0.17 115.3 ^ 0.5 117.3 ^ 2.7 158 ^ 28 115.2 ^ 0.5

28-16c 47 NA 0.0176 ^ 1.2 0.0924 ^ 6.0 26.258 ^ 5.9 0.21 112.4 ^ 1.4 89.7 ^ 5.1 2 479 ^ 77 113.8 ^ 1.4

28-17c 80 NA 0.0182 ^ 0.7 0.1121 ^ 3.4 22.338 ^ 3.4 0.20 116.0 ^ 0.8 107.9 ^ 3.5 2 68 ^ 41 116.5 ^ 0.8

28-18c 57 NA 0.0192 ^ 0.9 0.1467 ^ 3.6 18.025 ^ 3.5 0.25 122.4 ^ 1.1 139.0 ^ 4.7 432 ^ 39 121.4 ^ 1.1

28-19c 24 NA 0.0185 ^ 0.6 0.1087 ^ 2.0 23.464 ^ 1.9 0.32 118.1 ^ 0.7 104.8 ^ 2.0 2 190 ^ 24 119.0 ^ 0.8

28-20c 40 NA 0.0186 ^ 0.5 0.1159 ^ 7.5 22.120 ^ 7.4 0.07 118.8 ^ 0.6 111.4 ^ 7.8 2 44 ^ 90 119.3 ^ 0.8

28-21c 36 NA 0.0187 ^ 0.5 0.1155 ^ 8.1 22.364 ^ 8.1 0.06 119.7 ^ 0.6 111.0 ^ 8.5 2 71 ^ 98 120.2 ^ 0.8

28-22c 68 NA 0.0184 ^ 0.4 0.1414 ^ 3.8 17.914 ^ 3.8 0.11 117.3 ^ 0.5 134.2 ^ 4.7 445 ^ 42 116.2 ^ 0.6

5a1c 24 NA 0.0190 ^ 1.1 0.0940 ^ 13.8 27.826 ^ 13.7 0.08 121.2 ^ 1.3 91.2 ^ 12.0 2 636 ^ 188 123.1 ^ 1.5

5a2c 24 NA 0.0185 ^ 1.0 0.1592 ^ 6.9 16.040 ^ 6.8 0.15 118.3 ^ 1.2 150.0 ^ 9.5 686 ^ 73 116.2 ^ 1.3

5a3c 23 NA 0.0248 ^ 0.5 0.1744 ^ 14.4 19.614 ^ 14.4 0.03 158.0 ^ 0.7 163.3 ^ 21.5 240 ^ 166 157.7 ^ 1.6

5a4c 31 NA 0.0187 ^ 0.8 0.1084 ^ 7.6 23.806 ^ 7.6 0.11 119.5 ^ 1.0 104.5 ^ 7.5 2 226 ^ 95 120.5 ^ 1.1

5a5c 27 NA 0.0180 ^ 0.9 0.0856 ^ 17.8 28.890 ^ 17.8 0.05 114.7 ^ 1.1 83.4 ^ 14.2 2 740 ^ 249 116.6 ^ 1.4

5a6c 27 NA 0.0176 ^ 1.0 0.0858 ^ 11.4 28.296 ^ 11.3 0.08 112.5 ^ 1.1 83.6 ^ 9.1 2 682 ^ 156 114.4 ^ 1.2

5a7c 28 NA 0.0179 ^ 0.6 0.1245 ^ 7.9 19.820 ^ 7.9 0.08 114.3 ^ 0.7 119.1 ^ 8.9 216 ^ 92 114.0 ^ 0.9

5a8c 31 NA 0.0174 ^ 1.0 0.1216 ^ 6.8 19.771 ^ 6.7 0.14 111.4 ^ 1.1 116.5 ^ 7.4 222 ^ 77 111.1 ^ 1.2

5a9c 39 NA 0.0178 ^ 1.0 0.1153 ^ 7.3 21.276 ^ 7.2 0.14 113.7 ^ 1.1 110.8 ^ 7.6 49 ^ 86 113.9 ^ 1.2

5a10c 27 NA 0.0182 ^ 0.9 0.1258 ^ 10.7 19.889 ^ 10.6 0.09 115.9 ^ 1.0 120.3 ^ 12.0 208 ^ 123 115.7 ^ 1.3

5a11c 35 NA 0.0180 ^ 0.5 0.0627 ^ 13.6 39.460 ^ 13.6 0.03 114.7 ^ 0.5 61.8 ^ 8.1 2 1711 ^ 237 118.0 ^ 0.7

5a12c 39 NA 0.0181 ^ 0.6 0.0992 ^ 12.9 25.174 ^ 12.9 0.04 115.7 ^ 0.7 96.0 ^ 11.8 2 369 ^ 168 116.9 ^ 1.0

5a13c 22 NA 0.0180 ^ 0.8 0.0975 ^ 9.5 25.394 ^ 9.4 0.08 114.8 ^ 0.9 94.5 ^ 8.5 2 391 ^ 123 116.0 ^ 1.1

5a14c 48 NA 0.0178 ^ 0.8 0.1025 ^ 5.4 23.944 ^ 5.4 0.15 113.8 ^ 0.9 99.1 ^ 5.1 2 241 ^ 68 114.7 ^ 1.0

5a15c 24 NA 0.0186 ^ 0.8 0.1403 ^ 6.8 18.239 ^ 6.8 0.11 118.5 ^ 0.9 133.3 ^ 8.5 405 ^ 76 117.6 ^ 1.0

5a16c 42 NA 0.0176 ^ 0.9 0.0996 ^ 6.6 24.380 ^ 6.6 0.13 112.6 ^ 1.0 96.4 ^ 6.1 2 287 ^ 84 113.6 ^ 1.0

5a17c 20 NA 0.0186 ^ 0.9 0.1028 ^ 12.1 24.929 ^ 12.0 0.07 118.7 ^ 1.0 99.4 ^ 11.4 2 344 ^ 155 120.0 ^ 1.3

5a18c 32 NA 0.0185 ^ 0.9 0.1570 ^ 9.4 16.198 ^ 9.4 0.09 117.8 ^ 1.0 148.1 ^ 12.9 665 ^ 101 115.9 ^ 1.3

5a19c 36 NA 0.0194 ^ 1.1 0.1317 ^ 9.3 20.285 ^ 9.2 0.11 123.8 ^ 1.3 125.7 ^ 10.9 162 ^ 108 123.6 ^ 1.5

5a20c 25 NA 0.0188 ^ 0.9 0.1471 ^ 8.4 17.584 ^ 8.4 0.11 119.8 ^ 1.1 139.3 ^ 10.9 486 ^ 92 118.5 ^ 1.3

221K-1t 267 18051 0.0206 ^ 0.7 0.1309 ^ 2.8 21.669 ^ 2.7 0.25 131.3 ^ 0.9 124.9 ^ 3.7 5 ^ 33 NA

221K-2t 351 29570 0.0200 ^ 0.5 0.1267 ^ 1.9 21.757 ^ 1.8 0.26 127.6 ^ 0.6 121.2 ^ 2.4 2 4 ^ 22 NA

SG

14

56

—3

0/9

/20

03

—1

4:5

8—

SF

OR

ST

ER

—8

24

96

—M

OD

EL

5–

br,ed

,cor

K.A

.K

lepeis

eta

l./

Jou

rna

lo

fS

tructu

ral

Geo

log

yxx

(00

00

)xxx–

xxx1

8

AR

TIC

LE

INP

RE

SS

19

05

19

06

19

07

19

08

19

09

19

10

19

11

19

12

19

13

19

14

19

15

19

16

19

17

19

18

19

19

19

20

19

21

19

22

19

23

19

24

19

25

19

26

19

27

19

28

19

29

19

30

19

31

19

32

19

33

19

34

19

35

19

36

19

37

19

38

19

39

19

40

19

41

19

42

19

43

19

44

19

45

19

46

19

47

19

48

19

49

19

50

19

51

19

52

19

53

19

54

19

55

19

56

19

57

19

58

19

59

19

60

19

61

19

62

19

63

19

64

19

65

19

66

19

67

19

68

19

69

19

70

19

71

19

72

19

73

19

74

19

75

19

76

19

77

19

78

19

79

19

80

19

81

19

82

19

83

19

84

19

85

19

86

19

87

19

88

19

89

19

90

19

91

19

92

19

93

19

94

19

95

19

96

19

97

19

98

19

99

20

00

20

01

20

02

20

03

20

04

20

05

20

06

20

07

20

08

20

09

20

10

20

11

20

12

20

13

20

14

20

15

20

16

Page 19: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Table 2 (continued)

Sample U

(ppm)

206Pb/204Pb

ratio

206Pb/238U ^ ratio

(%)

207Pb/235U ^ ratio

(%)

206Pb/207P ^ ratio

(%)

Error

correction

206Pb/238U ^ age

(Ma)

207Pb/235U ^ age

(Ma)

206Pb/207Pb ^ age

(Ma)

207Pb correction ^ age

(Ma)

221K-3t 428 36509 0.0198 ^ 0.6 0.1297 ^ 1.6 21.040 ^ 1.4 0.41 126.3 ^ 0.8 123.8 ^ 2.0 76 ^ 17 NA

221K-4c 262 43226 0.0364 ^ 1.4 0.2498 ^ 1.9 20.096 ^ 1.4 0.71 230.5 ^ 3.2 226.4 ^ 4.9 184 ^ 16 NA

221K-5c 40 4497 0.0577 ^ 0.9 0.4339 ^ 4.0 18.347 ^ 3.9 0.23 361.8 ^ 3.3 365.9 ^ 17.3 392 ^ 43 NA

221K-6c 44 15551 0.0560 ^ 1.0 0.4324 ^ 4.7 17.856 ^ 4.6 0.22 351.2 ^ 3.7 364.8 ^ 20.4 453 ^ 51 NA

221K-7c 15 7790 0.0569 ^ 1.2 0.3726 ^ 16.6 21.039 ^ 16.6 0.07 356.4 ^ 4.4 321.6 ^ 60.9 76 ^ 197 NA

221K-8c 90 38799 0.0580 ^ 0.8 0.4120 ^ 3.5 19.416 ^ 3.4 0.23 363.6 ^ 3.0 350.3 ^ 14.5 263 ^ 39 NA

221K-9t 118 12730 0.0202 ^ 1.5 0.1271 ^ 6.8 21.893 ^ 6.6 0.23 128.8 ^ 2.0 121.5 ^ 8.7 2 19 ^ 80 NA

221K-10t 56 6095 0.0261 ^ 1.3 0.2049 ^ 9.5 17.538 ^ 9.5 0.14 165.9 ^ 2.2 189.3 ^ 19.6 492 ^ 104 NA

221K-11c 94 18096 0.0502 ^ 1.2 0.3895 ^ 7.1 17.782 ^ 7.0 0.17 315.9 ^ 3.9 334.0 ^ 27.7 462 ^ 77 NA

221K-12c 101 16611 0.0570 ^ 1.0 0.4402 ^ 2.0 17.864 ^ 1.8 0.49 357.5 ^ 3.6 370.4 ^ 9.0 452 ^ 19 NA

221K-13c 99 26804 0.0583 ^ 1.0 0.4178 ^ 3.1 19.225 ^ 3.0 0.32 365.0 ^ 3.8 354.5 ^ 13.1 286 ^ 34 NA

221K-14c 74 22844 0.0575 ^ 1.1 0.4156 ^ 2.2 19.089 ^ 1.9 0.51 360.6 ^ 4.2 352.9 ^ 9.3 302 ^ 22 NA

221K-15t 67 11424 0.0205 ^ 1.8 0.1253 ^ 11.0 22.504 ^ 10.8 0.16 130.5 ^ 2.4 119.8 ^ 13.9 2 86 ^ 133 NA

221K-16c 152 29614 0.0513 ^ 2.9 0.3736 ^ 4.9 18.927 ^ 4.0 0.58 322.4 ^ 9.4 322.3 ^ 18.4 322 ^ 45 NA

221K-17c 95 23260 0.0496 ^ 1.1 0.3565 ^ 3.6 19.191 ^ 3.4 0.31 312.2 ^ 3.5 309.6 ^ 12.8 290 ^ 39 NA

221K-18c 100 18092 0.0535 ^ 3.2 0.3782 ^ 7.7 19.504 ^ 7.0 0.42 336.0 ^ 11.1 325.7 ^ 29.1 253 ^ 80 NA

221K-19c 491 23329 0.0574 ^ 0.9 0.4329 ^ 2.3 18.274 ^ 2.1 0.40 360.1 ^ 3.4 365.8 ^ 10.3 401 ^ 24.0 NA

221K-20c 196 9298 0.0567 ^ 0.6 0.4178 ^ 2.6 18.723 ^ 2.5 0.22 356.3 ^ 2.0 354.9 ^ 11.7 346 ^ 28.0 NA

221K-21c 121 4240 0.0422 ^ 0.6 0.2875 ^ 3.1 20.220 ^ 3.0 0.19 266.2 ^ 1.5 257.2 ^ 9.2 169 ^ 35.0 NA

221K-22c 827 8487 0.0190 ^ 3.0 0.1263 ^ 3.7 20.737 ^ 2.1 0.82 121.8 ^ 3.7 121.4 ^ 5.2 110 ^ 25.0 NA

221K-23t 964 8645 0.0206 ^ 0.3 0.1402 ^ 2.1 20.257 ^ 2.1 0.16 131.5 ^ 0.4 133.7 ^ 3.6 165 ^ 24.0 NA

221K-24c 370 19432 0.0501 ^ 1.0 0.3744 ^ 2.5 18.451 ^ 2.3 0.40 315.3 ^ 3.1 323.2 ^ 2.7 379 ^ 25.0 NA

221K-25c 548 33147 0.0594 ^ 0.9 0.4508 ^ 2.2 18.153 ^ 2.0 0.41 372.1 ^ 3.5 378.3 ^ 10.3 416 ^ 23.0 NA

221K-26c 882 4216 0.0862 ^ 1.7 3.1545 ^ 4.6 3.770 ^ 4.2 0.38 533.8 ^ 9.6 1446.0 ^ 136.8 3278 ^ 33.0 NA

221K-27t 166 1708 0.0591 ^ 0.9 0.6847 ^ 5.1 11.890 ^ 5.0 0.18 370.0 ^ 3.6 530.6 ^ 35.1 1295 ^ 49.0 NA

221K-28t 480 22242 0.0448 ^ 1.6 0.3304 ^ 2.6 18.677 ^ 2.1 0.59 282.2 ^ 4.5 290.1 ^ 9.1 352 ^ 24.0 NA

221K-29c 1150 30148 0.0480 ^ 2.2 0.3685 ^ 3.0 17.952 ^ 2.0 0.74 302.2 ^ 6.8 319.8 ^ 11.9 441 ^ 22.0 NA

221K-30c 829 29551 0.0594 ^ 0.5 0.4506 ^ 2.1 18.183 ^ 2.0 0.22 372.4 ^ 1.8 378.4 ^ 9.0 412 ^ 22.0 NA

221K-31c 108 7676 0.0589 ^ 0.9 0.4254 ^ 2.6 19.100 ^ 2.5 0.33 369.0 ^ 3.3 360.5 ^ 11.1 301 ^ 28.0 NA

221K-32c 179 10102 0.0580 ^ 0.4 0.4291 ^ 2.4 18.619 ^ 2.4 0.16 363.5 ^ 1.4 363.1 ^ 10.7 359 ^ 27.0 NA

221K-33c 819 33551 0.0596 ^ 0.8 0.4537 ^ 2.2 18.100 ^ 2.0 0.35 373.1 ^ 2.9 380.8 ^ 10.4 422 ^ 23.0 NA

221K-34t 143 13634 0.0184 ^ 4.5 0.1306 ^ 11.9 19.389 ^ 11.0 0.38 117.7 ^ 5.3 125.9 ^ 16.3 267 ^ 126.0 NA

221K-35t 68 2536 0.0560 ^ 0.5 0.4308 ^ 4.6 17.934 ^ 4.6 0.10 351.8 ^ 1.7 364.2 ^ 20.2 443 ^ 51.0 NA

221K-36c 136 4135 0.0574 ^ 0.7 0.4827 ^ 2.8 16.388 ^ 2.7 0.24 360.2 ^ 2.5 400.0 ^ 14.8 640 ^ 29.0 NA

221K-37t 278 24793 0.0561 ^ 0.7 0.4215 ^ 2.3 18.361 ^ 2.2 0.32 352.4 ^ 2.6 357.4 ^ 10.1 390 ^ 24.0 NA

221K-38t 234 8850 0.0233 ^ 0.9 0.1544 ^ 3.6 20.767 ^ 3.5 0.25 148.9 ^ 1.3 146.6 ^ 6.2 107 ^ 41.0 NA

221K-39c 265 22663 0.0606 ^ 0.3 0.4561 ^ 2.2 18.310 ^ 2.1 0.16 379.0 ^ 1.3 382.3 ^ 10.7 396 ^ 24.0 NA

221K-40t 155 4240 0.0537 ^ 1.1 0.3853 ^ 3.0 19.218 ^ 2.8 0.36 337.3 ^ 3.8 331.7 ^ 12.6 287 ^ 32.0 NA

221K-41t 51 16902 0.0586 ^ 1.1 0.4362 ^ 3.7 18.535 ^ 3.6 0.30 367.1 ^ 4.2 368.1 ^ 16.6 369 ^ 40.0 NA

221K-42t 49 4837 0.0584 ^ 1.1 0.4686 ^ 5.4 17.185 ^ 5.3 0.21 366.0 ^ 4.2 0.3 ^ 26.3 537 ^ 58.0 NA

221K-43t 42 2299 0.0595 ^ 1.1 0.4142 ^ 5.2 19.809 ^ 5.1 0.21 373.3 ^ 4.3 352.7 ^ 22.0 217 ^ 59.0 NA

221K-44t 192 6125 0.0203 ^ 2.3 0.1206 ^ 4.0 23.251 ^ 3.3 0.57 130.6 ^ 3.0 116.5 ^ 5.5 2 167 ^ 41.0 NA

221K-45t 187 45340 0.0576 ^ 0.5 0.4275 ^ 2.3 18.578 ^ 2.3 0.23 361.3 ^ 2.0 361.0 ^ 10.1 364 ^ 26.0 NA

221K-46t 35 4453 0.0588 ^ 1.7 0.4522 ^ 8.9 17.926 ^ 8.8 0.19 368.5 ^ 6.3 379.2 ^ 40.3 444 ^ 98.0 NA

221K-47c 59 10614 0.0573 ^ 0.9 0.4391 ^ 4.5 17.975 ^ 4.5 0.20 359.1 ^ 3.0 370.5 ^ 20.6 438 ^ 49.0 NA

(continued on next page)

SG

14

56

—3

0/9

/20

03

—1

4:5

8—

SF

OR

ST

ER

—8

24

96

—M

OD

EL

5–

br,ed

,cor

K.A

.K

lepeis

eta

l./

Jou

rna

lo

fS

tructu

ral

Geo

log

yxx

(00

00

)xxx–

xxx1

9

AR

TIC

LE

INP

RE

SS

20

17

20

18

20

19

20

20

20

21

20

22

20

23

20

24

20

25

20

26

20

27

20

28

20

29

20

30

20

31

20

32

20

33

20

34

20

35

20

36

20

37

20

38

20

39

20

40

20

41

20

42

20

43

20

44

20

45

20

46

20

47

20

48

20

49

20

50

20

51

20

52

20

53

20

54

20

55

20

56

20

57

20

58

20

59

20

60

20

61

20

62

20

63

20

64

20

65

20

66

20

67

20

68

20

69

20

70

20

71

20

72

20

73

20

74

20

75

20

76

20

77

20

78

20

79

20

80

20

81

20

82

20

83

20

84

20

85

20

86

20

87

20

88

20

89

20

90

20

91

20

92

20

93

20

94

20

95

20

96

20

97

20

98

20

99

21

00

21

01

21

02

21

03

21

04

21

05

21

06

21

07

21

08

21

09

21

10

21

11

21

12

21

13

21

14

21

15

21

16

21

17

21

18

21

19

21

20

21

21

21

22

21

23

21

24

21

25

21

26

21

27

21

28

Page 20: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Table 2 (continued)

Sample U

(ppm)

206Pb/204Pb

ratio

206Pb/238U ^ ratio

(%)

207Pb/235U ^ ratio

(%)

206Pb/207P ^ ratio

(%)

Error

correction

206Pb/238U ^ age

(Ma)

207Pb/235U ^ age

(Ma)

206Pb/207Pb ^ age

(Ma)

207Pb correction ^ age

(Ma)

221K-48c 67 9742 0.0580 ^ 1.0 0.4275 ^ 4.6 18.699 ^ 4.4 0.23 363.8 ^ 4.0 361.1 ^ 20.4 349 ^ 50.0 NA

221K-49c 90 18019 0.0591 ^ 1.0 0.4423 ^ 4.0 18.413 ^ 3.9 0.25 370.4 ^ 4.0 372.8 ^ 18.9 384 ^ 44.0 NA

221K-50c 26 2991 0.0601 ^ 1.1 0.4130 ^ 4.4 20.052 ^ 4.3 0.24 376.4 ^ 4.0 351.1 ^ 18.0 189 ^ 50.0 NA

ADA2-1c 314 16959 0.0323 ^ 0.9 0.2236 ^ 1.6 19.907 ^ 1.3 0.54 204.8 ^ 1.8 204.9 ^ 3.6 206 ^ 15 NA

ADA2-2c 729 3605 0.0319 ^ 0.5 0.2051 ^ 1.0 21.437 ^ 0.9 0.49 202.3 ^ 1.0 189.4 ^ 2.0 31 ^ 10 NA

ADA2-3c 125 6567 0.0315 ^ 1.0 0.2113 ^ 4.2 20.563 ^ 4.0 0.25 200.0 ^ 2.1 194.6 ^ 8.9 130 ^ 47 NA

ADA2-4c 128 4641 0.0320 ^ 0.4 0.1987 ^ 3.2 22.200 ^ 3.2 0.12 203.0 ^ 0.8 184.0 ^ 6.5 2 53 ^ 39 NA

ADA2-5c 53 17623 0.0302 ^ 0.6 0.1870 ^ 7.4 22.282 ^ 7.4 0.07 191.9 ^ 1.1 174.0 ^ 14.0 2 62 ^ 90 NA

ADA2-6c 70 2532 0.0319 ^ 0.6 0.2188 ^ 4.8 20.080 ^ 4.7 0.13 202.2 ^ 1.3 200.9 ^ 10.6 186 ^ 55 NA

ADA2-7c 30 15567 0.0323 ^ 1.5 0.2033 ^ 16.4 21.890 ^ 16.3 0.09 204.8 ^ 3.1 187.9 ^ 33.3 2 19 ^ 197 NA

ADA2-8c 81 2624 0.0318 ^ 0.6 0.1967 ^ 5.6 22.305 ^ 5.6 0.11 201.9 ^ 1.3 182.3 ^ 11.1 2 65 ^ 68 NA

ADA2-9c 135 212500 0.0316 ^ 2.0 0.2363 ^ 3.8 18.429 ^ 3.2 0.53 200.4 ^ 4.1 215.4 ^ 9.0 382 ^ 36 NA

ADA2-10c 59 176999 0.0317 ^ 0.9 0.2102 ^ 12.0 20.796 ^ 11.9 0.08 201.2 ^ 1.9 193.7 ^ 25.2 103 ^ 141 NA

ADA2-11c 54 6500 0.0271 ^ 1.1 0.1906 ^ 11.4 19.602 ^ 11.3 0.10 172.3 ^ 2.0 177.1 ^ 21.8 242 ^ 131 NA

ADA2-12c 94 3222 0.0322 ^ 0.5 0.2014 ^ 6.7 22.062 ^ 6.7 0.07 204.4 ^ 1.0 186.3 ^ 13.7 2 38 ^ 82 NA

ADA2-13c 48 1570 0.0258 ^ 0.7 0.1465 ^ 14.3 24.243 ^ 14.3 0.05 164.0 ^ 1.1 138.8 ^ 21.0 2 272 ^ 181 NA

ADA2-14c 43 13125 0.0236 ^ 1.4 0.1503 ^ 18.9 21.688 ^ 18.9 0.07 150.6 ^ 2.1 142.2 ^ 28.4 3 ^ 227 NA

ADA2-15c 162 32272 0.0325 ^ 1.0 0.2277 ^ 2.9 19.675 ^ 2.7 0.34 206.2 ^ 2.0 208.3 ^ 6.6 233 ^ 31 NA

ADA2-16c 132 12743 0.0327 ^ 0.6 0.2176 ^ 3.7 20.728 ^ 3.6 0.16 207.5 ^ 1.2 199.9 ^ 8.1 111 ^ 43 NA

ADA2-17c 90 18709 0.0328 ^ 0.5 0.2286 ^ 5.4 19.775 ^ 5.4 0.09 208.0 ^ 1.1 209.0 ^ 12.5 221 ^ 62 NA

ADA2-18c 125 3783 0.0322 ^ 0.8 0.2325 ^ 3.9 19.085 ^ 3.8 0.20 204.2 ^ 1.6 212.2 ^ 9.1 303 ^ 43 NA

ADA2-19c 63 3726 0.0325 ^ 1.1 0.2299 ^ 6.5 19.478 ^ 6.4 0.18 206.0 ^ 2.4 210.1 ^ 14.9 256 ^ 73 NA

ADA2-20c 38 13947 0.0279 ^ 1.1 0.1673 ^ 13.1 22.960 ^ 13.1 0.08 177.1 ^ 2.0 157.0 ^ 22.1 2 136 ^ 162 NA

ADA2-21c 44 8748 0.0240 ^ 0.7 0.1499 ^ 12.3 22.026 ^ 12.3 0.06 152.6 ^ 1.1 141.9 ^ 18.5 2 34 ^ 149 NA

ADA2-22c 99 26510 0.0332 ^ 0.6 0.2388 ^ 4.4 19.175 ^ 4.4 0.14 210.6 ^ 1.3 217.5 ^ 10.7 292 ^ 50 NA

ADA2-23c 162 37708 0.0316 ^ 0.6 0.2342 ^ 2.6 18.621 ^ 2.6 0.23 200.8 ^ 1.2 213.7 ^ 6.3 359 ^ 29 NA

ADA2-24c 330 48018 0.0329 ^ 0.8 0.2430 ^ 1.6 18.657 ^ 1.4 0.48 208.5 ^ 1.6 220.8 ^ 4.0 354 ^ 16 NA

ADA2-25c 39 7083 0.0328 ^ 1.7 0.2370 ^ 12.2 19.084 ^ 12.1 0.14 208.1 ^ 3.5 216.0 ^ 28.9 303 ^ 137 NA

ADA2-26t 145 2506 0.0177 ^ 6.2 0.1220 ^ 9.3 19.989 ^ 7.0 0.66 113.0 ^ 7.0 116.8 ^ 11.5 196 ^ 82 NA

ADA2-27t 43 1987 0.0252 ^ 1.3 0.1605 ^ 14.2 21.628 ^ 14.1 0.09 160.2 ^ 2.0 151.1 ^ 22.8 10 ^ 170 NA

ADA2-28t 44 5893 0.0187 ^ 1.0 0.1249 ^ 4.4 20.597 ^ 4.3 0.22 119.1 ^ 1.2 119.5 ^ 5.6 126 ^ 51 NA

ADA2-29t 107 4837 0.0196 ^ 0.5 0.1382 ^ 1.1 19.529 ^ 1.0 0.42 124.9 ^ 0.6 131.4 ^ 1.6 250 ^ 12 NA

ADA2-30t 77 1348 0.0180 ^ 1.1 0.1040 ^ 2.3 23.825 ^ 2.1 0.45 114.8 ^ 1.2 100.4 ^ 2.5 2 228 ^ 26 NA

ADA2-31t 75 20322 0.0175 ^ 1.1 0.1416 ^ 6.4 17.042 ^ 6.3 0.17 111.8 ^ 1.2 134.4 ^ 9.1 555 ^ 69 NA

ADA2-32t 205 1111 0.0175 ^ 0.9 0.0982 ^ 1.9 24.544 ^ 1.6 0.49 111.7 ^ 1.0 95.1 ^ 1.9 2 304 ^ 21 NA

ADA2-33t 107 1503 0.0190 ^ 1.3 0.1139 ^ 1.7 23.000 ^ 1.2 0.74 121.4 ^ 1.6 109.6 ^ 2.0 2 140 ^ 14 NA

ADA2-34t 20 33533 0.0196 ^ 0.7 0.1456 ^ 2.2 18.590 ^ 2.1 0.32 125.3 ^ 0.9 138.0 ^ 3.2 362 ^ 23 NA

ADA2-35t 79 7649 0.0187 ^ 4.9 0.1440 ^ 13.3 17.874 ^ 12.3 0.37 119.2 ^ 5.9 136.6 ^ 19.2 450 ^ 137 NA

ADA2-36t 363 4293 0.0181 ^ 1.2 0.1192 ^ 6.7 20.934 ^ 6.6 0.18 115.6 ^ 1.4 114.3 ^ 8.1 88 ^ 78 NA

ADA2-37t 27 1387 0.0188 ^ 2.4 0.1489 ^ 6.6 17.411 ^ 6.1 0.37 120.1 ^ 2.9 140.9 ^ 9.9 508 ^ 67 NA

ADA2-38t 20 1362 0.0183 ^ 3.7 0.1107 ^ 7.1 22.817 ^ 6.0 0.53 117.0 ^ 4.4 106.6 ^ 7.9 2 120 ^ 74 NA

ADA2-39t 37 3061 0.0182 ^ 2.0 0.1087 ^ 6.3 23.083 ^ 6.0 0.32 116.3 ^ 2.4 104.8 ^ 7.0 2 149 ^ 74 NA

ADA2-40t 143 18329 0.0180 ^ 3.7 0.1280 ^ 11.8 19.412 ^ 11.2 0.31 115.2 ^ 4.3 122.3 ^ 15.2 264 ^ 129 NA

ADA2-41t 644 15254 0.0184 ^ 0.9 0.0870 ^ 3.6 29.202 ^ 3.5 0.24 117.7 ^ 1.0 84.7 ^ 3.2 2 770 ^ 49 NA

ADA2-42t 30 5509 0.0189 ^ 0.8 0.1384 ^ 1.6 18.837 ^ 1.4 0.50 120.8 ^ 1.0 131.6 ^ 2.3 333 ^ 16 NA

SG

14

56

—3

0/9

/20

03

—1

4:5

8—

SF

OR

ST

ER

—8

24

96

—M

OD

EL

5–

br,ed

,cor

K.A

.K

lepeis

eta

l./

Jou

rna

lo

fS

tructu

ral

Geo

log

yxx

(00

00

)xxx–

xxx2

0

AR

TIC

LE

INP

RE

SS

21

29

21

30

21

31

21

32

21

33

21

34

21

35

21

36

21

37

21

38

21

39

21

40

21

41

21

42

21

43

21

44

21

45

21

46

21

47

21

48

21

49

21

50

21

51

21

52

21

53

21

54

21

55

21

56

21

57

21

58

21

59

21

60

21

61

21

62

21

63

21

64

21

65

21

66

21

67

21

68

21

69

21

70

21

71

21

72

21

73

21

74

21

75

21

76

21

77

21

78

21

79

21

80

21

81

21

82

21

83

21

84

21

85

21

86

21

87

21

88

21

89

21

90

21

91

21

92

21

93

21

94

21

95

21

96

21

97

21

98

21

99

22

00

22

01

22

02

22

03

22

04

22

05

22

06

22

07

22

08

22

09

22

10

22

11

22

12

22

13

22

14

22

15

22

16

22

17

22

18

22

19

22

20

22

21

22

22

22

23

22

24

22

25

22

26

22

27

22

28

22

29

22

30

22

31

22

32

22

33

22

34

22

35

22

36

22

37

22

38

22

39

22

40

Page 21: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Daniel. In addition, the steep upper amphibolite facies

foliations (SSZ) of the Indecision Creek Shear Zone cut

across the lower (and eastern) contact of the WFO east of

Mt. Daniel (Figs. 9b and 10d). These relationships indicate

that once the WFO crystallized deformation was no longer

concentrated in a narrow melt-enhanced zone at the base of

the batholith. By ,116 Ma (and possibly earlier) this zone

had been replaced by the steep shear zones up to 15 km

wide. This represents a progressive widening of the zone of

deformation below the batholith to include the entire lower

crustal section between Milford and Caswell Sounds.

Following dehydration of the lower crust and abandon-

ment of the Mt. Daniel Shear Zone physical and kinematic

links developed above and below the WFO. The Caswell

fold–thrust belt continued to evolve in the middle crust

following crystallization of the batholith and records arc-

normal contraction and vertical thickening with a sinistral

component of arc-parallel displacement. The George Sound

and Indecision Creek Shear Zones also record subhorizontal

arc-normal contraction and vertical (layer-perpendicular)

thickening with components of sinistral arc-parallel dis-

placement. The simultaneous or near simultaneous devel-

opment of these contractional structures confirms indicate

that the middle and lower crust were coupled kinematically

above and below the batholith.

The style of contractional deformation affecting the

upper crust prior to ,105 Ma resembles the style of

deformation in the Caswell Thrust Belt (Fig. 9b). Narrow

thrust zones and steeply dipping foliations with down-dip

lineations were localized within the contact aureoles of

plutons on the western side of the Median Batholith in

Westland and on the eastern side east of Fiordland

(Bradshaw, 1989a; Tulloch and Challis, 2000). On Stewart

Island, dipping thrust faults formed inside the batholith

during ,125–105 Ma magmatism (Allibone and Tulloch,

1997; Tulloch and Kimbrough, 2003). These structural

patterns where deformation was concentrated in a narrow

(50–75 km wide) zone inside and on both sides of the

Median Batholith are remarkably similar to those preserved

in Fiordland where thrusts deform both pluton and country

rock in the contact aureoles of the WFO (Figs. 9a and 10d).

These relationships provide evidence that the upper, middle

and lower crusts were coupled kinematically during ,116–

105 Ma contraction as plutons were emplaced into the upper

crust.

9. Discussion: controls on vertical coupling and

decoupling in the crust

9.1. Changing strength and rheology of the lower crust

Structural and metamorphic relationships within north-

ern Fiordland suggest that decoupling of the middle and

lower crust above and below the WFO occurred as large

volumes of magma (.10 km thickness) were emplaced intoTab

le2

(co

nti

nued

)

Sam

ple

U (pp

m)

206P

b/2

04P

b

rati

o

206P

b/2

38U^

rati

o

(%)

207P

b/2

35U^

rati

o

(%)

206P

b/2

07P^

rati

o

(%)

Err

or

corr

ecti

on

206P

b/2

38U^

age

(Ma)

207P

b/2

35U^

age

(Ma)

206P

b/2

07P

b^

age

(Ma)

207P

bco

rrec

tion^

age

(Ma)

AD

A2

-43

t2

92

47

142

0.0

17

8^

0.7

0.1

25

5^

2.3

19

.51

8^

2.2

0.3

11

13

.5^

0.8

12

0.1

^2

.92

51^

25

NA

AD

A2

-44

t2

21

10

947

0.0

18

0^

1.6

0.1

19

8^

5.1

20

.73

3^

4.9

0.3

01

15

.1^

1.8

11

4.9

^6

.21

11^

58

NA

AD

A2

-45

t5

26

59

50

.01

80^

1.2

0.1

30

9^

2.8

18

.95

1^

2.5

0.4

21

15

.0^

1.3

12

4.9

^3

.73

19^

28

NA

Anal

yse

sin

ital

ics

are

not

use

din

age

calc

ula

tions.

Posi

tion

of

spot:

core

,t¼

tip

.N

info

rmat

ion

no

tav

aila

ble

.U

con

cen

trat

ion

has

anu

nce

rtai

nty

of,

25

%.

206P

b/2

38U

and

206P

b/2

07P

bra

tio

sh

ave

bee

nco

rrec

ted

for

frac

tionat

ion

usi

ng

afa

ctor

det

erm

ined

from

anal

ysi

sof

azi

rcon

cryst

alof

know

nag

e.U

nce

rtai

nti

esar

esh

ow

nat

1s

lev

el.

207P

bco

rrec

tio

nag

eis

det

erm

ined

by

adju

stin

gth

eco

mm

on

Pb

corr

ecti

on

tom

ake

rad

iogen

ic206P

b/2

38U

and

206P

b/2

07P

bag

esco

nco

rdan

t.F

or

sam

ple

s9

93

5,

22

1K

,an

dA

DA

2,

com

mon

Pb

corr

ecti

on

isfr

om

mea

sure

d206P

b/2

04P

b.

SG 1456—30/9/2003—14:58—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 21

ARTICLE IN PRESS

2241

2242

2243

2244

2245

2246

2247

2248

2249

2250

2251

2252

2253

2254

2255

2256

2257

2258

2259

2260

2261

2262

2263

2264

2265

2266

2267

2268

2269

2270

2271

2272

2273

2274

2275

2276

2277

2278

2279

2280

2281

2282

2283

2284

2285

2286

2287

2288

2289

2290

2291

2292

2293

2294

2295

2296

2297

2298

2299

2300

2301

2302

2303

2304

2305

2306

2307

2308

2309

2310

2311

2312

2313

2314

2315

2316

2317

2318

2319

2320

2321

2322

2323

2324

2325

2326

2327

2328

2329

2330

2331

2332

2333

2334

2335

2336

2337

2338

2339

2340

2341

2342

2343

2344

2345

2346

2347

2348

2349

2350

2351

2352

Page 22: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

and moved through a lower crust weakened by melt and

heat. Two of the dominant processes controlling decoupling

during this period were the rates of emplacement and

crystallization of the WFO and the duration of the thermal

pulse that accompanied the magmatism. Rapid rates of

emplacement and cooling are implied by available geo-

chronologic and metamorphic data (Fig. 10a). Crystal-

lization ages (Tulloch and Kimbrough, 2003; Hollis et al.,

2004) suggest that most of the WFO was emplaced into the

lower crust within the period ,126–120 Ma. By ,120 Ma,

garnet granulite facies metamorphism at temperatures of

750 8C , T , 850 8C affected large regions of the batholith

and its lower crustal host (Clarke et al., 2000; Tulloch et al.,

2000; Hollis et al., 2003). By ,116 Ma kyanite- and

paragonite-bearing assemblages replaced the older garnet-,

clinopyroxene-bearing granulite facies assemblages indi-

cating that the lower crust below the batholith had cooled to

temperatures of T ¼ 650–700 8C (Daczko et al., 2002c).

The age of ,116 Ma is an average of the metamorphic rims

dates obtained from the Caswell Sound fold–thrust belt and

the Indecision Creek Shear Zone (samples 995a and Ada2,

Fig. 8).

Simple one-dimensional models of conductive heat loss

using standard solutions to the heat flow equation (Spear,

1993, p. 43) support the interpretation of rapid cooling of the

WFO by ,116 Ma. Magma compositions and the liquidus

temperature of basalt suggest that initial intrusions tem-

peratures were close to T ¼ 1200 8C following the estimates

of Petford and Gallagher (2001). Advective removal of heat

via fluid and melt transfer and the thrusting of pre-existing

arc crust above and below the batholith during emplacement

increases the rate of cooling. Burial and crustal thickening

tend to slow the rate of cooling. However, metamorphic and

geochronologic data from northern Fiordland indicate that

most of the burial of the Arthur River Complex occurred

prior to ,120 Ma granulite facies metamorphism (Fig.

10c). Calculations using metamorphic data to constrain

ambient temperatures at ,120 Ma (T ¼ 750–800 8C in

crust below the batholith and T ¼ 550–600 8C in crust

above it) suggest that the thermal pulse accompanying WFO

emplacement decayed to temperatures of T ¼ 650–700 8C

within ,3–4 million years. These results agree well with

the mineral reactions that record cooling of the lower crust

below the batholith (Daczko et al., 2002c). The well-known

dependence of lower crustal strength and rheology on

temperature, fluid activity and melt fraction imply that the

lower crust experienced rapid changes in strength (effective

viscosity) during the ,120–105 Ma period.

In addition to rates of cooling and magma emplacement,

deformation patterns also were influenced by rheological

contrasts and mechanical anisotropies created by the layered

architecture of the lower crust. The influence of these

features are revealed by the formation of shear zones in the

contact aureoles of the batholith where steep temperature

gradients and melt created strength contrasts and their

parallelism with compositional layering (S1) in host rock.

The importance of inherited anisotropies has been observed

in other orogens also (Klepeis and Crawford, 1999; Miller

and Paterson, 2001).

Unlike the Mt. Daniel Shear Zone, the Caswell fold–

thrust belt at the top of the WFO continued to develop as

steep shear zones formed in the lowermost crust (Table 1).

This heterogeneity appears to reflect differences in lithology

at these two locations. As the WFO cooled, the strength

contrast between the dioritic rock of the WFO and

metasedimentary host rock at Caswell Sound continued to

focus deformation. However, at Mt. Daniel similar bulk

compositions between the WFO and its host rock meant that

once the batholith had cooled to ambient temperatures, the

contrast in strength across this zone was small. This

interpretation explains differences in the thickness and

duration of the shear zones. The relationships illustrate that

crustal decoupling and the formation of subhorizontal or

dipping shear zones in the deep crust strongly reflects local

rheological contrasts and temperature gradients (see also

Karlstrom and Williams, 2002).

The period of orogenesis characterized by high degrees

of vertical coupling (Fig. 10d) coincided with the develop-

ment of a strong, dry, mafic lower crustal root within

northernmost Fiordland. A strengthening lower crust is

implied in part by cooling. However, other processes also

greatly influenced lower crustal strength. Piston cylinder

experiments performed on unmelted samples of dioritic

gneiss from the Pembroke Valley (Antignano, 2002)

indicate that melt fractions remained low (#10 vol%) at

all temperatures up to T ¼ 975 8C. Fluid absent melting of

the gneiss was controlled by the decomposition of

hornblende ^ clinozoisite to produce garnet þ melt and

resulted in low melt volumes (Antignano, 2002; Klepeis

et al., 2003). Klepeis et al. (2003) pointed out that although

partial melting occurred in large parts of the lower crustal

section (Fig. 10c), the total volumes of melt probably

remained low. Low melt volumes would have helped the

lower crust remain strong even as it partially melted. In

addition, the formation of melt-induced fractures and vein

networks by ,120 Ma aided melt extraction and escape

from the lower crust (Daczko et al., 2001b). Klepeis and

Clarke (2003) showed that ductile deformation in shear

zones also aided melt escape. Davidson et al. (1994),

Roering et al. (1995) and Rushmer (1995) report similar

mechanisms of melt-enhanced embrittlement in the deep

crust. These processes helped prevent the wholesale

weakening of the lower crust following batholith

emplacement.

In summary, the relationships we describe suggest that

the development of a strong lower crustal root was aided by

the following processes: (1) rapid cooling and crystal-

lization of the WFO as pre-existing arc crust was thrust

above and below the batholith; (2) the dehydration of the

lower crust at granulite facies conditions; (3) the inherently

high strength of mafic lower crustal rocks even as they

underwent mineral reactions involving partial melting; (4)

SG 1456—30/9/2003—14:58—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx22

ARTICLE IN PRESS

2353

2354

2355

2356

2357

2358

2359

2360

2361

2362

2363

2364

2365

2366

2367

2368

2369

2370

2371

2372

2373

2374

2375

2376

2377

2378

2379

2380

2381

2382

2383

2384

2385

2386

2387

2388

2389

2390

2391

2392

2393

2394

2395

2396

2397

2398

2399

2400

2401

2402

2403

2404

2405

2406

2407

2408

2409

2410

2411

2412

2413

2414

2415

2416

2417

2418

2419

2420

2421

2422

2423

2424

2425

2426

2427

2428

2429

2430

2431

2432

2433

2434

2435

2436

2437

2438

2439

2440

2441

2442

2443

2444

2445

2446

2447

2448

2449

2450

2451

2452

2453

2454

2455

2456

2457

2458

2459

2460

2461

2462

2463

2464

Page 23: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

low volumes of partial melt in the lower crust controlled by

the decomposition of hornblende and clinozoisite; and (5)

the rapid extraction of partial melt via a regionally extensive

networks of fractures, veins and ductile shear zones.

The interplay among these processes illustrates the time-

dependent nature of lower crustal rheology. This variability

is consistent with data derived from rock deformation

experiments that indicate melt volume, temperature, the

mobilization of melt, and crustal composition cause

variations in rock strength (Dell’Angelo and Tullis, 1988;

Rushmer, 1995; Huerta et al., 1996; Rutter, 1997). The close

relationships we observed among magmatism, elevated

temperatures, melt-assisted strain localization, and vertical

decoupling during the ,126–120 Ma period also are

consistent with numerical and analytical models indicating

that crust–mantle decoupling is characteristic of orogens

with a weak lower crust. A weak, low viscosity zone in the

lower crust causes stresses arising from mantle convergence

preferentially to be transmitted horizontally rather than

vertically through the crust (Royden, 1996; Ellis et al., 1998;

Willett, 1998). The disparate strain paths we observed

between different parts of the crustal section in Fiordland

during WFO emplacement appear to reflect this process.

Decoupling of the crust above and below the WFO is

consistent with experimental data that show stress cannot be

transmitted across zones containing more than ,30–45%

volume melt (Rutter and Neumann, 1995).

The Fiordland example also illustrates spatially (both

vertical and horizontal) variability of lower crustal strength

profiles. The development of a relatively cool (T ¼ 650–

700 8C) lower crust in northernmost Fiordland by ,116 Ma

contrasts with the elevated temperatures (T . 800 8C) at

,107.5 Ma recorded in granulite facies mineral assem-

blages ,100 km to the south in Doubtful Sound (Gibson

and Ireland, 1995). Geochronologic and geochemical data

also indicate that Na-rich magmas that reflect lower crustal

sources continued to be emplaced into the middle and upper

crust until ,105 Ma (Tulloch and Kimbrough, 2003). This

magmatic activity at shallow crustal levels is compatible

with data that suggest dikes and deformation ductile shear

zones helped extract melt from the lower crust during the

,116–105 Ma period (Klepeis and Clarke, 2003; Klepeis

et al., 2003). High degrees of vertical coupling thus occurred

even as melt moved out of the lower crust and was emplaced

into the upper crust until ,105 Ma. Together these

observations imply that the effects of magmatism and

partial melting on strength profiles and rheological tran-

sitions in the lower crust are much more transient and

spatially heterogeneous than previously believed.

Finally, the features in the upper and lower crusts of the

Fiordland–Westland orogen suggest that a highly viscous

(strong, dry, mafic) lower crust influenced the overall

structural style and partitioning of deformation vertically

within the orogen. The narrow (50–75 km wide) focused

style of contraction in the upper crust matches that predicted

for orogens characterized by a highly viscous lower crust,

where stresses derived from convergence and flow in the

upper mantle preferentially are transmitted vertically

through the crust (Royden, 1996; Ellis et al., 1998). In

these types of orogens, the lower crust is not sufficiently

weak to allow channelized lower crustal flow and crust–

mantle decoupling during convergence.

9.2. Conditions influencing late-orogenic extension

The Fiordland–Westland orogen records the onset of

extension (Fig. 10d) and the cessation of contraction by

,108–105 Ma (Tulloch and Kimbrough, 1989; Gibson and

Ireland, 1995; Spell et al., 2000). This pattern invites

comparison with other orogens that also exhibit evidence of

late orogenic extension following a period of crustal

thickening. The mechanisms that can contribute to this

process are variable, including thermal relaxation of

thickened lower crust (Vanderhaeghe and Teyssier, 1997,

2001), changes in convergence rate (England and House-

man, 1989), and the convective removal of a lithospheric

root (Molnar et al., 1993). Below, we evaluate the possible

roles of these mechanisms for the Fiordland–Westland

orogen.

In vertically coupled orogens characterized by a highly

viscous lower crust, contraction and extension can be

balanced between different layers of the crust (e.g. Burg

et al., 1984; Burchfield et al., 1992; Burg and Ford, 1997).

Focused extension in the middle and upper crusts can occur

simultaneously with shortening in the lower crust as a result

of return flow away from the zone of mantle convergence

(Royden, 1996). However, the timing of normal faulting in

both the upper and lower crusts in the Fiordland–Westland

eliminates this as a possible mechanism for this orogen. An

alternative mechanism of late orogenic extension is a

thermal weakening of previously thickened lower crust. In

the Canadian Cordillera and the Himalaya–Tibet region, the

onset of late orogenic extension has been linked to the

thermal weakening of the middle or lower crust and an

increased role of buoyancy forces due to partial melting

(Royden, 1996; Vanderhaeghe and Teyssier, 1997; Ellis

et al., 1998; Jamieson et al., 1998; Willett, 1998). This

mechanism explains crosscutting relationships and differ-

ences in timing between contractional structures and normal

faulting. However, we suggest that because we have

evidence for a viscous, cooling, mafic lower crust and

strong vertical coupling prior to the onset of extension (Fig.

10), we also can eliminate this as a possible mechanism of

late orogenic extension. The relationships we observe in

Fiordland suggest that late orogenic extension was caused

by changes in horizontal forces derived from changing plate

boundary dynamics (e.g. England and Houseman, 1989)

rather than a changing lower crustal rheology or an increase

in buoyancy forces arising from a thermal weakening of the

lower crust during convergence.

SG 1456—30/9/2003—14:58—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 23

ARTICLE IN PRESS

2465

2466

2467

2468

2469

2470

2471

2472

2473

2474

2475

2476

2477

2478

2479

2480

2481

2482

2483

2484

2485

2486

2487

2488

2489

2490

2491

2492

2493

2494

2495

2496

2497

2498

2499

2500

2501

2502

2503

2504

2505

2506

2507

2508

2509

2510

2511

2512

2513

2514

2515

2516

2517

2518

2519

2520

2521

2522

2523

2524

2525

2526

2527

2528

2529

2530

2531

2532

2533

2534

2535

2536

2537

2538

2539

2540

2541

2542

2543

2544

2545

2546

2547

2548

2549

2550

2551

2552

2553

2554

2555

2556

2557

2558

2559

2560

2561

2562

2563

2564

2565

2566

2567

2568

2569

2570

2571

2572

2573

2574

2575

2576

Page 24: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

10. Conclusions

A viscous, dry, mafic lower crust resulted in strong

vertical coupling between the upper and lower crusts during

Early Cretaceous contraction and orogenesis in western

New Zealand. Vertical coupling lasted from ,116–105 Ma

and may have begun by ,120 Ma. Transient vertical

decoupling of the crust coincided with the emplacement of

.10 km (thickness) of mafic–intermediate magma into the

lower crust during the interval ,126–120 Ma. The

development of a strong lower crustal root by ,116 Ma

reflected the following conditions and processes: (1) cooling

and crystallization of magma as pre-existing arc crust was

thrust above and below the WFO batholith; (2) extensive

dehydration of the lower crust at granulite facies conditions;

(3) the inherently high strength of mafic lower crustal rocks

as they underwent mineral reactions involving partial

melting; (4) low volumes of partial melt in the lower crust

controlled by the decomposition of hornblende ^

clinozoisite in mafic crust; and (5) the rapid extraction of

partial melt via melt-enhanced fracture networks, veins and

ductile shear zones. Rheological contrasts and a viscous

lower crust controlled the partitioning of deformation

vertically within the orogen and produced a narrow (50–

75 km wide), focused structural style in the upper crust.

High degrees of vertical coupling occurred even as partial

melts and magma moved out of the lower crust and were

emplaced into the middle and upper crusts. The results

imply that the onset of late orogenic extension at ,108–

105 Ma was caused by a change in plate boundary dynamics

rather than lateral flow of a weak lower crust during

convergence. The Fiordland exposures provide an important

natural example of the time-dependent, heterogeneous

nature of lower crustal strength and rheology.

11. Uncited References

Karlstrom and Williams, 1995. Landis et al., 1999.

Oliver, 1980. White and Jiang, 1994.

Acknowledgements

Funding to support this work was provided by National

Science Foundation funding to KAK (EAR-0087323);

Australian Research Council funding to KAK and GLC

(ARC-A10009053); and funding from the Geological

Society of America. We thank A. Tulloch for a helpful

review; and N. Mortimer, I. Turnbull, T. Rushmer and W.C.

Collins for discussions and assistance. We thank the

Department of Land Conservation in Te Anau for

permission to visit and sample localities in the Fiordland

National Park. J. Hollis and T. Ireland provided assistance in

compiling geochronologic data from northern Fiordland.

J. Stevenson, A. Papadakis, A. Claypool, S. Marcotte and

W.C. Simonson provided assistance in the field.

Appendix A. Analytical procedures for U–Pb isotopic

analyses

Zircons were analyzed at the University of Arizona with

a Micromass Isoprobe multicollector ICPMS equipped with

9 faraday collectors, an axial Daly detector, and 4 ion-

counting channels. The Isoprobe is equipped with a DUV

193 laser ablation system from New Wave Research. The

laser is a Compex 102 ArF Excimer laser, manufactured by

Lamda Physik, with an emission wavelength of 193 nm. The

analyses were conducted on 25–50 micron spots with an

output energy of ,32 mJ and a repetition rate of 8 Hz. Each

analysis consisted of one 30-s integration on the back-

grounds (on peaks with no laser firing) and twenty 1-s

integrations on peaks with the laser firing. The depth of each

ablation pit is ,20 microns. The collector configuration

allows simultaneous measurement of 204Pb in a secondary

electron multiplier while 206Pb, 207Pb, 208Pb, 232Th, and238U are measured with Faraday detectors. All analyses

were conducted in static mode.

Correction for common Pb for most samples was

performed by measuring 206Pb/204Pb, with the composition

of common Pb from Stacey and Kramers (1975) and

uncertainties of 1.0 for 206Pb/204Pb and 0.3 for 207Pb/204Pb.

For two of the samples analyzed (995a and 9928), 204Pb was

not measured reliably, and common Pb was accounted for

with a 207Pb correction (Ludwig, 2001).

Fractionation of 206Pb/238U and 206Pb/207Pb during

ablation was monitored by analyzing fragments of a large

concordant zircon crystal that has a known (ID-TIMS) age

of 564 ^ 4 Ma (2s; G.E. Gehrels, unpublished data).

Typically this reference zircon was analyzed once for every

four unknowns. The uncertainty arising from this calibration

correction contributed ,2% systematic error to the206Pb/238U and 206Pb/207Pb ages (2s level). The uncertainty

from decay constants and common Pb composition yielded

an additional ,1% error to each analysis.

The reported ages are based primarily on 206Pb/238U

ratios because the errors of the 207Pb/235U and 206Pb/207Pb

ratios are significantly greater (Fig. 8; Table 2). This is due

in large part to the low intensity (commonly ,1 mv) of the207Pb signal from these young grains. For each of the five

samples (Fig. 8), the age data are plotted on a Pb/U

concordia diagram, and the critical 206Pb/238U ages are

shown on a separate age plot. The final age calculations are

based on the weighted mean of the cluster of 206Pb/238U

ages, with the error expressed both as the uncertainty of this

mean and as the error of the age. The age error is based on

the quadratic sum of the weighted mean error and the

systematic error. Both are expressed at the 2s level.

SG 1456—30/9/2003—14:58—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx24

ARTICLE IN PRESS

2577

2578

2579

2580

2581

2582

2583

2584

2585

2586

2587

2588

2589

2590

2591

2592

2593

2594

2595

2596

2597

2598

2599

2600

2601

2602

2603

2604

2605

2606

2607

2608

2609

2610

2611

2612

2613

2614

2615

2616

2617

2618

2619

2620

2621

2622

2623

2624

2625

2626

2627

2628

2629

2630

2631

2632

2633

2634

2635

2636

2637

2638

2639

2640

2641

2642

2643

2644

2645

2646

2647

2648

2649

2650

2651

2652

2653

2654

2655

2656

2657

2658

2659

2660

2661

2662

2663

2664

2665

2666

2667

2668

2669

2670

2671

2672

2673

2674

2675

2676

2677

2678

2679

2680

2681

2682

2683

2684

2685

2686

2687

2688

Page 25: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

References

Allibone, A.H., Tulloch, A.J., 1997. Metasedimentary, granitoid and

gabbroic rocks from central Stewart Island, New Zealand. New Zealand

Journal of Geology and Geophysics 40, 53–68.

Antignano A., IV, 2002, Experimental constraints on granitoid compo-

sitions in convergent regimes: a geochemical study. M.Sc. Thesis,

University of Vermont.

Axen, G.J., Selverstone, J., Byrne, T., Fletcher, J.M., 1998. If the strong

crust leads, will the weak crust follow? GSA Today 8, 1–8.

Bishop, D.G., Bradshaw, J.D., Landis, C.A., 1985. Provisional terrain map

of South Island, New Zealand. In: Howell, D.G., Jones, D.L., Cox, A.,

Nur, A. (Eds.), Tectonostratigraphic Terranes of the Circum–Pacific

Region, Circum–Pacific Council for Energy and Resources, Houston,

Texas, pp. 512–522.

Blattner, P., 1976. Replacement of hornblende by garnet in granulite facies

assemblages near Milford Sound, New Zealand. Contributions to

Mineralogy and Petrology 55, 181–190.

Blattner, P., 1991. The North Fiordland transcurrent convergence. New

Zealand Journal of Geology and Geophysics 34, 543–553.

Blattner, P., Graham, I.G., 2000. New Zealand’s Darran Complex and

Mackay Intrusives—Rb/Sr whole-rock isochrons in the Median

Tectonic Zone. American Journal of Science 300, 603–629.

Bradshaw, J.D., 1989. Cretaceous geotectonic patterns in the New Zealand

region. Tectonics 8, 803–820.

Bradshaw, J.D., 1993. A review of the Median Tectonic Zone: terrane

boundaries and terrane amalgamation near the Median Tectonic Line.

New Zealand Journal of Geology and Geophysics 36, 117–125.

Bradshaw, J.Y., 1985. Geology of the northern Franklin Mountains,

northern Fiordland, New Zealand, with emphasis on the origin and

evolution of Fiordland granulites. Ph.D. thesis, University of Otago,

New Zealand.

Bradshaw, J.Y., 1989a. Origin and metamorphic history of an Early

Cretaceous polybaric granulite terrain, Fiordland, southwest New

Zealand. Contribution to Mineralogy and Petrology 103, 346–360.

Bradshaw, J.Y., 1989b. Early Cretaceous vein-related garnet granulite in

Fiordland, southwest New Zealand: a case for infiltration of mantle-

derived CO2-rich fluids. Journal of Geology 97, 697–717.

Bradshaw, J.Y., 1990. Geology of crystalline rocks of northern Fiordland:

details of the granulite facies Western Fiordland Orthogneiss and

associated rock units. New Zealand Journal of Geology and Geophysics

33, 465–484.

Bradshaw, J.Y., Kimbrough, D.L., 1989. Enriched: age constraints on

metamorphism and the development of a metamorphic core complex in

Fiordland, southern New Zealand; discussion. Geology 17, 380–381.

Bradshaw, J.Y., Kimbrough, D.L., 1991. Mid-Paleozoic age of granitoids in

enclaves within Early Cretaceous granulites, Fiordland, southwest New

Zealand. New Zealand Journal of Geology and Geophysics 34,

455–469.

Brown, E.H., 1996. High-pressure metamorphism caused by magma

loading in Fiordland. New Zealand. Journal of Metamorphic Geology

14, 441–452.

Burchfield, B.C., Chen, Z., Hodges, K.V., Liu, Y., Royden, L.H., Deng, C.,

Xu, J., 1992. The South Tibetan detachment system, Himalayan orogen:

extension contemporaneous with and parallel to shortening in a

collisional mountain belt. Geological Society of America Special

Paper 269, 1–41.

Burg, J.-P., Ford, M., 1997. Orogeny through time: an overview. In: Burg,

J,-P., Ford, M. (Eds.), Orogeny Through Time. Geological Society of

London, Special Publications 121, pp. 1–17.

Burg, J.-P., Brunel, M., Gapais, D., Chen, G.M., Li, G.H., 1984.

Deformation of leucogranites of the crystalline Main Central Sheet in

southern Tibet (China). Journal of Structural Geology 6, 535–542.

Clarke, G.L., Klepeis, K.A., Daczko, N.R., 2000. Cretaceous high-P

granulites at Milford Sound, New Zealand: metamorphic history and

emplacement in a convergent margin setting. Journal of Metamorphic

Geology 18, 359–374.

Claypool, A., Klepeis, K.A., Dockrill, B., Clarke, G.L., Zwingmann, H.,

Tulloch, A., 2002. Structure and kinematics of oblique continental

convergence in Northern Fiordland, New Zealand. Tectonophysics 359,

329–358.

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.

Cooper, R.A., 1979. Lower Paleozoic Rocks of New Zealand. Journal of the

Royal Society of New Zealand 9, 29–84.

Cooper, R.A., Tulloch, A.J., 1992. Early Paleozoic terranes in New Zealand

and their relationship to the Lachlan Fold Belt. Tectonophysics 214,

129–144.

Daczko, N.R., Klepeis, K.A., Clarke, G.L., 2001a. Evidence of Early

Cretaceous collisional-style orogenesis in northern Fiordland, New

Zealand and its effects on the evolution of the lower crust. Journal of

Structural Geology 23, 693–713.

Daczko, N.R., Clarke, G.L., Klepeis, K.A., 2001b. Transformation of two-

pyroxene hornblende granulite to garnet granulite involving simul-

taneous melting and fracturing of the lower crust, Fiordland, New

Zealand. Journal of Metamorphic Geology 19, 547–560.

Daczko, N.R., Klepeis, K.A., Clarke, G.L., 2002a. Thermomechanical

evolution of the crust during convergence and deep crustal pluton

emplacement in the Western Province of Fiordland, New Zealand.

Tectonics 21, 1–18.

Daczko, N.R., Stevenson, J.A., Clarke, G.L., Klepeis, K.A., 2002b.

Successive hydration and dehydration of high-P mafic granofels

involving clinopyroxene–kyanite symplectites, Mt. Daniel, Fiordland,

New Zealand. Journal of Metamorphic Geology 20, 669–682.

Daczko, N.R., Klepeis, K.A., Clarke, G.L., 2002c. Kyanite–paragonite-

bearing assemblages, northern Fiordland, New Zealand: rapid cooling

at the lower crustal root of a Cretaceous magmatic arc. Journal of

Metamorphic Geology 20, 887–902.

Davidson, C., Schmid, S.M., Hollister, L.S., 1994. Role of melt during

deformation in the deep crust. Terra Nova 6, 133–142.

Dell’Angelo, L.N., Tullis, J., 1988. Experimental deformation of partially

melted granitic aggregates. Journal of Metamorphic Geology 6,

495–516.

Ellis, S., Beaumont, C., Jamieson, R.A., Quinlan, G., 1998. Continental

collision including a weak zone: the vise model and its application to

the Newfoundland Appalachians. Canadian Journal of Earth Sciences

35, 1323–1346.

England, P., Houseman, G.A., 1989. Extension during continental

convergence, with application to the Tibetan Plateau. Journal of

Geophysical Research 94, 561–571.

Gibson, G.M., 1990. Uplift and exhumation of middle and lower crustal

rocks in an extensional tectonic setting, Fiordland, New Zealand. In:

Salisbury, M.H., Fountain, D.M. (Eds.), Exposed Cross-Sections of the

Continental Crust. NATO ASI Series 317, Dordrecht, pp. 71–101.

Gibson, G.M., Ireland, T.R., 1995. Granulite formation during continental

extension in Fiordland. Nature 375, 479–482.

Gibson, G.M., McDougall, I., Ireland, T.R., 1988. Age constraints on

metamorphism and the development of a metamorphic core complex in

Fiordland, southern New Zealand. Geology 16, 405–408.

Grindley, G.W., 1980. Geological map of New Zealand. Department of

Scientific and Industrial Research Sheet, Wellington, New Zealand

S13-Cobb, scale 1:63360.

Harrison, T.M., McDougall, I., 1980. Investigations of an intrusive contact,

northwest Nelson, New Zealand. Thermal, chronological and isotopic

constraints. Geochemica et Cosmochemica Acta 44, 1985–2003.

Harry, D.L., Oldow, J.S., Sawyer, D.S., 1995. The growth of orogenic belts

and the role of crustal heterogeneities in decollement tectonics.

Geological Society of America Bulletin 107, 1411–1426.

Hill, E.J., 1995. A deep crustal shear zone exposed in western Fiordland,

New Zealand. Tectonics 14, 1172–1181.

Hollis, J.A., Clarke, G.L., Klepeis, K.A., Daczko, N.R., Ireland, T.R., 2003.

SG 1456—30/9/2003—14:58—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 25

ARTICLE IN PRESS

2689

2690

2691

2692

2693

2694

2695

2696

2697

2698

2699

2700

2701

2702

2703

2704

2705

2706

2707

2708

2709

2710

2711

2712

2713

2714

2715

2716

2717

2718

2719

2720

2721

2722

2723

2724

2725

2726

2727

2728

2729

2730

2731

2732

2733

2734

2735

2736

2737

2738

2739

2740

2741

2742

2743

2744

2745

2746

2747

2748

2749

2750

2751

2752

2753

2754

2755

2756

2757

2758

2759

2760

2761

2762

2763

2764

2765

2766

2767

2768

2769

2770

2771

2772

2773

2774

2775

2776

2777

2778

2779

2780

2781

2782

2783

2784

2785

2786

2787

2788

2789

2790

2791

2792

2793

2794

2795

2796

2797

2798

2799

2800

Page 26: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

Geochronology and geochemistry of high-pressure granulites of the

Arthur River Complex, Fiordland, New Zealand: Cretaceous magma-

tism and metamorphism on the Palaeo–Pacific margin. Journal of

Metamorphic Geology 21, 299–313.

Hollis, J.A., Clarke, G.L., Klepeis, K.A., Daczko, N.R., Ireland, T.R., 2004.

U–Pb zircon geochronology of Cretaceous granulites from Fiordland,

New Zealand: rapid burial and uplift along the Mesozoic Pacific

Gondwana margin. Journal of Metamorphic Geology, in press.

Huerta, A.D., Royden, L.H., Hodges, K.V., 1996. The interdependence of

deformational and thermal processes in mountain belts. Science 273,

637–639.

Ireland, T.R., Gibson, G.M., 1998. SHRIMP monazite and zircon

geochronology of high-grade metamorphism in New Zealand. Journal

of Metamorphic Geology 16, 149–167.

Jamieson, R.A., Beaumont, C., Ellis, S., 1998. Partial melting in convergent

orogens; thermal constraints and mechanical consequences. Geological

Society of America Annual Meeting 30, 297.

Jiang, D., Williams, P.F., 1998. High-strain zones: a unified model. Journal

of Structural Geology 20, 1105–1120.

Karlstrom, K.E., Williams, M.L., 1995. The case of simultaneous

deformation, metamorphism and plutonism: an example from Proter-

ozoic rocks in central Arizona. Journal of Structural Geology 17,

59–81.

Karlstrom, K.E., Williams, M.L., 2002. Nature of the middle crust-

heterogeneity of structure and process due to pluton-enhanced

tectonism: an example from Proterozoic rocks of the North American

Southwest. In: Brown, M., Rushmer, T. (Eds.), Evolution and

Differentiation of the Continental Crust. Cambridge University Press,

in press.

Kimbrough, D.L., Tulloch, A.J., Coombs, D.S., Landis, C.A., Johnston,

M.R., Mattinson, J.M., 1994. Uranium-lead zircon ages from the

Median Tectonic Zone, New Zealand. New Zealand Journal of Geology

andGeophysics 37, 393–419.

Klepeis, K.A., Crawford, M.L., 1999. High temperature, arc, parallel

normal faulting at the roots of an obliquely convergent orogen. Geology

27, 7–10.

Klepeis, K.A., Daczko, N.R., Clarke, G.L., 1999. Kinematic vorticity and

tectonic significance of superposed mylonites in a major lower crustal

shear zone, northern Fiordland, New Zealand. Journal of Structural

Geology 21, 1385–1405.

Klepeis, K.A., Clarke, G.L., Rushmer, T., 2003. Magma transport and

coupling between deformation and magmatism in the continental

lithosphere. GSA Today 13, 4–11.

Lamarche, G., Collot, J.-Y., Wood, R.A., Sossom, M., Sutherland, R.,

Delteil, J., 1997. The Oligocene–Miocene Pacific–Australia plate

boundary, south of New Zealand: evolution from oceanic spreading to

strike-slip faulting. Earth and Planetary Science Letters 148, 129–139.

Landis, C., Coombs, D.S., 1967. Metamorphic belts and orogenesis in

southern New Zealand. Tectonophysics 4, 501–518.

Landis, C.A., Campbell, H.J., Aslund, T., Caywood, P.A., Douglas, A.,

Kimbrough, D.L., Pillai, D.D.L., Raine, J.I., Willsman, A., 1999.

Permian–Jurassic strata at Productus Creek, Southland, New Zealand:

implications for terrane dynamics of the eastern Gondwanaland margin.

New Zealand Journal of Geology and Geophysics 42, 255–278.

Lin, S., Jiang, D., Williams, P.F., 1998. Transpression (or transtension)

zones of triclinic symmetry: natural example and theoretical modelling.

In: Holdsworth, R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental

Transpressional and Transtensional Tectonics. Geological Society of

London, Special Publications 135, pp. 41–47.

Ludwig, K.J., 2001. Isoplot/Ex (rev. 2.49). Berkeley Geochronology Center

Special Publication No. 1a, 56pp.

Mattinson, J.L., Kimbrough, D.L., Bradshaw, J.Y., 1986. Western Fiord-

land orthogneiss: Early Cretaceous arc magmatism and granulite facies

metamorphism, New Zealand. Contributions to Mineralogy and

Petrology 92, 383–392.

Mayer, G., Mai, P.M., Plenefisch, T., Echtler, H., Luschen, E., Wehrle, V.,

Muller, B., Bonjer, K.P., Prodehl, C., Fuchs, K., 1997. The deep crust of

the southern Rhine Graben: reflectivity and seismicity as images of

dynamic processes. Tectonophysics 275, 15–40.

McCulloch, M.T., Bradshaw, J.Y., Taylor, S.R., 1987. Sm–Nd and Rb–Sr

isotopic and geochemical systematics in Phanerozoic granulites from

Fiordland, Southwest New Zealand. Contributions to Mineralogy and

Petrology 97, 183–195.

McKenzie, D., Nimmo, F., Jackson, J.A., 2000. Characteristics and

consequences of flow in the lower crust. Journal of Geophysical

Research 105, 11,029–11,046.

Miller, R.B., Paterson, S.R., 2001. Influence of lithological heterogeneity,

mechanical anisotropy, and magmatism on the rheology of an arc,

North Cascades, Washington. Tectonophysics 342, 351–370.

Molnar, P., England, P., Martinod, J., 1993. Mantle dynamics, uplift of the

Tibetan Plateau, and the Indian Monsoon. Reviews of Geophysics 31,

357–396.

Molnar, P., Anderson, H.J., Audoine, E., Eberhart-Phillips, D., Gledhill,

K.R., Klosko, E.R., Mcevilly, T.V., Okaya, D., Savage, M.K., Stern, T.,

Wu, F.T., 1999. Continuous deformation versus faulting through the

continental lithosphere of New Zealand. Science 286, 516–519.

Mortimer, N., Tulloch, A., 1996. The Mesozoic basement of New Zealand.

Geological Society of Australia Extended Abstracts 43, 391–399.

Mortimer, N., Tulloch, A.J., Spark, R., Walker, N., Ladley, E., Kimbrough,

D.L., Allibone, A.H., 1999a. Overview of the Median Batholith, New

Zealand: a new interpretation of the geology of the Median Tectonic

Zone and adjacent rocks. Journal of African Earth Sciences 29,

257–268.

Mortimer, N., Gans, P.B., Calvert, A., Walker, N., 1999b. Geology and

thermochronometry of the east edge of the Median Batholith (Median

Tectonic Zone); a new perspective on Permian to Cretaceous crustal

growth of New Zealand. The Island Arc 8, 404–425.

Muir, R.J., Ireland, T.R., Weaver, S.D., Bradshaw, J.D., 1994. Ion

microprobe U–Pb zircon geochronology of granitic magmatism in

the Western Province of the South Island, New Zealand. Chemical

Geology 113, 171–189.

Muir, R.J., Weaver, S.D., Bradshaw, J.D., Eby, G.N., Evans, J.A., 1995.

The Cretaceous Separation Point Batholith, New Zealand; granitoid

magmas formed by melting of mafic lithosphere. Journal of the

Geological Society of London 152, 689–701.

Muir, R.J., Weaver, S.D., Bradshaw, J.D., Eby, G.N., Evans, J.A., Ireland,

T.R., 1996. Geochemistry of the Karamea Batholith, New Zealand, and

comparisons with the Lachlan Fold Belt granites of SE Australia. Lithos

39, 1–20.

Muir, R.J., Ireland, T.R., Weaver, S.D., Bradshaw, S.D., Evans, J.A., Eby,

G.N., Shelley, D., 1998. Geochronology and geochemistry of a

Mesozoic magmatic arc system, Fiordland, New Zealand. Journal of

the Geological Society of London 155, 1037–1053.

Nathan, S., Thurlow, C., Warnes, P., Zucchetto, R., 2000. Geochronology

database for New Zealand rocks (2nd edition): 1961–1999. Institute of

Geological and Nuclear Sciences Report 11, p. 51.

Nemes, R., Neubauer, F., Cloetingh, S., Genser, J., 1997. The Klagenfurt

Basin in the Eastern Alps: an intra-orogenic decoupled flexural basin?

Tectonophysics 282, 189–203.

Norris, R.J., Cooper, A.F., 2001. Late Quaternary slip rates and slip

partitioning on the Alpine Fault, New Zealand. Journal of Structural

Geology 23, 507–520.

Oldow, J.S., Bally, A.W., Ave Lallemant, H.G., 1990. Transpression,

orogenic float and lithospheric balance. Geology 18, 991–994.

Oliver, G.J.H., 1977. Feldspathic hornblende and garnet granulites and

associated anorthosite pegmatites from Doubtful Sound, Fiordland,

New Zealand. Contributions to Mineralogy and Petrology 65, 111–121.

Oliver, G.J.H., 1980. Geology of the granulite and amphibolite facies

gneisses of Doubtful Sound, Fiordland, New Zealand. New Zealand

Journal Geology and Geophysics 1, 27–41.

Oliver, G.J.H., 1990. An exposed cross-section of continental crust,

Doubtful Sound, Fiordland, New Zealand; geophysical and geological

setting. In: Salisbury, M.H., Fountain, D.M. (Eds.), Exposed Cross-

SG 1456—30/9/2003—14:58—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx26

ARTICLE IN PRESS

2801

2802

2803

2804

2805

2806

2807

2808

2809

2810

2811

2812

2813

2814

2815

2816

2817

2818

2819

2820

2821

2822

2823

2824

2825

2826

2827

2828

2829

2830

2831

2832

2833

2834

2835

2836

2837

2838

2839

2840

2841

2842

2843

2844

2845

2846

2847

2848

2849

2850

2851

2852

2853

2854

2855

2856

2857

2858

2859

2860

2861

2862

2863

2864

2865

2866

2867

2868

2869

2870

2871

2872

2873

2874

2875

2876

2877

2878

2879

2880

2881

2882

2883

2884

2885

2886

2887

2888

2889

2890

2891

2892

2893

2894

2895

2896

2897

2898

2899

2900

2901

2902

2903

2904

2905

2906

2907

2908

2909

2910

2911

2912

Page 27: Processes controlling vertical coupling and …kklepeis/publications/Klepeis et al. 2004.pdfProcesses controlling vertical coupling and decoupling between the upper and lower crust

UNCORRECTED PROOF

sections of the Continental Crust. NATO ASI Series 317, Dordrecht, pp.

43–69.

Oliver, G.J.H., Coggon, J.H., 1979. Crustal structure of Fiordland, New

Zealand. Tectonophysics 54, 253–292.

Petford, N., Gallagher, K., 2001. Partial melting of mafic (amphibolitic)

lower crust by periodic influx of basaltic magma. Earth and Planeteary

Science Letters 193, 483–499.

Roering, C., van Reenen, D.D., Smit, C.A., Du Toit, R., 1995. Deep crustal

embrittlement and fluid flow during granulite metamorphism in the

Limpopo Belt, South Africa. Journal of Geology 103, 673–686.

Royden, L., 1996. Coupling and decoupling of crust and mantle in

convergent orogens: implications for strain partitioning in the crust.

Journal of Geophysical Research 101, 17,679–17,705.

Rushmer, T., 1995. An experimental deformation study of partially molten

amphibolite: Applications to low-melt fraction segregation. In: Brown,

M., Rushmer, T., Saywer, E.W. (Eds.), Mechanisms and Consequences

of Melt Segregation from Crustal Protoliths. Journal of Geophysical

Research 100, pp. 15,681–15,696.

Rutter, E.H., 1997. The influence of deformation on the extraction of crustal

melts: a consideration of the role of melt-assisted granular flow. In:

Holness, M. (Ed.), Deformation-enhanced Melt Segregation and

Metamorphic Fluid Transport. The Mineralogical Society Series 8,

pp. 82–110.

Rutter, E.H., Neumann, D.H.K., 1995. Experimental deformation of

partially molten Westerly granite under fluid-absent conditions, with

implications for the extraction of granitic magmas. In: Brown, M.,

Rushmer, T., Saywer, E.W. (Eds.), Mechanisms and Consequences of

Melt Segregation from Crustal Protoliths. Journal of Geophysical

Research 100, pp. 15,697–15,715.

Sisson, V.B., Pavlis, T.L., 1993. Geologic consequences of plate

reorganization; an example from the Eocene Southern Alaska fore

arc. Geology 21, 913–916.

Spear, S., 1993. Metamorphic Phase Equilibria and Pressure–Tempera-

ture–Time Paths. Mineralogical Society of America Monograph. pp.

1–799.

Spell, T.L., McDougall, I., Tulloch, A.J., 2000. Thermochronologic

constraints on the breakup of the Pacific Gondwana margin: the

Paparoa metamorphic core complex, South Island, New Zealand.

Tectonics 19, 433–451.

Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead isotope

evolution by a two-stage model. Earth and Planetary Science Letters 26,

207–221.

Sutherland, R., 1995. The Australia–Pacific boundary and Cenozoic plate

motions in the SW Pacific: some constraints from Geosat data.

Tectonics 14, 819–831.

Sutherland, R., Davey, F., Beavan, J., 2000. Plate boundary deformation in

South Island, New Zealand is related to inherited lithospheric structure.

Earth and Planetary Science Letters 177, 141–151.

Teyssier, C., Tikoff, B., Weber, J., 2002. Attachment between brittle and

ductile crust at wrenching plate boundaries. European Geophysical

Society Special Publication 1, 119–144.

Tulloch, A.J., 1979. Plutonic and metamorphic rocks of the Victoria Range

segment of the Karamea Batholith, southwest Nelson, New Zealand.

Ph.D. thesis, University of Otago, New Zealand.

Tulloch, A.J., Challis, G.A., 2000. Emplacement depths of Paleozoic–

Mesozoic plutons from western New Zealand estimated by hornblende-

Al geobarometry. New Zealand Journal of Geology and Geophysics 43,

555–567.

Tulloch, A.J., Kimbrough, D.L., 1989. The Paparoa metamorphic core

complex, New Zealand: Cretaceous extension associated with frag-

mentation of the Pacific margin of Gondwana. Tectonics 8, 1217–1234.

Tulloch, A.J., Kimbrough, D.L., 2003. Paired plutonic belts in convergent

margins and the development of high Sr/Y magmatism: the Peninsular

Ranges Batholith of California and the Median Batholith of New

Zealand. Geological Society of America Special Paper, in press.

Tulloch, A.J., Ireland, T.R., Walker, N.W., Kimbrough, D.L., 2000. U–Pb

zircon ages from the Milford Orthogneiss, Milford Sound, northern

Fiordland: Paleozoic igneous emplacement and Early Cretaceous

metamorphism. Institute of Geological and Nuclear Sciences Report

6, pp. 1–17.

Turnbull, I.M., 2000. Geology of the Wakatipu area. Institute of Geological

and Nuclear Sciences geologic map 18, 1 sheet þ pp. 72, Lower Hutt,

New Zealand, scale 1:250,000.

Vanderhaeghe, O., Teyssier, C., 1997. Formation of the Shuswap

metamorphic complex during late-orogenic collapse of the Canadian

Cordillera: Role of ductile thinning and partial melting of the mid- to

lower crust. Geodynamica Acta 10, 41–58.

Vanderhaeghe, O., Teyssier, C., 2001. Crustal-scale rheological transitions

during late orogenic collapse. Tectonophysics 335, 211–228.

Waight, T.E., Weaver, S.D., Muir, R.J., 1998. Mid-Cretaceous granitic

magmatism during the transition from subduction to extension in

southern New Zealand: a chemical and tectonic synthesis. Lithos 45,

469–482.

Warner, M., 1990. Basalts, water, or shear zones in the lower continental

crust? Tectonophysics 173, 163–174.

Wellman, H.W., 1953. Data for the study of recent and late Pleistocene

faulting in the South Island of New Zealand. New Zealand Journal of

Science and Technology B34, 270–288.

White, J.C., Jiang, D., 1994. High-strength ductile behavior of ancient

lower continental crust: evidence from natural deformation. EOS

Transactions of the American Geophysical Union 75, 329.

Wilks, K.R., Carter, N.L., 1990. Rheology of some continental lower

crustal rocks. Tectonophysics 182, 57–77.

Willett, S.D., 1998. Geodynamic modeling and insight into deep crustal

processes. Geological Society of America Annual Meeting 30, 243.

Williams, J.G., 1978. Eglinton Volcanics—stratigraphy, petrography and

metamorphism. New Zealand Journal of Geology and Geophysics 21,

713–732.

Williams, J.G., Harper, C.T., 1978. Age and status of the Mackay Intrusives

in the Eglinton–Upper Hollyford area. New Zealand Journal of

Geology and Geophysics 21, 733–742.

Wood, B.L., 1972. Metamorphosed ultramafites and associated formations

near Milford Sound, New Zealand. New Zealand Journal of Geology

and Geophysics 15, 88–128.

SG 1456—30/9/2003—14:58—SFORSTER—82496— MODEL 5 – br,ed,cor

K.A. Klepeis et al. / Journal of Structural Geology xx (0000) xxx–xxx 27

ARTICLE IN PRESS

2913

2914

2915

2916

2917

2918

2919

2920

2921

2922

2923

2924

2925

2926

2927

2928

2929

2930

2931

2932

2933

2934

2935

2936

2937

2938

2939

2940

2941

2942

2943

2944

2945

2946

2947

2948

2949

2950

2951

2952

2953

2954

2955

2956

2957

2958

2959

2960

2961

2962

2963

2964

2965

2966

2967

2968

2969

2970

2971

2972

2973

2974

2975

2976

2977

2978

2979

2980

2981

2982

2983

2984

2985

2986

2987

2988

2989

2990

2991

2992

2993

2994

2995

2996

2997

2998

2999

3000

3001

3002

3003

3004

3005

3006

3007

3008

3009

3010

3011

3012

3013

3014

3015

3016

3017

3018

3019

3020

3021

3022

3023

3024