29.12.2014 Views

title of the thesis - Department of Geology - Queen's University

title of the thesis - Department of Geology - Queen's University

title of the thesis - Department of Geology - Queen's University

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

FROM CESSATION OF SOUTH-DIRECTED MID-CRUST<br />

EXTRUSION TO ONSET OF OROGEN-PARALLEL EXTENSION,<br />

NW NEPAL HIMALAYA<br />

by<br />

Carl Nagy<br />

A <strong>the</strong>sis submitted to <strong>the</strong> <strong>Department</strong> <strong>of</strong> Geological Sciences and Geological Engineering<br />

In conformity with <strong>the</strong> requirements for<br />

<strong>the</strong> degree <strong>of</strong> Master’s <strong>of</strong> Science<br />

Queen’s <strong>University</strong><br />

Kingston, Ontario, Canada<br />

(September, 2012)<br />

Copyright © Carl Nagy, 2012


View <strong>of</strong> <strong>the</strong> nor<strong>the</strong>rn Karnali valley near <strong>the</strong> village <strong>of</strong> Chala. Dr. Borja Antolín is leading <strong>the</strong><br />

way, followed by our trusted guide Dawa Tamang. The foreground is Tethyan sedimentary<br />

sequence and <strong>the</strong> distant peaks are Greater Himalayan sequence. The contact between <strong>the</strong>se<br />

domains is roughly 200 meters towards <strong>the</strong> peaks from <strong>the</strong> photograph position.<br />

ii


Abstract<br />

Field mapping and, structural, microstructural, and chronological analyses confirm <strong>the</strong><br />

existence <strong>of</strong> a segment <strong>of</strong> <strong>the</strong> Gurla-Mandhata-Humla fault, an orogen-parallel strike-slip<br />

dominated shear zone in <strong>the</strong> upper Karnali valley <strong>of</strong> northwestern Nepal. This shear zone forms<br />

<strong>the</strong> upper contact <strong>of</strong>, and cuts obliquely across <strong>the</strong> Greater Himalayan Sequence (GHS). Data<br />

from this study reveal two phases <strong>of</strong> GHS deformation.<br />

Phase 1 is characterized by U-Th-Pb monazite crystallization ages (~26–12 Ma, peak<br />

~18–15 Ma), consistent with typical Neohimalayan metamorphic ages, and <strong>the</strong> final stages <strong>of</strong><br />

south-directed extrusion <strong>of</strong> <strong>the</strong> GHS.<br />

Phase 2 is characterized by south-dipping high-strain foliations and intensely developed<br />

ESE-WNW trending, shallowly plunging mineral elongation lineations, indicating orogen-parallel<br />

extension. Thermochronology <strong>of</strong> muscovite defining <strong>the</strong>se fabrics implies that <strong>the</strong> area was<br />

cooling and experiencing orogen-parallel extension by ~15–9 Ma. Mineral deformation<br />

mechanisms and quartz c-axis patterns <strong>of</strong> <strong>the</strong>se fabrics record a rapid increase in temperature<br />

from ~350°C along <strong>the</strong> shear zone, to ~650°C at ~2.5 structural km below <strong>the</strong> shear zone. Such<br />

temperature gradients may be remnants <strong>of</strong> telescoped and/or flattened iso<strong>the</strong>rms generated during<br />

south-directed extrusion <strong>of</strong> <strong>the</strong> GHS.<br />

Overprinting ESE-WNW fabrics record progressive deformation <strong>of</strong> <strong>the</strong> GHS at lower<br />

temperatures. Progressive deformation included a significant component <strong>of</strong> pure shear, as<br />

indicated by symmetric high-temperature quartz c-axis fabrics and a lower-temperature vorticity<br />

estimate (~59% pure shear). A transition in c-axis fabrics from type I to type II cross-girdles at ~<br />

1.2 km below <strong>the</strong> fault could indicate a transition from plane strain towards constriction.<br />

Toge<strong>the</strong>r, <strong>the</strong>se data suggest orogen-parallel extension was occurring as a result <strong>of</strong> transtension.<br />

This study reveals a transition from south-directed extrusion <strong>of</strong> <strong>the</strong> GHS to orogenparallel<br />

extension between ~15–13 Ma. Comparing <strong>the</strong>se data with tectonic events across <strong>the</strong><br />

iii


Himalaya reveals an orogen-wide middle Miocene transition, coeval with <strong>the</strong> uplift <strong>of</strong> eastern<br />

Tibet. This is consistent with interpretations invoking radial spreading <strong>of</strong> Tibet and east-directed<br />

lower-crustal flow to explain orogen-parallel extension. Our study leads to <strong>the</strong> suggestion that a<br />

transition affecting mid- to lower-crustal processes may be responsible for <strong>the</strong> cessation <strong>of</strong> southdirected<br />

extrusion <strong>of</strong> <strong>the</strong> GHS and onset <strong>of</strong> east-directed lower-crustal flow.<br />

iv


Co-Authorship<br />

This <strong>the</strong>sis is my own work. The following contributors are co-authors on Chapter 2: My<br />

supervisor, L. Godin, who provided scientific direction, thought-provoking discussion, and<br />

editorial assistance; J. Cottle who provided direction, expertise, and performed <strong>the</strong> final stage <strong>of</strong><br />

U-Th-Pb geochronology; D. Archibald who provided direction, expertise, and guided <strong>the</strong><br />

40 Ar/ 39 Ar <strong>the</strong>rmochronology; B. Antolín who provided valuable field assistance, discussion, and<br />

editorial comments.<br />

v


Acknowledgements<br />

It is with utmost pleasure that I am able to acknowledge those that have made <strong>the</strong> past<br />

years a success, both academically and personally.<br />

Firstly, I must thank Dr. Laurent Godin for <strong>the</strong> opportunity to work in <strong>the</strong> Himalaya and<br />

for guidance in <strong>the</strong> initial stages <strong>of</strong> my geologic career. His constant enthusiasm and support over<br />

<strong>the</strong> past years are sincerely appreciated. Borja Antolín is thanked for an exciting 2011 field<br />

season, numerous scientific discussions, and a thorough review <strong>of</strong> Chapter 2. I must acknowledge<br />

<strong>the</strong> work <strong>of</strong> John Cottle at <strong>the</strong> <strong>University</strong> <strong>of</strong> California, Santa Barbara, with whom I collaborated<br />

to perform <strong>the</strong> U-Th-Pb geochronology, and Doug Archibald, who guided <strong>the</strong> 40 Ar/ 39 Ar<br />

<strong>the</strong>rmochronology. I would also like to thank <strong>the</strong> following people for <strong>the</strong>ir scientific and<br />

analytical assistance: Brian Joy for guidance with electron microprobe analyses, Dan Gibson<br />

(Simon Fraser <strong>University</strong>) for his assistance with <strong>the</strong> presentation <strong>of</strong> geochronology results, Al<br />

Grant for assistance on <strong>the</strong> scanning electron microscope, Ashley K. Rudy for her master GIS<br />

assistance, and Martin Wong (Colgate <strong>University</strong>) for assistance and accommodation during<br />

electron backscatter diffraction analyses. Gratitude is also expressed towards my examining<br />

committee <strong>of</strong> Drs. Herb Helmstaedt and John Dixon for <strong>the</strong>ir constructive reviews <strong>of</strong> this <strong>the</strong>sis.<br />

Fieldwork was made exciting and possible by Dawa Tamang, <strong>the</strong> infamous Birval and too many<br />

o<strong>the</strong>r porters to name.<br />

I am beyond grateful for my family, and <strong>the</strong> friends I have made at <strong>Queen's</strong> <strong>University</strong><br />

and throughout Kingston. Our experiences toge<strong>the</strong>r have made <strong>the</strong> last two years much more than<br />

an academic success.<br />

This <strong>the</strong>sis was funded by a Natural Sciences and Engineering Research Council<br />

(NSERC) discovery grant to L. Godin, a NSERC MSc. Canadian Graduate Scholarship, an<br />

Ontario Graduate Scholarship in Science and Technology, and a Geological Society <strong>of</strong> America<br />

Student Research Grant awarded to C. Nagy.<br />

vi


Table <strong>of</strong> Contents<br />

Abstract .......................................................................................................................................... iii<br />

Co-Authorship ................................................................................................................................ v<br />

Acknowledgements ....................................................................................................................... vi<br />

Abbreviations .............................................................................................................................. xiii<br />

Chapter 1 Introduction.................................................................................................................. 1<br />

1.1 Introduction ............................................................................................................................ 1<br />

1.2 Evolution <strong>of</strong> <strong>the</strong> Himalayan-Tibetan Orogen ......................................................................... 4<br />

1.3 Transition from south-directed extrusion <strong>of</strong> <strong>the</strong> mid-crust to onset <strong>of</strong> orogen-parallel<br />

extension ...................................................................................................................................... 8<br />

1.4 This study ............................................................................................................................... 9<br />

1.4.1 <strong>Geology</strong> <strong>of</strong> northwestern Nepal and <strong>the</strong> upper Karnali valley ........................................ 9<br />

1.4.2 Purpose <strong>of</strong> this study ..................................................................................................... 10<br />

Chapter 2 From cessation <strong>of</strong> south-directed mid-crust extrusion to onset <strong>of</strong> orogen-parallel<br />

extension, NW Nepal Himalaya .................................................................................................. 13<br />

2.1 Abstract ................................................................................................................................ 13<br />

2.2 Introduction .......................................................................................................................... 14<br />

2.3 Geologic Setting................................................................................................................... 18<br />

2.3.1 Himalayan Overview .................................................................................................... 18<br />

2.3.1.1 South-directed extrusion <strong>of</strong> <strong>the</strong> GHS ..................................................................... 18<br />

2.3.1.2 Orogen-parallel extension ...................................................................................... 19<br />

2.4 <strong>Geology</strong> <strong>of</strong> <strong>the</strong> upper Karnali Valley ................................................................................... 20<br />

2.4.1 Background ................................................................................................................... 20<br />

2.4.2 This study ...................................................................................................................... 21<br />

2.4.2.1 Domain I – Folded TSS ......................................................................................... 23<br />

2.4.2.2 Domain II – Transposed TSS ................................................................................. 23<br />

2.4.2.3 Domain III – Ultra-high-strain zone ...................................................................... 25<br />

2.4.2.4 Domain IV – High-strain quartzite and pelitic GHS .............................................. 28<br />

2.4.2.5 Domain V – High-strain gneissic GHS .................................................................. 30<br />

2.4.2.6 Domain VI – Migmatitic GHS ............................................................................... 33<br />

2.5 Microstructural analyses ...................................................................................................... 35<br />

2.5.1 Crystallographic preferred orientation <strong>of</strong> quartz ........................................................... 35<br />

2.5.1.1 Methodology .......................................................................................................... 36<br />

vii


2.5.1.2 Descriptions <strong>of</strong> quartz CPO patterns ...................................................................... 38<br />

2.5.1.3 Quartz CPO deformation temperatures .................................................................. 45<br />

2.5.2 Rigid clast vorticity analysis ......................................................................................... 47<br />

2.5.3 Summary <strong>of</strong> microstructural analyses ........................................................................... 50<br />

2.6 Geochronology and <strong>the</strong>rmochronology ............................................................................... 52<br />

2.6.1 U-Th-Pb monazite geochronology ................................................................................ 53<br />

2.6.1.1 Methodology .......................................................................................................... 54<br />

2.6.1.2 Results .................................................................................................................... 55<br />

2.6.1.3 Summary <strong>of</strong> U-Th-Pb geochronology .................................................................... 69<br />

2.6.2 40 Ar/ 39 Ar <strong>the</strong>rmochronology .......................................................................................... 71<br />

2.6.2.1 Methodology .......................................................................................................... 72<br />

2.6.2.2 Results .................................................................................................................... 74<br />

2.6.2.3 Summary <strong>of</strong> 40 Ar/ 39 Ar <strong>the</strong>rmochronology .............................................................. 79<br />

2.7 Discussion ............................................................................................................................ 81<br />

2.7.1 South-directed extrusion <strong>of</strong> <strong>the</strong> GHS ............................................................................ 83<br />

2.7.1.1 Initial south-directed extrusion .............................................................................. 83<br />

2.7.1.2 Timing <strong>of</strong> south-directed extrusion ........................................................................ 84<br />

2.7.2 Orogen-parallel extension ............................................................................................. 85<br />

2.7.2.1 ESE-WNW deformation and exhumation .............................................................. 85<br />

2.7.2.2 Timing <strong>of</strong> ESE-WNW deformation and cooling ................................................... 87<br />

2.7.3 Transition from south-directed extrusion to orogen-parallel extension ........................ 88<br />

2.7.3.1 Upper Karnali valley .............................................................................................. 88<br />

2.7.3.2 Himalayan transition .............................................................................................. 90<br />

2.7.4 Tectonic model explaining orogen-parallel extension .................................................. 94<br />

2.8 Conclusions .......................................................................................................................... 97<br />

Chapter 3 Discussion ................................................................................................................... 99<br />

3.1 Summary .............................................................................................................................. 99<br />

3.1.1 Middle Miocene orogen-parallel extension and east-directed lower-crustal flow ........ 99<br />

3.2 Future research considerations ........................................................................................... 103<br />

3.2.1 Why did south-directed mid-crustal extrusion cease, and why did east-directed lower<br />

crustal flow initiate ............................................................................................................. 103<br />

3.2.2 Is <strong>the</strong> Karakoram fault linked to faults <strong>of</strong> <strong>the</strong> upper Karnali valley .......................... 104<br />

Chapter 4 Microstructural analyses: proposals for advancing and validating current<br />

techniques ................................................................................................................................... 106<br />

viii


4.1 Introduction ........................................................................................................................ 106<br />

4.2 Modern quartz CPO analyses: future research opportunities ............................................. 106<br />

4.2.1 Evolution <strong>of</strong> CPO fabrics during progressive strain development .............................. 109<br />

4.2.2 Lateral variations in strain during extrusion <strong>of</strong> <strong>the</strong> Greater Himalayan sequence ...... 112<br />

4.3 Evaluation <strong>of</strong> <strong>the</strong> rigid grain method for estimating kinematic vorticity numbers; future<br />

research opportunities .............................................................................................................. 113<br />

4.3.1 Analogue modeling <strong>of</strong> rigid clasts in a 3D flowing matrix ......................................... 117<br />

4.4 Conclusions ........................................................................................................................ 119<br />

Appendix A Station locations and structural data .................................................................. 120<br />

Appendix B Quartz CPO data .................................................................................................. 125<br />

Appendix C Monazite images and elemental maps ................................................................ 149<br />

Appendix D Muscovite 40 Ar/ 39 Ar <strong>the</strong>rmochronology ............................................................. 219<br />

References ................................................................................................................................... 263<br />

ix


List <strong>of</strong> Figures<br />

Figure 1.1 Color relief map <strong>of</strong> <strong>the</strong> Himalaya, Tibet, and surrounding area. .................................... 2<br />

Figure 1.2 <strong>Geology</strong> <strong>of</strong> <strong>the</strong> Himalaya and NW Nepal / SW Tibet. ................................................... 3<br />

Figure 1.3 Extensional and compressional structures in <strong>the</strong> Himalaya, Tibet, and NW Nepal / SW<br />

Tibet ................................................................................................................................................. 6<br />

Figure 2.1 <strong>Geology</strong> <strong>of</strong> <strong>the</strong> Himalaya and NW Nepal / SW Tibet. ................................................. 15<br />

Figure 2.2 Extensional and compressional structures in <strong>the</strong> Himalaya, Tibet, and NW Nepal / SW<br />

Tibet ............................................................................................................................................... 17<br />

Figure 2.3 Lithotectonic map <strong>of</strong> <strong>the</strong> upper Karnali valley ............................................................. 22<br />

Figure 2.4 Structures and rocks <strong>of</strong> <strong>the</strong> Tethyan Sedimentary sequence......................................... 24<br />

Figure 2.5 Structures and rocks <strong>of</strong> <strong>the</strong> ultra-high-strain zone ........................................................ 26<br />

Figure 2.6 Structures and rocks <strong>of</strong> <strong>the</strong> high-strain pelitic Greater Himalayan sequence ............... 29<br />

Figure 2.7 Structures and rocks <strong>of</strong> <strong>the</strong> gneissic and migmatitic Greater Himalayan sequence. .... 31<br />

Figure 2.8 Brittle structures <strong>of</strong> <strong>the</strong> upper Karnali valley ............................................................... 34<br />

Figure 2.9 Summary <strong>of</strong> EBSD data processing ............................................................................. 37<br />

Figure 2.10 Summary <strong>of</strong> quartz CPO fabrics ................................................................................. 39<br />

Figure 2.11 Samples analyzed for quartz CPOs. ........................................................................... 40<br />

Figure 2.12 Quartz CPO fabrics. .................................................................................................... 41<br />

Figure 2.13 C-axis fabric opening angles versus deformation temperature plot. .......................... 46<br />

Figure 2.14 Vorticity sample and analysis ..................................................................................... 49<br />

Figure 2.15 Summary <strong>of</strong> microstructural results ........................................................................... 51<br />

Figure 2.16 Monazite images and elemental maps ........................................................................ 56<br />

Figure 2.17 U-Th-Pb concordia and probability diagrams. ........................................................... 60<br />

Figure 2.18 Muscovite 40Ar/39Ar step-heating plateaus. ............................................................. 75<br />

Figure 2.19 Plots <strong>of</strong> muscovite cooling ages versus distance from shear zone. ............................ 80<br />

Figure 2.20 Summary <strong>of</strong> CPO, vorticity, <strong>the</strong>rmochronology and geochronology analyses .......... 82<br />

Figure 2.21 Transition from south-directed extrusion to orogen-parallel extension, upper Karnali<br />

valley. ............................................................................................................................................. 89<br />

Figure 2.22 Orogen-wide transition from south-directed extrusion to orogen-parallel extension . 92<br />

Figure 2.23 Numerical model <strong>of</strong> Himalyan-Tibetan lower crustal flow ........................................ 96<br />

Figure 3.1 Numerical model <strong>of</strong> Himalayan lower crustal flow ................................................... 100<br />

Figure 4.1 Summary <strong>of</strong> quartz CPO fabrics ................................................................................. 108<br />

Figure 4.2 Progressive development <strong>of</strong> quartz CPO fabrics ........................................................ 110<br />

x


Figure 4.3 Microscale CPO variability ........................................................................................ 111<br />

Figure 4.4 Deformation and vorticity analysis planes ................................................................. 115<br />

Figure 4.5 Analogue model <strong>of</strong> vorticity experiment .................................................................... 118<br />

xi


List <strong>of</strong> Tables<br />

Table 2-1 U-Th-Pb geochronologic data ....................................................................................... 61<br />

Table 2-2 40 Ar/ 39 Ar <strong>the</strong>rmochronological data ............................................................................... 77<br />

xii


Abbreviations<br />

Mineral<br />

Biotite<br />

Calcite<br />

Garnet<br />

Feldspar<br />

Muscovite<br />

Pyrite<br />

Sillimanite<br />

Tourmaline<br />

Quartz<br />

Abbreviation<br />

Bt<br />

Cal<br />

Grt<br />

Fsp<br />

Ms<br />

Py<br />

Sil<br />

Tur<br />

Qtz<br />

Element<br />

Argon<br />

Cerium<br />

Lanthanum<br />

Lead<br />

Neodymium<br />

Oxygen<br />

Phosphorus<br />

Thorium<br />

Uranium<br />

Yttrium<br />

Abbreviation<br />

Ar<br />

Ce<br />

La<br />

Pb<br />

Nd<br />

O<br />

P<br />

Th<br />

U<br />

Y<br />

Feature<br />

Ama Drime Massif<br />

Chuwa granite<br />

Greater Himalayan sequence<br />

Gurla Mandhata core complex<br />

Gurla Mandhata-Humla fault system<br />

Indus-Yalu suture zone<br />

Karakoram Fault<br />

Kiogar-Jungbwa mantle-type rocks<br />

Leo Pargil gneiss dome<br />

Main Boundary thrust<br />

Main Central Thrust<br />

Main Frontal Thrust<br />

Namche Barwa<br />

Nanga Parbat<br />

South Tibetan Detachment system<br />

Tethyan sedimentary sequence<br />

Thakkhola graben<br />

Abbreviation<br />

AD<br />

CG<br />

GHS<br />

GM<br />

GMH<br />

IYSZ<br />

KF<br />

KJO<br />

LP<br />

MBT<br />

MCT<br />

MFT<br />

NB<br />

NP<br />

STDS<br />

TSS<br />

TG<br />

xiii


Chapter 1<br />

Introduction<br />

1.1 Introduction<br />

The Himalaya-Karakoram mountain belt and adjacent Tibetan plateau are <strong>the</strong> result <strong>of</strong> <strong>the</strong><br />

ongoing convergence and collision <strong>of</strong> India and Asia, which initiated approximately 55-50<br />

million years ago (Green et al., 2008; Najman et al., 2010). The Tibetan plateau is <strong>the</strong> world’s<br />

largest and most elevated plateau, residing at a mean elevation <strong>of</strong> over 5 kilometers (Fielding et<br />

al., 1994), and spanning from <strong>the</strong> Kunlun Mountains in central China to <strong>the</strong> Himalayan front (Fig.<br />

1.1). The topographic front <strong>of</strong> <strong>the</strong> Tibetan plateau, also known as <strong>the</strong> Himalaya, lies between <strong>the</strong><br />

Indian shield to <strong>the</strong> south and <strong>the</strong> Indus-Yalu Suture to <strong>the</strong> north (Fig. 1.1). The mountain belt<br />

trends WNW-ESE, spanning 2500 kilometers between two structural syntaxes, Nanga Parbat in<br />

nor<strong>the</strong>rn Pakistan and Namche Barwa in sou<strong>the</strong>astern Tibet (Fig. 1.1; Hodges, 2000).<br />

The geology <strong>of</strong> <strong>the</strong> Himalaya consists <strong>of</strong> four distinct lithotectonic domains bounded by<br />

crustal-scale fault systems, all <strong>of</strong> which are laterally continuous across <strong>the</strong> length <strong>of</strong> <strong>the</strong> orogen<br />

(Fig. 1.2). From north to south, <strong>the</strong>se domains are <strong>the</strong> Tethyan sedimentary, Greater Himalayan,<br />

Lesser Himalayan, and Sub-Himalayan sequences. Their bounding fault systems include <strong>the</strong><br />

South Tibetan detachment system, <strong>the</strong> Main Central thrust, <strong>the</strong> Main Boundary thrust, and <strong>the</strong><br />

Main Frontal thrust (Fig. 1.2). This study focuses on an area <strong>of</strong> northwestern Nepal, within <strong>the</strong><br />

two nor<strong>the</strong>rnmost tectonostratigraphic domains, <strong>the</strong> Tethyan sedimentary sequence and Greater<br />

Himalayan sequence, <strong>the</strong>ir bounding faults, <strong>the</strong> South Tibetan Detachment system (STDS), and<br />

<strong>the</strong> Gurla Mandhata-Humla fault system (Fig. 1.2).<br />

1


Figure 1.1 Color relief map <strong>of</strong> <strong>the</strong> Himalaya, Tibet, and surrounding area constructed from <strong>the</strong><br />

ETOPO2v2 (2006) database with a Cylindrical Equidistant projection. The upper Karnali valley<br />

(black box) is <strong>the</strong> study area. Dashed line indicates nor<strong>the</strong>rn boundary <strong>of</strong> <strong>the</strong> Himalaya (IYSZ).<br />

Abbreviations: IYSZ – Indus-Yalu suture zone; KF – Karakoram Fault; NB - Namche Barwa;<br />

NP - Nanga Parbat.<br />

2


Figure 1.2 Simplified geologic and structural map, and cross-section <strong>of</strong> (a) <strong>the</strong><br />

Himalaya, modified after Yin (2006) and Hintersberger et al. (2010), and (b) and<br />

(c), <strong>the</strong> upper Karnali valley and surrounding regions, modified after Murphy<br />

and Copeland (2005), Robinson et al. (2006) and Yakymchuk and Godin (2012).<br />

Abbreviations: AD - Ama Drime Massif, CG – Chuwa granite, GM - Gurla<br />

Mandhata gneiss dome, GMH – Gurla Mandhata-Humla fault, IYSZ - Indus-<br />

Yalu suture zone, KF - Karakorum fault, KJO - Kiogar-Jungbwa ophiolitic and<br />

mantle-type rocks, LP - Leo Pargil gneiss dome, MBT - Main Boundary thrust,<br />

MCT - Main Central thrust, MFT - Main Frontal thrust, STDS - South Tibetan<br />

detachment system, TG - Thakkhola graben.<br />

3


1.2 Evolution <strong>of</strong> <strong>the</strong> Himalayan-Tibetan Orogen<br />

The majority <strong>of</strong> sou<strong>the</strong>rn Tibet and <strong>the</strong> nor<strong>the</strong>rn Himalaya comprise <strong>the</strong> Tethyan<br />

sedimentary sequence (TSS), a wedge <strong>of</strong> Cambrian to Eocene supracrustal sedimentary strata<br />

deposited on <strong>the</strong> nor<strong>the</strong>rn margin <strong>of</strong> <strong>the</strong> Indian craton (Gansser, 1964; Garzanti, 1999). These<br />

sedimentary rocks accommodated crustal shortening and were metamorphosed at low grades<br />

from <strong>the</strong> onset <strong>of</strong> collision through to <strong>the</strong> late Oligocene/early Miocene (Godin, 2003; Murphy<br />

and Yin, 2003). Significant thickening <strong>of</strong> <strong>the</strong> TSS occurred during Eohimalayan deformation<br />

from roughly 45 to 32 Ma, accompanied by metamorphism at high pressure and moderate<br />

temperature conditions (Hodges and Silverberg, 1988; Godin et al., 1999, 2001; Aikman et al.,<br />

2008; Kellett and Godin, 2009; Carosi et al., 2010).<br />

Laterally throughout <strong>the</strong> orogen, <strong>the</strong> Greater Himalayan sequence (GHS) typically crops<br />

out south <strong>of</strong> <strong>the</strong> TSS. The GHS is <strong>the</strong> high metamorphic grade and anatectic core <strong>of</strong> <strong>the</strong> Himalaya<br />

bounded on its base by <strong>the</strong> Main Central Thrust (MCT), a south-verging thrust fault, and on its<br />

upper surface by <strong>the</strong> South Tibetan detachment system (STDS), a low-angle top-to-<strong>the</strong>-north<br />

normal fault (Fig. 1.2; Burchfiel and Royden, 1985; Burchiel et al., 1992; Hodges, 2000).<br />

Neohimalayan metamorphism is recorded throughout <strong>the</strong> GHS from early to middle Miocene<br />

(Vannay and Hodges, 1996; Godin et al., 2001; Searle and Szulc, 2005; Cottle et al., 2009b), and<br />

is characterized by extensive anatexis and synchronous motion along <strong>the</strong> STDS and MCT (see<br />

Godin et al., 2006b for review). This metamorphism is contemporaneous with iso<strong>the</strong>rmal<br />

decompression <strong>of</strong> <strong>the</strong> GHS and south-directed extrusion during early to middle Miocene (Vannay<br />

and Hodges, 1996; Harris et al., 2004). These relationships prompted <strong>the</strong> development <strong>of</strong> a<br />

channel-flow model to explain <strong>the</strong> process through which <strong>the</strong> GHS was extruded (Nelson et al.,<br />

4


1996; Hodges et al., 2001; Beaumont et al., 2001, 2004, 2006; Jamieson et al., 2004; Godin et al.,<br />

2006a).<br />

The channel-flow model stipulates that ductile extrusion <strong>of</strong> <strong>the</strong> GHS occurred normal to<br />

<strong>the</strong> trend <strong>of</strong> <strong>the</strong> orogen towards <strong>the</strong> topographic front <strong>of</strong> <strong>the</strong> Himalaya via flow in a low-viscosity<br />

channel, potentially coupled to focused surface denudation (Beaumont et al., 2001). However,<br />

this process is controversial and adamantly debated (Grujic, 2006; Harrison, 2006; Harris, 2007;<br />

Kohn, 2008). Alternative models have employed wedge-type extrusion (Burchfiel and Royden,<br />

1985; Grujic et al., 1996; Kohn, 2008) and tectonic wedging (Yin, 2006; Webb et al., 2007, 2011)<br />

to explain <strong>the</strong> presence <strong>of</strong> <strong>the</strong> enigmatic GHS.<br />

By middle Miocene time, motion along <strong>the</strong> MCT and STDS waned, and shortening was<br />

accommodated south <strong>of</strong> <strong>the</strong> GHS, along <strong>the</strong> Main Boundary thrust and <strong>the</strong> Lesser Himalayan<br />

imbricate thrust belt (Fig. 1.2; Schelling, 1992; Meigs et al., 1995; Huyghe et al., 2001). From<br />

this time forward, deformation has progressed southward in a typical in-sequence foreland<br />

propagating style.<br />

The onset <strong>of</strong> orogen-parallel extension throughout <strong>the</strong> Himalaya and Tibetan plateau<br />

broadly coincides with <strong>the</strong> cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS (Jessup and Cottle,<br />

2010), and on-going orogen-perpendicular shortening in <strong>the</strong> foreland (Huyghe et al., 2001).<br />

Orogen-parallel extension is indicated by structures that accommodate extension and translation<br />

in an orientation roughly parallel to <strong>the</strong> Himalayan arc. These structures are typically east-west or<br />

ESE-WNW trending strike-slip faults and north-south or NNE-SSW striking normal faults.<br />

Extensional deformation has been active from middle Miocene up to present day and is<br />

commonly manifested throughout central and sou<strong>the</strong>rn Tibet as east-west oriented strike-slip<br />

faults and north-south oriented normal faults (Fig. 1.3; Armijo et al., 1986, 1989; Taylor et al.,<br />

5


Figure 1.3 Structures in <strong>the</strong> Himalaya and Tibet associated with orogen-parallel extension and<br />

north-south contraction. (a) Active structures in <strong>the</strong> Himalaya, modified after Styron et al.<br />

(2011) and references <strong>the</strong>rein. Red lines represent faults associated, at least in part, with<br />

orogen-parallel extension. The Main Frontal Thrust (black) is associated with ongoing N-S<br />

contraction. Numbers in grey, and italics, refer to initiation ages <strong>of</strong> structures, and sources,<br />

respectively. (B) and (c) Simplified structural map and cross-section <strong>of</strong> <strong>the</strong> upper Karnali and<br />

surrounding regions, constructed from SRTM DEM data sheets 53_06, 53_07, and modified<br />

after Murphy and Copeland (2005), Robinson et al. (2006) and Yakymchuk and Godin (2012).<br />

Structures in black and red are associated with south-directed extrusion <strong>of</strong> <strong>the</strong> GHS, and<br />

orogen-parallel extension, respectively. Abbreviations and sources: GMH – Gurla<br />

Mandhata-Humla fault, ITSZ – Indus-Tsangpo suture zone, KC – Kung Co rift,LK – Lopukangri rift, MCT - Main Central thrust, STDS - South Tibetan<br />

detachment system, TG – Thakkhola graben, TY – Tangra-Yumco rift, 1 – Phillips et al. (2004), 2 – Thiede et al. (2006), 3 – Murphy et al. (2002), 4 – Murphy<br />

and Copeland (2005), 5 – Garzione et al. (2003), 6 – Murphy et al. (2010), 7 – Kali et al. (2010), 8 – Lee et al. (2011), 9 - Williams et al. (2001), 10 – Dewane<br />

et al. (2006).<br />

6


2003; Taylor and Yin, 2009). These structures are currently limited to upper crustal levels (see<br />

Fig. 2 from Styron et al., 2011 for comprehensive review).<br />

Within <strong>the</strong> Himalaya, extensional deformation is expressed as rare strike-slip faults,<br />

normal faults, and migmatite-cored gneiss domes, all exhibiting orogen-parallel slip directions<br />

(Fig. 1.3; Nakata, 1989; Murphy et al., 2002, 2009; Murphy and Copeland, 2005; Thiede et al.,<br />

2006; Jessup et al., 2008; Li and Yin, 2008, Jessup and Cottle, 2010). Gneiss domes related to<br />

orogen-parallel extension (Fig. 1.3) are <strong>the</strong> Leo Pargil dome in <strong>the</strong> Western Himalaya (Thiede et<br />

al., 2006), <strong>the</strong> Gurla Mandhata core complex in <strong>the</strong> central-west Himalaya (Murphy et al., 2002),<br />

and <strong>the</strong> Ama Drime Massif in <strong>the</strong> central-east Himalaya (Jessup and Cottle, 2010 and references<br />

<strong>the</strong>rein). These domes exhume upper amphibolite to granulite facies rocks in settings that are<br />

kinematically and contemporaneously linked to extension throughout <strong>the</strong> Himalaya and Tibetan<br />

plateau. Note that <strong>the</strong> terms gneiss dome and core complex, used interchangeably in this study,<br />

refer to migmatite-cored domes, and do not strictly follow <strong>the</strong> definition <strong>of</strong> Eskola (1949).<br />

An array <strong>of</strong> mechanisms have been proposed to explain orogen-parallel deformation,<br />

including lateral extrusion <strong>of</strong> a rigid Tibet (Tapponnier et al., 1982), oroclinal bending <strong>of</strong> <strong>the</strong><br />

Himalayan arc (Klootwijk et al., 1985; Ratschbacher et al., 1994; Bendick and Bilham, 2001),<br />

gravitational instability <strong>of</strong> <strong>the</strong> Tibetan plateau (Molnar and Tapponnier, 1978; Molnar and Chen<br />

1983; Coleman and Hodges, 1995), right-lateral oblique convergence (McCaffrey and Nabelek,<br />

1998; Seeber and Pêcher, 1998), and radial spreading <strong>of</strong> <strong>the</strong> Tibetan plateau (England and<br />

Houseman, 1988; Copley and McKenzie, 2007; Copley 2008; Cook and Royden, 2008). To date,<br />

<strong>the</strong>re is little consensus on which model best explains orogen-parallel extension. The kinematics<br />

and spatial distribution <strong>of</strong> strike-slip and normal faults are commonly used to test predictions <strong>of</strong><br />

7


<strong>the</strong> various models. However, this approach is problematic as a variety <strong>of</strong> models make similar<br />

predictions (Chapters 2, 3).<br />

1.3 Transition from south-directed extrusion <strong>of</strong> <strong>the</strong> mid-crust to onset <strong>of</strong> orogenparallel<br />

extension<br />

The cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS and <strong>the</strong> onset <strong>of</strong> orogen-parallel<br />

extension expose a fundamental middle Miocene tectonic transition in Himalayan evolution (e.g.,<br />

Jessup and Cottle, 2010). Characterizing <strong>the</strong> kinematics, geographic distribution, and timing <strong>of</strong><br />

features intrinsic to this transition are key in determining <strong>the</strong> processes that govern it. Ultimately,<br />

<strong>the</strong> determination <strong>of</strong> <strong>the</strong>se processes has a direct bearing on <strong>the</strong> fashion in which <strong>the</strong> Himalaya<br />

and o<strong>the</strong>r large, hot collisional orogens are perceived and understood.<br />

Thorough constraints have been established for individual components <strong>of</strong> this transition,<br />

including cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> mid-crust and initiation <strong>of</strong> orogen-parallel<br />

extension; however, limited studies have focused on <strong>the</strong> transition itself. Such studies require<br />

geologic settings that expose rocks and fabrics associated with both <strong>the</strong> GHS and orogen-parallel<br />

deformation, such as <strong>the</strong> Leo Pargil gneiss dome, <strong>the</strong> Gurla Mandhata core complex, and <strong>the</strong> Ama<br />

Drime massif.<br />

A compilation <strong>of</strong> data from <strong>the</strong> Ama Drime massif and Everest region reveals a cessation<br />

<strong>of</strong> motion along <strong>the</strong> STDS between 16-12 Ma and exhumation <strong>of</strong> melt along <strong>the</strong> Ama Drime and<br />

o<strong>the</strong>r detachments in a setting kinematically linked to extension between 12-8 Ma (Jessup and<br />

Cottle, 2010). Mapping and chronology <strong>of</strong> <strong>the</strong> Gurla Mandhata core complex reveals east-west<br />

extension along an evolving low-angle normal-fault system, juxtaposing mid-crustal rocks above<br />

<strong>the</strong> TSS (Murphy et al., 2002). Orogen-parallel extension along this fault system may have<br />

initiated post 17 Ma, was potentially active by 13 Ma, and was certainly active by 9 Ma. Data<br />

8


from <strong>the</strong> Leo Pargil dome also reveal <strong>the</strong> exhumation and juxtaposition <strong>of</strong> mid-crustal rocks<br />

against <strong>the</strong> TSS via sustained normal displacement along <strong>the</strong> west-dipping western flank <strong>of</strong> <strong>the</strong><br />

dome (Thiede et al., 2006). Chronological constraints suggest dome exhumation initiated in a<br />

setting linked to orogen-parallel extension by 16 Ma (Thiede et al., 2006), and potentially post 19<br />

Ma (Leech, 2008).<br />

1.4 This study<br />

The purpose <strong>of</strong> this study is to characterize <strong>the</strong> cessation <strong>of</strong> south-directed mid-crust<br />

extrusion and <strong>the</strong> onset <strong>of</strong> orogen-parallel deformation. The area selected to perform this study is<br />

located within <strong>the</strong> upper Karnali valley <strong>of</strong> northwestern Nepal (Figs. 1.2b, 2.2b), where a<br />

northwest-striking, strike-slip-dominated shear zone defines <strong>the</strong> upper contact and cuts obliquely<br />

across <strong>the</strong> GHS (Murphy and Copeland, 2005).<br />

1.4.1 <strong>Geology</strong> <strong>of</strong> northwestern Nepal and <strong>the</strong> upper Karnali valley<br />

The upper Karnali valley lies to <strong>the</strong> northwest <strong>of</strong> <strong>the</strong> village <strong>of</strong> Simikot, located in farnorthwestern<br />

Nepal (Fig. 1.2b). The regional geological framework <strong>of</strong> <strong>the</strong> area was pioneered by<br />

Fuchs (1977) and Shrestha (1987), and subsequently refined by Murphy and Copeland (2005),<br />

Robinson et al. (2006), and Yakymchuk and Godin (2012).<br />

A compilation <strong>of</strong> data reveal a high-strain south-dipping contact, which juxtaposes<br />

weakly deformed and metamorphosed TSS over strongly deformed, high-metamorphic grade<br />

GHS (Fig. 1.2b; Murphy and Copeland, 2005). This lithotectonic relationship is consistent with<br />

<strong>the</strong> STDS, however <strong>the</strong> kinematics along <strong>the</strong> high-strain zone are not. Well-developed down-dip<br />

mineral elongation lineations indicating a significant dip-slip component are expected along <strong>the</strong><br />

STDS in order to juxtapose units <strong>of</strong> contrasting metamorphic grades. Contrastingly, shallow ESE-<br />

WNW lineations are observed along <strong>the</strong> high-strain zone, indicating a dominant strike-slip sense<br />

9


<strong>of</strong> shear (Murphy and Copeland, 2005). Additionally, this high strain zone is not limited to <strong>the</strong><br />

TSS-GHS interface, but also cuts obliquely across <strong>the</strong> GHS and possibly into <strong>the</strong> Lesser<br />

Himalayan sequence to <strong>the</strong> southwest (Fig. 1.2b; Murphy and Copeland, 2005). These<br />

observations, in conjunction with middle Miocene chronologic constraints, prompted <strong>the</strong><br />

interpretation that this contact is a component <strong>of</strong> <strong>the</strong> Gurla Mandhata-Humla fault system (GMH),<br />

a northwest-striking shear zone associated with orogen-parallel extension, postdating southdirected<br />

extrusion <strong>of</strong> <strong>the</strong> GHS (Murphy and Copeland, 2005).<br />

The presence <strong>of</strong> <strong>the</strong> GMH strike-slip-dominated fault system atop and along <strong>the</strong> GHS<br />

provides an ideal geologic context in which to investigate <strong>the</strong> relationship between south-directed<br />

extrusion <strong>of</strong> <strong>the</strong> GHS and orogen-parallel extension.<br />

1.4.2 Purpose <strong>of</strong> this study<br />

The purpose <strong>of</strong> this study is to characterize <strong>the</strong> transition from cessation <strong>of</strong> south-directed<br />

extrusion <strong>of</strong> <strong>the</strong> GHS to onset <strong>of</strong> orogen-parallel extension. Results from this <strong>the</strong>sis clarify this<br />

fundamental Himalayan tectonic transition and provide insight on large-scale processes governing<br />

<strong>the</strong> evolution <strong>of</strong> large, hot collisional orogens. This problem is approached by addressing <strong>the</strong><br />

following questions:<br />

1. What are <strong>the</strong> relationships between <strong>the</strong> structures, kinematics, temperatures, and<br />

timing <strong>of</strong> deformation associated with south-directed extrusion <strong>of</strong> <strong>the</strong> GHS and<br />

orogen-parallel extension<br />

2. What do <strong>the</strong>se data imply for tectonic processes governing (1) south-directed<br />

mid-crustal flow, (2) orogen-parallel extension, and (3) orogen-wide evolution <strong>of</strong><br />

<strong>the</strong> Himalayan-Tibetan orogen<br />

10


These questions are addressed through <strong>the</strong> following approaches: field mapping,<br />

structural, microstructural, and chronological analyses.<br />

Field mapping and sampling took place over two field seasons (Spring 2010, 2011), and<br />

focused on detailed transects across and along high strain zones <strong>of</strong> <strong>the</strong> upper Karnali valley (a<br />

short film <strong>of</strong> <strong>the</strong> expedition can be seen at: http://vimeo.com/47755261). Structurally-oriented<br />

mapping provided an initial framework upon which additional data were incorporated.<br />

Representative samples were taken throughout <strong>the</strong> field area. Most samples were oriented, and<br />

optimized for microstructural and chronological analyses.<br />

Microstructural analyses, including petrographic analyses, traditional and modern quartz<br />

crystallographic preferred orientation analyses, and vorticity analysis, provide qualitative to<br />

quantitative assessments <strong>of</strong> <strong>the</strong> temperatures and kinematics <strong>of</strong> deformation. Muscovite 40 Ar/ 39 Ar<br />

<strong>the</strong>rmochronology and monazite U-Th-Pb geochronology provide constraints on <strong>the</strong> absolute<br />

timing <strong>of</strong> crystallization and cooling ages across and along <strong>the</strong> high-strain zone. Results,<br />

interpretations, and consequences <strong>of</strong> this study are presented in Chapters 2, 3, 4, and Appendices<br />

A through D.<br />

Chapter 2 presents and syn<strong>the</strong>sizes all meaningful data and is written as a stand-alone<br />

manuscript to be submitted for publication in Geological Society <strong>of</strong> America Bulletin. Chapter 3<br />

incorporates conclusions <strong>of</strong> this study with previously published models to generate orogen-wide<br />

tectonic conclusions. Conclusions drawn in this chapter involve a degree <strong>of</strong> interpretation not yet<br />

suitable for publication and thus are not fully incorporated into Chapter 2. Additionally, this<br />

chapter provides suggestions on fur<strong>the</strong>r work necessary to validate <strong>the</strong> proposed tectonic models.<br />

11


Chapter 4 outlines proposed future research in <strong>the</strong> field <strong>of</strong> microstructural analyses,<br />

including a critique <strong>of</strong> vorticity analyses and additional applications <strong>of</strong> quartz crystallographic<br />

preferred orientation studies.<br />

Appendix A presents a comprehensive list <strong>of</strong> geologic stations, including station<br />

locations and elevation, and structural data. Appendix B presents all quartz crystallographic<br />

preferred orientation data, including samples not suitable for publication. Appendix C presents<br />

thin section maps locating monazite grains, as well as chemical maps, secondary electron and<br />

electron backscatter images <strong>of</strong> all grains. Appendix D presents data <strong>of</strong> all muscovite 40 Ar/ 39 Ar<br />

analyses, including additional analyses not suitable for publication.<br />

12


Chapter 2<br />

From cessation <strong>of</strong> south-directed mid-crust extrusion to onset <strong>of</strong><br />

orogen-parallel extension, NW Nepal Himalaya<br />

2.1 Abstract<br />

Field mapping and, structural, microstructural, and chronological analyses confirm <strong>the</strong><br />

existence <strong>of</strong> a segment <strong>of</strong> <strong>the</strong> Gurla-Mandhata-Humla fault, an orogen-parallel strike-slip<br />

dominated shear zone in <strong>the</strong> upper Karnali valley <strong>of</strong> northwestern Nepal. This shear zone forms<br />

<strong>the</strong> upper contact <strong>of</strong>, and cuts obliquely across <strong>the</strong> Greater Himalayan Sequence (GHS). Data<br />

from this study reveal two phases <strong>of</strong> GHS deformation.<br />

Phase 1 is characterized by U-Th-Pb monazite crystallization ages (~26–12 Ma, peak<br />

~18–15 Ma), consistent with typical Neohimalayan metamorphic ages, and <strong>the</strong> final stages <strong>of</strong><br />

south-directed extrusion <strong>of</strong> <strong>the</strong> GHS.<br />

Phase 2 is characterized by south-dipping high-strain foliations and intensely developed<br />

ESE-WNW trending, shallowly plunging mineral elongation lineations, indicating orogen-parallel<br />

extension. Thermochronology <strong>of</strong> muscovite defining <strong>the</strong>se fabrics implies that <strong>the</strong> area was<br />

cooling and experiencing orogen-parallel extension by ~15–9 Ma. Mineral deformation<br />

mechanisms and quartz c-axis patterns <strong>of</strong> <strong>the</strong>se fabrics record a rapid increase in temperature<br />

from ~350°C along <strong>the</strong> shear zone, to ~650°C at ~2.5 structural km below <strong>the</strong> shear zone. Such<br />

temperature gradients may be remnants <strong>of</strong> telescoped and/or flattened iso<strong>the</strong>rms generated during<br />

south-directed extrusion <strong>of</strong> <strong>the</strong> GHS.<br />

Overprinting ESE-WNW fabrics record progressive deformation <strong>of</strong> <strong>the</strong> GHS at lower<br />

temperatures. Progressive deformation included a significant component <strong>of</strong> pure shear, as<br />

13


indicated by symmetric high-temperature quartz c-axis fabrics and a lower-temperature vorticity<br />

estimate (~59% pure shear). A transition in c-axis fabrics from type I to type II cross-girdles at ~<br />

1.2 km below <strong>the</strong> fault could indicate a transition from plane strain towards constriction.<br />

Toge<strong>the</strong>r, <strong>the</strong>se data suggest orogen-parallel extension was occurring as a result <strong>of</strong> transtension.<br />

These data reveal a transition within <strong>the</strong> upper Karnali valley from south-directed<br />

extrusion <strong>of</strong> <strong>the</strong> GHS to onset <strong>of</strong> orogen-parallel extension between ~15 -13 Ma. Comparing<br />

<strong>the</strong>se data with similar tectonic environments across <strong>the</strong> Himalaya reveal an orogen-wide middle<br />

Miocene transition. The timing <strong>of</strong> this transition is consistent with interpretations invoking<br />

models <strong>of</strong> radial spreading <strong>of</strong> Tibet and east-directed lower crustal flow to explain orogen-parallel<br />

extension.<br />

2.2 Introduction<br />

Middle Miocene evolution <strong>of</strong> <strong>the</strong> Himalaya and Tibetan plateau is characterized by <strong>the</strong><br />

progression from south-directed extrusion <strong>of</strong> <strong>the</strong> Greater Himalayan Sequence (GHS) to orogenparallel<br />

extension (Jessup and Cottle, 2010). This transition period attests to a fundamental<br />

evolution <strong>of</strong> tectonic processes governing <strong>the</strong> architecture <strong>of</strong> <strong>the</strong> Himalayan-Tibetan orogen.<br />

The first phase <strong>of</strong> middle Miocene evolution is characterized by <strong>the</strong> cessation <strong>of</strong> southdirected<br />

extrusion <strong>of</strong> <strong>the</strong> GHS. The GHS is <strong>the</strong> high-metamorphic grade and anatectic core <strong>of</strong> <strong>the</strong><br />

Himalaya, typically cropping out at <strong>the</strong> crest <strong>of</strong> <strong>the</strong> Himalaya, along <strong>the</strong> length <strong>of</strong> <strong>the</strong> orogen (Fig.<br />

2.1). The GHS is commonly interpreted to represent a sequence <strong>of</strong> rocks extruded southward and<br />

normal to <strong>the</strong> orogen, from mid-crustal depths (Hodges, 2000; Yin and Harrison, 2000; Yin,<br />

2006). South-directed extrusion <strong>of</strong> <strong>the</strong> GHS was active up to ~ 16 Ma in a setting linked to northsouth<br />

contraction and shortening (see review <strong>of</strong> ages in Godin et al., 2006b; Cottle et al., 2009b).<br />

14


Figure 2.1 Simplified geologic and structural map, and cross-section <strong>of</strong> (a) <strong>the</strong><br />

Himalaya, modified after Yin (2006) and Hintersberger et al. (2010), and (b) and<br />

(c), <strong>the</strong> upper Karnali valley and surrounding regions, modified after Murphy and<br />

Copeland (2005), Robinson et al. (2006) and Yakymchuk and Godin (2012).<br />

Abbreviations: AD - Ama Drime Massif, CG – Chuwa granite, GM - Gurla<br />

Mandhata gneiss dome, GMH – Gurla Mandhata-Humla fault, IYSZ - Indus-Yalu<br />

suture zone, KF - Karakorum fault, KJO - Kiogar-Jungbwa ophiolitic and mantletype<br />

rocks, LP - Leo Pargil gneiss dome, MBT - Main Boundary thrust, MCT -<br />

Main Central thrust, MFT - Main Frontal thrust, STDS - South Tibetan<br />

detachment system, TG - Thakkhola graben.<br />

15


The second phase <strong>of</strong> middle Miocene evolution is characterized by <strong>the</strong> onset <strong>of</strong> orogenparallel<br />

extension throughout <strong>the</strong> Tibetan plateau and Himalayan arc (Allègre et al., 1984;<br />

Mercier et al., 1987; Styron et al., 2011). With <strong>the</strong> exception <strong>of</strong> <strong>the</strong> enigmatic Karakoram fault<br />

(Zhang et al., 2010), structures related to orogen-parallel extension initiated between ~ 16 -11 Ma<br />

(Fig. 2.2; see Fig. 2 from Styron et al., 2011 for a comprehensive age review). To date, <strong>the</strong>re is<br />

little consensus on <strong>the</strong> mechanisms that initiated and continue to drive this extension.<br />

In this study we investigate <strong>the</strong> transition between <strong>the</strong> cessation <strong>of</strong> south-directed<br />

extrusion <strong>of</strong> <strong>the</strong> GHS and onset <strong>of</strong> orogen-parallel extension. This is accomplished by integrating<br />

geological, structural, microstructural, and chronological data from <strong>the</strong> upper Karnali valley <strong>of</strong><br />

northwest Nepal (Figs. 2.1, 2.2). Within this valley, a northwest-striking, strike-slip dominated<br />

shear zone, <strong>the</strong> Gurla Mandhata-Humla fault system (GMH), crops out along a segment <strong>of</strong> GHS-<br />

TSS interface, and cuts obliquely across <strong>the</strong> GHS (Figs. 2.1, 2.2; Murphy and Copeland, 2005).<br />

This provides an opportune geologic context to investigate <strong>the</strong> relationship between southdirected<br />

extrusion <strong>of</strong> <strong>the</strong> GHS and ESE-WNW deformation associated with orogen-parallel<br />

extension. To date, this transition has only been constrained in <strong>the</strong> Everest region <strong>of</strong> <strong>the</strong> Himalaya<br />

(Jessup and Cottle, 2010).<br />

Based on our results from <strong>the</strong> upper Karnali valley, and <strong>the</strong> integration <strong>of</strong> our results with<br />

o<strong>the</strong>r similar transitions across <strong>the</strong> Himalaya, we attempt to reconcile <strong>the</strong> following questions:<br />

(1) What are <strong>the</strong> relationships between <strong>the</strong> structures, kinematics, temperatures, and<br />

timing <strong>of</strong> deformation associated with south-directed extrusion <strong>of</strong> <strong>the</strong> GHS and orogenparallel<br />

extension<br />

(2) What do <strong>the</strong>se data imply for <strong>the</strong> tectonic processes governing <strong>the</strong> onset <strong>of</strong> orogenparallel<br />

extension<br />

16


Figure 2.2 Structures in <strong>the</strong> Himalaya and Tibet associated with orogen-parallel extension<br />

and north-south contraction. (a) Active structures in <strong>the</strong> Himalaya, modified after Styron et<br />

al. (2011) and references <strong>the</strong>rein. Red lines represent faults associated, at least in part, with<br />

orogen-parallel extension. The Main Frontal Thrust (black) is associated with ongoing N-S<br />

contraction. Numbers in grey, and italics, refer to initiation ages <strong>of</strong> structures, and sources,<br />

respectively. (B) and (c) Simplified structural map and cross-section <strong>of</strong> <strong>the</strong> upper Karnali<br />

and surrounding regions, constructed from SRTM DEM data sheets 53_06, 53_07, and<br />

modified after Murphy and Copeland (2005), Robinson et al. (2006) and Yakymchuk and<br />

Godin (2012). Structures in black and red are associated with south-directed extrusion <strong>of</strong> <strong>the</strong><br />

GHS, and orogen-parallel extension, respectively. Abbreviations and sources: GMH – Gurla<br />

Mandhata-Humla fault, ITSZ – Indus-Tsangpo suture zone, KC – Kung Co rift,LK – Lopukangri rift, MCT - Main Central thrust, STDS - South Tibetan<br />

detachment system, TG – Thakkhola graben, TY – Tangra-Yumco rift, 1 – Phillips et al. (2004), 2 – Thiede et al. (2006), 3 – Murphy et al. (2002), 4 – Murphy<br />

and Copeland (2005), 5 – Garzione et al. (2003), 6 – Murphy et al. (2010), 7 – Kali et al. (2010), 8 – Lee et al. (2011), 9 - Williams et al. (2001), 10 – Dewane et<br />

al. (2006).<br />

17


2.3 Geologic Setting<br />

2.3.1 Himalayan Overview<br />

The geology <strong>of</strong> <strong>the</strong> Himalaya consists <strong>of</strong> four distinct lithotectonic domains bounded by<br />

crustal-scale fault systems, all <strong>of</strong> which are laterally continuous across <strong>the</strong> length <strong>of</strong> <strong>the</strong> orogen<br />

(Fig. 2.1a). From north to south, <strong>the</strong>se domains are: <strong>the</strong> Tethyan sedimentary, Greater Himalayan,<br />

Lesser Himalayan, and Sub-Himalayan sequences. Their bounding fault systems include <strong>the</strong><br />

South Tibetan detachment system, <strong>the</strong> Main Central thrust, <strong>the</strong> Main Boundary thrust, and <strong>the</strong><br />

Main Frontal thrust (Fig.2.1a). This study focuses on an area <strong>of</strong> northwestern Nepal within <strong>the</strong><br />

two nor<strong>the</strong>rnmost lithotectonic domains, <strong>the</strong> Tethyan sedimentary sequence (TSS) and Greater<br />

Himalayan sequence (GHS), and <strong>the</strong>ir bounding faults, <strong>the</strong> South Tibetan Detachment system<br />

(STDS) and <strong>the</strong> Gurla Mandhata-Humla (GMH) fault system (Figs.2.1b, 2.2b).<br />

2.3.1.1 South-directed extrusion <strong>of</strong> <strong>the</strong> GHS<br />

Laterally throughout <strong>the</strong> orogen, <strong>the</strong> GHS typically crops out south <strong>of</strong> <strong>the</strong> TSS and north<br />

<strong>of</strong> <strong>the</strong> Lesser Himalayan sequence. The GHS is <strong>the</strong> high-metamorphic grade and anatectic core <strong>of</strong><br />

<strong>the</strong> Himalaya. It is bounded at its base by <strong>the</strong> Main Central thrust (MCT), a south-verging thrust<br />

fault, and at its upper surface by <strong>the</strong> STDS, a low-angle, top-to-<strong>the</strong>-north normal fault (Fig 2.1;<br />

Burchfiel and Royden, 1985; Burchiel et al., 1992; Hodges, 2000). Neohimalayan metamorphism<br />

is recorded throughout <strong>the</strong> GHS from early to middle Miocene (Vannay and Hodges, 1996; Godin<br />

et al., 1999; 2001; Searle and Szulc, 2005; Cottle et al., 2009a,b), and is characterized by<br />

extensive anatexis and synchronous motion along <strong>the</strong> STDS and MCT (see Godin et al., 2006b<br />

for review <strong>of</strong> ages). This metamorphism is contemporaneous with iso<strong>the</strong>rmal decompression <strong>of</strong><br />

<strong>the</strong> GHS and south-directed extrusion during Early to middle Miocene (e.g. Vannay and Hodges,<br />

18


1996; Godin et al., 2001; Harris et al., 2004). Motion along <strong>the</strong> STDS terminated by ~16 Ma in<br />

central and eastern Nepal (Godin et al., 2006b; Cottle et al., 2009a,b), implying that extrusion <strong>of</strong><br />

<strong>the</strong> GHS had also ceased.<br />

2.3.1.2 Orogen-parallel extension<br />

Orogen-parallel extension has been active from middle Miocene up to present day<br />

throughout <strong>the</strong> Himalaya and Tibetan plateau (see Fig.2.2 from Styron et al., 2011 for<br />

comprehensive review). Extensional deformation is commonly manifested throughout Tibet as<br />

east-west to ESE-WNW oriented strike-slip faults and north-striking rifts (Fig. 2.2a; Armijo et al.,<br />

1986, 1989; Taylor et al., 2003; Taylor and Yin, 2009). The majority <strong>of</strong> <strong>the</strong>se structures are<br />

located within central to sou<strong>the</strong>rn Tibet and restricted to <strong>the</strong> upper crust (Armijo et al., 1986,<br />

1989; Taylor et al., 2003; Taylor and Yin, 2009). Within <strong>the</strong> Himalaya, extensional deformation<br />

is expressed as orogen-perpendicular normal faults, orogen-parallel strike-slip faults, and<br />

migmatite-cored gneiss domes (Fig 2.2a; Nakata, 1989; Murphy et al., 2002; 2009; Murphy and<br />

Copeland, 2005; Thiede et al., 2006; Jessup et al., 2008; Li and Yin, 2008; Jessup and Cottle,<br />

2010). Three Himalayan gneiss domes (Leo Pargil, Gurla Mandhata, and Ama Drime – Fig.<br />

2.2a,b) exhume rocks <strong>of</strong> upper amphibolite to granulite metamorphic facies in settings<br />

kinematically linked to orogen-parallel extension (Murphy et al, 2002, 2009; Thiede et al., 2006;<br />

Jessup and Cottle, 2010). Structures related to orogen-parallel extension initiated between ~ 16 -<br />

11 Ma (Fig 2.2; see Fig. 2 from Styron et al., 2011 for a comprehensive age review).<br />

19


2.4 <strong>Geology</strong> <strong>of</strong> <strong>the</strong> upper Karnali Valley<br />

2.4.1 Background<br />

The upper Karnali valley lies to <strong>the</strong> northwest <strong>of</strong> <strong>the</strong> village <strong>of</strong> Simikot, located in farnorthwestern<br />

Nepal (Figs. 2.1, 2.2). The regional geological framework <strong>of</strong> <strong>the</strong> area was pioneered<br />

by Fuchs (1977) and Shrestha (1987), and subsequently refined by Murphy and Copeland (2005),<br />

Robinson et al. (2006), and Yakymchuk and Godin (2012).<br />

First-order mapping reveals low-metamorphic-grade TSS juxtaposed along a southdipping<br />

contact over strongly deformed and metamorphosed rocks <strong>of</strong> <strong>the</strong> GHS (Figs. 2.1b, 2.2b;<br />

Murphy and Copeland, 2005). The TSS <strong>of</strong> <strong>the</strong> upper Karnali valley is a sequence <strong>of</strong> lowmetamorphic-grade<br />

rocks interpreted to be Cambrian through Mesozoic based on correlative units<br />

to <strong>the</strong> northwest (Cheng and Xu, 1987; Murphy and Yin, 2003; Murphy and Copeland, 2005).<br />

The GHS <strong>of</strong> <strong>the</strong> upper Karnali valley consists <strong>of</strong> amphibolite metamorphic facies schist<br />

and gneiss (Murphy and Copeland, 2005; Yakymchuk and Godin, 2012), consistent with high<br />

metamorphic grade and anatectic assemblages observed across <strong>the</strong> entirety <strong>of</strong> <strong>the</strong> GHS (Hodges,<br />

2000; Yin and Harrison, 2000; Yin, 2006). These assemblages imply that <strong>the</strong> GHS was exhumed<br />

from mid-crustal depths (Yakymchuk and Godin, 2012). However, <strong>the</strong> processes controlling this<br />

extrusion within northwestern Nepal remain contentious (Robinson et al., 2006; Yakymchuk and<br />

Godin, 2012).<br />

Along <strong>the</strong> length <strong>of</strong> <strong>the</strong> orogen, <strong>the</strong> STDS is a north-dipping structure, juxtaposing TSS to<br />

<strong>the</strong> north and GHS to <strong>the</strong> south (Fig. 2.1a). Within <strong>the</strong> upper Karnali valley, <strong>the</strong> same<br />

juxtaposition <strong>of</strong> units is observed; however, <strong>the</strong> contact is dips southward, with <strong>the</strong> TSS to <strong>the</strong><br />

south and <strong>the</strong> GHS to <strong>the</strong> north (Figs. 2.1b, 2.2b; Murphy and Copeland, 2005). Typical northdipping<br />

STDS structures and relationships are observed south <strong>of</strong> <strong>the</strong> upper Karnali valley, along<br />

20


<strong>the</strong> Seti River (Fig. 2.1b, 2.2b; Robinson et al., 2006). This geometry suggests that <strong>the</strong> TSS south<br />

<strong>of</strong> <strong>the</strong> upper Karnali valley and north <strong>of</strong> <strong>the</strong> Seti River forms an approximately 35 km wide<br />

synclinorium plunging to <strong>the</strong> northwest, and closing to <strong>the</strong> sou<strong>the</strong>ast (Fig. 2.1b). This may be due<br />

to crustal-scale buckling <strong>of</strong> <strong>the</strong> GHS following south-directed extrusion, as documented in central<br />

Nepal (Godin et al., 2006b) and northwest Bhutan (Antolín et al., 2012; Kellett and Grujic, 2012).<br />

Alternatively, <strong>the</strong> current geometry may be associated in part with <strong>the</strong> formation <strong>of</strong> <strong>the</strong> Gurla<br />

Mandhata core complex to <strong>the</strong> northwest (Fig. 2.1b, 2.2b; Murphy and Copeland, 2005).<br />

Mineral elongation lineations and asymmetric fabric elements <strong>of</strong> <strong>the</strong> upper Karnali valley<br />

reveal predominant ESE-WNW strike-slip dominated deformation (Murphy and Copeland, 2005).<br />

This contrasts with <strong>the</strong> typical top-to-<strong>the</strong>-north STDS structural fabrics, and thus prompted <strong>the</strong><br />

interpretation <strong>of</strong> this shear zone as a component <strong>of</strong> <strong>the</strong> normal, right-slip, WNW-striking Gurla<br />

Mandhata-Humla (GMH) fault system (Murphy and Copeland, 2005). This fault system trends<br />

roughly parallel to <strong>the</strong> strike <strong>of</strong> <strong>the</strong> orogen, from <strong>the</strong> Gurla Mandhata core complex sou<strong>the</strong>ast to<br />

<strong>the</strong> upper Karnali valley, where it forms <strong>the</strong> contact between <strong>the</strong> TSS and <strong>the</strong> GHS (Figs. 2.1,<br />

2.2). Fur<strong>the</strong>r eastward, <strong>the</strong> GMH branches south and sou<strong>the</strong>ast, along <strong>the</strong> GHS-TSS contact and<br />

obliquely across <strong>the</strong> GHS, respectively (Figs. 2.1, 2.2). The GMH is interpreted to accommodate<br />

arc-parallel extension and translation between 15 and 7 Ma (Murphy and Copeland, 2005).<br />

2.4.2 This study<br />

Detailed mapping <strong>of</strong> transects across and along <strong>the</strong> upper Karnali valley reveal six<br />

distinct lithotectonic domains. They are, from south to north, (I) folded TSS, (II) transposed TSS,<br />

(III) ultra-high-strain zone, (IV) high-strain quartzite and pelitic GHS, (V) high-strain gneissic<br />

GHS, and (VI) migmatitic GHS (Fig. 2.3). Deformation associated with <strong>the</strong> GMH is moderate to<br />

21


Figure 2.3 Lithotectonic map and simplified cross-section <strong>of</strong> <strong>the</strong> upper Karnali valley, northwestern<br />

Nepal. D.I, II, III, IV, V and VI refer to lithotectonic domains, see text for descriptions. Inset is identical<br />

scale to main map and shows simplified geology and topography, constructed from SRTM DEM data<br />

sheets 53_06, 53_07. Stereonets were constructed using OSX Stereonet v. 1.3 developed by N. Cardozo<br />

and R. Allmendinger. Abbreviations: GHS – Greater Himalayan sequence, GMH – Gurla Mandhata-<br />

Humla fault system, TSS – Tethyan sedimentary sequence.<br />

_____________________________________________________________________________<br />

pervasive throughout domains II, III, IV, and V. Macro- and micro-scale observations from all<br />

domains are described below (see Appendix A for station locations and additional field data).<br />

2.4.2.1 Domain I – Folded TSS<br />

The folded TSS is <strong>the</strong> sou<strong>the</strong>rnmost mapped domain, outcropping between ~2.7 to 1 km<br />

structurally above <strong>the</strong> GHS-TSS interface (Figs. 2.3, 2.4a,b,c). This domain comprises<br />

interlayered tan to grey marbles and calc-silicates (Cal ± Qtz ± Fsp ± phyllosilicates ± Py). Along<br />

<strong>the</strong> sou<strong>the</strong>rn edge <strong>of</strong> <strong>the</strong> mapped area, domain I exposes a layer <strong>of</strong> unknown thickness <strong>of</strong> dull<br />

green to purple calcareous phyllite (Cal ± Qtz ± phyllosilicates ± Fsp). Calcite is very fine<br />

grained (~30 µm) and no twins are discernable. This may suggest a complete lack <strong>of</strong> microscale<br />

deformation (Fig. 2.4c; Burkhard, 1993). Both rock types are interpreted to represent <strong>the</strong> Cambro-<br />

Ordovician stratigraphic level mapped by Murphy and Copeland (2005).<br />

Original bedding is crudely preserved and defines macroscale open to tight folds with<br />

axial planes moderately inclined to <strong>the</strong> south. Folds verge towards <strong>the</strong> nor<strong>the</strong>ast and fold hinges<br />

plunge shallowly (~0° to 25°) to <strong>the</strong> ESE and WNW (Figs. 2.3, 2.4a). Rare C/S/C’ fabrics<br />

indicating dextral shear are observed within <strong>the</strong> purple calcareous phyllite (Fig. 2.4b).<br />

2.4.2.2 Domain II – Transposed TSS<br />

The transposed TSS is located within ~1 km structurally above <strong>the</strong> GHS-TSS interface<br />

(Figs. 2.3, 2.4d,e,f,g). Throughout this domain, bedding <strong>of</strong> <strong>the</strong> folded TSS is transposed planar to<br />

23


Figure 2.4 Representative structures, rock types and deformation mechanisms <strong>of</strong> domain I and II; <strong>the</strong><br />

folded, and transposed Tethyan Sedimentary sequence, respectively. They are from highest to lowest<br />

structural levels: (a) Nor<strong>the</strong>ast-verging moderately overturned folds <strong>of</strong> marble and calc-silicate <strong>of</strong> domain<br />

I, south <strong>of</strong> <strong>the</strong> village <strong>of</strong> Tumkot. (b) C/S/C’ fabrics indicating dextral shear within a purple calcareous<br />

phyllite <strong>of</strong> domain I. (c) Fine-grained calcite with no discernable twins within a calc-silicate <strong>of</strong> domain I.<br />

(d) Recumbant and isoclinal folds <strong>of</strong> domain II. (e) Calc-silicate <strong>of</strong> domain II exhibiting thin type I calcite<br />

twins (red arrows). (f) High-strain foliation (S 1 ) within calc-silicate <strong>of</strong> domain II. (g) Calc-silicate <strong>of</strong><br />

domain II exhibiting tabular type II calcite twins (thin red arrows) and curved and tapered type III calcite<br />

twins (thick red arrows). Inset is identical scale from elsewhere in <strong>the</strong> thin section. Abbreviations: cal -<br />

calcite, qtz - quartz.<br />

24


<strong>the</strong> main shear zone fabric. The lithology <strong>of</strong> this domain is identical to <strong>the</strong> overlying folded TSS,<br />

comprising interlayered tan to grey marbles and calc-silicates (Cal ± Qtz ± Fsp ± phyllosilicates ±<br />

Py).<br />

At upper structural levels within this domain, broad open folds <strong>of</strong> domain I transition to<br />

recumbent and isoclinal (Fig. 2.4d). Fold hinges plunge shallowly to moderately and trend ESE-<br />

WNW (Fig. 2.3). Axial planes <strong>of</strong> <strong>the</strong>se folds are roughly parallel to <strong>the</strong> main shear zone foliation.<br />

Calcite is deformed and shows thin type I twins, and rare type II twins (Fig. 2.4e). This implies<br />

deformation temperatures <strong>of</strong> less than ~200°C (Burkhard 1993; Ferrill et al., 2004). Quartz grains<br />

within calc-silicate layers are both undeformed and show patchy undulose extinction, suggesting<br />

temperatures <strong>of</strong> deformation below which quartz dynamically recrystallizes (


Figure 2.5 Representative structures, rock types and deformation mechanisms <strong>of</strong> domain III, <strong>the</strong> ultrahigh-strain<br />

zone. They are from highest to lowest structural levels: (a) Angular fragments <strong>of</strong> marble and<br />

calcite (outlined in white) within a chaotic carbonate matrix. (b) Well-rounded clasts <strong>of</strong> quartz, feldspar<br />

and phyllosilicate within a matrix <strong>of</strong> fine-grained calcite. (c) High-strain south-dipping tectonic foliation<br />

(S 1 ) within a pelitic schist. (d) High-strain schist composed <strong>of</strong> feldspar porphyroclasts exhibiting bent<br />

twins, patchy undulose extinction, and abundant grain-scale fractures, including anti<strong>the</strong>tic micro-faulting<br />

(red arrows), within a matrix dominated by fine-grained dynamically recrystallized quartz. (e) Dextrally<br />

sheared mica fish within a matrix <strong>of</strong> high-strain dynamically recrystallized quartz (viewed to <strong>the</strong> south). (f)<br />

Intensely attenuated and dynamically recrystallized quartz grains exhibiting subgrain extinction and<br />

bulging (red arrows). Dotted red line shows length <strong>of</strong> a single elongate quartz grain. Abbreviations: cal -<br />

calcite, fld - feldspar, ms – muscovite, qtz – quartz.<br />

26


South <strong>of</strong> <strong>the</strong> interface, <strong>the</strong> TSS is composed <strong>of</strong> mm- to cm-thick, interlayered dark grey,<br />

green, brown, and tan marbles and calc-slicates (Cal ± Qtz ± Fsp ± phyllosilicates). Calcite is<br />

very fine grained and deformation and recrystallization textures are not discernable. Flattened and<br />

elongate layers <strong>of</strong> carbonate minerals define <strong>the</strong> pervasive south-dipping foliation. Hinge lines <strong>of</strong><br />

micr<strong>of</strong>olds along foliation surfaces define a shallowly plunging ESE-WNW trending crenulation<br />

lineation. South <strong>of</strong> <strong>the</strong> village <strong>of</strong> Puiya (Fig. 2.3), this crenulation lineation overprints an older,<br />

moderately plunging, SE-NW trending mineral lineation defined by aligned mafic minerals.<br />

Within 10 m superjacent to <strong>the</strong> interface, angular fragments <strong>of</strong> marble and calcite are<br />

observed within a chaotic carbonate matrix on a centimeter scale (Fig. 2.5a), interpreted to be<br />

related to low temperature cataclastic deformation.<br />

Within 1 m <strong>of</strong> <strong>the</strong> interface, microscale observations <strong>of</strong> high-strain marble show wellrounded<br />

clasts <strong>of</strong> quartz, feldspar and phyllosilicate within a matrix <strong>of</strong> fine-grained calcite (Fig.<br />

2.5b). These clasts are internally deformed, showing undulatory extinction and internal brittle<br />

fracturing. Based on <strong>the</strong>se attributes, we interpret <strong>the</strong> clasts to have been incorporated into <strong>the</strong><br />

marble from <strong>the</strong> underlying GHS and rotated extensively during subsequent deformation.<br />

North <strong>of</strong> <strong>the</strong> GHS-TSS interface, <strong>the</strong> GHS consists <strong>of</strong> quartzite and pelitic schist (Qtz ±<br />

Fsp ± Ms ± Bt ± Grt). Mylonitic to ultramylonitic fabrics define <strong>the</strong> pervasive south-dipping<br />

tectonic foliation (Fig. 2.5c,d,e). Pervasive mineral lineations defined by elongation <strong>of</strong> quartz,<br />

feldspar, and mica plunge shallowly to moderately towards <strong>the</strong> ESE-WNW. A single leucogranite<br />

dyke (Fsp + Qtz + Ms + Bt ± Tur) crosscuts <strong>the</strong> ultra-high-strain zone north <strong>of</strong> <strong>the</strong> village <strong>of</strong><br />

Chala. The dyke is semi-concordant to <strong>the</strong> main foliation and exhibits a protomylonitic texture,<br />

suggestive <strong>of</strong> syndeformational intrusion.<br />

27


Quartz grains are dynamically recrystallized and intensely attenuated into elongate<br />

ribbons (Fig. 2.5e,f). Relict grains show subgrain extinction, bulges, and recrystallized subgrains<br />

and bulges (Fig. 2.5f). These textures suggest deformation temperatures approximating <strong>the</strong><br />

transition from bulging to subgrain rotation recrystallization at ~400°C (Stipp et al., 2002).<br />

Feldspar grains exhibit bent twins, patchy undulose extinction, and abundant grain-scale<br />

fractures, including anti<strong>the</strong>tic micro-faulting (Fig. 2.5d). Brittle deformation within <strong>the</strong> feldspar<br />

suggests a temperature <strong>of</strong> deformation below ~400°C (Tullis and Yund, 1991). Coeval dynamic<br />

recrystallization <strong>of</strong> quartz and brittle deformation <strong>of</strong> feldspar constrain deformation temperatures<br />

to 280 -400°C within <strong>the</strong> ultra-high-strain zone (Stipp et al., 2002; Tullis and Yund, 1991).<br />

Asymmetric fabric elements observed within <strong>the</strong> ultra-high-strain zone include delta-type<br />

rotated feldspar porphyroclasts and mica fish (Fig. 2.5d,e). Shear sense indicators show a near<br />

even distribution <strong>of</strong> dextral and sinistral senses <strong>of</strong> motion. Within <strong>the</strong> leucogranite dyke,<br />

numerous conjugate shear bands are observed. Additional microscale observations include garnet<br />

and tourmaline grains that are fractured, boudinaged, and pulled apart parallel to <strong>the</strong> dominant<br />

mineral elongation lineation.<br />

2.4.2.4 Domain IV – High-strain quartzite and pelitic GHS<br />

Highly strained quartzite and pelitic schist <strong>of</strong> <strong>the</strong> GHS are found ~1 km structurally<br />

underlying <strong>the</strong> GHS-TSS interface (Figs. 2.3, 2.6). The mineralogy <strong>of</strong> <strong>the</strong> quartzite and pelitic<br />

schist (Qtz ± Fsp ± Ms ± Bt ± Grt) is similar to <strong>the</strong> overlying ultra-high-strain GHS rocks;<br />

however, this domain has been subjected to lower deformation at higher temperatures. Mylonitic<br />

fabrics and mineral lineations defined by alignment and elongation <strong>of</strong> quartz, feldspar, and mica<br />

define a pervasive south-dipping foliation and shallowly to moderately plunging ESE-WNW<br />

28


WNW<br />

Figure 2.6 Representative structures, rock types and deformation mechanisms <strong>of</strong> domain IV, <strong>the</strong><br />

high-strain quartzite and pelitic Greater Himalayan sequence. They are from highest to lowest<br />

structural levels: (a) C/S/C’ fabrics and a sigma-type feldspar porphyroclast indicating dextral<br />

shear within high-strain pelitic schist. Note, late stage warping <strong>of</strong> <strong>the</strong> outcrop reoriented sample<br />

N-S. Orientation at which fabrics were formed is interpreted to be ESE-WNW. (b) C/S/C’ fabrics<br />

indicating sinistral shear within high-strain pelitic schist (red arrows indicate C’ planes). (c)<br />

Intensely fractured and pulled apart garnets parallel to an east-trending mineral elongation<br />

lineation, and C/S/C’ fabrics indicating sinistral shear within high-strain quartzite (red arrows<br />

indicate C’ planes). (d) Flattened and dynamically recrystallized quartz grains exhibiting<br />

extensive subgrain boundaries (red arrows). Abbreviations: fld - feldspar, grt – garnet, ms –<br />

muscovite, qtz – quartz.<br />

29


trending lineation (Fig. 2.3). Discrete pods and rare crosscutting to semi-concordant leucogranite<br />

dykes constitute < 10 vol% <strong>of</strong> <strong>the</strong> domain.<br />

Strongly flattened quartz grains exhibit core and mantle structures, recrystallized<br />

subgrains, and subgrain extinction (Fig. 2.6d), indicating deformation temperatures <strong>of</strong> ~400-<br />

500°C (Stipp et al., 2002). Garnets are inclusion-rich, intensely fractured and pulled apart parallel<br />

to <strong>the</strong> mineral elongation lineation (Fig. 2.6c).<br />

Asymmetric fabric elements, including C/S/C’ fabrics and sigma-type feldspar<br />

porphyroclasts, are abundant and well developed within high-strain quartzite and pelitic schist<br />

(Fig. 2.6a,b). Shear sense indicators show a near even distribution <strong>of</strong> dextral and sinistral senses<br />

<strong>of</strong> motion.<br />

2.4.2.5 Domain V – High-strain gneissic GHS<br />

High-strain gneiss <strong>of</strong> <strong>the</strong> GHS is exposed approximately 1 to 6 km structurally below <strong>the</strong><br />

GHS-TSS interface, and spans ESE-WNW across <strong>the</strong> extent <strong>of</strong> <strong>the</strong> field area (Figs. 2.3, 2.7).<br />

Rocks within <strong>the</strong> structurally uppermost section <strong>of</strong> this domain are highly strained, and<br />

progressively decrease in strain and increase in volume percent <strong>of</strong> melt with increasing structural<br />

depth. The high-strain gneiss (Fsp + Qtz ± Bt ± Ms ± Grt ± Tur ± Sil) contains up to 30 vol%<br />

leucosome and is fairly homogeneous, with <strong>the</strong> exception <strong>of</strong> rare meter sized quartz/quartzite<br />

layers and a ~100 m thick layer <strong>of</strong> pelitic schist (Qtz + Bt + Ms + Sil + Grt).<br />

Structurally above this pelitic schist, to east <strong>of</strong> <strong>the</strong> village <strong>of</strong> Yangar, a mesoscale low<br />

strain zone exposes augen gneiss <strong>of</strong> a similar mineralogy to <strong>the</strong> dominant high-strain gneiss. This<br />

may suggest that <strong>the</strong> augen gneiss is <strong>the</strong> protolith to <strong>the</strong> high-strain gneissic GHS.<br />

Garnets are inclusion-rich and intensely fractured but not consistently stretched parallel<br />

to <strong>the</strong> mineral elongation lineation. Sillimanite in <strong>the</strong> form <strong>of</strong> fibrolite is abundant within <strong>the</strong><br />

30


Figure 2.7 Representative structures, rock types and deformation mechanisms <strong>of</strong> domain V and VI; <strong>the</strong><br />

high-strain gneissic, and migmatitic Greater Himalayan sequence, respectively. They are from highest to<br />

lowest structural levels: (a) Delta-type rotated leucogranite pod indicating dextral shear within domain V.<br />

(b) Feldspar (fld) porphyroclasts bordered by micas and dynamically recrystallized quartz, in a core-andmantle<br />

sigma-type structure indicating dextral shear, within high-strain gneiss <strong>of</strong> domain V. (c) Sinistral<br />

shear band (red arrows) faulting and drawing into plane sillimanite (sil) and biotite (bt), within pelitic<br />

schist <strong>of</strong> domain V. (d) Quartz (qtz) grains showing interlobate to amoeboid quartz grain boundaries and<br />

subgrain extinction (red arrows), within gneissic GHS <strong>of</strong> domain V. (e) Chessboard-style extinction within<br />

quartz <strong>of</strong> domain VI. (f) Equigranular quartz and feldspar grains <strong>of</strong> domain VI, exhibiting uniform<br />

extinction, and interlobate grain boundaries.<br />

31


pelitic layer and rare throughout <strong>the</strong> gneiss. Within <strong>the</strong> pelitic layer, abundant shear bands both<br />

fault and draw into plane sillimanite and biotite (Fig. 2.7c).<br />

At upper structural levels within <strong>the</strong> domain, muscovite is common and quartz grains are<br />

mildly elongate and inequigranular. Along strike, feldspar porphyroclasts are bordered by<br />

recrystallized quartz and feldspar in core-and-mantle type structures (Fig. 2.7b), corresponding to<br />

deformation temperatures <strong>of</strong> ~400-500°C (Tullis and Yund, 1991).<br />

At progressively lower structural levels within <strong>the</strong> domain, muscovite decreases in<br />

abundance and quartz grain boundaries transition from interlobate to amoeboid, occasionally<br />

exhibiting zones <strong>of</strong> subgrain extinction (Fig. 2.7d). These textures represent recrystallization<br />

dominantly through grain boundary migration, with a considerable component <strong>of</strong> recrystallization<br />

by subgrain rotation. This may correspond to <strong>the</strong> transition between <strong>the</strong>se two recrystallization<br />

mechanisms at ~500°C (Stipp et al., 2002).<br />

These trends are consistent with <strong>the</strong> weakening <strong>of</strong> high-strain and asymmetric fabric<br />

elements with increasing structural depth throughout <strong>the</strong> entirety <strong>of</strong> <strong>the</strong> domain. Despite<br />

decreased intensity, tectonic foliations and mineral elongation lineations have a relatively<br />

consistent geometry (Fig. 2.3). Elongate quartz, feldspar, and phyllosilicates define a pervasive<br />

south-dipping foliation and a shallow to moderately plunging ESE-WNW trending lineation (Fig.<br />

2.3). Minor, broad, upright, open folds with fold axes roughly parallel to <strong>the</strong> mineral elongation<br />

lineation gently warp <strong>the</strong> domain. Centimeter- to meter-scale isoclinal folds with identical fold<br />

axes are also locally observed throughout <strong>the</strong> high-strain gneiss.<br />

Asymmetric fabric elements, including C/S/C’ fabrics, sigma-type feldspar<br />

porphyroclasts and delta-type rotated leucogranite pods are best developed within <strong>the</strong> upper<br />

structural section <strong>of</strong> <strong>the</strong> domain (Fig. 2.7a,b,c). Shear sense indicators show a near even<br />

32


distribution <strong>of</strong> dextral and sinistral senses <strong>of</strong> motion. Rare sets <strong>of</strong> conjugate extensional shear<br />

bands are also observed throughout <strong>the</strong> domain.<br />

Additionally throughout this domain, brittle structures and lower-temperature solid-state<br />

deformation overprint <strong>the</strong> ductilly deformed high-strain gneiss. Brittle structures include normal<br />

faulting and cataclastic deformation <strong>of</strong> ductilly deformed gneiss and leucogranite dykes and<br />

injection <strong>of</strong> pseudotachylite across weakly cataclastic foliation (Fig. 2.8a,b,c). Solid-state<br />

deformation includes <strong>the</strong> development <strong>of</strong> centimeter-wide mylonitic to ultramylonitic layers that<br />

cut obliquely across high-strain gneiss (Fig. 2.8d).<br />

2.4.2.6 Domain VI – Migmatitic GHS<br />

The migmatitic GHS is <strong>the</strong> nor<strong>the</strong>rnmost and structurally lowest domain mapped (Figs.<br />

2.3, 2.7e,f). The domain consists <strong>of</strong> migmatitic quartz<strong>of</strong>eldspathic gneiss (Fsp + Qtz + Bt + Ms ±<br />

Grt ± Tur ± Sil) with up to 70 vol% leucosome. The mineralogy <strong>of</strong> <strong>the</strong> migmatitic GHS is similar<br />

to <strong>the</strong> overlying high-strain gneissic GHS; however, <strong>the</strong> modal abundance <strong>of</strong> muscovite and<br />

leucosome, and intensity <strong>of</strong> strain are markedly different.<br />

Sillimanite, in <strong>the</strong> form <strong>of</strong> fibrolite, is rare and anastomoses around clusters <strong>of</strong> quartz and<br />

feldspar. Garnets are inclusion-rich and intensely fractured, but not in a systematic orientation.<br />

Rare muscovite grains amidst dynamically recrystallized quartz and feldspar show irregular grain<br />

boundaries, suggestive <strong>of</strong> non-equilibrium conditions. Quartz and feldspar grains are equigranular<br />

and show interlobate grain boundaries (Fig. 2.7e). This indicates recrystallization via grain<br />

boundary migration, which is associated with temperatures <strong>of</strong> deformation between ~500-700°C<br />

(Stipp et al., 2002). Additionally, quartz displays chessboard extinction (Fig. 2.7f), thought to<br />

represent a combination <strong>of</strong> slip along <strong>the</strong> basal and prism lattice planes (Blumenfeld et<br />

33


Figure 2.8 Representative brittle structures and lower temperature deformation within domain V.<br />

(a) Normal faulting <strong>of</strong> high-strain gneiss and leucogranite. (b) Pseudotachylite (outlined in white)<br />

crosscutting cataclasite. (c) Semi-ductile normal faulting <strong>of</strong> leucogranite dyke. (d) Ultramylonitic<br />

to mylonitic layers cutting obliquely across high-strain gneiss.<br />

34


al., 1986; Mainprice et al., 1986). This texture corresponds to deformation temperatures in excess<br />

<strong>of</strong> ~630°C (Stipp et al., 2002).<br />

Along <strong>the</strong> deepest structural level observed, ~10 km below <strong>the</strong> GHS-TSS interface, initial<br />

stages <strong>of</strong> diatexis are seen, in which biotite- and feldspar-rich rafts and leucosome form intermerging<br />

schlieren and nebulitic structures.<br />

Throughout <strong>the</strong> migmatitic GHS, macroscale anastomosing leucosome and melanosome<br />

define a weakly developed south-dipping foliation. At <strong>the</strong>se structural levels, no mineral<br />

lineations are observed.<br />

2.5 Microstructural analyses<br />

2.5.1 Crystallographic preferred orientation <strong>of</strong> quartz<br />

Quartz crystallographic preferred orientations (CPOs) are <strong>the</strong> result <strong>of</strong> deformation<br />

temperature, strain rate, non-coaxiality <strong>of</strong> flow, and distortional strain geometry (Sullivan and<br />

Beane, 2010 and references <strong>the</strong>rein). Consequently, <strong>the</strong> analysis <strong>of</strong> quartz CPOs has become a<br />

powerful tool in characterizing high-strain zones and exhumed metamorphic terranes, such as <strong>the</strong><br />

GHS and its bounding faults (Bouchez & Pêcher, 1976; Brunel and Kienast, 1986; Grujic et al.,<br />

1996; Grasemann et al., 1999; Bhattacharya and Weber, 2004; Law et al., 2004; Larson and<br />

Godin, 2009; Larson et al., 2010a; Yakymchuk and Godin, 2012).<br />

Historically, quartz CPO analysis has been focused on c-axis fabrics due to analytical<br />

limitations. However, development <strong>of</strong> <strong>the</strong> Electron Back Scatter Diffraction (EBSD) technique<br />

has provided an alternative methodology whereby complete crystallographic orientations <strong>of</strong> all<br />

lattice planes and axes can be determined (Prior et al., 1999). This allows for evaluation <strong>of</strong> both<br />

<strong>the</strong> quartz c- and a- axes. The analysis <strong>of</strong> a-axis fabrics provides an independent assessment <strong>of</strong><br />

plane strain, flattening, and constriction (Passchier and Trouw, 2005). Fur<strong>the</strong>rmore, Sullivan and<br />

35


Beane (2010) demonstrated that for deformation at high temperatures with non-coaxial flow, a-<br />

axis fabrics provide a more accurate representation <strong>of</strong> strain geometry. We <strong>the</strong>refore follow <strong>the</strong><br />

recommendation <strong>of</strong> <strong>the</strong> aforementioned study and record c- and a- axes for all samples.<br />

Additionally, c-axis fabrics <strong>of</strong> four samples were measured via a universal stage mounted on an<br />

optical microscope for comparative purposes with EBSD-derived data.<br />

2.5.1.1 Methodology<br />

Complete crystallographic orientation data <strong>of</strong> quartz for all samples were obtained using<br />

<strong>the</strong> EBSD method. Data were collected through an HKL EBSD detector installed on a JEOL<br />

JSM636OLV Scanning Electron Microscope (SEM) at Colgate <strong>University</strong>, located in Hamilton,<br />

New York, USA. Oxford Instruments HKL Channel 5.0 s<strong>of</strong>tware was used for acquisition and<br />

processing <strong>of</strong> data. The following SEM conditions were used: high-vacuum mode, accelerating<br />

voltage <strong>of</strong> 20kV and a working distance <strong>of</strong> 12-20mm (optimal at 15-20mm). Samples were<br />

initially polished to a standard probe polish <strong>the</strong>n subsequently polished on a vibrating table for 4-<br />

6 hours using a slightly basic solution to ensure both a mechanical and chemical final polish.<br />

Following this process, samples were mounted uncoated within <strong>the</strong> SEM.<br />

EBSD analyses were collected using automated mapping where individual point data for<br />

quartz were measured at established intervals (step sizes) and automatically indexed to determine<br />

crystallographic orientations. Step sizes were set at increments smaller than <strong>the</strong> smallest grain<br />

size in a given sample to ensure sampling <strong>of</strong> all grains. Following data acquisition, wild data<br />

spikes were extrapolated and clusters <strong>of</strong> proximal point data with crystallographic misorientations<br />

less than 10° were grouped and filled in using a nearest-neighbor algorithm (Fig. 2.9). This<br />

eliminated any erratic or misindexed crystallographic orientations. Following this,<br />

crystallographic misorientations greater than 10° were defined as grain boundaries; clusters <strong>of</strong><br />

36


Figure 2.9 Summary <strong>of</strong> CPO post-acquisition data processing. (a) SEM images from localities at which<br />

data was collected. (b) Overlay <strong>of</strong> s<strong>of</strong>tware-generated colors on SEM images indicating <strong>the</strong> identification<br />

<strong>of</strong> a quartz CPO for individual pixels. A lack <strong>of</strong> color indicates <strong>the</strong> s<strong>of</strong>tware did not identify a CPO<br />

associable with quartz. (c) Rasterized data image where colors represent c-axis orientations; similar colors<br />

represent similar orientation. (d) Rasterized data image after extrapolation <strong>of</strong> wild data spikes and<br />

application <strong>of</strong> nearest neighbor algorithm. Using this modified data set, a new data set was created where<br />

each grain (as seen by sets <strong>of</strong> pixels <strong>of</strong> like colors) represents an individual point.<br />

37


similar crystallographic orientations were classified as individual grains; and a new data set was<br />

created where each grain was represented as a single point (Fig. 2.9). This eliminated any bias<br />

created by oversampling large grains and allowed for direct comparison with data derived from<br />

<strong>the</strong> universal-stage (U-Stage) method. A summary <strong>of</strong> this post-acquisition data processing is<br />

shown in Figure 2.9.<br />

Quartz c-axis orientations were also measured for four samples via an optical microscope<br />

mounted with a universal stage. A single c-axis measurement was taken for each individual grain<br />

and attention was paid to not measure any grain more than once.<br />

2.5.1.2 Descriptions <strong>of</strong> quartz CPO patterns<br />

Quartz CPO patterns follow <strong>the</strong> definitions <strong>of</strong> Lister (1977), in which <strong>the</strong> plane normal to<br />

<strong>the</strong> lineation (Y-Z plane) is defined as <strong>the</strong> symmetry plane. Fabrics that do not mirror <strong>the</strong>mselves<br />

across this plane are described as asymmetric. A summary <strong>of</strong> <strong>the</strong> orientation <strong>of</strong> principal strain<br />

axes, active slip systems, fabric geometries in different strain fields, and evolution <strong>of</strong> CPO<br />

patterns during non-coaxial strain at various temperatures is portrayed in Figure 2.10.<br />

Photomicrographs <strong>of</strong> samples analyzed for quartz CPO patterns are seen in Figure 2.11. Quartz<br />

CPO c- and a- axes are presented in Figure 2.12 as lower-hemisphere equal-area projections<br />

oriented normal to <strong>the</strong> foliation and parallel to <strong>the</strong> lineation (see Appendix B for additional CPO<br />

data).<br />

Analyzed CPO fabrics were sampled from three domains: <strong>the</strong> ultra-high-strain zone<br />

(domain III), <strong>the</strong> high-strain quartzite and pelitic GHS (domain IV), and <strong>the</strong> high-strain gneissic<br />

GHS (domain V). CPO patterns from within each zone are described below.<br />

38


a<br />

c<br />

a<br />

c<br />

Figure 2.10 Summary <strong>of</strong> quartz CPO fabrics showing (a) <strong>the</strong> orientation <strong>of</strong> principle strain axes,<br />

active slip systems, and fabric geometries in different strain fields, and (b) <strong>the</strong> evolution <strong>of</strong> a<br />

CPO fabric with increasing temperature (T.) during non-coaxial deformation (assuming plane<br />

strain). Note that <strong>the</strong> letters ‘c’ and ‘a’ refer to <strong>the</strong> quartz crystallographic c- and a-axes. Adapted<br />

and modified after Schmid and Casey (1986); Passchier and Trouw (2005); Toy et al. (2008);<br />

Sullivan and Beane (2010).<br />

Figure 2.11 (Next page) Photomicrographs <strong>of</strong> samples analyzed for quartz CPOs. From<br />

structurally highest to lowest, <strong>the</strong>y are: Samples (a) HK124B2 and (b) HK125, from millimeter<br />

thick bands <strong>of</strong> extremely fine-grained and ribbon quartz within domain III. Samples (c)<br />

HK131A, (d) HK139D, (e) HK123, and (f) HK140, from high-strain quartzite and<br />

quartz<strong>of</strong>eldspathic schist <strong>of</strong> domain IV. Samples (g) HK145 and (h) HK105A, from quartzite and<br />

quartz<strong>of</strong>eldspathic schist <strong>of</strong> domain V. See text for description <strong>of</strong> quartz textures.<br />

39


Figure 2.12 Quartz CPO c- and a- axes. Columns from left to right are: sample number and distance from<br />

GHS-TSS interface measured perpendicular to <strong>the</strong> average foliation; EBDS-derived c-axes (),<br />

including shear sense when applicable, and lineation data; EBSD-derived a-axes (), including multiples<br />

<strong>of</strong> uniform density (mud) for EBSD-derived contours; U-Stage-derived c-axes () and number <strong>of</strong><br />

uniform contours. For each analysis, point and contoured data is shown; EBSD data were contoured using<br />

Channel 5 s<strong>of</strong>tware, and U-stage data were contoured using OSX Stereonet v.1.3 developed by N. Cardozo<br />

and R. Allmendinger. Number <strong>of</strong> grains measured corresponds to n. All stereonets are lower-hemisphere<br />

equal-area projections oriented normal to <strong>the</strong> foliation and parallel to <strong>the</strong> lineation. Note that certain<br />

EBSD-derived CPO fabrics are not centered on <strong>the</strong> stereonet due to limitations <strong>of</strong> <strong>the</strong> EBSD s<strong>of</strong>tware in<br />

rotating data <strong>of</strong> poorly-oriented samples.<br />

______________________________________________________________________________<br />

Domain III, ultra-high-strain quartz CPO patterns<br />

Ultra-high-strain quartz CPOs, samples HK124B2 and HK125, were measured from<br />

millimeter-thick bands <strong>of</strong> extremely fine-grained and ribbon quartz (~100 µm wide) exhibiting<br />

subgrain rotation and bulging recrystallization (Fig. 2.11a,b). Separating <strong>the</strong>se quartz bands are<br />

layers <strong>of</strong> attenuated micas showing undulose extinction, fine-grained quartz, and brittly deformed<br />

feldspar (Figs. 2.5e,f, 2.11a,b).<br />

CPO patterns analyzed within this domain (Fig. 2.12) are characterized by type I singlegirdle<br />

c-axis fabrics with dominant contributions <strong>of</strong> rhomb and prism slip and a lesser<br />

component <strong>of</strong> basal slip. The c-axis pattern <strong>of</strong> sample HK124B2 is roughly symmetric,<br />

whereas that <strong>of</strong> sample HK125 is slightly asymmetric indicating a sinistral sense <strong>of</strong> shear. A-axis<br />

fabrics are concentrated in five to six point maxima around <strong>the</strong> edge <strong>of</strong> <strong>the</strong> stereonet (Fig. 2.12).<br />

Domain IV, high-strain quartzite and pelitic GHS CPO patterns<br />

CPOs <strong>of</strong> domain IV were measured from high-strain quartzite and quartz<strong>of</strong>eldspathic<br />

schist across <strong>the</strong> entire structural extent <strong>of</strong> <strong>the</strong> high-strain quartzite and pelitic GHS. With<br />

increasing depth, CPO patterns record a progression from type I single-girdle c-axis fabrics with a<br />

component <strong>of</strong> basal slip, into a loss <strong>of</strong> basal and rhomb slip, an intensification <strong>of</strong> a<br />

Y-maximum, and <strong>the</strong> formation <strong>of</strong> pseudo type II cross-girdle c-axis fabrics (Fig. 2.12).<br />

42


Sample HK131A is a high-strain quartzite (Fig. 2.11c). Quartz accounts for ~85 vol%<br />

and is characterized by strongly flattened grains, subgrain extinction and recrystallization through<br />

subgrain rotation. Elongate and flattened mica define non-laterally consistent partings between<br />

<strong>the</strong> quartz layers (Fig. 2.11c). Contoured EBSD data <strong>of</strong> sample HK131A indicate a weak<br />

component <strong>of</strong> basal slip and a more symmetrical distribution <strong>of</strong> c-axis orientations not<br />

observed in contoured U-Stage data (Fig. 2.12). This is attributed to <strong>the</strong> quantity <strong>of</strong> grains<br />

measured; 426 grains were measured using U-Stage, versus 1042 grains using EBSD. Apart from<br />

this discrepancy, <strong>the</strong> EBSD data are consistent with <strong>the</strong> U-Stage method. A-axis fabrics are<br />

concentrated in six point maxima around <strong>the</strong> edge <strong>of</strong> <strong>the</strong> stereonet.<br />

Sample HK139D and HK123 are <strong>the</strong> next structurally lower samples analyzed for CPOs.<br />

Sample HK139D is a high-strain quartzite (Fig. 2.11d). Moderately coarse-grained<br />

porphyroclastic quartz (~300 µm wide) accounts for greater than 95 vol% and is characterized by<br />

flattened grains, serrate grain boundaries, subgrain extinction and recrystallization through<br />

subgrain rotation. Elongated and flattened micas define non-laterally consistent partings between<br />

quartz layers (Fig. 2.11d).<br />

Sample HK123 comes from a centimeter-thick mylonitic layer within a high-strain<br />

quartz<strong>of</strong>eldspathic schist. Within this sample, CPOs were measured from millimeter-thick<br />

mylonitic bands <strong>of</strong> quartz showing extensive dynamic recrystallization through subgrain rotation<br />

(Fig. 2.11e).<br />

CPO patterns <strong>of</strong> both samples reveal type I single-girdle c-axis fabrics approaching a Y-<br />

maxima. Slip systems are predominantly prism , and show weakened contributions <strong>of</strong> rhomb<br />

relative to overlying samples. A right-lateral asymmetry is observed in sample HK139D. A-<br />

axis fabrics are concentrated in four point maxima around <strong>the</strong> edge <strong>of</strong> <strong>the</strong> stereonet (<strong>the</strong>re would<br />

43


e six point maxima if <strong>the</strong> fabric was centered on <strong>the</strong> stereonet – see Fig. 2.12 and caption).<br />

EBSD and U-Stage c-axis fabrics <strong>of</strong> sample HK139D are consistent (Fig. 2.12).<br />

The next structurally lower sample, HK140, is a mylonitic quartzite (Fig. 2.11f). Coarsegrained<br />

porphyroclastic quartz (~1 mm wide) account for >95 vol% and are characterized by<br />

crudely interlocking interlobate and serrate grain boundaries, and extensive dynamic<br />

recrystallization through subgrain rotation and grain boundary migration. Rare and small (~20 µm<br />

wide) micas are dispersed throughout <strong>the</strong> sample (Fig. 2.11f). A crudely developed type II crossgirdle<br />

c-axis fabric dominated by prism slip characterizes <strong>the</strong> CPO pattern <strong>of</strong> this sample. A-<br />

axis fabrics are concentrated in six point maxima around <strong>the</strong> edge <strong>of</strong> <strong>the</strong> stereonet. EBSD and U-<br />

Stage c-axis fabrics are consistent (Fig. 2.12).<br />

Domain V, high-strain gneissic GHS CPO patterns<br />

CPOs <strong>of</strong> Domain V were measured from quartzite and quartz<strong>of</strong>eldspathic schist (Fig.<br />

2.11g). Sample HK145A is a mylonitic quartz<strong>of</strong>eldspathic schist. CPOs were measured from<br />

quartz clusters comprising ~40 vol%. Well-developed interlobate interlocking quartz grain<br />

boundaries imply recrystallization through grain boundary migration. The remainder <strong>of</strong> <strong>the</strong> matrix<br />

comprises perthitic feldspar and rare fractured and inclusion-rich garnets (Fig. 2.11g). The<br />

subsequent structurally lower specimen, HK105A, was sampled from a 10 m thick quartzite layer.<br />

Coarse- to fine-grained quartz represents >95 vol% and is extensively recrystallized (Fig. 2.11h).<br />

Coarse-grained quartz grains show well-developed chessboard style extinction (Fig. 2.7e).<br />

Both samples yield moderately defined type II cross-girdle c-axis fabrics with ambiguous<br />

slip systems, potentially dominated by rhomb with lesser components <strong>of</strong> prism slip. This<br />

may be consistent with <strong>the</strong> hypo<strong>the</strong>sis that type II cross-girdle c-axis patterns form during<br />

44


constriction, where rhomb slip dominates over prism slip (Bouchez 1978; Schmid and Casey,<br />

1986).<br />

A-axis fabrics within <strong>the</strong>se samples deviate from <strong>the</strong> typical six point maxima around <strong>the</strong><br />

edge <strong>of</strong> <strong>the</strong> stereonet (Fig. 2.12). This deviation may also indicate a departure from plane strain<br />

conditions (Fig. 2.10). EBSD and U-Stage c-axis fabrics from sample HK105A are consistent<br />

(Fig. 2.12).<br />

2.5.1.3 Quartz CPO deformation temperatures<br />

Assuming natural strain rates and plane strain, temperatures <strong>of</strong> deformation can be<br />

constrained through <strong>the</strong> following two methods using quartz c-axis fabrics: (1) interpretation <strong>of</strong><br />

dominant slip systems active during deformation (Law, 1991; Toy et al, 2008), and (2) evaluation<br />

<strong>of</strong> quartz c-axis opening angles (Kruhl, 1998; Law et al., 2004).<br />

The first method relies on <strong>the</strong> transition <strong>of</strong> c-axis slip systems with varying deformation<br />

temperatures (Sullivan and Beane, 2010 and references <strong>the</strong>rein). However, this methodology can<br />

be problematic as <strong>the</strong> absolute deformation temperatures <strong>of</strong> <strong>the</strong>se transitions are poorly<br />

constrained (Toy et al., 2008 and references <strong>the</strong>rein). The second method utilizes an empirical<br />

linear relationship between temperatures <strong>of</strong> deformation and angles between opposing arms <strong>of</strong><br />

quartz c-axis fabric skeletons, referred to as <strong>the</strong> c-axis opening angle (Fig. 2.13; Kruhl, 1998;<br />

Law et al., 2004). The opening angle method <strong>of</strong>fers a quantitative approach to qualifying<br />

deformation temperatures and has thus been applied in various Himalayan studies (Law et al.,<br />

2004; Larson and Godin, 2009; Langille et al., 2010; Larson et al., 2010a,b; Yakymchuk and<br />

Godin, 2012). This relationship is calibrated on empirical data and subject to deviations from<br />

natural strain rates. Consequently, estimates <strong>of</strong> deformation temperature are subject to an error <strong>of</strong><br />

45


Figure 2.13 Plot <strong>of</strong> c-axis fabric opening angles from <strong>the</strong> upper<br />

Karnali valley against deformation temperature. For simplicity,<br />

<strong>the</strong> prefix “HK” has been removed from sample numbers within<br />

this figure. Modified after Law et al. (2004).<br />

46


± 50°C (Law et al., 2004). Fur<strong>the</strong>rmore, <strong>the</strong> relationship is valid only up to ~650°C, above which<br />

prism slip becomes increasingly active and <strong>the</strong> linear relationship deteriorates (Law et al.,<br />

2004).<br />

Quartz CPOs from domain III to domain V show an increase in CPO opening angles<br />

from ~45° to 85° (Fig. 2.12, 2.13). This corresponds to an increase in temperatures <strong>of</strong><br />

deformation from ~350°C along <strong>the</strong> GHS-TSS interface within domain III, to ~650°C at depths<br />

<strong>of</strong> ~2.6km below <strong>the</strong> interface within domain V (Fig. 2.13). Sample HK105A is an exception to<br />

this trend, exhibiting a lower than expected opening angle relative to its depth. However, this<br />

sample exhibits pervasive chessboard extinction (Fig. 2.11h), resulting from simultaneous slip<br />

along <strong>the</strong> basal and prism lattice planes (Blumenfeld et al., 1986; Mainprice et al., 1986).<br />

Due to slip along <strong>the</strong> prism plane, this sample would not be expected to respect <strong>the</strong> opening<br />

angle linear trend (Law et al., 2004).<br />

2.5.2 Rigid clast vorticity analysis<br />

Vorticity analyses quantify <strong>the</strong> internal rotational component <strong>of</strong> flow to estimate <strong>the</strong><br />

relative contributions <strong>of</strong> pure and simple shear throughout deformation (Passchier and Trouw,<br />

2005). Consequently, vorticity analysis has become a powerful tool in characterizing shear zones<br />

and strain regimes throughout <strong>the</strong> Himalaya (Grasemann et al., 1999; Law et al., 2004; Carosi et<br />

al., 2006; Jessup et al., 2006, 2007; Larson and Godin, 2009; Lee and Wagner, 2009; Langille et<br />

al., 2010; Larson et al., 2010a,b).<br />

For plane strain deformation, <strong>the</strong> relative proportions <strong>of</strong> pure and simple shear can be<br />

directly estimated from <strong>the</strong> kinematic vorticity number (W k ). In a strictly pure shear system W k =<br />

0, whereas in a strictly simple shear system, W k = 1 (Means et al., 1980). The relationship<br />

between pure and simple shear is not linear, with equal contributions <strong>of</strong> both at W k =0.71 (Law et<br />

47


al., 2004). The W k value refers to instantaneous kinematic vorticity and is only representative <strong>of</strong><br />

steady-state deformation, a rare situation in natural systems (Bailey et al., 2004). The mean<br />

kinematic vorticity (W m ) averages <strong>the</strong> W k value over time and allows for a more accurate<br />

estimation <strong>of</strong> pure and simple shear throughout non-steady-state deformation (Fossen and Tik<strong>of</strong>f,<br />

1997, 1998; Jiang, 1998). Vorticity analyses are performed in two dimensions; however <strong>the</strong><br />

motion <strong>of</strong> rigid clasts during deformation is a three dimensional problem. Consequently, error<br />

associated with this method may be significant (Tik<strong>of</strong>f and Fossen, 1995; Forte and Bailey, 2007;<br />

Iacopini et al., 2008; Li and Jiang, 2011).<br />

W m values were estimated using <strong>the</strong> rigid grain method (Jessup et al., 2007). Only<br />

samples that abided by <strong>the</strong> following criteria, as stipulated by Passchier (1987), were chosen for<br />

analysis: (1) presence <strong>of</strong> abundant rigid pre-deformational porphyroclasts, (2) porphyroclasts are<br />

significantly larger than host matrix grains, (3) no mechanical interaction between porphyroclasts,<br />

(4) significant quantity <strong>of</strong> porphyroclasts with a wide range <strong>of</strong> aspect ratios, (5) a homogenously<br />

deforming matrix, (6) having experienced high and protracted strains allowing grains to rotate to<br />

stable-sink positions.<br />

Sample HK146 was <strong>the</strong> only specimen that respected <strong>the</strong> above criteria and defined a<br />

clear W m threshold. This sample is located within <strong>the</strong> lower-temperature mylonitic component <strong>of</strong><br />

domain V, <strong>the</strong> high-strain gneissic GHS (Fig. 2.3). The specimen analyzed was taken from a


Figure 2.14 (a) Photomicrograph <strong>of</strong> sample analyzed for vorticity showing rigid clasts <strong>of</strong><br />

feldspar (fld) and quartz (qtz) in a fine-grained deforming matrix. (b) Results <strong>of</strong> rigid clast<br />

vorticity analysis displayed on a rigid grain net plot (modified from Jessup et al., 2007),<br />

and <strong>the</strong> calculation from W m to percent pure shear.<br />

49


temperature. Therefore <strong>the</strong> calculated vorticity estimate should reflect this stage <strong>of</strong> deformation.<br />

Analysis <strong>of</strong> 208 quartz and feldspar porphyroclasts yielded W m values <strong>of</strong> 0.58-0.61,<br />

corresponding to ~59% pure shear (Fig. 2.14b).<br />

2.5.3 Summary <strong>of</strong> microstructural analyses<br />

A summary <strong>of</strong> mineral deformation mechanisms, quartz CPO fabrics, peak metamorphic<br />

assemblages, and temperatures <strong>of</strong> deformation across <strong>the</strong> GHS-TSS interface is presented in<br />

Figure 2.15. A vertical pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> shear zone, from domain I through to domain VI, reveals a<br />

systematic increase in deformation temperatures, and a potential deviation <strong>of</strong> quartz CPO fabrics<br />

from plane strain towards constriction.<br />

From domain I to <strong>the</strong> base <strong>of</strong> domain II, calcite microstructures transition from (1) fine<br />

grained, non-deformed calcite lacking twins (Fig. 2.4c), to (2) moderately deformed calcite<br />

bearing dominantly type I twins (Fig. 2.4e), to (3) recrystallized and deformed calcite bearing<br />

dominantly type II twins, with occasional type I and III twins (Fig. 2.4g).<br />

From <strong>the</strong> ultra-high-strain zone <strong>of</strong> domain III, through to domain VI, quartz and feldspar<br />

deformation mechanisms transition from (1) strongly attenuated ribbon quartz exhibiting bulging<br />

and subgrain rotation recrystallization, and brittle feldspar (Fig. 2.5c,d,e,f), to (2) flattened quartz<br />

recrystallized via predominantly subgrain rotation (Fig. 2.6d), to (3) interlocking quartz grains<br />

recrystallized through grain boundary migration (Fig. 2.7d), and (4) interlobate dynamically<br />

recrystallized feldspar and quartz exhibiting chessboard extinction (Fig. 2.7e,f).<br />

The above corresponds to an increase in deformation temperatures from less than 200°C at <strong>the</strong><br />

uppermost structural levels <strong>of</strong> domain I, to ~350-400°C within domain III along <strong>the</strong> GHS-TSS<br />

interface, and greater than ~630°C within domain IV and V, at depths <strong>of</strong> ~2.5 -5.5 km below <strong>the</strong><br />

GHS-TSS interface (Fig. 2.15).<br />

50


Figure 2.15 Summary <strong>of</strong> lithotectonic domains (D. I, … VI), mineral deformation mechanisms, quartz<br />

CPO fabrics and slip systems, peak metamorphic assemblages and temperatures <strong>of</strong> deformation across <strong>the</strong><br />

GHS-TSS interface. Numbered boxes refer to mineral deformation mechanisms: (1) predominant type I<br />

calcite twins ( 630 ± 30°C;<br />

Stipp et al., 2002). Distance from GHS-TSS interface measured perpendicular to <strong>the</strong> average foliation.<br />

Type I, type II, 6-point (pt) maxima, basal , rhomb , and prism refer to c- and a- axis fabrics<br />

and slip systems as described in <strong>the</strong> text and seen in Figure 2.7. Abbreviations: Bt – biotite, Ms –<br />

muscovite, Sil – sillimanite.<br />

51


Quartz c-axis fabrics show a systematic progression with increasing depth, corresponding<br />

to an increase in temperature and a potential deviation from plane strain toward constriction<br />

(Figs. 2.12, 2.15). From domain III through to domain V, quartz c-axis fabrics transition from (1)<br />

type I cross-girdles with a component <strong>of</strong> basal slip, to (2) type I cross-girdles lacking basal<br />

slip and approaching a Y-maxima, to (3) moderately-defined type II cross-girdles. C-axis<br />

opening angles increase with depth from ~45° to 85°, corresponding to an increase in<br />

temperatures <strong>of</strong> deformation from ~350°C within domain III, to ~650°C within domain V (Figs.<br />

2.13, 2.15).<br />

The transition in c-axis fabrics from type I to type II cross-girdles may correspond to an<br />

increase in constriction (Fig. 2.10; Bouchez 1978; Schmid and Casey, 1986). This transition also<br />

corresponds with a minor deviation from <strong>the</strong> six point maxima a-axis fabrics (Figs. 2.12, 2.15),<br />

which may also indicate a departure from plane strain at <strong>the</strong> deeper structural levels (Fig. 2.10).<br />

Minor asymmetry, indicating dextral and sinistral motion, is observed in only two quartz<br />

CPO fabrics, both located within ~700 m <strong>of</strong> <strong>the</strong> GHS-TSS interface. The remaining CPO fabrics<br />

reveal roughly symmetric c- and a- axis patterns (Fig. 2.12). This attests to an important, and<br />

potentially dominant component <strong>of</strong> pure shear throughout deformation.<br />

A W m estimate <strong>of</strong> 0.58-0.61, corresponding to ~59% pure shear, was derived from a lowtemperature<br />

crosscutting mylonitic layer within domain V (Fig. 2.15, 2.14). This attests to <strong>the</strong><br />

ongoing importance <strong>of</strong> pure shear throughout lower-temperature deformation.<br />

2.6 Geochronology and <strong>the</strong>rmochronology<br />

U-Th-Pb geochronology and 40 Ar/ 39 Ar <strong>the</strong>rmochronology were performed on a suite <strong>of</strong><br />

samples collected from traverses across and along <strong>the</strong> GHS portion <strong>of</strong> <strong>the</strong> upper Karnali valley<br />

(domains III, IV, V and VI). Integrated with structural analyses and detailed analysis <strong>of</strong> textural<br />

52


elationships, <strong>the</strong>se data constrain <strong>the</strong> absolute timing and cooling history <strong>of</strong> metamorphism and<br />

deformation in <strong>the</strong> upper Karnali valley.<br />

2.6.1 U-Th-Pb monazite geochronology<br />

Polished thin sections <strong>of</strong> samples likely to contain monazite grains were examined using<br />

backscatter electron imaging in a scanning electron microprobe (SEM) at Queen’s <strong>University</strong>.<br />

The presence <strong>of</strong> heavy elements within monazite grains (Ce, La, Nd, Th, Y) produces an intense<br />

electron backscatter signal allowing for identification <strong>of</strong> monazite-rich samples. Grains identified<br />

as monazite were subsequently confirmed using high-speed energy dispersive X-ray spectroscopy<br />

analyses. Samples HK109, KH18 and HK117 were deemed to have sufficient numbers <strong>of</strong> large<br />

monazite grains and were selected for U-Th-Pb in situ monazite geochronology. Detailed textural<br />

relationships between monazites and <strong>the</strong>ir host fabrics were investigated by examining highresolution<br />

electron backscatter images <strong>of</strong> entire thin sections for each selected sample.<br />

Samples KH18a and HK109 are high-strain Bt + Ms + Sil ± Grt pelitic schist. Sample<br />

KH18a is located within <strong>the</strong> high-strain quartzite and pelitic schist <strong>of</strong> domain IV, and sample<br />

HK109 is located in a thin pelitic layer within <strong>the</strong> high-strain gneissic GHS <strong>of</strong> domain V (Fig.<br />

2.3). In both samples, garnet porphyroclasts are pre-deformational, exhibiting extensive brittle<br />

fracturing and stretching parallel to <strong>the</strong> mineral elongation lineation. Quartz is recrystallized<br />

dominantly through subgrain rotation with lesser amounts <strong>of</strong> grain boundary migration. Pervasive<br />

C/S/C’ fabrics are defined by micas ± sillimanite, and within sample HK109, both minerals are<br />

faulted by and drawn into plane along shear bands (Fig. 2.7c). Textures seen in Figure 2.6 are<br />

representative <strong>of</strong> <strong>the</strong>se samples.<br />

Sample HK117a is a Bt + Ms + Sil ± Grt migmatitic gneiss located within <strong>the</strong> migmatitic<br />

GHS <strong>of</strong> domain VI (Fig. 2.3). Garnet porphyroclasts are rare, pre-deformational and fractured in a<br />

53


non-systematic fashion. Quartz is recrystallized through grain boundary migration. Muscovite,<br />

biotite, and sillimanite are present in low quantities and define a moderate anastomosing foliation.<br />

Textures seen in Figure 2.7e,f are representative <strong>of</strong> this sample.<br />

2.6.1.1 Methodology<br />

Between 8 and 11 monazite grains per sample were compositionally mapped for Th, Y, U<br />

and Nd / La using a JEOL JXA-8230 electron microprobe at Queen’s <strong>University</strong>. This was<br />

completed to identify chemical zonation within <strong>the</strong> grains to facilitate targeting individual<br />

compositional domains. The following electron microprobe conditions were used: accelerating<br />

voltage <strong>of</strong> 15 kV, a beam current <strong>of</strong> ~350-500 nA, a dwell time <strong>of</strong> 100 ms and a step size <strong>of</strong> 0.5-<br />

1.2 μm. High- and low-magnification backscatter-electron and secondary-electron images were<br />

also collected to establish <strong>the</strong> textural context and physical condition <strong>of</strong> <strong>the</strong> monazite grains.<br />

Monazites were analyzed using a laser ablation multi collector inductively coupled<br />

plasma mass spectrometer (LA-MC-ICPMS) system housed at <strong>the</strong> <strong>University</strong> <strong>of</strong> California, Santa<br />

Barbara (UCSB). Instrumentation consists <strong>of</strong> a Nu Plasma MC-ICPMS (Nu Instruments,<br />

Wrexham, UK) and a 193 nm ArF laser ablation system equipped with a two-volume ‘HelEx’<br />

ablation cell that facilitates rapid transfer and washout <strong>of</strong> ablated material (Photon Machines, San<br />

Diego, USA). Analytical protocol is similar to that described by Cottle et al. (2009a,b, 2011)<br />

with <strong>the</strong> modification that <strong>the</strong> collector arrangement on <strong>the</strong> Nu Plasma at UCSB allows for<br />

simultaneous determination <strong>of</strong> 232 Th and 238 U on high-mass side Faraday cups equipped with 10 11<br />

ohm resistors and 208 Pb, 207 Pb, 206 Pb and 204 Pb on four low-mass side ETP discrete dynode<br />

secondary electron multipliers.<br />

U-Th-Pb analyses were conducted for 20 seconds each using a 7 μm spot diameter, a<br />

frequency <strong>of</strong> 4 Hz and 1.2 J/cm 2 fluence (equating to crater depths <strong>of</strong> approximately 5- 6 μm). A<br />

54


primary reference material, “44096” (424 Ma Pb/U ID-TIMS age, Aleinik<strong>of</strong>f et al., 2006) was<br />

employed to monitor and correct for mass bias as well as Pb/U and Pb/Th down-hole<br />

fractionation. To monitor data accuracy, two secondary reference monazites “FC-1” (55.7 Ma<br />

Pb/U ID-TIMS age, Horstwood et al., 2003), and “Managotry” monazite (554 Ma Pb/U ID-TIMS<br />

age, Paquette et al. 1994) and “554” monazite were analyzed concurrently (once every 5<br />

unknowns) and mass bias- and fractionation-corrected based on measured isotopic ratios <strong>of</strong> <strong>the</strong><br />

primary reference material. During <strong>the</strong> analytical period, repeat analyses <strong>of</strong> FC-1 gave a weighted<br />

mean 206 Pb/ 238 U age <strong>of</strong> 56.5 ± 0.4 Ma, MSWD = 0.7, and a weighted mean 208 Pb/ 232 Th age <strong>of</strong> 54.0<br />

± 0.3 Ma, MSWD = 0.8 (2σ) (n = 15). Repeat analyses <strong>of</strong> Managotry yield a weighted mean<br />

206 Pb/ 238 U age <strong>of</strong> 553.5 ± 3.3 Ma, MSWD = 0.5, and a weighted mean 208 Pb/ 232 Th age <strong>of</strong> 557.8 ±<br />

3.1 Ma, MSWD = 0.7 (2σ) (n = 15). Data reduction, including corrections for baseline,<br />

instrumental drift, mass bias, down-hole fractionation and uncorrected age calculations were<br />

carried out using Iolite version 2.1.2. Full details <strong>of</strong> <strong>the</strong> data reduction methodology can be found<br />

in Paton et al. (2010). All uncertainties are quoted at 2σ and include contributions from <strong>the</strong><br />

external reproducibility <strong>of</strong> <strong>the</strong> primary reference material for <strong>the</strong> 207 Pb/ 206 Pb, 206 Pb/ 238 U ratios and<br />

208 Pb/ 232 Th ratios. Concordia diagrams were constructed using Isoplot 3.75 (Ludwig, 2012).<br />

2.6.1.2 Results<br />

Monazite grains exhibit a variety <strong>of</strong> textures, including elongate, rounded and<br />

comminuted into masses <strong>of</strong> fine-grained fragments (Fig. 2.16, see Appendix C for additional BSE<br />

images). Grains are typically between 20-200 μm in diameter or width/ length. The majority <strong>of</strong><br />

monazite grains are strongly deformed and characterized by rounded and ragged edges, brittle<br />

fractures, and occasional pitted interiors. Rounded grains occasionally show a greater degree <strong>of</strong><br />

internal deformation, consisting <strong>of</strong> intense internal fracturing and pitting <strong>of</strong> <strong>the</strong> grain. All elongate<br />

55


Figure 2.16 Backscatter electron images (BSE image) <strong>of</strong> monazite showing textural relationships with<br />

adjacent minerals and <strong>the</strong> dominant fabric (S), and microprobe generated yttrium and backscatter electron<br />

maps (BSE map) for samples (a) KH18a, (b) HK109, and (c) HK117a. 208 Pb/ 232 Th ages ± 2σ are shown on<br />

yttrium maps for all points analyzed. Diameter <strong>of</strong> analyzed points (black circles) are 7 μm. “N/A” refers to<br />

analyses not used in <strong>the</strong> final age interpretation. Note that yttrium color scales are not all identical, and that<br />

warmer and cooler colors indicate higher and lower elemental abundances, respectively.<br />

56


Figure 2.16 – continued.<br />

57


Figure 2.16 – continued.<br />

58


monazite grains are aligned parallel to <strong>the</strong> main foliation and located adjacent to or within pristine<br />

muscovite and biotite grains defining <strong>the</strong> dominant foliation (Fig. 2.16).<br />

Characteristic yttrium (Y) and thorium (Th) compositional zoning is present in all<br />

monazite grains. The vast majority <strong>of</strong> <strong>the</strong>se grains exhibit Y-poor / Th-rich cores and Y-rich / Thpoor<br />

rims (Fig. 2.16, see Appendix C for additional elemental maps). Rare grains show subtle<br />

deviations from <strong>the</strong> above trend, including <strong>the</strong> absence <strong>of</strong> Y / Th zoning, a random distribution <strong>of</strong><br />

Y / Th zoning, opposite Y / Th zoning, and an additional Y-rich / Th-poor core. Note, <strong>the</strong><br />

chemical formula for <strong>the</strong> analyzed monazite is (Ce, La, Nd, Th, Y)PO 4 , <strong>the</strong>refore it is expected<br />

that Y and Th are inversely related, as <strong>the</strong>y may substitute for one ano<strong>the</strong>r. For all monazite<br />

grains, locations <strong>of</strong> <strong>the</strong> spot analyses were based on Y / Th compositional zoning and were<br />

selected to constrain ages associated with different chemical domains (Fig. 2.16). The full U-Th-<br />

Pb data set is presented in Table 2.1, and graphically presented as 208 Pb/ 232 Th vs. 206 Pb/ 238 U<br />

concordia and 208 Pb/ 232 Th age probability distribution diagrams in Figure 2.17.<br />

For sample KH18a, a total <strong>of</strong> 65 meaningful analyses were obtained on six monazite<br />

grains (Fig. 2.17, Table 2.1). These data yield minor peaks in age distribution at ~26 Ma and ~24<br />

-22 Ma, with <strong>the</strong> bulk <strong>of</strong> ages occurring between ~21 -15 Ma. The oldest grouping <strong>of</strong> ages at ~26<br />

Ma corresponds to a minor group <strong>of</strong> Y-poor cores. The next youngest age population corresponds<br />

to a minor peak <strong>of</strong> Y-rich rims between ~24 -21 Ma. The bulk <strong>of</strong> <strong>the</strong> ages record a secondary<br />

Figure 2.17 (Next page) U-Th-Pb concordia (left column) and probability (right column)<br />

diagrams for dated monazite. Pink and green color coding in both concordia and probability cures<br />

correspond to analyses <strong>of</strong> Y-high rims, and Y-low cores, respectively. Blue analyses are<br />

interpreted as mixtures between Y domains. Color-coding within probability diagrams is not<br />

exact and only intended to show general trends. The low frequency area <strong>of</strong> HK117a, between 12 -<br />

14 Ma, is not color-coded as it does not show a distinct distribution <strong>of</strong> high/low Y analyses. Total<br />

number <strong>of</strong> analyses per sample (n) is recorded in <strong>the</strong> upper right corner <strong>of</strong> each probability plot.<br />

Concordia diagrams plot 206 Pb/ 238 U against 208 Pb/ 232 Th to avoid complications from 235 U- 207 Pb<br />

disequilibrium. All plots were constructed with Isoplot 3.75 (Ludwig, 2012).<br />

59


Table 2-1 U-Th-Pb geochronologic data<br />

Analysis Position<br />

KH18a<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK18a_2_01 H-Y,R 45 3050 30100 9.8 0.359 6.13 0.00377 5.57 0.192 11.46 0.95 0.001092 4.85 24.2 1.4 22.1 1.1<br />

HK18a_2_02 H-Y,R 59 4860 51700 10.4 0.0929 2.80 0.0027 1.96 0.0345 3.19 0.38 0.000817 2.08 17.4 0.3 16.5 0.3<br />

HK18a_2_03 H-Y,R 66 6390 55700 8.4 0.0864 1.85 0.0028 2.00 0.033 2.61 0.74 0.000856 1.87 18.0 0.4 17.3 0.3<br />

HK18a_2_04 H-Y,R 62 5180 54000 10.0 0.0878 1.94 0.0027 2.11 0.03311 2.84 0.72 0.000831 2.05 17.4 0.4 16.8 0.3<br />

HK18a_2_05 H-Y,R 64 5590 54200 9.4 0.0844 2.13 0.00277 1.77 0.03223 2.79 0.66 0.000854 1.87 17.9 0.3 17.3 0.3<br />

HK18a_2_06 H-Y,R 67 6510 56600 8.4 0.0852 1.88 0.00272 2.06 0.03165 2.91 0.73 0.000854 2.22 17.5 0.4 17.3 0.4<br />

HK18a_2_07 H-Y,R 87 11200 67800 5.9 0.0848 1.13 0.00291 1.99 0.03391 2.12 0.84 0.000921 2.06 18.7 0.4 18.6 0.4<br />

HK18a_2_08 H-Y,R 63 6910 48100 6.9 0.092 1.74 0.00303 1.98 0.0381 2.89 0.73 0.000938 2.13 19.5 0.4 19.0 0.4<br />

HK18a_2_09 H-Y,R 71 7090 58700 8.2 0.0916 1.75 0.00282 2.17 0.03605 2.75 0.70 0.00086 2.21 18.1 0.4 17.4 0.4<br />

HK18a_2_10 H-Y,R 72 7260 54800 7.6 0.1259 1.59 0.00305 1.87 0.0527 2.28 0.73 0.000929 1.94 19.7 0.4 18.8 0.4<br />

HK18a_2_11 L-Y,C 84 13970 64200 4.8 0.0854 1.29 0.00292 1.82 0.03367 2.11 0.77 0.000921 1.95 18.8 0.3 18.6 0.4<br />

HK18a_2_12 L-Y,C 91 13820 67700 5.2 0.1049 1.14 0.00301 2.36 0.0433 2.54 0.87 0.00095 2.32 19.4 0.5 19.2 0.4<br />

HK18a_2_13 L-Y,C 92 10120 70700 7.6 0.0998 1.70 0.00304 2.30 0.041 2.68 0.81 0.000927 2.16 19.6 0.5 18.7 0.4<br />

HK18a_2_14 L-Y,C 85 13290 62400 5.2 0.1167 1.20 0.00302 2.15 0.0487 2.46 0.88 0.000948 2.11 19.5 0.4 19.2 0.4<br />

HK18a_2_15 L-Y,C 82 11810 60200 5.7 0.1185 1.43 0.00308 1.92 0.0498 2.01 0.78 0.00096 1.98 19.8 0.4 19.4 0.4<br />

HK18a_2_16 L-Y,C 66 9500 50100 5.8 0.0937 1.49 0.00299 2.41 0.0383 2.87 0.82 0.000939 2.56 19.2 0.5 19.0 0.5<br />

HK18a_2_17 H-Y,R 103 16470 74500 5.0 0.1459 4.46 0.00311 2.16 0.0642 5.92 0.69 0.000977 2.05 20.0 0.4 19.7 0.4<br />

HK18a_2_18 H-Y,R 59 8230 47100 6.3 0.1182 3.98 0.00287 2.16 0.0471 5.10 0.64 0.000875 2.29 18.5 0.4 17.7 0.4<br />

HK18a_2_19 L-Y,C 58 8490 44300 5.7 0.1082 2.59 0.00296 1.93 0.0443 3.16 0.62 0.00091 2.09 19.0 0.4 18.4 0.4<br />

HK18a_2_20 H-Y,R 71 7670 44100 6.2 0.297 4.38 0.00381 3.41 0.156 7.69 0.91 0.001128 3.28 24.5 0.9 22.8 0.7<br />

HK18a_4_01 L-Y,R 68 5480 39000 7.4 0.3788 1.06 0.00438 1.49 0.2274 2.11 0.85 0.001253 1.44 28.2 0.4 25.3 0.4<br />

HK18a_4_02 L-Y,R 35 2660 24500 8.7 0.3675 1.58 0.0043 2.79 0.217 3.87 0.90 0.001085 4.06 27.7 0.8 21.9 0.9<br />

HK18a_4_03 L-Y,R 146 5870 36740 6.4 0.579 2.59 0.0099 6.46 0.803 8.97 0.99 0.00276 6.16 63.5 4.1 55.8 3.5<br />

HK18a_4_04 L-Y,R 66 4340 41200 9.8 0.4059 1.06 0.00445 3.15 0.2496 3.37 0.95 0.001109 2.71 28.6 0.9 22.4 0.6<br />

HK18a_4_05 L-Y,R 102 4150 32290 7.9 0.6066 0.74 0.00866 2.08 0.728 2.61 0.96 0.002248 2.18 55.6 1.1 45.4 1.0<br />

HK18a_4_06 H-Y,C 120 8240 34240 4.2 0.5096 1.06 0.0062 1.94 0.439 2.73 0.94 0.002513 2.71 39.9 0.8 50.7 1.4<br />

HK18a_4_07 H-Y,C 151 7630 34160 4.5 0.5743 0.68 0.00792 1.77 0.635 2.20 0.96 0.003176 1.64 50.8 0.9 64.1 1.0<br />

HK18a_4_08 H-Y,C 107 9970 32280 3.3 0.4271 0.96 0.00519 1.93 0.3076 2.54 0.94 0.002397 2.29 33.4 0.6 48.4 1.1<br />

HK18a_4_09 L-Y,R 92 10280 37030 3.6 0.3375 2.90 0.00437 2.52 0.205 4.88 0.93 0.001814 2.65 28.1 0.7 36.6 1.0<br />

61


Analysis Position<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK18a_4_10 L-Y,R 140 4840 41700 8.6 0.6163 1.17 0.00919 3.26 0.791 4.17 0.98 0.002423 2.81 58.9 1.9 48.9 1.4<br />

HK18a_4_11 L-Y,R 59 10410 45410 4.3 0.0825 1.21 0.00286 1.72 0.03323 2.29 0.83 0.000943 1.59 18.4 0.3 19.1 0.3<br />

HK18a_4_12 L-Y,R 114 8260 50630 6.1 0.428 3.74 0.00518 4.83 0.309 8.09 0.98 0.001646 3.28 33.3 1.6 33.2 1.1<br />

HK18a_4_13 L-Y,R 148 8330 47900 5.6 0.5347 0.47 0.00695 2.16 0.516 2.33 0.98 0.002262 2.08 44.6 1.0 45.7 0.9<br />

HK18a_4_14 L-Y,R 106 10640 44300 4.1 0.3559 1.85 0.00458 2.05 0.2279 3.33 0.88 0.00172 2.33 29.4 0.6 34.7 0.8<br />

HK18a_4_15 L-Y,R 217 9590 35300 3.6 0.6372 0.42 0.00998 2.00 0.892 2.24 0.99 0.0045 3.11 64.0 1.3 90.8 2.7<br />

HK18a_5_01 Mix 25 2950 12500 4.3 0.415 3.13 0.00401 3.49 0.232 6.03 0.92 0.001559 4.55 25.8 0.9 31.5 1.4<br />

HK18a_5_02 Mix 35 3780 22900 6.4 0.273 1.68 0.00327 2.69 0.1253 3.35 0.83 0.001092 2.75 21.0 0.6 22.1 0.6<br />

HK18a_5_03 Mix 90 9540 71900 7.9 0.1599 2.06 0.00296 1.76 0.0658 2.89 0.68 0.000911 1.87 19.1 0.3 18.4 0.3<br />

HK18a_5_04 Mix 86 9730 61600 6.7 0.1717 3.38 0.0032 2.00 0.0763 4.59 0.77 0.001009 2.08 20.6 0.4 20.4 0.4<br />

HK18a_5_05 Mix 136 12930 85700 7.0 0.3046 1.84 0.00402 2.11 0.1703 3.23 0.86 0.001156 2.08 25.9 0.6 23.4 0.5<br />

HK18a_5_06 Mix 80 9230 56700 6.6 0.2168 2.91 0.00334 2.07 0.1005 4.18 0.82 0.00102 1.96 21.5 0.4 20.6 0.4<br />

HK18a_5_07 Mix 78 10020 62600 6.6 0.0886 2.26 0.0028 1.72 0.0346 2.89 0.65 0.000898 1.67 18.0 0.3 18.1 0.3<br />

HK18a_5_08 Mix 83 8080 56700 7.3 0.3051 2.88 0.00399 3.26 0.17 5.88 0.94 0.001061 2.73 25.7 0.8 21.4 0.6<br />

HK18a_5_09 Mix 84 9670 61780 6.6 0.1323 2.72 0.00309 1.78 0.0564 3.19 0.66 0.000985 1.83 19.9 0.4 19.9 0.4<br />

HK18a_5_10 Mix 79 9610 60100 6.4 0.1926 2.75 0.0031 2.19 0.0817 4.04 0.75 0.000949 2.00 20.0 0.4 19.2 0.4<br />

HK18a_6_01 H-Y,R 69 8750 60700 6.9 0.0877 2.39 0.00265 1.73 0.03195 2.97 0.62 0.000821 1.83 17.1 0.3 16.6 0.3<br />

HK18a_6_02 H-Y,R 69 9730 55900 5.7 0.0782 2.05 0.00276 1.63 0.03022 2.78 0.59 0.000884 1.70 17.8 0.3 17.9 0.3<br />

HK18a_6_03 H-Y,R 90 9960 72000 7.1 0.088 1.82 0.00287 1.74 0.0352 3.13 0.70 0.00089 1.69 18.5 0.3 18.0 0.3<br />

HK18a_6_04 H-Y,R 69 9300 55600 5.9 0.0793 2.14 0.00274 1.94 0.03011 2.99 0.64 0.000877 1.94 17.6 0.3 17.7 0.3<br />

HK18a_6_05 H-Y,R 63 8290 51700 6.1 0.0775 2.32 0.00273 1.69 0.02906 2.99 0.61 0.000867 1.61 17.6 0.3 17.5 0.3<br />

HK18a_6_06 H-Y,R 63 8280 56600 6.8 0.0781 2.43 0.00251 1.71 0.02727 2.97 0.54 0.000794 1.64 16.2 0.3 16.0 0.3<br />

HK18a_6_07 H-Y,R 44 4400 32700 7.3 0.277 5.78 0.00317 5.05 0.123 9.76 0.89 0.000948 4.64 20.4 1.0 19.1 0.9<br />

HK18a_6_08 H-Y,R 79 7160 48100 6.7 0.293 2.66 0.00379 2.01 0.153 4.12 0.88 0.001167 2.06 24.4 0.5 23.6 0.5<br />

HK18a_6_09 H-Y,R 61 7540 49300 6.5 0.0773 1.55 0.00275 1.78 0.02959 2.84 0.79 0.000876 1.83 17.7 0.3 17.7 0.3<br />

HK18a_6_10 H-Y,R 64 7850 50400 6.4 0.0788 1.90 0.0028 1.61 0.03056 2.59 0.63 0.000898 1.78 18.0 0.3 18.2 0.3<br />

HK18a_6_11 L-Y,C 54 7260 45420 6.3 0.0762 2.36 0.0026 2.11 0.02697 3.23 0.69 0.000831 2.17 16.8 0.4 16.8 0.4<br />

HK18a_6_12 L-Y,C 48 6850 40030 5.8 0.1417 4.23 0.00263 1.97 0.0508 4.92 0.55 0.000849 2.12 17.0 0.3 17.2 0.4<br />

HK18a_6_13 L-Y,C 44 4150 32500 7.9 0.224 5.80 0.00309 3.24 0.098 8.78 0.92 0.00094 2.66 19.9 0.7 19.0 0.5<br />

HK18a_6_14 H-Y,R 71 6720 55500 8.3 0.1282 2.57 0.00293 1.95 0.0518 3.09 0.62 0.000895 1.79 18.9 0.4 18.1 0.3<br />

HK18a_6_15 H-Y,R 66 6400 49500 7.8 0.0856 2.34 0.00296 1.72 0.035 3.14 0.62 0.000945 1.59 19.1 0.3 19.1 0.3<br />

HK18a_7_01 H-Y,R 29 1881 14540 7.8 0.4426 1.15 0.00439 1.98 0.2671 2.58 0.89 0.001446 2.21 28.2 0.6 29.2 0.7<br />

62


Analysis Position<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK18a_7_02 H-Y,R 53 5760 47100 8.3 0.0932 1.82 0.00255 1.69 0.03317 2.80 0.71 0.000792 1.64 16.4 0.3 16.0 0.3<br />

HK18a_7_03 H-Y,R 53 5056 43900 8.8 0.1357 2.21 0.00284 1.73 0.0528 3.03 0.63 0.000861 1.63 18.3 0.3 17.4 0.3<br />

HK18a_7_04 H-Y,R 57 5940 47700 8.1 0.1035 1.93 0.00272 1.98 0.0387 2.84 0.69 0.000844 2.01 17.5 0.4 17.1 0.3<br />

HK18a_7_05 L-Y,C 65 6490 49000 7.6 0.1034 1.64 0.00303 1.91 0.04259 2.30 0.69 0.00096 1.88 19.5 0.4 19.4 0.4<br />

HK18a_7_06 L-Y,C 76 7130 55600 7.8 0.141 1.49 0.00313 1.92 0.0601 2.66 0.81 0.000976 1.95 20.1 0.4 19.7 0.4<br />

HK18a_7_07 H-Y,R 70 6390 50000 7.9 0.1674 2.09 0.00326 1.66 0.0751 2.93 0.75 0.001006 1.69 21.0 0.4 20.3 0.4<br />

HK18a_7_08 H-Y,R 71 6280 45590 7.3 0.2647 2.68 0.0036 2.03 0.1299 4.39 0.86 0.001113 1.98 23.2 0.5 22.5 0.4<br />

HK18a_7_09 H-Y,R 71 6140 48000 7.8 0.2285 0.83 0.00349 1.43 0.1101 1.91 0.89 0.001063 1.41 22.5 0.3 21.5 0.3<br />

HK18a_7_10 H-Y,R 97 4780 37100 7.8 0.5796 1.02 0.00662 2.42 0.528 3.22 0.97 0.001896 2.48 42.5 1.0 38.3 1.0<br />

HK18a_7_11 H-Y,R 96 8730 76200 8.6 0.0902 1.22 0.00294 1.87 0.03651 2.14 0.83 0.000896 1.90 18.9 0.4 18.1 0.3<br />

HK18a_7_12 H-Y,R 86 8500 54000 6.3 0.2558 2.78 0.00374 2.08 0.1317 4.40 0.87 0.001152 1.91 24.1 0.5 23.3 0.5<br />

HK18a_7_13 H-Y,R 58 5450 34300 6.2 0.3861 1.74 0.00398 2.36 0.2127 3.67 0.89 0.001195 2.18 25.6 0.6 24.1 0.5<br />

HK18a_7_14 H-Y,R 88 7470 43600 5.8 0.3797 2.45 0.00469 2.56 0.248 4.84 0.94 0.001432 2.65 30.2 0.8 28.9 0.8<br />

HK18a_7_15 H-Y,R 82 8050 49110 6.1 0.3657 1.86 0.00406 1.99 0.2043 3.28 0.91 0.001175 1.96 26.1 0.5 23.7 0.5<br />

HK18a_8_01 L-Y 100 8580 55700 6.6 0.4096 1.05 0.00472 1.78 0.2662 2.40 0.92 0.001267 1.74 30.4 0.5 25.6 0.4<br />

HK18a_8_02 L-Y 78 10520 58600 5.8 0.1334 3.37 0.00303 1.91 0.0556 4.32 0.71 0.000941 1.91 19.5 0.4 19.0 0.4<br />

HK18a_8_03 L-Y 44 4210 29000 7.0 0.313 7.03 0.00382 4.19 0.165 10.91 0.95 0.001076 3.90 24.6 1.1 21.7 0.8<br />

HK18a_8_04 L-Y 108 13340 83300 6.3 0.0993 1.31 0.00295 1.80 0.04022 2.21 0.78 0.000916 1.86 19.0 0.3 18.5 0.3<br />

HK18a_8_05 L-Y 76 6350 41800 6.7 0.3438 1.05 0.00423 1.73 0.1998 2.10 0.89 0.001272 1.81 27.2 0.5 25.7 0.5<br />

HK18a_8_06 L-Y 133 13730 95500 7.0 0.149 1.88 0.00321 1.65 0.0661 2.57 0.73 0.00098 1.63 20.7 0.3 19.8 0.3<br />

HK18a_8_07 L-Y 39 3130 19800 6.3 0.3235 1.89 0.00428 2.57 0.1875 3.25 0.83 0.001466 2.66 27.5 0.7 29.6 0.8<br />

HK18a_8_08 H-Y 69 3272 39670 12.2 0.453 2.65 0.00521 3.84 0.319 6.27 0.96 0.001228 2.52 33.5 1.3 24.8 0.6<br />

HK18a_8_09 H-Y 68 4404 57600 13.0 0.1058 2.84 0.00278 1.94 0.0407 3.69 0.63 0.000836 1.79 17.9 0.4 16.9 0.3<br />

HK18a_8_10 H-Y 127 10540 86100 8.2 0.2044 4.84 0.00347 1.90 0.0987 6.28 0.82 0.00104 1.83 22.4 0.4 21.0 0.4<br />

HK18a_8_11 H-Y 69 6190 56700 9.2 0.0836 2.15 0.00277 1.66 0.03174 2.65 0.61 0.000874 1.72 17.8 0.3 17.7 0.3<br />

HK18a_8_12 H-Y 67 5520 52500 9.5 0.139 1.65 0.00298 2.05 0.0577 2.77 0.75 0.000919 1.96 19.2 0.4 18.6 0.4<br />

HK18a_8_13 H-Y 86 5120 49400 9.6 0.3432 2.16 0.00442 1.99 0.2076 3.61 0.89 0.001261 1.90 28.4 0.6 25.5 0.5<br />

HK18a_8_14 H-Y 83 5635 58500 10.4 0.239 5.86 0.00352 3.13 0.119 8.40 0.91 0.001017 2.46 22.6 0.7 20.6 0.5<br />

HK18a_8_15 H-Y 71 3376 50280 15.0 0.3438 2.12 0.004 2.45 0.1889 4.02 0.92 0.001016 2.26 25.8 0.6 20.5 0.5<br />

HK18a_8_16 H-Y 77 5160 55000 10.7 0.258 6.20 0.0035 3.43 0.122 9.84 0.95 0.001003 2.49 22.5 0.8 20.3 0.5<br />

HK18a_8_17 H-Y 87 1492 22340 15.0 0.72 0.68 0.01593 3.64 1.57 4.01 0.99 0.002808 3.17 101.9 3.7 56.7 1.8<br />

HK18a_8_18 H-Y 73 5810 57500 9.9 0.1332 3.75 0.00297 1.65 0.0545 4.77 0.71 0.000915 1.64 19.1 0.3 18.5 0.3<br />

63


Analysis Position<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK18a_8_19 H-Y 115 10410 89400 8.6 0.0909 1.21 0.00296 1.83 0.03699 2.27 0.80 0.000921 1.63 19.0 0.4 18.6 0.3<br />

HK18a_8_20 H-Y 62 4690 53500 11.4 0.0944 1.69 0.00272 2.02 0.03555 2.62 0.73 0.000838 2.03 17.5 0.4 16.9 0.3<br />

HK109<br />

HK109_2_01 H-Y,R 36 5867 34488 5.88 0.08409 2.97 0.00236 2.16 0.02710 3.46 0.57 0.000708 2.42 15.2 0.3 14.3 0.3<br />

HK109_2_02 H-Y,R 39 6970 36542 5.25 0.08084 2.61 0.00242 2.44 0.02719 3.48 0.73 0.000746 2.43 15.6 0.4 15.1 0.4<br />

HK109_2_03 H-Y,R 42 7859 38601 4.93 0.07828 2.82 0.00246 1.98 0.02649 2.92 0.48 0.000759 2.13 15.8 0.3 15.3 0.3<br />

HK109_2_04 H-Y,R 41 7076 39225 5.61 0.08228 3.05 0.00243 2.29 0.02746 3.64 0.59 0.000740 2.47 15.7 0.4 14.9 0.4<br />

HK109_2_05 H-Y,R 45 7227 41952 5.77 0.07755 2.74 0.00248 2.52 0.02657 3.50 0.60 0.000751 2.73 16.0 0.4 15.2 0.4<br />

HK109_2_06 H-Y,R 49 8150 48583 6.02 0.08200 3.39 0.00234 2.60 0.02683 4.70 0.70 0.000713 2.66 15.1 0.4 14.4 0.4<br />

HK109_2_07 H-Y,R 52 10055 49174 4.93 0.07728 2.81 0.00245 2.54 0.02589 3.51 0.65 0.000760 2.53 15.8 0.4 15.4 0.4<br />

HK109_2_08 H-Y,R 38 6694 40851 6.15 0.09938 3.50 0.00219 2.47 0.02943 3.96 0.55 0.000665 2.13 14.1 0.3 13.4 0.3<br />

HK109_2_09 L-Y,C 77 9813 62883 6.46 0.08099 2.31 0.00280 2.48 0.03128 3.21 0.73 0.000869 2.52 18.0 0.4 17.5 0.4<br />

HK109_2_10 L-Y,C 81 10218 65197 6.37 0.08466 1.92 0.00285 2.58 0.03316 3.34 0.82 0.000884 2.59 18.4 0.5 17.9 0.5<br />

HK109_2_11 L-Y,C 78 9474 64086 6.76 0.08304 2.22 0.00283 2.45 0.03245 3.29 0.74 0.000864 2.50 18.2 0.4 17.4 0.4<br />

HK109_3_01 L-Y 69 11710 61451 5.19 0.07357 2.81 0.00261 2.51 0.02678 4.09 0.80 0.000803 2.78 16.8 0.4 16.2 0.4<br />

HK109_3_02 L-Y 68 11470 60350 5.20 0.07432 2.66 0.00266 3.16 0.02705 3.69 0.72 0.000814 3.46 17.1 0.5 16.4 0.6<br />

HK109_3_03 L-Y 67 11405 59740 5.25 0.07489 2.21 0.00264 2.79 0.02698 3.41 0.75 0.000803 2.67 17.0 0.5 16.2 0.4<br />

HK109_3_04 L-Y 67 11379 59548 5.20 0.07386 2.11 0.00266 3.50 0.02729 4.24 0.88 0.000821 3.46 17.1 0.6 16.6 0.6<br />

HK109_3_05 L-Y 75 11210 62640 5.58 0.07935 2.30 0.00277 3.27 0.03015 4.08 0.80 0.000858 3.39 17.8 0.6 17.3 0.6<br />

HK109_3_06 L-Y 87 12143 72044 5.90 0.08758 2.20 0.00284 3.29 0.03445 3.90 0.82 0.000869 3.12 18.3 0.6 17.6 0.5<br />

HK109_4_01 H-Y,R 39 6869 39827 5.73 0.13569 4.58 0.00234 2.86 0.04434 5.86 0.63 0.000722 3.05 15.1 0.4 14.6 0.4<br />

HK109_4_02 H-Y,R 39 6274 34357 5.40 0.15993 8.06 0.00259 6.49 0.05437 11.50 0.76 0.000834 6.50 16.7 1.1 16.9 1.1<br />

HK109_4_03 H-Y,R 46 9983 50358 4.96 0.09315 4.04 0.00207 2.72 0.02679 4.99 0.62 0.000662 2.75 13.3 0.4 13.4 0.4<br />

HK109_4_04 H-Y,R 54 9944 57213 5.75 0.08418 2.63 0.00225 2.72 0.02632 3.46 0.69 0.000669 2.64 14.5 0.4 13.5 0.4<br />

HK109_4_05 H-Y,R 57 13529 60859 4.48 0.07607 2.36 0.00222 3.16 0.02308 3.84 0.83 0.000673 3.35 14.3 0.5 13.6 0.5<br />

HK109_4_06 L-Y,C 70 9381 60399 6.43 0.07447 2.79 0.00266 3.08 0.02758 4.04 0.73 0.000817 2.93 17.2 0.5 16.5 0.5<br />

HK109_4_07 L-Y,C 68 8584 58399 6.85 0.07471 2.25 0.00266 2.56 0.02737 3.15 0.78 0.000818 2.70 17.1 0.4 16.5 0.4<br />

HK109_4_08 L-Y,C 67 7730 57545 7.41 0.07607 2.36 0.00271 2.75 0.02868 3.42 0.74 0.000839 2.73 17.5 0.5 17.0 0.5<br />

HK109_4_09 L-Y,C 60 7446 52272 7.08 0.07421 2.40 0.00267 3.03 0.02731 3.66 0.75 0.000812 2.80 17.2 0.5 16.4 0.5<br />

HK109_4_10 H-Y,R 63 7041 52058 7.62 0.07419 2.22 0.00279 2.53 0.02860 3.16 0.71 0.000853 2.56 17.9 0.5 17.2 0.4<br />

HK109_5_01 H-Y,R 40 5150 40823 8.15 0.08141 3.37 0.00230 3.21 0.02605 4.68 0.71 0.000695 2.99 14.8 0.5 14.0 0.4<br />

HK109_5_02 H-Y,R 40 5334 40577 7.78 0.08141 3.04 0.00232 3.06 0.02622 4.05 0.68 0.000710 3.16 14.9 0.5 14.3 0.5<br />

64


Analysis Position<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK109_5_03 H-Y,R 35 5934 39858 6.84 0.09178 4.75 0.00203 4.63 0.02549 6.09 0.71 0.000628 4.83 13.1 0.6 12.7 0.6<br />

HK109_5_04 H-Y,R 43 6121 43391 7.18 0.08018 3.26 0.00237 4.04 0.02610 5.10 0.79 0.000719 4.00 15.3 0.6 14.5 0.6<br />

HK109_5_05 H-Y,R 43 6068 43384 7.20 0.07952 3.42 0.00232 3.25 0.02551 4.65 0.69 0.000702 3.39 15.0 0.5 14.2 0.5<br />

HK109_5_06 L-Y 136 7741 53854 6.85 0.06516 1.70 0.00599 3.64 0.05342 3.86 0.91 0.001823 3.52 38.5 1.4 36.8 1.3<br />

HK109_5_07 L-Y 66 7660 51408 6.64 0.07722 2.60 0.00281 3.53 0.02971 4.17 0.82 0.000907 3.39 18.1 0.6 18.3 0.6<br />

HK109_5_08 H-Y,R 58 7943 51836 6.50 0.07634 2.71 0.00252 3.93 0.02652 4.96 0.84 0.000787 3.89 16.2 0.6 15.9 0.6<br />

HK109_5_09 H-Y,R 348 5671 40765 7.14 0.06162 1.30 0.01554 4.58 0.13254 4.67 0.95 0.006171 4.46 99.4 4.5 124.3 5.5<br />

HK109_5_10 M-Y,C 622 3261 23161 6.95 0.05881 0.98 0.06527 3.66 0.52320 3.45 0.96 0.019476 4.11 407.2 ## 389.7 ##<br />

HK109_6_01 H-Y,R 53 7831 50546 6.69 0.07684 2.81 0.00237 3.82 0.02574 4.36 0.80 0.000722 3.52 15.3 0.6 14.6 0.5<br />

HK109_6_02 H-Y,R 55 8356 53705 6.66 0.07975 2.66 0.00239 3.72 0.02638 4.14 0.82 0.000738 3.55 15.4 0.6 14.9 0.5<br />

HK109_6_03 H-Y,R 58 8768 55526 6.64 0.07890 2.45 0.00242 3.70 0.02603 4.04 0.85 0.000737 3.54 15.6 0.6 14.9 0.5<br />

HK109_6_04 H-Y,R 56 8435 54456 6.84 0.08103 3.02 0.00240 4.10 0.02682 4.89 0.79 0.000725 4.06 15.5 0.6 14.7 0.6<br />

HK109_6_05 H-Y,R 64 10273 59909 6.24 0.07919 2.73 0.00250 3.46 0.02779 4.28 0.79 0.000771 3.34 16.1 0.6 15.6 0.5<br />

HK109_6_06 L-Y,C 69 16150 61746 4.09 0.07015 2.16 0.00258 3.04 0.02433 4.01 0.84 0.000813 3.19 16.6 0.5 16.4 0.5<br />

HK109_6_07 L-Y,C 70 15297 61477 4.29 0.06847 2.35 0.00256 3.43 0.02408 4.05 0.83 0.000805 3.31 16.5 0.6 16.3 0.5<br />

HK109_6_08 H-Y,R 58 9043 56834 6.67 0.08013 2.36 0.00240 3.62 0.02593 4.35 0.86 0.000745 3.67 15.4 0.6 15.1 0.6<br />

HK109_6_09 H-Y,R 62 9204 58088 6.70 0.08204 2.43 0.00248 3.29 0.02828 4.39 0.82 0.000763 3.32 16.0 0.5 15.4 0.5<br />

HK109_6_10 H-Y,R 50 7279 49743 7.07 0.08405 2.77 0.00236 3.31 0.02716 3.95 0.73 0.000731 3.55 15.2 0.5 14.8 0.5<br />

HK109_9_01 H-Y,R 56 8315 53777 6.47 0.08691 4.39 0.00244 3.84 0.02973 6.04 0.63 0.000740 3.99 15.7 0.6 15.0 0.6<br />

HK109_9_02 H-Y,R 53 7283 51789 7.07 0.08960 2.59 0.00246 3.94 0.03029 4.45 0.79 0.000738 3.61 15.8 0.6 14.9 0.5<br />

HK109_9_03 H-Y,R 67 11173 57095 4.96 0.07526 2.72 0.00274 5.22 0.02873 5.50 0.86 0.000857 5.04 17.7 0.9 17.3 0.9<br />

HK109_9_04 L-Y,C 69 10626 58723 5.47 0.08263 2.43 0.00276 4.20 0.03147 5.29 0.87 0.000843 4.28 17.8 0.7 17.0 0.7<br />

HK109_9_05 L-Y,C 56 8767 44249 5.00 0.08638 2.33 0.00294 4.16 0.03449 4.49 0.84 0.000926 4.24 18.9 0.8 18.7 0.8<br />

HK109_9_06 L-Y,C 62 9834 51784 5.24 0.08787 2.26 0.00277 3.37 0.03311 3.89 0.79 0.000855 3.60 17.8 0.6 17.3 0.6<br />

HK109_10_01 L-Y 80 14739 71543 4.71 0.07048 2.74 0.00259 3.04 0.02484 4.16 0.77 0.000813 2.87 16.7 0.5 16.4 0.5<br />

HK109_10_02 H-Y 77 8491 63171 7.28 0.08345 2.49 0.00291 3.72 0.03354 4.21 0.82 0.000882 3.67 18.8 0.7 17.8 0.7<br />

HK109_10_03 H-Y 66 8296 54042 6.40 0.07404 3.00 0.00289 3.55 0.02939 4.51 0.76 0.000900 3.68 18.6 0.7 18.2 0.7<br />

HK109_10_04 H-Y 65 7777 54415 6.79 0.07879 2.94 0.00284 4.53 0.03091 5.10 0.82 0.000886 4.36 18.4 0.8 17.9 0.8<br />

HK109_10_05 L-Y 74 10226 65381 6.30 0.07243 2.60 0.00335 6.82 0.03360 6.41 0.91 0.000848 4.59 21.5 1.5 17.1 0.8<br />

HK109_10_06 L-Y 82 10508 73086 6.88 0.07701 2.68 0.00270 3.49 0.02875 3.97 0.81 0.000832 3.45 17.4 0.6 16.8 0.6<br />

HK109_10_07 L-Y 76 8288 65151 7.85 0.07481 2.85 0.00275 4.41 0.02854 4.76 0.83 0.000837 4.23 17.7 0.8 16.9 0.7<br />

65


Analysis Position<br />

HK117A<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK117A_1_1 H-Y 37 6640 30690 4.6 0.2952 3.15 0.00248 4.03 0.1001 5.99 0.85 0.000834 4.20 16.0 0.7 16.9 0.7<br />

HK117A_1_2 H-Y 101 10140 94300 9.4 0.0635 1.89 0.00244 1.80 0.02163 2.40 0.53 0.000747 1.74 15.7 0.3 15.1 0.3<br />

HK117A_1_3 L-Y 51 5520 44300 8.1 0.2521 2.74 0.00302 2.29 0.1052 3.99 0.80 0.0008 2.25 19.4 0.4 16.2 0.4<br />

HK117A_1_4 H-Y 95 10100 88700 8.8 0.0625 1.76 0.00245 1.47 0.021 2.38 0.63 0.000751 1.46 15.7 0.2 15.2 0.2<br />

HK117A_1_5 H-Y 83 10160 73100 7.2 0.2046 2.79 0.00285 2.07 0.0798 4.01 0.76 0.000782 2.17 18.3 0.4 15.8 0.3<br />

HK117A_1_6 L-Y 89 8370 83300 9.9 0.0624 1.92 0.00246 1.75 0.02111 2.79 0.67 0.000745 1.61 15.8 0.3 15.1 0.3<br />

HK117A_1_7 H-Y 79 11120 68300 6.2 0.2093 2.82 0.00281 2.78 0.0825 4.97 0.83 0.000805 2.73 18.1 0.5 16.3 0.5<br />

HK117A_1_8 L-Y 89 8050 82300 10.2 0.0642 2.02 0.00249 1.57 0.02239 2.72 0.52 0.000761 1.71 16.1 0.3 15.4 0.3<br />

HK117A_1_9 H-Y 89 10440 87400 8.3 0.0625 1.92 0.00236 1.66 0.02028 2.51 0.60 0.000717 1.67 15.2 0.3 14.5 0.3<br />

HK117A_1_10 L-Y 86 7850 84100 10.5 0.0637 1.88 0.00247 1.74 0.02169 2.54 0.61 0.000739 1.62 15.9 0.3 14.9 0.3<br />

HK117A_1_11 L-Y 77 6780 71100 10.4 0.0633 1.90 0.00249 1.61 0.02165 2.31 0.56 0.000767 1.56 16.0 0.3 15.5 0.2<br />

HK117A_1_12 L-Y 119 13430 107800 7.9 0.1302 2.84 0.00269 2.60 0.0482 3.94 0.74 0.000804 2.86 17.3 0.5 16.2 0.5<br />

HK117A_1_13 L-Y 79 9850 75300 7.6 0.1815 2.75 0.00269 2.52 0.0677 3.99 0.70 0.000765 2.35 17.3 0.4 15.5 0.4<br />

HK117A_1_14 H-Y 87 7170 83000 11.8 0.066 2.27 0.00253 1.78 0.02284 3.02 0.62 0.000755 1.72 16.3 0.3 15.3 0.3<br />

HK117A_1_15 H-Y 94 6920 88100 12.6 0.0637 1.88 0.0025 1.64 0.02203 2.63 0.67 0.000759 1.58 16.1 0.3 15.3 0.2<br />

HK117A_2_1 H-Y,R 66 10380 68600 6.6 0.0605 1.82 0.00225 1.74 0.01861 2.36 0.60 0.000702 1.85 14.5 0.3 14.2 0.3<br />

HK117A_2_2 H-Y,R 72 7920 69800 8.8 0.0636 2.20 0.00239 1.38 0.02106 2.52 0.41 0.000743 1.62 15.4 0.2 15.0 0.2<br />

HK117A_2_3 H-Y,R 58 4350 48000 10.9 0.3212 1.56 0.00339 2.18 0.1496 3.07 0.89 0.00089 2.47 21.8 0.5 18.0 0.4<br />

HK117A_2_4 H-Y,R 103 7680 91400 12.0 0.1161 1.55 0.00276 1.70 0.0448 2.68 0.63 0.000809 1.61 17.8 0.3 16.3 0.3<br />

HK117A_2_5 H-Y,R 111 8730 105400 12.1 0.0648 1.70 0.00252 1.67 0.02248 2.40 0.63 0.000761 1.71 16.2 0.3 15.4 0.3<br />

HK117A_2_6 H-Y,R 109 7840 102700 13.2 0.0721 1.66 0.00256 1.84 0.02527 2.22 0.56 0.00077 1.82 16.5 0.3 15.6 0.3<br />

HK117A_2_7 H-Y,R 115 8000 104400 13.2 0.1053 2.56 0.00266 1.69 0.0384 3.13 0.58 0.000781 1.79 17.1 0.3 15.8 0.3<br />

HK117A_2_8 H-Y,R 77 5570 67700 12.2 0.1938 2.79 0.00293 1.81 0.0785 3.31 0.71 0.000811 1.85 18.9 0.3 16.4 0.3<br />

HK117A_2_9 L-Y,C 90 6200 81800 13.3 0.0837 2.75 0.00263 1.90 0.0302 3.31 0.51 0.000796 1.88 16.9 0.3 16.1 0.3<br />

HK117A_2_10 L-Y,C 99 6380 83500 13.2 0.1767 1.53 0.00303 1.65 0.0742 2.43 0.72 0.00085 1.53 19.5 0.3 17.2 0.3<br />

HK117A_2_11 L-Y,C 80 6350 73900 11.7 0.0649 2.31 0.0025 1.72 0.02256 2.97 0.49 0.000776 1.80 16.1 0.3 15.7 0.3<br />

HK117A_2_12 L-Y,C 74 8510 71800 8.5 0.0634 2.05 0.00237 1.52 0.02089 2.73 0.66 0.000746 1.47 15.2 0.2 15.1 0.2<br />

HK117A_2_13 H-Y,R 52 5540 44800 8.1 0.1642 3.84 0.00276 2.03 0.0631 4.75 0.65 0.000834 2.04 17.8 0.4 16.8 0.3<br />

HK117A_2_14 L-Y,C 85 6880 80600 11.7 0.0655 2.29 0.00248 1.61 0.0226 3.05 0.56 0.000772 1.55 16.0 0.3 15.6 0.3<br />

HK117A_2_15 L-Y,C 39 3520 35000 9.9 0.2518 1.67 0.00294 2.25 0.1018 2.75 0.82 0.000822 2.07 18.9 0.4 16.6 0.4<br />

HK117A_2_16 L-Y,C 97 6530 78400 11.9 0.2521 1.98 0.00342 2.25 0.117 3.59 0.84 0.000908 2.31 22.0 0.5 18.4 0.4<br />

66


Analysis Position<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK117A_2_17 L-Y,C 100 7930 95700 12.0 0.0677 2.22 0.00242 1.69 0.02301 3.17 0.57 0.000768 1.69 15.6 0.3 15.5 0.3<br />

HK117A_2_18 L-Y,C 95 8330 90700 10.8 0.0668 2.25 0.00243 1.57 0.02229 2.96 0.53 0.000782 1.53 15.6 0.3 15.8 0.2<br />

HK117A_2_19 L-Y,C 82 8390 77300 9.1 0.104 1.73 0.00251 1.87 0.03584 2.51 0.66 0.00079 1.77 16.2 0.3 16.0 0.3<br />

HK117A_2_20 L-Y,C 122 11690 113700 9.7 0.068 2.35 0.00236 1.91 0.02235 3.49 0.60 0.000771 1.95 15.2 0.3 15.6 0.3<br />

HK117A_3_1 H-Y,R 33 3100 26600 8.5 0.226 1.59 0.0027 1.89 0.0836 2.75 0.75 0.00093 1.94 17.4 0.3 18.8 0.4<br />

HK117A_3_2 H-Y,R 34 2999 28320 9.3 0.2621 2.40 0.00302 2.82 0.1086 4.05 0.88 0.000894 2.68 19.4 0.6 18.1 0.5<br />

HK117A_3_3 H-Y,R 40 3466 31890 9.1 0.2864 1.78 0.00319 1.85 0.1245 2.81 0.88 0.000937 1.92 20.5 0.4 18.9 0.4<br />

HK117A_3_4 L-Y,C 110 12030 99790 8.3 0.0649 2.31 0.00226 1.19 0.02005 2.59 0.44 0.000799 1.38 14.6 0.2 16.1 0.2<br />

HK117A_3_5 L-Y,C 118 13120 106600 8.1 0.0677 2.51 0.00226 1.51 0.02089 2.82 0.40 0.000798 1.63 14.5 0.2 16.1 0.3<br />

HK117A_3_6 L-Y,C 116 14620 103700 7.1 0.0661 2.27 0.00224 1.34 0.02039 3.04 0.35 0.000802 1.37 14.5 0.2 16.2 0.2<br />

HK117A_3_7 L-Y,C 114 13220 101500 7.7 0.0665 2.71 0.00225 1.69 0.02031 3.45 0.51 0.0008 1.88 14.5 0.3 16.2 0.3<br />

HK117A_3_8 L-Y,C 104 10550 90800 8.6 0.0678 2.51 0.00227 1.41 0.02089 2.87 0.32 0.000813 1.48 14.6 0.2 16.4 0.2<br />

HK117A_3_9 L-Y,C 130 12600 116900 9.3 0.0757 1.98 0.00222 1.31 0.02282 2.19 0.43 0.000785 1.13 14.3 0.2 15.9 0.2<br />

HK117A_3_10 H-Y,R 125 14760 108900 7.3 0.0767 1.56 0.0022 1.54 0.02286 2.01 0.58 0.000794 1.64 14.2 0.2 16.0 0.3<br />

HK117A_4_1 L-Y,C 154 9430 140300 14.9 0.0613 4.40 0.00239 1.76 0.02001 4.80 0.24 0.000765 1.83 15.4 0.3 15.5 0.3<br />

HK117A_4_2 L-Y,C 146 8900 132500 15.0 0.1352 5.47 0.0026 1.69 0.0476 6.51 0.63 0.000774 1.42 16.7 0.3 15.6 0.2<br />

HK117A_4_3 L-Y,C 109 10680 101900 9.4 0.0568 1.94 0.00222 1.31 0.01716 2.16 0.26 0.000757 1.31 14.3 0.2 15.3 0.2<br />

HK117A_4_4 L-Y,C 154 9600 145100 15.2 0.0527 2.09 0.00231 1.34 0.01679 2.62 0.33 0.00076 1.30 14.9 0.2 15.4 0.2<br />

HK117A_4_5 L-Y,C 147 9350 139500 15.0 0.0527 2.66 0.00231 1.21 0.01667 2.76 0.17 0.000746 1.29 14.9 0.2 15.1 0.2<br />

HK117A_4_6 H-Y,R 90 11740 84500 7.2 0.076 4.34 0.00227 1.54 0.0235 4.68 0.49 0.000759 1.45 14.6 0.2 15.3 0.2<br />

HK117A_4_7 H-Y,R 119 16630 115100 6.9 0.05816 1.44 0.00225 1.69 0.01792 1.62 0.60 0.000749 1.47 14.5 0.2 15.1 0.2<br />

HK117A_4_8 L-Y,C 119 8570 114900 13.4 0.058 2.41 0.00234 1.33 0.01883 2.66 0.31 0.000751 1.33 15.0 0.2 15.2 0.2<br />

HK117A_4_9 L-Y,C 115 8730 110700 12.8 0.063 1.90 0.00235 1.36 0.02031 2.36 0.56 0.000753 1.46 15.2 0.2 15.2 0.2<br />

HK117A_4_10 L-Y,C 116 8680 110900 12.9 0.0743 2.42 0.00239 1.46 0.02476 3.31 0.65 0.000758 1.45 15.4 0.2 15.3 0.2<br />

HK117A_4_11 L-Y,C 63 4356 57900 13.3 0.1852 3.67 0.00284 1.69 0.0724 4.97 0.75 0.000785 1.53 18.3 0.3 15.9 0.2<br />

HK117A_4_12 L-Y,C 113 14490 108300 7.5 0.0849 5.42 0.00245 1.39 0.0289 6.23 0.56 0.00076 1.45 15.8 0.2 15.4 0.2<br />

HK117A_4_13 L-Y,C 139 10750 131700 12.3 0.0544 2.02 0.00244 1.31 0.01849 2.16 0.41 0.000749 1.47 15.7 0.2 15.1 0.2<br />

HK117A_4_14 L-Y,C 130 13100 125400 9.6 0.06393 1.42 0.00241 1.33 0.02144 2.10 0.65 0.000742 1.35 15.5 0.2 15.0 0.2<br />

HK117A_4_15 L-Y,C 133 12510 127200 10.2 0.0612 1.80 0.00242 1.36 0.0205 2.34 0.57 0.000746 1.34 15.6 0.2 15.1 0.2<br />

HK117A_6_1 H-Y,R 69 20490 79800 4.0 0.05789 1.59 0.00194 1.39 0.01563 2.11 0.53 0.000611 1.46 12.5 0.2 12.4 0.2<br />

HK117A_6_2 H-Y,R 76 20060 86500 4.4 0.0599 2.00 0.00199 1.31 0.01649 2.49 0.35 0.000627 1.36 12.8 0.2 12.7 0.2<br />

HK117A_6_3 H-Y,R 82 19660 87500 4.5 0.0613 1.63 0.00209 1.39 0.01769 2.15 0.55 0.000662 1.51 13.5 0.2 13.4 0.2<br />

67


Analysis<br />

Position<br />

Pb U Th<br />

Th/U<br />

207 Pb/ 206 2σ<br />

Pb<br />

206 Pb/ 238 2σ<br />

U<br />

207 Pb/ 235 U b 2σ<br />

Rho c 208 Pb/ 232 2σ<br />

206 Pb/ 238 U 2σ<br />

208 Pb/ 232 Th 2σ<br />

Th<br />

(ppm) a (ppm) a (ppm) a (%) (%) (%) (%) (Ma) d abs (Ma) e abs<br />

HK117A_6_4 H-Y,R 108 14880 101900 7.0 0.0627 2.07 0.00242 1.32 0.02116 2.41 0.41 0.000743 1.48 15.6 0.2 15.0 0.2<br />

HK117A_6_5 H-Y,R 109 15870 103100 6.6 0.0661 2.27 0.00241 1.29 0.02233 2.96 0.35 0.000744 1.32 15.5 0.2 15.0 0.2<br />

HK117A_6_6 L-Y,C 75 19060 82400 4.4 0.1051 2.47 0.00204 1.52 0.02943 2.85 0.48 0.000641 1.56 13.2 0.2 13.0 0.2<br />

HK117A_6_7 H-Y,R 71 16850 75600 4.6 0.1395 2.44 0.00209 1.39 0.0398 2.51 0.59 0.00066 1.41 13.5 0.2 13.3 0.2<br />

HK117A_6_8 H-Y,R 49 10080 44500 4.5 0.2441 1.31 0.0024 1.63 0.0814 2.09 0.82 0.000795 1.76 15.4 0.3 16.1 0.3<br />

HK117A_6_9 L-Y,C 132 21580 129200 6.1 0.0591 1.69 0.0023 1.43 0.01892 2.27 0.44 0.000723 1.52 14.8 0.2 14.6 0.2<br />

HK117A_6_10 L-Y,C 105 13490 98500 7.4 0.0653 1.99 0.00238 1.60 0.02134 2.39 0.52 0.000747 1.47 15.3 0.2 15.1 0.2<br />

HK117A_7_1 H-Y,R 78 6800 22000 3.3 0.0453 7.95 0.0072 22.22 0.045 24.44 0.98 0.00225 ### 46.0 ## 45.0 ##<br />

HK117A_7_2 H-Y,R 34 3150 20600 6.5 0.468 3.21 0.00373 4.29 0.243 4.94 0.77 0.001233 5.68 24.0 1.1 24.9 1.4<br />

HK117A_7_3 H-Y,R 155 12910 154900 11.9 0.0658 1.82 0.00226 1.37 0.02055 2.29 0.58 0.000709 1.41 14.5 0.2 14.3 0.2<br />

HK117A_7_4 H-Y,R 225 15220 220200 14.3 0.0578 1.45 0.00238 1.34 0.01906 1.99 0.43 0.000732 1.22 15.4 0.2 14.8 0.2<br />

HK117A_7_5 H-Y,R 140 9000 140600 15.4 0.069 2.46 0.00231 1.73 0.02197 2.64 0.42 0.000711 1.69 14.9 0.3 14.4 0.3<br />

HK117A_7_6 L-Y,C 105 12870 124100 9.5 0.0813 2.46 0.0019 1.84 0.02151 2.65 0.52 0.000617 1.94 12.3 0.2 12.5 0.3<br />

HK117A_7_7 L-Y,C 114 9370 98800 10.5 0.262 1.49 0.00293 2.12 0.1059 2.64 0.82 0.00083 2.05 18.9 0.4 16.8 0.3<br />

HK117A_7_8 L-Y,C 179 12040 176400 14.6 0.0596 1.85 0.00237 1.69 0.01928 2.54 0.68 0.000725 1.52 15.3 0.3 14.7 0.2<br />

HK117A_7_9 L-Y,C 164 11300 155500 13.7 0.2 3.50 0.00271 1.96 0.0752 4.92 0.86 0.000761 1.58 17.5 0.3 15.4 0.3<br />

HK117A_7_10 L-Y,C 121 9780 126400 12.9 0.0668 2.25 0.00213 1.60 0.01952 2.72 0.39 0.000683 1.61 13.7 0.2 13.8 0.2<br />

HK117A_9_1 H-Y,R 75 2790 16100 5.7 0.6544 0.76 0.01054 2.28 0.951 2.52 0.97 0.00351 3.70 67.6 1.5 70.8 2.5<br />

HK117A_9_2 H-Y,R 37 3610 23800 6.6 0.3195 2.57 0.00324 2.99 0.1426 4.56 0.82 0.001101 3.00 20.9 0.6 22.2 0.7<br />

HK117A_9_3 H-Y,R 44 4590 31500 6.9 0.298 1.58 0.00302 2.81 0.1237 2.75 0.82 0.001015 3.05 19.5 0.5 20.5 0.6<br />

HK117A_9_4 H-Y,R 20 901 4290 4.8 0.5802 1.55 0.00847 4.84 0.677 5.61 0.95 0.0034 6.47 54.4 2.7 68.6 4.4<br />

HK117A_9_5 H-Y,R 112 9440 106900 11.4 0.06275 1.51 0.00235 1.79 0.02053 2.19 0.67 0.000736 1.77 15.1 0.3 14.9 0.3<br />

HK117A_9_6 H-Y,R 122 10470 118400 11.4 0.0608 1.97 0.00237 1.31 0.01989 2.51 0.50 0.000731 1.35 15.3 0.2 14.8 0.2<br />

HK117A_9_7 L-Y,C 122 10360 113100 11.0 0.0604 1.82 0.00243 1.69 0.02015 2.73 0.65 0.000758 1.58 15.7 0.3 15.3 0.2<br />

HK117A_9_8 L-Y,C 114 9950 105400 10.6 0.0631 2.06 0.00244 1.56 0.02123 2.64 0.62 0.000762 1.57 15.7 0.3 15.4 0.2<br />

HK117A_9_9 L-Y,C 116 10470 108800 10.5 0.0629 1.75 0.00242 1.45 0.02102 2.09 0.44 0.000748 1.60 15.6 0.2 15.1 0.2<br />

HK117A_9_10 L-Y,C 117 10530 108200 10.3 0.0624 1.60 0.00246 1.63 0.02154 2.46 0.65 0.000754 1.72 15.8 0.3 15.2 0.3<br />

HK117A_9_11 L-Y,C 104 10470 97400 9.3 0.0633 1.90 0.00242 1.69 0.02089 2.35 0.58 0.000742 1.62 15.6 0.3 15.0 0.3<br />

HK117A_9_12 L-Y,C 117 11150 108900 9.8 0.0651 1.84 0.00251 1.71 0.02243 2.32 0.61 0.000763 1.57 16.2 0.3 15.4 0.2<br />

_______________________________________________________________________________________________________________________________________________<br />

Analysis column sample numbers refer to: SAMPLE_GRAIN_SPOT. H-Y,M-Y,L-Y refer to high, medium, and low yttrium, respectively. C and R indicate core<br />

and rim, respectively. Analyses stroked-through were not included in final age interpretation. (See next page for additional table details).<br />

68


a concentration data are normalized to <strong>the</strong> primary reference material and are accurate to ~10%<br />

b 207Pb/235U calculated assuming a natural 235U/238U ratio <strong>of</strong> 137.88<br />

c Rho value is calculated following <strong>the</strong> method outlined in Paton et al. (2010)<br />

d Age calculations based on <strong>the</strong> decay constants <strong>of</strong> Jaffey et al. (1971)<br />

e Age calculations based on <strong>the</strong> decay constant <strong>of</strong> Amelin & Zaytsev (2002)<br />

____________________________________________________________________________________<br />

metamorphic event with a maximum age population at ~19 Ma. These ages are roughly<br />

characterized by a ~ 21 -18 Ma peak <strong>of</strong> Y-poor cores and a ~18 -16 Ma peak <strong>of</strong> Y-rich rims.<br />

For sample HK109, a total <strong>of</strong> 54 meaningful analyses were obtained on seven monazite<br />

grains (Fig. 2.17, Table 2.1). These data span between ~19 -13 Ma, and outline two distinct age<br />

populations associated with Y / Th compositional zoning. A peak between ~16 -19 Ma consists <strong>of</strong><br />

Y-poor cores, and a second peak <strong>of</strong> similar intensity between ~13 -16 Ma is characterized by Y-<br />

rich rims. The only exception to this are three outlying ages obtained from monazite grain<br />

HK109_5, at ~37 Ma, 124 Ma and 390 Ma (Table 2.1). These data are not displayed in Figure<br />

2.17.<br />

For sample HK117, a total <strong>of</strong> 54 meaningful analyses were obtained on seven monazite<br />

grains (Fig. 2.17, Table 2.1). These data yield a main age population at ~15 Ma, and a minor<br />

distribution <strong>of</strong> ages between ~14 -12 Ma. The main age population consists <strong>of</strong> a high frequency<br />

<strong>of</strong> 16 -15 Ma Y-poor cores and a lower frequency <strong>of</strong> 16 -14 Ma Y-rich rims. Ages between 14 -12<br />

Ma are derived from a mix <strong>of</strong> Y-rich rims and Y-poor cores.<br />

2.6.1.3 Summary <strong>of</strong> U-Th-Pb geochronology<br />

Our data reveal a span <strong>of</strong> 208 Pb/ 232 Th ages from ~26 to 12 Ma, with peaks at 19 Ma and<br />

15 Ma (Fig. 2.17). All samples yield one dominant age range. Sample KH18a shows an additional<br />

minor set <strong>of</strong> older ages, and sample HK117a reveals a low frequency <strong>of</strong> younger ages. All ages<br />

69


are derived from intensely deformed, elongate and/or rounded monazite grains that are typically<br />

aligned parallel to <strong>the</strong> dominant ESE-WNW trending lineation (Fig. 2.16).<br />

For samples KH18a and HK109, distinct older and younger age groups correspond to Y-<br />

poor cores and Y-rich rims, respectively (Figs. 2.16, 2.17). As garnet has a higher partition<br />

coefficient for Y than does monazite, it essentially acts as a Y sink during crystallization (Foster<br />

et al., 2002). Therefore, older age groups associated with Y-poor monazite cores are interpreted to<br />

correspond to a period <strong>of</strong> simultaneous garnet growth and monazite crystallization. Younger age<br />

groups associated with Y-rich rims are interpreted to reflect monazite crystallization concurrent<br />

with garnet breakdown.<br />

Although <strong>the</strong> correlation between Y zonation and age domains is generally consistent,<br />

older and younger age groups do not always correspond with analyses <strong>of</strong> Y-poor cores and Y-rich<br />

rims, respectively (Fig. 2.17, Table 2.1). In certain cases, <strong>the</strong> analysis <strong>of</strong> one Y domain may yield<br />

ages expected from <strong>the</strong> o<strong>the</strong>r. This crossover <strong>of</strong> data is due to mixing <strong>of</strong> Y zones and associated<br />

age domains during analysis. Laser ablation generates craters with a diameter <strong>of</strong> ~7 μm and a<br />

depth <strong>of</strong> ~5-6 μm. Thus, when analyzing elemental zones within in situ grains, it would be<br />

expected that analyses occasionally include underlying zones <strong>of</strong> different composition and age.<br />

This would lead to a mixing <strong>of</strong> age domains and a continuous spread <strong>of</strong> data between age peaks.<br />

Sample HK109 yielded significantly older Eocene, Mesozoic and Paleozoic ages (Table<br />

2.1). The Eocene age <strong>of</strong> ~37 Ma is interpreted to reflect Eo-Himalayan metamorphism associated<br />

with Eocene-Oligocene crustal thickening (Godin et al., 1999; 2001). The Paleozoic age <strong>of</strong> ~390<br />

Ma is interpreted to represent a pre-Himalayan metamorphic event, or a mixture between<br />

Paleozoic and younger age domains (e.g. Godin et al., 2001). The Mesozoic age <strong>of</strong> ~124 Ma is<br />

70


derived from a Y-rich monazite rim, and is interpreted as a mixture between Paleozoic and<br />

younger age domains. These older ages suggest that a portion <strong>of</strong> this grain was inherited.<br />

2.6.2 40 Ar/ 39 Ar <strong>the</strong>rmochronology<br />

Samples rich in muscovite were collected from traverses across and along <strong>the</strong> upper<br />

Karnali valley. Mineral separates were obtained through standard crushing and sieving techniques<br />

and muscovites were hand picked under a binocular macroscope. Only grains with good clarity,<br />

greater than or equal to 500 μm in diameter, and void <strong>of</strong> inclusions and intergrowth were selected<br />

for <strong>the</strong>rmochronology. The 40 Ar/ 39 Ar <strong>the</strong>rmochronology was performed at <strong>Queen's</strong> <strong>University</strong><br />

Argon Geochronology laboratory on muscovite grains from 19 samples located throughout<br />

domain III, IV, V and VI.<br />

Sample HK 125 is taken from an ultra-high-strain pelitic schist <strong>of</strong> domain III (Fig. 2.3).<br />

Within this sample, intensely stretched and deformed muscovite represent ~10 vol%, and define<br />

<strong>the</strong> mineral elongation lineation and foliation. Muscovite porphyroclasts show abundant<br />

deformation structures, including mica fish, micro-boudinaging, and undulatory extinction (Fig.<br />

2.5e). No grains within this sample show any indication <strong>of</strong> recrystallization.<br />

Sample KH13 and samples KH18, KH35, HK131c, HK138 are taken from a quartzite<br />

and from pelitic schist <strong>of</strong> domain IV, respectively (Fig. 2.3). Within <strong>the</strong>se samples well-formed<br />

muscovite grains represent ~25 vol%, define a pervasive mineral elongation lineation and<br />

foliation, and form distinct C/S/C’ fabrics. Shear bands <strong>of</strong>fset coarse-grained muscovite, while<br />

fine-grained muscovites grow along <strong>the</strong>se same shear planes (Fig. 2.6d). Muscovite located<br />

within <strong>the</strong>se shear bands may be somewhat recrystallized; however by only selecting grains with<br />

diameters ≥ 500 μm we ensured that <strong>the</strong>rmochronology was not performed on any recrystallized<br />

grains.<br />

71


Samples KH10, HK102a, HK105, HK115, HK142 are taken from schist and gneiss <strong>of</strong><br />

domain V (Fig. 2.3). Within <strong>the</strong>se samples, well-formed muscovite grains and muscovite<br />

aggregates represent ~15 vol% and define a pervasive to moderate mineral elongation lineation<br />

and foliation. Locally, muscovite grains form C/S/C’ fabrics and mantle rotated sigma-style<br />

quartz and feldspar porphyroclasts (Fig. 2.7b). Shear bands <strong>of</strong>fset coarse-grained muscovite,<br />

while minor amounts <strong>of</strong> fine-grained and potentially recrystallized muscovite grow along <strong>the</strong>se<br />

same shear planes. Again, only muscovite grains ≥ 500 μm in diameter were selected to avoid<br />

analyzing recrystallized grains.<br />

Samples HK117a, HK118a and HK119a are taken from migmatitic gneiss <strong>of</strong> domain VI<br />

(Fig. 2.3). Within <strong>the</strong>se samples, muscovite represents ~5 vol% and is characterized by ragged<br />

and serrate edges. Muscovite grains are both randomly oriented and aligned roughly parallel to<br />

<strong>the</strong> foliation. Aggregates <strong>of</strong> muscovite are moderately aligned parallel to <strong>the</strong> mineral lineation<br />

when present. These textural relationships suggest that muscovite grains within <strong>the</strong>se three<br />

samples are not in equilibrium with <strong>the</strong>ir host rocks.<br />

2.6.2.1 Methodology<br />

Muscovite separates and standards (used to monitor flux) were packaged in Al-foil,<br />

loaded into an 8.5 cm long and 2.0 cm diameter irradiation capsule and irradiated with fast<br />

neutrons in position 5C for a duration <strong>of</strong> 5 hours at 3 MWh at <strong>the</strong> McMaster Nuclear Reactor<br />

(Hamilton, Ontario). Packets <strong>of</strong> flux monitors were located at ~ 1 cm intervals along <strong>the</strong><br />

irradiation container and J-values for individual samples were determined by second-order<br />

polynomial interpolation between replicate analyses <strong>of</strong> splits for each monitor position in <strong>the</strong><br />

capsule. Typically, J-values vary by < 10% over <strong>the</strong> length <strong>of</strong> <strong>the</strong> capsule. No attempt is made to<br />

monitor horizontal flux gradients as <strong>the</strong>se are considered to be minor in <strong>the</strong> core <strong>of</strong> <strong>the</strong> reactor.<br />

72


Analyses were performed in two separate sessions. For each analytical session, a CO 2<br />

laser system was used for total fusion <strong>of</strong> monitors and step-heating <strong>of</strong> samples. Between 4 and 15<br />

irradiated muscovite grains per sample were loaded into pits <strong>of</strong> a copper sample-holder, which<br />

was <strong>the</strong>n mounted beneath a ZnS view-port <strong>of</strong> a small, stainless-steel chamber connected to an<br />

ultra-high vacuum purification system. For step-heating in <strong>the</strong> first analytical session, <strong>the</strong> laser<br />

beam <strong>of</strong> a 30W New Wave Research MIR 10-30 CO 2 laser was defocused to 2 mm to cover <strong>the</strong><br />

entire sample. For step-heating in <strong>the</strong> second analytical session, <strong>the</strong> laser beam <strong>of</strong> an identical<br />

laser was equipped with a faceted lens to allow for uniform heating <strong>of</strong> <strong>the</strong> entire sample. Heating<br />

periods were ~3 -4 minutes at increasing percent power settings. The evolved gas, after<br />

purification using an SAES C50 getter (~ 5 minutes), was admitted to an on-line, MAP 216 mass<br />

spectrometer, with a Bäur Signer source and an electron multiplier (set to a gain <strong>of</strong> 100 over <strong>the</strong><br />

Faraday). Blanks, measured routinely, were subtracted from <strong>the</strong> subsequent sample gas-fractions.<br />

The extraction blanks are typically < 10 x 10 -13 , < 0.5 x 10 -13 , < 0.5 x 10 -13 , and < 0.5 x 10 -13 cm -3<br />

STP for masses 40, 39, 37, and 36, respectively.<br />

Measured argon-isotope peak heights were extrapolated to zero-time, normalized to <strong>the</strong><br />

40 Ar/ 36 Ar atmospheric ratio (295.5) using measured values <strong>of</strong> atmospheric argon, and corrected<br />

for neutron-induced 40 Ar from potassium, 39 Ar and 36 Ar from calcium, and 36 Ar from chlorine<br />

(Roddick, 1983). Dates and errors were calculated using ISOPLOT 3.75 (Ludwig, 2012), and <strong>the</strong><br />

constants <strong>of</strong> Steiger and Jäger (1977). Errors shown in <strong>the</strong> tables and on <strong>the</strong> age plateau diagrams<br />

represent <strong>the</strong> analytical precision at 2σ, assuming that <strong>the</strong> errors in <strong>the</strong> ages <strong>of</strong> <strong>the</strong> flux monitors<br />

are zero. This is suitable for comparing within-spectrum variation and determining which steps<br />

form a plateau (e.g., McDougall and Harrison, 1988, p, 89). A conservative estimate <strong>of</strong> this error<br />

in <strong>the</strong> J-value is 0.5% and can be added for inter-sample comparison. The dates and J-values for<br />

73


<strong>the</strong> intralaboratory standard (MAC-83 biotite at 24.36 Ma; Sandeman et al., 1999) are referenced<br />

to FCT sanidine at 28.02 Ma (Renne et al., 1998) for <strong>the</strong> first analytical session, and TCR<br />

sanidine at 28.34 Ma for <strong>the</strong> second analytical session. Sample KH35 was used as an internal<br />

standard in both analytical sessions, and yielded near identical ages.<br />

Well-defined, and moderately-defined age plateaus are defined as a minimum <strong>of</strong> three<br />

consecutive steps with ages <strong>of</strong> overlapping error, releasing 90% or more and 40% or more, <strong>of</strong> <strong>the</strong><br />

total 39 Ar, respectively.<br />

2.6.2.2 Results<br />

Of <strong>the</strong> 19 samples selected for 40 Ar/ 39 Ar age determination, 10 yielded well-defined age<br />

plateaus, and 4 yielded moderately-defined age plateaus (Fig. 2.18, Table 2.2, see Appendix B for<br />

a complete data set, including inverse isochron diagrams). The remaining samples did not yield<br />

acceptable age plateaus, resulting in a mixture <strong>of</strong> chaotic plateaus, plateaus with insufficient total<br />

percent argon released, and continuously rising and or descending step-heating pr<strong>of</strong>iles<br />

(Appendix D).<br />

Sample HK125 <strong>of</strong> domain III yielded a well-defined plateau age <strong>of</strong> 13.01 ± 0.30 Ma<br />

corresponding to 92.8% <strong>of</strong> released 39 Ar (Fig. 2.18, Table 2.2). Analyzed muscovite grains define<br />

a shallow west plunging lineation and show extensive post-crystallization deformation.<br />

Samples KH13 and KH18 <strong>of</strong> domain IV yielded moderately-defined plateau ages (Fig. 2.18,<br />

Table 2.2). Analyzed muscovite grains define a shallow ESE-WNW lineation and are <strong>of</strong>fset by<br />

shear bands. Sample KH13 yielded an age <strong>of</strong> 14.25 ± 0.23 Ma corresponding to 42.3% <strong>of</strong><br />

released 39 Ar. Sample KH18 was analyzed in two separate analytical sessions and yielded ages <strong>of</strong><br />

13.52 ± 0.16 Ma and 12.79 ± 0.53 Ma, corresponding to 43.0% and 69.5% <strong>of</strong> released 39 Ar,<br />

74


Figure 2.18 Muscovite 40 Ar/ 39 Ar step-heating plateaus grouped toge<strong>the</strong>r within <strong>the</strong>ir respective<br />

lithotectonic domains. Plateau ages are calculated from grey steps. All errors are reported at 2σ. Plots were<br />

constructed with Isoplot 3.75 (Ludwig, 2012). Additional details can be found in Table 2.2.<br />

75


Figure 2.18 – continued.<br />

76


Table 2-2 40 Ar/ 39 Ar <strong>the</strong>rmochronological data<br />

Sample Domain Rock Type Condition <strong>of</strong> Muscovite<br />

Plateau Age<br />

(Ma)<br />

Error<br />

(Ma, 1s)<br />

Percent 39 Ar<br />

defining<br />

plateau<br />

MSDW<br />

Probability<br />

HK125 III<br />

Ultra-high-strain pelitic<br />

schist<br />

Deformed: faulted, fractured,<br />

undulatory extinction 13.01 0.30 92.8 0.33 0.96<br />

KH13 IV High-strain quartzite Well-formed, <strong>of</strong>fset by shear bands 14.25 0.23 42.3 1.06 0.38<br />

KH18 (1)* IV High-strain pelitic schist Well-formed, <strong>of</strong>fset by shear bands 13.52 0.16 43.0 0.71 0.62<br />

KH18 (2)* IV High-strain pelitic schist Well-formed, <strong>of</strong>fset by shear bands 12.79 0.53 69.5 0.65 0.62<br />

KH35 (1)* IV High-strain pelitic schist<br />

Well-formed, <strong>of</strong>fset by shear bands,<br />

minor undulatory extinction 13.03 0.17 100.0 0.40 0.96<br />

KH35 (2)* IV High-strain pelitic schist<br />

Well-formed, <strong>of</strong>fset by shear bands,<br />

minor undulatory extinction 13.00 0.45 100.0 0.41 0.90<br />

HK131c IV High-strain pelitic schist Well-formed, <strong>of</strong>fset by shear bands 11.61 0.25 100.0 0.65 0.73<br />

HK138 IV High-strain pelitic schist<br />

Well-formed, <strong>of</strong>fset by shear bands,<br />

minor undulatory extinction 12.29 0.30 100.0 0.27 0.98<br />

HK102a V High-strain gneiss Well-formed 9.27 0.19 54.2 2.10 0.12<br />

HK115 V High-strain gneiss<br />

Well-formed, occasionally <strong>of</strong>fset by<br />

shear bands 15.74 0.28 60.7 0.73 0.48<br />

KH10 V High-strain gneiss<br />

Well-formed, mantles fsp and qtz<br />

porphyroclasts 12.08 0.09 99.4 1.30 0.22<br />

HK105b V High-strain pelitic schist Well-formed 9.90 0.28 100.0 0.20 0.99<br />

HK142 V High-strain gneiss Well-formed 10.16 0.33 100.0 0.19 0.99<br />

HK117a VI Migmatitic gneiss<br />

Ragged and serrate grain boundaries,<br />

possible replacement textures 11.57 0.25 100.0 1.60 0.14<br />

HK118a VI Migmatitic gneiss<br />

HK119a VI Migmatitic gneiss<br />

Small, ragged and serrate grain<br />

boundaries, possible replacement<br />

textures 10.81 0.16 100.0 0.38 0.93<br />

Small, ragged and serrate grain<br />

boundaries, possible replacement<br />

textures 11.39 0.25 100.0 0.46 0.88<br />

*Multiple analyses <strong>of</strong> individual samples performed in separate analytical sessions have different J-value and are <strong>the</strong>refore reported as independent analyses (i.e. 1, 2)<br />

Data in italics indicate moderately-defined plateaus corresponding to ~ 40-70% <strong>of</strong> total 39 Ar released<br />

77


espectively. Note that <strong>the</strong> plateaus <strong>of</strong> KH18 cannot be combined into a single plateau age<br />

because <strong>the</strong>y are associated with J-values <strong>of</strong> <strong>the</strong>ir respective analytical session.<br />

Samples KH35, HK131c, and HK138 <strong>of</strong> domain IV yielded well-defined plateau ages<br />

(Fig. 2.18, Table 2.2). Analyzed muscovite grains define a shallow ESE-WNW lineation and are<br />

<strong>of</strong>fset by shear bands. Sample KH35 was analyzed in both analytical sessions as an internal<br />

laboratory<br />

standard, and yielded ages <strong>of</strong> 13.03 ± 0.17 Ma and 13.00 ± 0.45 Ma, both corresponding to 100%<br />

<strong>of</strong> released 39 Ar. Note that plateaus <strong>of</strong> KH35 cannot be combined into a single plateau age<br />

because <strong>the</strong>y are associated with different J-values <strong>of</strong> <strong>the</strong>ir respective analytical session. Samples<br />

HK131c and HK138 yielded ages <strong>of</strong> 11.61 ± 0.25 Ma and 12.29 ± 0.30 Ma, respectively. Both<br />

ages correspond to 100% <strong>of</strong> released 39 Ar.<br />

Samples HK102a and HK115 <strong>of</strong> domain V yield moderately-defined plateau ages (Fig.<br />

2.18, Table 2.2). Analyzed muscovite grains define a shallow ESE-WNW lineation and are <strong>of</strong>fset<br />

by occasional shear bands. Sample HK102a and HK115 yield ages <strong>of</strong> 9.27 ± 0.19 Ma and 15.74 ±<br />

0.28 Ma, corresponding to 54.2% and 60.7% <strong>of</strong> released 39 Ar, respectively.<br />

Samples KH10, HK105, and HK142 <strong>of</strong> domain V yield well-defined plateau ages (Fig.<br />

2.18, Table 2.2). Analyzed muscovite and muscovite aggregates are present in non-laterally<br />

continuous layers and define a shallow ESE-WNW lineation. From WNW to ESE, samples<br />

KH10, HK142 and HK105 yield ages <strong>of</strong> 12.08 ± 0.09 Ma, 10.16 ± 0.33 Ma and 9.90 ± 0.28 Ma,<br />

corresponding to 99.4%, 100% and 100% <strong>of</strong> released 39 Ar, respectively.<br />

Samples HK117a, HK118a and HK119a <strong>of</strong> domain VI yield well-defined plateau ages<br />

(Fig. 2.18, Table 2.2). Analyzed muscovite grains are not pristine and exhibit ragged and serrate<br />

boundaries, possibly indicative <strong>of</strong> disequilibrium conditions. From south to north, samples<br />

78


HK117a, HK119a and HK118a yield ages <strong>of</strong> 11.57 ± 0.25 Ma, 11.39 ± 0.25 Ma and 10.81 ± 0.16<br />

Ma, respectively. All ages correspond to 100% <strong>of</strong> released 39 Ar.<br />

2.6.2.3 Summary <strong>of</strong> 40 Ar/ 39 Ar <strong>the</strong>rmochronology<br />

Our <strong>the</strong>rmochronology results provide ages at which muscovite passed through <strong>the</strong> argon<br />

closure temperature, commonly reported to be 350°C (Hodges, 1991). However, recent studies<br />

have shown argon diffusion in muscovite to cease at much higher temperatures, corresponding to<br />

hotter closure temperatures generally in <strong>the</strong> range <strong>of</strong> ~425°C (Harrison et al., 2009), and up to<br />

500-600°C (see review in Villa, 2004). The closure temperature <strong>of</strong> muscovite in <strong>the</strong> upper Karnali<br />

valley is estimated to be ~450°C (for pressures <strong>of</strong> ~5 kbar, muscovite grain diameters <strong>of</strong> ~500<br />

μm, and a cooling rate <strong>of</strong> ~60°C/Ma, based <strong>of</strong> Fig. 7 in Harrison et al., 2009). Based upon <strong>the</strong><br />

lack <strong>of</strong> recrystallization textures observed in <strong>the</strong> analyzed muscovite <strong>of</strong> <strong>the</strong> upper Karnali valley,<br />

we interpret plateau ages obtained in this study to represent muscovite cooling ages.<br />

Based exclusively upon well-defined plateau ages, <strong>the</strong> GHS <strong>of</strong> <strong>the</strong> upper Karnali valley<br />

cooled through ~450°C between ~13 -10 Ma (Fig. 2.18, Table 2.2). If moderately-defined age<br />

plateaus are also considered, <strong>the</strong>n <strong>the</strong> cooling age <strong>of</strong> <strong>the</strong> GHS in this region may span from ~15 -9<br />

Ma (Fig. 2.18, Table 2.2).<br />

Plots <strong>of</strong> cooling ages along and across <strong>the</strong> GHS-TSS interface reveal younging trends to<br />

<strong>the</strong> nor<strong>the</strong>ast and ESE (Fig. 2.19). If considering only well-defined age plateaus, cooling ages<br />

young from ~12.5 Ma along <strong>the</strong> GHS-TSS interface, to ~10.5 Ma at 10 km structurally below <strong>the</strong><br />

interface towards <strong>the</strong> nor<strong>the</strong>ast (Fig. 2.19a). A linear regression <strong>of</strong> this trend yield a R 2 value <strong>of</strong><br />

~0.36. This R 2 value may appear low, however, considering <strong>the</strong> low frequency <strong>of</strong> data points, and<br />

<strong>the</strong> distribution <strong>of</strong> data points along strike where o<strong>the</strong>r types <strong>of</strong> variation would also be expected,<br />

79


Figure 2.19 Plots <strong>of</strong> muscovite cooling ages against (a) increasing structural distance below <strong>the</strong> GHS-TSS<br />

interface (measured perpendicular to average foliation), and (b) distance parallel to <strong>the</strong> GHS-TSS<br />

interface. The line <strong>of</strong> origin for plot (b) is perpendicular to <strong>the</strong> GHS-TSS interface and intersects <strong>the</strong><br />

village <strong>of</strong> Tumkot (e.g. plot portrays evolution <strong>of</strong> muscovite cooling ages towards <strong>the</strong> sou<strong>the</strong>ast). Village<br />

names are provided as reference. See text for description <strong>of</strong> well-defined and moderately-plateaus.<br />

80


this correlation may be significant. A best-fit line through well-defined and moderately-defined<br />

age plateaus exhibits a similar trend with a lower R 2 value.<br />

Cooling ages also show a younging trend parallel to <strong>the</strong> GHS-TSS interface (Fig. 2.19b).<br />

Cooling ages <strong>of</strong> all samples progress from ~14 -13 Ma along <strong>the</strong> WNW edge <strong>of</strong> <strong>the</strong> field area, to<br />

~9 Ma within 25 km to <strong>the</strong> ESE. Linear regressions <strong>of</strong> well-defined plateaus, and <strong>of</strong> all plateaus,<br />

yield R 2 values <strong>of</strong> ~0.38 and ~0.44, respectively. Although <strong>the</strong>se R 2 values are low, considering<br />

variations that would arise from <strong>the</strong> low frequency <strong>of</strong> data points, and <strong>the</strong> distribution <strong>of</strong><br />

data points from various structural levels along strike, this correlation may be significant.<br />

Four <strong>of</strong> <strong>the</strong> samples analyzed yield moderately-defined age plateaus. Of <strong>the</strong>se samples,<br />

KH13 yields an upward-stepping pr<strong>of</strong>ile, KH18 and HK115 yield downward-stepping pr<strong>of</strong>iles,<br />

and HK102a yields a subtle U-shaped pr<strong>of</strong>ile. The upward-stepping pr<strong>of</strong>ile is interpreted to be <strong>the</strong><br />

result <strong>of</strong> overloading <strong>of</strong> <strong>the</strong> sample pit with an excess number <strong>of</strong> muscovite grains, leading to<br />

uneven heating <strong>of</strong> all grains during analysis. As a consequence, 39 Ar that should have been<br />

released during initial steps was released in <strong>the</strong> final steps, generating an upward stepping pr<strong>of</strong>ile.<br />

Downward-stepping and U-shaped pr<strong>of</strong>iles are attributed to <strong>the</strong> disruption <strong>of</strong> 39 Ar retention since<br />

initial cooling. This may be <strong>the</strong> result <strong>of</strong> continued deformation below <strong>the</strong> argon closure<br />

temperature in <strong>the</strong> upper Karnali valley.<br />

2.7 Discussion<br />

Here, data from this study are discussed with regard to south-directed extrusion <strong>of</strong> <strong>the</strong><br />

GHS, orogen-parallel extension and <strong>the</strong> transition between <strong>the</strong>se two processes. A summary <strong>of</strong><br />

results is provided in Figure 2.20. These results are <strong>the</strong>n combined with prior studies to assess <strong>the</strong><br />

applicability <strong>of</strong> <strong>the</strong> proposed transition on an orogen-wide scale, and to evaluate models<br />

attempting to explain <strong>the</strong> processes responsible for orogen-parallel extension.<br />

81


Figure 2.20 Summary <strong>of</strong> results from quartz CPO, vorticity, muscovite <strong>the</strong>rmochronology and monazite geochronology analyses. D. I, II, III, IV, V and VI<br />

refer to lithotectonic domains. Map area is identical to Figure 2.2.<br />

82


2.7.1 South-directed extrusion <strong>of</strong> <strong>the</strong> GHS<br />

2.7.1.1 Initial south-directed extrusion<br />

Geologic mapping <strong>of</strong> <strong>the</strong> upper Karnali valley reveals low-grade sedimentary rocks <strong>of</strong> <strong>the</strong><br />

TSS juxtaposed along a south-dipping contact over strongly deformed and metamorphosed rocks<br />

<strong>of</strong> <strong>the</strong> GHS (Figs. 2.3, 2.20). Mineral elongation lineations throughout <strong>the</strong> high-strain zones<br />

mantling <strong>the</strong> GHS-TSS interface plunge shallowly to <strong>the</strong> ESE and WNW. The kinematics <strong>of</strong><br />

<strong>the</strong>se zones show dominant strike-slip and minor dip-slip kinematics (Fig. 2.3). The high-strain<br />

zones are focused along <strong>the</strong> GHS-TSS interface, but also run obliquely across <strong>the</strong> GHS<br />

throughout <strong>the</strong> eastern segment <strong>of</strong> <strong>the</strong> field area (Fig. 2.3). These fabrics are inconsistent with <strong>the</strong><br />

STDS (Burchfiel and Royden, 1985; Burchfiel et al., 1992; Hodges, 2000), and confirm <strong>the</strong><br />

presence <strong>of</strong> <strong>the</strong> orogen-parallel, strike-slip-dominated GMH fault system described by Murphy<br />

and Copeland (2005).<br />

However, it is likely that exhumation <strong>of</strong> <strong>the</strong> GHS occurred along <strong>the</strong> STDS prior to<br />

orogen-parallel deformation for two reasons:<br />

(1) The kinematics <strong>of</strong> <strong>the</strong> upper Karnali are dominated by strike-slip and minor dip-slip<br />

geometries, which do not provide an effective mechanism for exhuming <strong>the</strong> mid-crustal<br />

rocks <strong>of</strong> <strong>the</strong> GHS. Considering an average lineation plunging ~17° towards 115˚, and an<br />

initial depth <strong>of</strong> ~26 km for <strong>the</strong> GHS <strong>of</strong> <strong>the</strong> upper Karnali valley (calculated from P-T data<br />

from Yakymchuk and Godin (2012) from samples adjacent to <strong>the</strong> village <strong>of</strong> Simikot, with<br />

a geo<strong>the</strong>rmal gradient <strong>of</strong> ~25°C/km), <strong>the</strong> exhumation <strong>of</strong> mid-crustal rocks via exclusively<br />

orogen-parallel geometries would have required <strong>the</strong> exhumation <strong>of</strong> rocks from ~85<br />

surficial km to <strong>the</strong> WNW to recreate <strong>the</strong> GHS-TSS contact currently observed in <strong>the</strong><br />

upper Karnali valley.<br />

83


(2) Previously published chronological constraints suggest that <strong>the</strong> GHS was undergoing<br />

south-directed extrusion prior to <strong>the</strong> ~15 Ma onset <strong>of</strong> orogen-parallel extension (see<br />

Godin, 2006b for review <strong>of</strong> ages, Cottle et al., 2009a,b). This is discussed in <strong>the</strong><br />

following section.<br />

With this logic, <strong>the</strong> final juxtaposition <strong>of</strong> units in <strong>the</strong> upper Karnali valley should be a<br />

combined result <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS, followed by exhumation during orogenparallel<br />

extension.<br />

2.7.1.2 Timing <strong>of</strong> south-directed extrusion<br />

Metamorphic monazite ages within <strong>the</strong> upper Karnali span from ~26 to 12 Ma, with <strong>the</strong><br />

highest frequency residing between 19 -15 Ma (Figs. 2.17, 2.20). Neohimalayan metamorphism<br />

throughout <strong>the</strong> central to central-eastern Himalaya is bracketed between 25 -16 Ma and is<br />

associated with sillimanite-grade metamorphism, extensive anatexis, and south-directed extrusion<br />

<strong>of</strong> <strong>the</strong> GHS (Vannay and Hodges, 1996; Godin et al., 1999; 2001; Searle and Szulc, 2005; Cottle<br />

et al., 2009b). Metamorphic ages within <strong>the</strong> GHS <strong>of</strong> <strong>the</strong> upper Karnali valley are consistent with<br />

this time frame. We <strong>the</strong>refore interpret <strong>the</strong> monazite crystallization ages from <strong>the</strong> GHS <strong>of</strong> <strong>the</strong><br />

upper Karnali valley to be associated with south-directed extrusion <strong>of</strong> <strong>the</strong> GHS.<br />

The majority <strong>of</strong> monazite grains show older low-Y cores mantled by younger high-Y<br />

rims (Fig. 2.16, 2.17). As garnet has a high affinity for Y during crystallization, <strong>the</strong> quantity <strong>of</strong> Y<br />

in <strong>the</strong> system (and hence in <strong>the</strong> monazite) can be considered as an inverse analogue for garnet<br />

growth and prograde-metamorphism. In this regard, we interpret <strong>the</strong> transition from older low-Y<br />

cores to younger high-Y rims to record <strong>the</strong> progression from peak to retrograde metamorphism<br />

between ~18 -15 Ma. This time frame is likely concurrent with <strong>the</strong> final stages <strong>of</strong> south-directed<br />

GHS extrusion, prior to <strong>the</strong> onset <strong>of</strong> orogen-parallel deformation. A study by Kellett et al. (2010)<br />

84


along <strong>the</strong> STDS in Bhutan documented a similar age progression, in which <strong>the</strong> growth <strong>of</strong> 16-15<br />

Ma Y-rich monazite rims are interpreted to record retrograde GHS metamorphism during final<br />

stages <strong>of</strong> south-directed extrusion.<br />

A minor amount <strong>of</strong> monazite crystallization occurred between 14 -12 Ma (Fig. 2.17).<br />

These ages are derived from <strong>the</strong> structurally deepest sample (HK117) located in <strong>the</strong> migmatitic<br />

gneiss <strong>of</strong> domain VI (Fig. 2.17, 2.20). High temperatures recorded at <strong>the</strong>se structural levels (Fig.<br />

2.15) may have resulted in a longer cooling history, allowing for protracted crystallization <strong>of</strong><br />

monazite, likely post-dating <strong>the</strong> cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS.<br />

2.7.2 Orogen-parallel extension<br />

2.7.2.1 ESE-WNW deformation and exhumation<br />

Structural fabrics observed in <strong>the</strong> upper Karnali valley define a high-strain ESE-WNW<br />

trending shear zone, associated with orogen-parallel extension (Figs. 2.3, 2.20). Deformation<br />

temperatures record a progression from high-temperature ductile to lower-temperature ductile and<br />

brittle deformation (Fig. 2.15). These data imply that <strong>the</strong> GHS was deformed during exhumation<br />

at progressively shallower levels. Kinematic data indicate a weak dominance <strong>of</strong> dextral fabric<br />

elements, a significant contribution <strong>of</strong> pure shear, and a potential deviation from plane strain<br />

towards constriction (Fig. 2.15). Integrating <strong>the</strong>se data suggests that within <strong>the</strong> upper Karnali<br />

valley, ESE-WNW oriented extension at progressively lower temperatures was related to ongoing<br />

transtension and orogen-parallel extension.<br />

Overprinting relationships show a progression from high-temperature ductile to lowertemperature<br />

ductile and brittle deformation, indicating progressive exhumation <strong>of</strong> <strong>the</strong> GHS at<br />

lower structural levels. This relationship is best observed within <strong>the</strong> high-strain gneiss <strong>of</strong> domain<br />

V. Within this domain, <strong>the</strong> dominant ESE-WNW oriented fabric (deformation T. <strong>of</strong> ~500-650°C;<br />

85


Figs. 2.7b,d, 2.15) is locally overprinted by centimeter-thick mylonitic layers (deformation T. <strong>of</strong><br />

~ 350°C; Figs. 2.8d, 2.14), and brittle structures (Fig. 2.8a,b,c). Cataclastic deformation <strong>of</strong><br />

dynamically recrystallized quartz confirms a progression towards lower temperature, while <strong>the</strong><br />

injection <strong>of</strong> crosscutting pseudotachylite (Fig. 2.8b) implies high-strain seismic events occurred<br />

during exhumation.<br />

A temperature progression is also observed with increasing distance normal to <strong>the</strong> GHS-<br />

TSS interface. Microstructural data show an increase in deformation temperatures <strong>of</strong> ESE-WNW<br />

oriented fabrics from ~350°C along <strong>the</strong> GHS-TSS interface to ~650°C at structural depths below<br />

~2.5 km (Figs. 2.15, 2.20). Similar increases in temperature are commonly associated with<br />

telescoping and/or flattening <strong>of</strong> iso<strong>the</strong>rms as a result <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS (Law<br />

et al., 2011 and references <strong>the</strong>rein). Telescoping <strong>of</strong> iso<strong>the</strong>rms may be possible during normal<br />

faulting and core complex formation; however, it would require ei<strong>the</strong>r penetrative pure shear<br />

resulting in extensive vertical shortening or heterogeneous simple shear resulting in <strong>the</strong><br />

juxtaposition <strong>of</strong> different particle paths, or some combination <strong>of</strong> both (Law et al., 2011). These<br />

processes are typically documented in channelized flow type scenarios, as <strong>the</strong>y require high<br />

temperatures and significant transport displacement (Law et al., 2011 and references <strong>the</strong>rein).<br />

We <strong>the</strong>refore interpret <strong>the</strong> rapid increase and high deformation temperatures recorded<br />

within ESE-WNW oriented fabrics to be a remnant <strong>of</strong> telescoped and/or flattened iso<strong>the</strong>rms<br />

generated during south-directed extrusion <strong>of</strong> <strong>the</strong> GHS. This <strong>the</strong>rmal inheritance attests to <strong>the</strong><br />

intimate temporal link between <strong>the</strong>se two tectonic processes.<br />

Ductile ESE-WNW oriented extension is characterized by significant contributions <strong>of</strong><br />

pure and simple shear. This is revealed through, (1) fabric elements that show a near even<br />

distribution <strong>of</strong> shear senses, with a minor dominance <strong>of</strong> dextral strike-slip motion (Figs. 2.4, 2.5,<br />

86


2.6, 2.7), and (2) dominantly symmetric quartz CPO fabrics (Figs. 2.12, 2.15). The presence <strong>of</strong><br />

asymmetric CPO patterns proximal to <strong>the</strong> GHS-TSS interface may imply that simple shear was<br />

predominant along this boundary and pure shear, which resulted in WNW-ESE oriented<br />

extension, dominated at lower structural levels. Transitions in CPO fabrics at depths <strong>of</strong> ~1 km<br />

below <strong>the</strong> GHS-TSS interface may correspond to a deviation from plane strain towards<br />

constriction (Figs. 2.12, 2.15).<br />

Deformation during exhumation at progressively lower temperatures is also characterized<br />

by pure shear (~59%), as revealed from a vorticity analysis <strong>of</strong> a crosscutting lower temperature<br />

mylonitic layer (deformation T. <strong>of</strong> ~ 350°C; Figs. 2.8d, 2.14).<br />

Although sparse, <strong>the</strong>se data are consistent with transtensional deformation, in which noncoaxial<br />

deformation parallel to <strong>the</strong> fault zone is coupled with coaxial extension oblique to <strong>the</strong><br />

fault zone. This simultaneous deformation would generate a bulk constrictional strain (Dewey,<br />

2002), consistent with our observations. This detailed interpretation is also consistent with <strong>the</strong><br />

regional network <strong>of</strong> transtension proposed by Murphy and Copeland (2005).<br />

2.7.2.2 Timing <strong>of</strong> ESE-WNW deformation and cooling<br />

Intense internal deformation and alignment <strong>of</strong> monazite grains parallel to <strong>the</strong> main<br />

lineation (Fig. 2.16) suggest that monazite underwent extensive post-crystallization deformation<br />

and transposition. These grains are typically aligned parallel to, or overgrown by, ESE-WNW<br />

oriented micas. Previously described age relationships suggest that <strong>the</strong> crystallization <strong>of</strong> monazite<br />

rims was concurrent with <strong>the</strong> final stages <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS. Therefore, <strong>the</strong><br />

deformation and alignment <strong>of</strong> monazite grains, and growth <strong>of</strong> ESE-WNW oriented micas should<br />

post-date south-directed extrusion <strong>of</strong> <strong>the</strong> GHS and be synchronous with orogen-parallel<br />

extension.<br />

87


Thermochronology <strong>of</strong> muscovite grains defining <strong>the</strong> ESE-WNW lineation reveal that by<br />

~15 to 9 Ma, <strong>the</strong> upper Karnali valley had cooled through ~450°C in a setting linked to orogenparallel<br />

extension (Fig. 2.18, Table 2.2). If only well-defined muscovite age plateaus are<br />

considered, <strong>the</strong>n <strong>the</strong> cooling age <strong>of</strong> <strong>the</strong> upper Karnali valley can be constrained between ~13 to<br />

10 Ma. Undulatory extinction and brittle deformation <strong>of</strong> <strong>the</strong>se muscovite grains (Fig. 2.5e)<br />

indicate that orogen-parallel extension was ongoing at progressively cooler temperatures.<br />

Cooling ages <strong>of</strong> <strong>the</strong> upper Karnali valley show a mild younging trend with increasing<br />

structural depth below <strong>the</strong> GHS-TSS interface (Fig. 2.19a). This is consistent with slightly<br />

younger monazite crystallization ages at depth (sample HK117, Fig. 2.17), suggesting a<br />

protracted cooling history at deep structural levels due to high temperatures.<br />

Muscovite cooling ages also show a younging trend towards <strong>the</strong> sou<strong>the</strong>ast <strong>of</strong> <strong>the</strong> field<br />

area (Fig. 2.19b), potentially due to a sou<strong>the</strong>asterly propagating fault system. This hypo<strong>the</strong>sis was<br />

also proposed by Murphy and Copeland (2005) and is consistent with transtensional deformation<br />

initiating near <strong>the</strong> Gurla Mandhata core complex and propagating towards <strong>the</strong> sou<strong>the</strong>ast.<br />

2.7.3 Transition from south-directed extrusion to orogen-parallel extension<br />

2.7.3.1 Upper Karnali valley<br />

Our data reveal that <strong>the</strong> GHS <strong>of</strong> <strong>the</strong> upper Karnali valley has been progressively extruded<br />

/exhumed in two distinct tectonic phases (Fig. 2.21):<br />

1. Pre ~15 Ma south-directed extrusion. Monazite ages associated with this stage are<br />

consistent with typical Neohimalayan metamorphism throughout <strong>the</strong> central to centraleastern<br />

Himalaya (~25 -16 Ma, see prior discussion for references).<br />

2. Orogen-parallel extension resulting in exhumation via oblique strike-slip faulting,<br />

initiating at ~15 -13 Ma. Structures and deformation mechanisms suggest that this phase<br />

88


Figure 2.21 Compilation <strong>of</strong> chronologic data and deformation temperatures from <strong>the</strong> upper Karnali valley.<br />

Red (HK117a), and purple (KH18) boxes and arrows show cooling history <strong>of</strong> individual samples. Grey<br />

vertical box is interpreted as <strong>the</strong> transition period between south-directed extrusion <strong>of</strong> <strong>the</strong> GHS and<br />

orogen-parallel extension.<br />

Box (1) represents previously published metamorphic ages associated with south-directed extrusion <strong>of</strong> <strong>the</strong><br />

GHS (see text for references). Monazite ages from this study are found in <strong>the</strong> youngest section <strong>of</strong> this<br />

period.<br />

Boxes (2) through to (6) represent WNW-ESE oriented fabrics and ages associated with orogen-parallel<br />

deformation and extension. Deformation temperatures <strong>of</strong> <strong>the</strong>se boxes are calculated via microstructures,<br />

and ages are determined relative to one ano<strong>the</strong>r (i.e. fabrics <strong>of</strong> box 5 deforms fabrics <strong>of</strong> box 4 at lower<br />

temperatures). Thus, <strong>the</strong>se boxes are intended to illustrate a trend, and <strong>the</strong>ir absolute temperature and age<br />

constraints may vary from what is shown.<br />

Box (2) is constrained between 500 -700°C through quartz recrystallization mechanisms (grain boundary<br />

migration, chessboard extinction) and quart CPO opening angle deformation temperatures (samples<br />

HK140, HK105). These fabrics formed at temperatures lower <strong>the</strong>n <strong>the</strong> crystallization <strong>of</strong> monazite, host<br />

deformed monazite, and are hotter <strong>the</strong>n <strong>the</strong> closure temperature <strong>of</strong> muscovite. Therefore this box is older<br />

than <strong>the</strong> cooling <strong>of</strong> muscovite and younger than <strong>the</strong> crystallization <strong>of</strong> monazite. Caption continued on<br />

next page.<br />

89


Figure 2.21 – continued –<br />

Box (3) represents muscovite that define WNW-ESE fabrics. Closure temperatures are estimated at 450 ±<br />

25°C (see text for reference), and ages calculated by 40 Ar/ 39 Ar <strong>the</strong>rmochronology.<br />

Box (4) is constrained between 350 -500°C through quartz recrystallization mechanisms (subgrain<br />

rotation) and quart CPO opening angle deformation temperatures (samples HK131A, HK139D). These<br />

fabrics formed at temperatures similar to <strong>the</strong> muscovite closure temperature, and <strong>the</strong>refore should be<br />

roughly coeval with muscovite cooling ages.<br />

Box (5) is constrained between 280 -400°C through quartz recrystallization mechanisms (bulging), quartz<br />

CPO opening angle deformation temperature (sample HK124B), and brittly deformed muscovite.<br />

Deformed muscovites indicate that <strong>the</strong>se fabrics must be younger than muscovite cooling ages.<br />

Box (6) represents brittle deformation at < 280 ± 30°C. Normal faults and cataclasite post-date and<br />

deform all ductile fabrics, <strong>the</strong>refore <strong>the</strong>y must be <strong>the</strong> youngest event.<br />

_____________________________________________________________________________<br />

initiated at high temperatures, and progressed to lower temperatures during ongoing<br />

orogen-parallel deformation and exhumation (Fig. 2.21). Monazites grains associated<br />

with south-directed extrusion are typically deformed and transposed parallel to orogenparallel<br />

lineations throughout this phase.<br />

Taken toge<strong>the</strong>r, <strong>the</strong>se data from <strong>the</strong> upper Karnali valley reveal a fundamental tectonic<br />

transition from <strong>the</strong> cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS to onset <strong>of</strong> orogen-parallel<br />

extension (Fig. 2.21). Based upon chronological constraints associated with <strong>the</strong>se two phases, we<br />

suggest this transition occurred between ~15 -13 Ma.<br />

2.7.3.2 Himalayan transition<br />

To characterize <strong>the</strong> transition between cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> midcrust<br />

and onset <strong>of</strong> orogen-parallel extension, <strong>the</strong> analysis <strong>of</strong> field areas containing both styles <strong>of</strong><br />

deformation is required. The majority <strong>of</strong> orogen-parallel deformation is manifested as upper<br />

crustal east-trending strike-slip and N-S striking normal faults throughout Tibet and <strong>the</strong> nor<strong>the</strong>rn<br />

Himalaya (Armijo et al., 1986, 1989; Taylor et al., 2003; Taylor and Yin, 2009; Ratschbacher et<br />

al., 2012). There are however, three migmatite-cored domes across <strong>the</strong> Himalaya that exhume<br />

90


mid-crustal rocks in settings kinematically linked to extension. These domes are: <strong>the</strong> Leo Pargil<br />

dome (Thiede et al., 2006; Leech, 2008), <strong>the</strong> Gurla Mandhata core complex (Murphy et al.,<br />

2002), and <strong>the</strong> Ama Drime Massif (Jessup et al., 2008, Cottle et al., 2009a).<br />

A compilation <strong>of</strong> chronologic data from across <strong>the</strong>se domes with ages from <strong>the</strong> upper<br />

Karnali valley suggests a common middle Miocene transition from cessation <strong>of</strong> south-directed<br />

extrusion <strong>of</strong> <strong>the</strong> GHS to onset <strong>of</strong> orogen-parallel extension (Fig. 2.22). This transition is well<br />

constrained in upper Karnali valley and in <strong>the</strong> Everest region between ~15 – 13 Ma (this study;<br />

Jessup and Cottle, 2010). In contrast, <strong>the</strong> transitions in <strong>the</strong> Leo Pargil dome (data from Thiede et<br />

al., 2006; Leech, 2008) and <strong>the</strong> Gurla Mandhata core complex (data from Murphy et al., 2002)<br />

are less well constrained, but can still be estimated to be between ~19 – 16 Ma and ~17 – 13 Ma,<br />

respectively.<br />

The age <strong>of</strong> <strong>the</strong> transition is oldest in <strong>the</strong> Leo Pargil dome, located in <strong>the</strong> western<br />

Himalaya. Although contentious, it can be argued that <strong>the</strong> collision between India and Asia was<br />

diachronous, having initiated in <strong>the</strong> west and progressed eastward (Rowley, 1996; 1998). If this is<br />

true, <strong>the</strong>n it may be possible that this transition follows a similar temporal trend. Thus, <strong>the</strong><br />

transition within <strong>the</strong> Leo Pargil dome may be expected to initiate earlier than within gneiss-domes<br />

located in <strong>the</strong> central and eastern Himalaya.<br />

Based upon <strong>the</strong> following three criteria, we suggest that <strong>the</strong> transition from south-directed<br />

extrusion <strong>of</strong> <strong>the</strong> GHS to onset <strong>of</strong> orogen-parallel extension occurred in <strong>the</strong> middle Miocene (~ 16<br />

Ma) across <strong>the</strong> length <strong>of</strong> <strong>the</strong> orogen.<br />

1. A common middle Miocene transition within migmatite-cored domes across <strong>the</strong><br />

Himalaya.<br />

91


92<br />

Figure 2.22 (a) Location <strong>of</strong> study area<br />

and migmatite cored domes associated<br />

with orogen-parallel extension. Red<br />

lines represent faults associated with<br />

exhumation <strong>of</strong> <strong>the</strong>se domes, and<br />

orogen-parallel extensional structures.<br />

Modified after Styron et al. (2011). (b)<br />

Compilation from west to east <strong>of</strong><br />

chronologic data from across <strong>the</strong>se<br />

domes and <strong>the</strong> upper Karnali valley.<br />

See text for chronology references.<br />

Abbreviations: ITSZ – Indus-Tsangpo<br />

suture zone, MFT – Main Frontal<br />

Thrust.


2. The cessation <strong>of</strong> south-directed GHS metamorphism at ~16 Ma in <strong>the</strong> central to centraleastern<br />

Himalaya (see prior discussion for references).<br />

3. The onset <strong>of</strong> orogen-parallel and east-west extension at ≤ 16 Ma (see prior discussion for<br />

references). (An exception to this is a set <strong>of</strong> north-south striking dykes in sou<strong>the</strong>rn Tibet<br />

dated between 18.3 ± 2.7 Ma to 13.3 ± 0.8 Ma (Williams et al., 2001). If <strong>the</strong> large error<br />

associated with <strong>the</strong> older age <strong>of</strong> onset is considered, <strong>the</strong>n <strong>the</strong>ses age constraints are in<br />

agreement with <strong>the</strong> proposed transition).<br />

There are, however, some constraints that do not agree with our proposed transition.<br />

Notably, timing constraints from <strong>the</strong> eastern GHS within Bhutan indicate protracted hightemperature<br />

metamorphism until ~10 Ma (Warren et al., 2012). According to our transition, <strong>the</strong><br />

cessation <strong>of</strong> south-directed extrusion and metamorphism <strong>of</strong> <strong>the</strong> GHS should have ceased several<br />

million years prior.<br />

Warren et al. (2012) attribute this young high-temperature metamorphism to <strong>the</strong><br />

exhumation <strong>of</strong> deeper structural levels not exposed elsewhere in <strong>the</strong> Himalaya. Initial stages <strong>of</strong><br />

this exhumation path are constrained by 14 -13 Ma monazite and 15 -14 Ma zircon (Warren et al.,<br />

2012). These initial age constraints are roughly coeval with <strong>the</strong> onset <strong>of</strong> <strong>the</strong> orogen-parallel<br />

extension throughout <strong>the</strong> Himalaya. Therefore, it may be possible that exhumation from deep<br />

structural levels was in part assisted by <strong>the</strong> onset <strong>of</strong> orogen-parallel extension and exhumation. If<br />

this were true, <strong>the</strong>n a middle Miocene transition may have occurred in <strong>the</strong> eastern GHS.<br />

Alternatively, <strong>the</strong>se young ages may be due to a diachronous transition initiating in <strong>the</strong><br />

west and younging towards <strong>the</strong> east <strong>of</strong> <strong>the</strong> orogen (Rowley, 1996; 1998). This would be<br />

consistent with older ages observed in <strong>the</strong> Leo Pargil dome, and <strong>the</strong> sou<strong>the</strong>ast-directed younging<br />

<strong>of</strong> orogen-parallel deformation observed in <strong>the</strong> upper Karnali valley (Fig. 2.19b). However, <strong>the</strong>re<br />

93


is little data to support both aforementioned hypo<strong>the</strong>ses. Therefore, we cannot discount <strong>the</strong><br />

possibility that young GHS ages within Bhutan may indicate that our interpretation <strong>of</strong> a middle<br />

Miocene transition is not applicable on an orogen-wide scale.<br />

2.7.4 Tectonic model explaining orogen-parallel extension<br />

Numerous studies have attempted to explain orogen-parallel extension throughout <strong>the</strong><br />

Himalaya and Tibet, invoking models <strong>of</strong> lateral extrusion <strong>of</strong> Tibet (Tapponnier et al. 1982),<br />

oroclinal bending <strong>of</strong> <strong>the</strong> Himalayan arc (Klootwijk et al., 1985; Rataschbacher et al., 1994;<br />

Bendick and Bilham, 2001), gravitational instability <strong>of</strong> <strong>the</strong> Tibetan plateau (Molnar and<br />

Tapponnier, 1978; Molnar and Chen 1983; Coleman and Hodges, 1995), right-lateral oblique<br />

convergence (McCaffrey and Nabelek, 1998; Seeber and Pêcher, 1998) and radial spreading <strong>of</strong><br />

<strong>the</strong> Tibetan plateau (England and Houseman, 1988; Copley and McKenzie, 2007; Copley 2008;<br />

Cook and Royden, 2008). These models are commonly based upon <strong>the</strong> kinematics and orientation<br />

<strong>of</strong> faulting with respect to <strong>the</strong> orogen (e.g. Murphy and Copeland, 2005; Thiede et al., 2006), and<br />

more recently on present-day GPS velocity vectors (e.g. Shen et al., 2000; 2001; Zhang et al.,<br />

2004; Allmendinger et al., 2007; Gan et al., 2007; Styron et al., 2011).<br />

The high-strain zone <strong>of</strong> <strong>the</strong> upper Karnali valley defines an approximately orogenparallel<br />

right-lateral fault system in <strong>the</strong> central-western Himalaya. The geographic position and<br />

sense <strong>of</strong> motion along this fault are consistent with radial spreading <strong>of</strong> <strong>the</strong> Tibetan plateau,<br />

oblique convergence and possibly lateral extrusion <strong>of</strong> Tibet. Therefore, we cannot rely upon <strong>the</strong>se<br />

data alone to evaluate <strong>the</strong> mechanism driving orogen-parallel extension.<br />

By integrating chronological data we are able to constrain <strong>the</strong> onset <strong>of</strong> orogen-parallel<br />

extension to ~15 -13 Ma in <strong>the</strong> upper Karnali valley, and potentially to <strong>the</strong> middle Miocene<br />

94


throughout <strong>the</strong> Himalaya (Figs. 2.21, 2.22). This high-resolution time frame provides a constraint<br />

that can be used to test <strong>the</strong> various models seeking to explain orogen-parallel extension.<br />

The only model that predicts a major tectonic change in <strong>the</strong> middle Miocene leading to<br />

<strong>the</strong> onset <strong>of</strong> orogen-parallel extension is a numerical model <strong>of</strong> modified radial spreading <strong>of</strong> <strong>the</strong><br />

lower crust <strong>of</strong> Tibet (Fig. 2.23; Royden et al., 1997; Cook and Royden, 2008). This numerical<br />

model simulates three-dimensional deformation during continental convergence during which <strong>the</strong><br />

lower crust decreases viscosity, decouples from <strong>the</strong> overlying upper crust, and flows radially. By<br />

incorporating pre-collisional crustal heterogeneities, including <strong>the</strong> strong Tarim and Sichuan<br />

Basin, <strong>the</strong> model predicts that after ~40 million years <strong>of</strong> convergence <strong>the</strong> lower crust underlying<br />

<strong>the</strong> Tibetan plateau would begin to spread laterally to <strong>the</strong> east and rotate clockwise around <strong>the</strong><br />

eastern syntaxis (Fig. 2.23). Orogen-parallel extension throughout central and sou<strong>the</strong>rn Tibet is<br />

interpreted as a result <strong>of</strong> <strong>the</strong> onset <strong>of</strong> lower-crustal east-directed flow (Cook and Royden, 2008).<br />

Assuming onset <strong>of</strong> collision between India and Asia at ~55-50 Ma (Najman et al., 2010), <strong>the</strong> end<br />

<strong>of</strong> 40 million years <strong>of</strong> convergence corresponds to ~15-10 Ma. This timing is consistent with <strong>the</strong><br />

onset <strong>of</strong> orogen-parallel extension recorded in <strong>the</strong> upper Karnali valley (Fig. 2.21) and potentially<br />

orogen-wide (Fig. 2.22).<br />

If this model is correct, <strong>the</strong>n <strong>the</strong> middle Miocene should also correspond to <strong>the</strong> uplift <strong>of</strong><br />

eastern Tibet due to <strong>the</strong> buttressing <strong>of</strong> east-flowing lower crust against <strong>the</strong> strong Sichuan basin.<br />

This rise in mean elevation <strong>of</strong> eastern Tibet is confirmed via two independent methods: (1)<br />

initiation <strong>of</strong> rapid river incision at 13 -9 Ma in sou<strong>the</strong>astern Tibet (Clark et al., 2005), and (2) a<br />

significant change in <strong>the</strong> oxygen isotopic composition <strong>of</strong> meteoric water between 13 -12 Ma in<br />

nor<strong>the</strong>astern Tibet, as inferred from <strong>the</strong> δ 18 O <strong>of</strong> lacustrine carbonates (Dettman et al., 2003). A<br />

major shift in carbonate δ 18 O between 13-12 Ma indicates a shift to more arid conditions, thought<br />

95


Figure 2.23 Numerical model <strong>of</strong> modified radial spreading <strong>of</strong> <strong>the</strong> lower crust <strong>of</strong> Tibet,<br />

after Cook and Royden (2008). (a) Modeled elevation and surface velocity <strong>of</strong> a<br />

Himalayan-Tibetan sized orogen after 40 million years (m.yr.) <strong>of</strong> convergence. (b)<br />

Present day topography colored to match elevation scale, and select normal and strikeslip<br />

faults. Numbers correspond to <strong>the</strong> following events synchronous during <strong>the</strong> middle<br />

Miocene: (1) onset <strong>of</strong> orogen-parallel extension, (2) east-directed spreading and rotation<br />

<strong>of</strong> <strong>the</strong> Tibetan plateau, and (3) cessation <strong>of</strong> south-directed mid-crustal flow (not<br />

stipulated by <strong>the</strong> model, but suggested by our results).<br />

96


to represent <strong>the</strong> time at which <strong>the</strong> nor<strong>the</strong>astern Tibetan plateau reached <strong>the</strong> necessary elevation to<br />

block <strong>the</strong> penetration <strong>of</strong> moisture from <strong>the</strong> surrounding oceans (Dettman et al., 2003).<br />

Fur<strong>the</strong>rmore, most present-day GPS velocity vector studies suggest that <strong>the</strong> Tibetan plateau may<br />

be deforming as a continuous medium and extending towards <strong>the</strong> ESE (e.g. Zhang et al., 2004;<br />

Gan et al., 2007), consistent with a model <strong>of</strong> radial spreading.<br />

Actual chronologic constraints characterizing <strong>the</strong> onset <strong>of</strong> orogen-parallel extension<br />

(Figs. 2.2a, 2.21) are consistent with those predicted in <strong>the</strong> Cook and Royden (2008) model <strong>of</strong><br />

radial spreading <strong>of</strong> <strong>the</strong> Tibetan plateau (Fig. 2.23). This does not discount o<strong>the</strong>r models<br />

explaining orogen-parallel extension; however, it does <strong>of</strong>fer additional support for models <strong>of</strong><br />

radial spreading.<br />

The transition we studied suggests that cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS<br />

occurred immediately prior to <strong>the</strong> onset <strong>of</strong> orogen-parallel extension. Therefore within <strong>the</strong> context<br />

<strong>of</strong> radial-spreading models, our data provide <strong>the</strong> following additional constraint: initiation <strong>of</strong> eastdirected<br />

lower-crustal flow and orogen-parallel extension immediately postdates <strong>the</strong> cessation <strong>of</strong><br />

south-directed extrusion <strong>of</strong> <strong>the</strong> GHS (Fig. 2.23). It may be possible that <strong>the</strong>re is an intimate link<br />

between cessation <strong>of</strong> south-directed mid-crustal flow and onset <strong>of</strong> east-directed lower crustal<br />

flow; however, current state <strong>of</strong> knowledge has not yet clarified this issue.<br />

2.8 Conclusions<br />

Integrated geological, structural, microstructural, and chronological data reveal a twophase<br />

extrusion /exhumation history <strong>of</strong> <strong>the</strong> GHS <strong>of</strong> <strong>the</strong> upper Karnali valley. The first phase is<br />

associated with south-directed extrusion <strong>of</strong> <strong>the</strong> GHS. Dominant U-Th-Pb monazite ages <strong>of</strong> 19 -15<br />

Ma are consistent with <strong>the</strong> final stages <strong>of</strong> Neohimalayan metamorphism and <strong>the</strong> cessation <strong>of</strong><br />

south-directed extrusion.<br />

97


The second phase is associated with orogen-parallel extension. Structural and kinematic<br />

data suggest that this phase <strong>of</strong> deformation represents orogen-parallel deformation and<br />

exhumation at progressively lower temperatures, possibly within a larger network <strong>of</strong> transtension.<br />

Muscovite <strong>the</strong>rmochronology indicate that by ~15 -13 Ma, orogen-parallel extension was active<br />

in <strong>the</strong> upper Karnali valley.<br />

Assembling <strong>the</strong>se data delineates a transition from cessation <strong>of</strong> south-directed extrusion<br />

<strong>of</strong> <strong>the</strong> GHS to onset <strong>of</strong> orogen-parallel extension. This transition initiated at high temperatures,<br />

progressed into <strong>the</strong> brittle realm, and is bracketed between ~15 -13 Ma within <strong>the</strong> upper Karnali<br />

valley.<br />

A compilation <strong>of</strong> our chronologic data with three migmatite-cored domes across <strong>the</strong><br />

Himalaya that exhume mid-crustal rocks in settings kinematically linked to extension, reveals a<br />

common middle Miocene orogen-wide transition.<br />

A middle Miocene transition supports a model <strong>of</strong> modified radial spreading (Cook and<br />

Royden, 2008), to explain <strong>the</strong> onset and continuation <strong>of</strong> orogen-parallel extension throughout <strong>the</strong><br />

Himalaya and Tibet.<br />

98


Chapter 3<br />

Discussion<br />

3.1 Summary<br />

Integrating geological, structural, microstructural, and chronological data from <strong>the</strong> upper<br />

Karnali valley reveals a fundamental tectonic transition between 15 -13 Ma, from cessation <strong>of</strong><br />

south-directed extrusion <strong>of</strong> <strong>the</strong> mid-crust to onset <strong>of</strong> orogen-parallel extension. A compilation <strong>of</strong><br />

our data with three migmatite-cored domes that exhume mid-crustal rocks in settings<br />

kinematically linked to extension reveals a common middle Miocene orogen-wide transition (Ch.<br />

2).<br />

3.1.1 Middle Miocene orogen-parallel extension and east-directed lower-crustal flow<br />

The studied transition is broadly contemporaneous with <strong>the</strong> cessation <strong>of</strong> mid-crustal flow<br />

and onset <strong>of</strong> orogen-parallel extension across <strong>the</strong> entire Himalayan-Tibetan orogen during <strong>the</strong><br />

middle Miocene (Ch. 2). These processes are also roughly synchronous with initiation <strong>of</strong> uplift <strong>of</strong><br />

<strong>the</strong> eastern Tibetan plateau (Ch. 2). These observations can be explained by a numerical model <strong>of</strong><br />

radial spreading <strong>of</strong> Tibet, in which <strong>the</strong> lower crust becomes <strong>the</strong>rmally weak, decouples from <strong>the</strong><br />

upper crust and flows radially and southward during <strong>the</strong> early to middle Miocene, and eastward<br />

<strong>the</strong>reafter (Royden et al., 1997; Cook and Royden, 2008).<br />

The numerical models <strong>of</strong> Royden et al. (1997), and Cook and Royden (2008) simulate<br />

three-dimensional deformation <strong>of</strong> a viscous crust during continental convergence (Fig. 3.1). The<br />

model parameters include an idealized crust with depth-dependent viscosity and a convergence<br />

rate approximating <strong>the</strong> velocity <strong>of</strong> <strong>the</strong> mantle. Modeling results show that after 4 -16 Myr <strong>of</strong><br />

convergence, <strong>the</strong> lower crust becomes so <strong>the</strong>rmally weak that it decoupled from <strong>the</strong> overlying<br />

99


yr<br />

Figure 3.1 Numerical model <strong>of</strong> Cook and Royden (2008) simulating <strong>the</strong> formation <strong>of</strong> a<br />

plateau after (a) ~16 Myr <strong>of</strong> convergence, and (b) after ~40 Myr <strong>of</strong> convergence. (c)<br />

Present day topography and select normal and strike-slip faults (black lines), also<br />

modified after Cook and Royden, 2008. Elevation scale is same for all figures.<br />

Numbers correspond to tectonic and metamorphic events, see text for references.<br />

100


upper crust and begins to flow radially, forming a plateau (Fig. 3.1; Royden et al., 1997; Cook<br />

and Royden, 2008). Assuming onset <strong>of</strong> collision <strong>of</strong> India and Eurasia at ~55-50 Ma (Green et al.,<br />

2008; Najman et al., 2010), this timing is consistent with <strong>the</strong> earliest metamorphic ages obtained<br />

from <strong>the</strong> GHS at ~39 Ma, corresponding to crustal thickening and prograde metamorphism<br />

(Godin et al., 2001; Prince et al., 2001; Cottle et al., 2009b), and <strong>the</strong> oldest ages obtained for <strong>the</strong><br />

surface <strong>of</strong> central Tibet at ~35 Ma, corresponding to plateau formation (Rowley and Currie,<br />

2006).<br />

Following ~40 Myr <strong>of</strong> convergence, modeling results <strong>of</strong> Cook and Royden (2008) predict<br />

that <strong>the</strong> Tibetan plateau begins to flow laterally to <strong>the</strong> east and rotates clockwise around <strong>the</strong><br />

eastern syntaxis (Fig. 3.1). This results in orogen-parallel extension across <strong>the</strong> topographic front<br />

and plateau areas. Fur<strong>the</strong>rmore, this model, and a study by Clark and Royden (2000), predict that<br />

buttressing <strong>of</strong> east-flowing lower crust against <strong>the</strong> strong Sichuan basin causes thickening <strong>of</strong> <strong>the</strong><br />

lower crust adjacent to <strong>the</strong> basin, resulting in uplift <strong>of</strong> eastern Tibet (Fig. 3.1). These predictions<br />

and <strong>the</strong>ir timing are consistent with <strong>the</strong> onset <strong>of</strong> orogen-parallel extension (~16 -11 Ma), and <strong>the</strong><br />

uplift (~13-12 Ma) <strong>of</strong> and lack <strong>of</strong> deformation within <strong>the</strong> upper crust <strong>of</strong> eastern Tibet (Burchfiel<br />

et al., 1995).<br />

This numerical model <strong>of</strong> east-directed lower crustal flow is also consistent with middle<br />

Miocene to present day evolution <strong>of</strong> <strong>the</strong> Himalaya and Tibet. Within sou<strong>the</strong>rn and central Tibet,<br />

high plateau elevations have remained relatively stable (Harris, 2006). In contrast, nor<strong>the</strong>rn Tibet<br />

and eastern Tibet continue to be uplifted (Yang and Liu, 2002; Dettman et al., 2003; Clark et al.,<br />

2005), forming in some regions <strong>of</strong> eastern Tibet a topographic front somewhat comparable to <strong>the</strong><br />

Himalaya (Fig. 3.1; Burchfiel et al., 1995). Despite this gain in elevation, eastern Tibet continues<br />

to show a lack <strong>of</strong> Cenozoic upper-crustal shortening features (Burchfiel et al., 1995). These<br />

101


phenomena are consistent with ongoing east-directed lower-crustal flow and thickening <strong>of</strong> <strong>the</strong><br />

lower-crust adjacent to <strong>the</strong> Sichuan Basin.<br />

This time period also coincides with significant regional tectonic changes (Fig. 3.1),<br />

including <strong>the</strong> cessation <strong>of</strong> spreading in <strong>the</strong> South China Sea at ~15 Ma (Li et al., 2006), <strong>the</strong><br />

initiation <strong>of</strong> a major stage <strong>of</strong> extension within <strong>the</strong> Andaman Sea at ~15 Ma (Curray, 2005; Hall,<br />

2002), and <strong>the</strong> termination <strong>of</strong> east-directed subduction <strong>of</strong> India below <strong>the</strong> Burman plates at ~11<br />

Ma (Everett et al., 1990). These tectonic events were proceeded by major environmental changes<br />

including <strong>the</strong> initiation <strong>of</strong> <strong>the</strong> Asian monsoon at 9 -8 Ma (Zhisheng et al., 2001), and 8 -7 Ma<br />

changes in climate observed in <strong>the</strong> Siwalik foreland basin (Quade et al., 1989; 1995), Arabian Sea<br />

(Kroon et al., 1991; Prell et al., 1992) and Bay <strong>of</strong> Bengal (Burbank et al., 1993; Derry and<br />

France-Lanord, 1996). Although <strong>the</strong>se tectonic and environmental events are not directly related<br />

to <strong>the</strong> numerical model <strong>of</strong> Cook and Royden (2008), it may be significant that <strong>the</strong>y are roughly<br />

coeval with <strong>the</strong> studied transition and <strong>the</strong> onset <strong>of</strong> east-directed flow.<br />

A syn<strong>the</strong>sis <strong>of</strong> middle Miocene evolution <strong>of</strong> <strong>the</strong> Himalaya and Tibet suggest that<br />

cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS occurred immediately prior to <strong>the</strong> onset <strong>of</strong><br />

orogen-parallel extension and east-directed lower-crustal flow. If we interpret <strong>the</strong> GHS as <strong>the</strong><br />

result <strong>of</strong> an extruded low-viscosity channel from mid-crustal depths (Beaumont et al., 2001;<br />

2004; 2006; Jamieson et al., 2004; Godin et al., 2006a), <strong>the</strong>n it may be appropriate to make <strong>the</strong><br />

following hypo<strong>the</strong>sis: a middle Miocene transition affecting mid- to lower- crustal processes is<br />

responsible for <strong>the</strong> cessation <strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS and onset <strong>of</strong> east-directed<br />

lower-crustal flow.<br />

102


3.2 Future research considerations<br />

Extensive research opportunities exist throughout <strong>the</strong> Himalayan-Tibetan orogen. Our<br />

research in particular, has highlighted two important tectonic issues that merit fur<strong>the</strong>r<br />

investigation.<br />

3.2.1 Why did south-directed mid-crustal extrusion cease, and why did east-directed lower<br />

crustal flow initiate<br />

Results <strong>of</strong> this study shown an intimate link between cessation <strong>of</strong> south-directed midcrust<br />

extrusion and <strong>the</strong> onset <strong>of</strong> orogen-parallel extension, interpreted to correspond to <strong>the</strong><br />

initiation <strong>of</strong> east-directed lower crustal flow. However, it has not adequately addressed <strong>the</strong> issue<br />

<strong>of</strong> why south-directed extrusion <strong>of</strong> <strong>the</strong> mid-crust ceased, and why east-directed lower crustal flow<br />

initiated. We briefly hypo<strong>the</strong>sized that a transition affecting mid- to lower-crustal processes may<br />

have been responsible, however we cannot verify this.<br />

The model <strong>of</strong> Cook and Royden (2008) suggesst that <strong>the</strong> onset <strong>of</strong> east-directed lowercrustal<br />

flow may be a result <strong>of</strong> modified lower crustal flow after ~40 Myr <strong>of</strong> convergence.<br />

Alternatively, <strong>the</strong> cessation <strong>of</strong> spreading in <strong>the</strong> South China Sea (Li et al., 2006), and <strong>the</strong><br />

initiation <strong>of</strong> a major rifting phase in <strong>the</strong> Andaman Sea (Hall et al., 2002; Curray, 2005), suggest<br />

that far-field plate reconfigurations may also be partly responsible for middle Miocene transitions<br />

in Himalayan-Tibetan tectonics.<br />

An additional viable hypo<strong>the</strong>sis includes a middle Miocene modification <strong>of</strong> horizontal<br />

gradients in lithostatic pressure across <strong>the</strong> Himalaya and Tibet, due to south-directed extrusion <strong>of</strong><br />

<strong>the</strong> GHS. Extrusion <strong>of</strong> mid-crustal material via a channel-flow model requires a significant<br />

contrast in crustal thickness between <strong>the</strong> plateau and <strong>the</strong> foreland (Beaumont et al., 2001; Godin<br />

et al., 2006a). It may be possible that by middle Miocene, south-directed extrusion <strong>of</strong> <strong>the</strong> GHS<br />

103


towards <strong>the</strong> topographic front <strong>of</strong> <strong>the</strong> Himalaya had modified horizontal gradients in lithostatic<br />

pressure between <strong>the</strong> Himalayan foreland and <strong>the</strong> Tibetan plateau. A diminished contrast in<br />

horizontal pressure gradients between <strong>the</strong> Himalaya and Tibet could have resulted in <strong>the</strong> cessation<br />

<strong>of</strong> south-directed extrusion <strong>of</strong> <strong>the</strong> GHS. However, <strong>the</strong> mid- to lower-crust would have remained<br />

hot and retained low-viscosity due to its depth below <strong>the</strong> elevated Tibetan plateau (Harris, 2006).<br />

Subsequently, east-directed mid- to lower-crustal flow may have initiated due to lateral pressure<br />

gradients between <strong>the</strong> elevated Tibetan plateau and <strong>the</strong> low-lying areas south and north <strong>of</strong> <strong>the</strong><br />

Sichuan basin (Figs. 3.1, 3.2; Clark and Royden, 2000).<br />

Solving this problem is complex and may require detailed investigation and modeling <strong>of</strong><br />

tectonic and environmental changes throughout <strong>the</strong> transition period. Results <strong>of</strong> this work are<br />

critical to understanding <strong>the</strong> evolution <strong>of</strong> lower-crustal processes throughout continent-continent<br />

convergence. Fur<strong>the</strong>rmore, considering <strong>the</strong> chronologic resolution that may be required to resolve<br />

<strong>the</strong>se issues, solutions may be only be possible in modern-day mountain belts such as <strong>the</strong><br />

Himalaya-Tibetan orogen.<br />

3.2.2 Is <strong>the</strong> Karakoram fault linked to faults <strong>of</strong> <strong>the</strong> upper Karnali valley<br />

Understanding <strong>the</strong> link between strike-slip dominated faults <strong>of</strong> <strong>the</strong> upper Karnali valley<br />

and larger-scale fault systems is fundamental to determining <strong>the</strong> role <strong>of</strong> strike-slip faulting during<br />

continent-continent convergence. The following two contrasting interpretations exist: (1)<br />

deformation is localized along very few large-scale faults that accommodate considerable slip<br />

(e.g. Tapponnier et al., 1982; Armijo et al., 1989; Replumaz and Tapponnier, 2003), and (2)<br />

deformation is distributed amongst many small faults, that each accommodate minor amounts <strong>of</strong><br />

slip (e.g. Ro<strong>the</strong>ry and Drury, 1984; Taylor et al., 2003).<br />

104


Determining <strong>the</strong> relationship between faults <strong>of</strong> <strong>the</strong> upper Karnali valley and <strong>the</strong><br />

Karakoram fault will provide additional support for one <strong>of</strong> <strong>the</strong>se interpretations. Data from this<br />

<strong>the</strong>sis, and from Yakymchuk (2010), suggest that faults <strong>of</strong> <strong>the</strong> Karnali valley young towards <strong>the</strong><br />

sou<strong>the</strong>ast. Murphy and Copeland (2005) suggest that <strong>the</strong> faults <strong>of</strong> <strong>the</strong> Karnali are kinematically<br />

and chronologically linked to <strong>the</strong> crustal-scale Karakoram fault.<br />

Thermochronologic data from a wider area, coupled with micro-seismicity and slip<br />

estimates <strong>of</strong> active faults, will provide a basis for evaluating and comparing <strong>the</strong> kinematic,<br />

chronologic and present-day evolution <strong>of</strong> <strong>the</strong> Karnali valley and <strong>the</strong> Karakoram fault.<br />

105


Chapter 4<br />

Microstructural analyses: proposals for advancing and validating<br />

current techniques<br />

4.1 Introduction<br />

Quartz crystallographic preferred orientation (CPO) and rigid clast vorticity analyses are<br />

microstructural techniques used to enhance our understanding <strong>of</strong> flow kinematics and ductile<br />

deformation (Passchier and Trouw, 2005). Both styles <strong>of</strong> analyses were performed with variable<br />

degrees <strong>of</strong> success on samples from <strong>the</strong> upper Karnali valley (Chapter 2.5).<br />

Quartz CPO analyses from <strong>the</strong> upper Karnali valley were performed by classic<br />

(universal-stage) and modern methods (electron backscatter diffraction - EBSD). Both<br />

methodologies yielded meaningful results (Chapter 2.5.1), however, <strong>the</strong> EBSD method <strong>of</strong>fered<br />

additional potential, including <strong>the</strong> analysis <strong>of</strong> microscale domains and quartz a-axes. In contrast<br />

to CPO analyses, only a single rigid clast vorticity analysis from <strong>the</strong> upper Karnali valley yielded<br />

a meaningful result (Chapter 2.5.2).<br />

This chapter outlines two studies made possible with modern CPO acquisition s<strong>of</strong>tware,<br />

and a third study evaluating <strong>the</strong> practical application <strong>of</strong> rigid clast vorticity analyses.<br />

4.2 Modern quartz CPO analyses: future research opportunities<br />

Quartz crystallographic preferred orientations (CPOs) are <strong>the</strong> result <strong>of</strong> deformation<br />

temperature, strain rate, non-coaxiality <strong>of</strong> flow, and distortional strain geometry (Sullivan and<br />

Beane, 2010, and references <strong>the</strong>rein). Thus, <strong>the</strong> analysis <strong>of</strong> quartz CPOs has become a powerful<br />

tool in characterizing high-strain zones and exhumed metamorphic terranes, particularly within<br />

<strong>the</strong> Greater Himalayan Sequence and its bounding faults (e.g. Bouchez and Pêcher, 1976;<br />

106


Bouchez, 1977; Brunel and Kienast, 1986; Grujic et al., 1996; Bhattacharya and Weber, 2004;<br />

Law et al., 2004; Larson and Godin, 2009; Larson et al., 2010a; Yakymchuk and Godin, 2012).<br />

Historically, analysis <strong>of</strong> quartz CPO fabrics was limited to <strong>the</strong> c-axis due to analytical<br />

restrictions <strong>of</strong> traditional methods (e.g. universal stage microscope). However, <strong>the</strong> development<br />

<strong>of</strong> <strong>the</strong> Electron Back Scatter Diffraction (EBSD) technique (Prior et al., 1999), and Fabric<br />

Analyzer (FA) method (Wilson et al., 2003; 2007) have provided alternative and improved<br />

approaches.<br />

Both methods allow for <strong>the</strong> automatic recognition <strong>of</strong> quartz c-axes. C-axis fabrics are<br />

commonly used to assess shear sense, active slip systems, deformation temperatures and<br />

deviations from plane strain towards constriction or flattening (Fig. 4.1). The EBSD method<br />

allows for <strong>the</strong> additional recognition <strong>of</strong> <strong>the</strong> complete crystallographic orientations <strong>of</strong> all lattice<br />

planes and axes. With regard to quartz, this allows for evaluation <strong>of</strong> both <strong>the</strong> c- and a- axes. A-<br />

axes provide an additional independent assessment <strong>of</strong> plane strain, flattening, and constriction<br />

(Fig. 4.1; Passchier and Trouw, 2005). Fur<strong>the</strong>rmore, Sullivan and Beane (2010) demonstrate that<br />

at high temperatures and non-coaxial flow, a-axis fabrics are more consistent and betterdeveloped<br />

than c-axis fabrics, and may provide a more accurate representation <strong>of</strong> strain geometry.<br />

Additional advantages <strong>of</strong> modern CPO analyses (EBSD and FA) include high-resolution<br />

acquisition <strong>of</strong> CPO data from rock-forming mineral grains as small as ~1 µm (EBSD) and ~2.8<br />

µm (FA), and fully automated data acquisition and analysis. This enables rapid comparison <strong>of</strong><br />

statistically significant data from different microstructural domains. For a full description <strong>of</strong><br />

EBSD and FA techniques, and <strong>the</strong>ir application to geologic problems, see Prior et al. (1999;<br />

2009) and Wilson et al. (2003; 2007), respectively.<br />

107


a<br />

c<br />

a<br />

c<br />

Figure 4.1 Summary <strong>of</strong> quartz CPO fabrics showing (a) <strong>the</strong> orientation <strong>of</strong><br />

principle strain axes, active slip systems, and fabric geometries in different<br />

strain fields, and (b) <strong>the</strong> evolution <strong>of</strong> a CPO fabric during progressive noncoaxial<br />

deformation (assuming plane strain). Note that <strong>the</strong> letters ‘c’ and ‘a’<br />

refer to <strong>the</strong> quartz crystallographic c- and a-axes. Adapted and modified after<br />

Schmid and Casey, 1986; Passchier and Trouw, 2005; Toy et al., 2008;<br />

Sullivan and Beane, 2010.<br />

108


The following section outlines two experiments that use modern quartz CPO analyses to<br />

study (1) <strong>the</strong> evolution <strong>of</strong> CPO fabrics throughout progressive strain development, and (2) lateral<br />

variations in strain during extrusion <strong>of</strong> <strong>the</strong> Greater Himalayan sequence (GHS).<br />

4.2.1 Evolution <strong>of</strong> CPO fabrics during progressive strain development<br />

The goal <strong>of</strong> this study is to document how quartz CPO fabrics evolve at <strong>the</strong> microscale<br />

during progressive deformation. An understanding <strong>of</strong> this evolution will provide a context for<br />

evaluating CPO fabrics from extruded and exhumed high strain zones, such as <strong>the</strong> Greater<br />

Himalayan sequence.<br />

Toy et al. (2008) show that c-axis fabrics <strong>of</strong> exhumed rocks are strongly influenced by<br />

strain rates at depth (Fig. 4.2). They demonstrate that fabrics formed at depth are inherited by, and<br />

directly influence, lower temperature fabric transitions (Fig. 4.2). CPOs from this study were<br />

measured from entire thin sections.<br />

Peternell et al. (2010) illustrate heterogeneities in CPO fabrics on <strong>the</strong> scale <strong>of</strong> a thin<br />

section (Fig. 4.3). They show that bulk CPO fabrics (i.e. those measured from entire thin sections)<br />

are equivalent to <strong>the</strong> addition <strong>of</strong> all microscale domains (Fig. 4.3).<br />

The proposed study builds upon <strong>the</strong> conclusions <strong>of</strong> Toy et al. (2008) and Peternell et al.<br />

(2010), by examining quartz CPOs <strong>of</strong> different microstructural domains, and entire thin sections,<br />

at various stages <strong>of</strong> exhumation. Additionally, quartz CPOs will be obtained from veins formed<br />

during <strong>the</strong>se different stages <strong>of</strong> exhumation.<br />

Results will provide insight on <strong>the</strong> evolution <strong>of</strong> quartz CPO fabrics <strong>of</strong> different<br />

microstructural domains during exhumation, and <strong>the</strong>ir relationship to bulk CPOs. Additionally,<br />

careful observation <strong>of</strong> microstructural domains could reveal, (1) CPO fabrics that may o<strong>the</strong>rwise<br />

have not been observed due to low abundance or a rapid fabric transition, and (2) controlling<br />

109


Figure 4.2 Results <strong>of</strong> study by Toy et al. (2008) showing <strong>the</strong> evolution <strong>of</strong> bulk c-axis fabrics for<br />

different exhumation paths. Numbers in bubbles correspond to total strain experienced by that<br />

sample; bold black line corresponds to <strong>the</strong> fault, and bold arrow points arrow from <strong>the</strong> fault in<br />

<strong>the</strong> direction <strong>of</strong> decreasing strain. The different shaded regions correspond to protomylonitic and<br />

ultramylonitic exhumation paths. Note that <strong>the</strong> strong Y-maxima developed at depth in <strong>the</strong><br />

ultramylonite (typically indicative <strong>of</strong> high deformation temperatures) is preserved during lower<br />

temperature fabric transitions. This is not <strong>the</strong> case for <strong>the</strong> protomylonite, as it was unable to<br />

develop a Y-maxima at high temperatures due to lower strains. Modified after Toy et al. (2008).<br />

110


Figure 4.3 Results <strong>of</strong> study by Peternell et al. (2010) showing (a) different quartz c-axis fabrics<br />

from five micro-domains within a single thin section, and (b) <strong>the</strong> addition <strong>of</strong> <strong>the</strong>se five domains<br />

yielding <strong>the</strong> final bulk c-axis fabric. Note how <strong>the</strong> asymmetric CPO from domain (ii) is barely<br />

visible on <strong>the</strong> final addition. Modified after Peternell et al. (2010).<br />

111


factors on <strong>the</strong> formation <strong>of</strong> different CPOs and slip systems.<br />

CPOs derived from veins emplaced at various stages <strong>of</strong> exhumation should record an<br />

unbiased CPO pattern, with no prior inheritance. CPOs <strong>of</strong> <strong>the</strong>se veins can be compared against<br />

CPOs <strong>of</strong> <strong>the</strong> entire thin section and different microstructural domains, to better understand how<br />

preexisting CPOs influence ongoing fabric development.<br />

Lastly, <strong>the</strong>se data will provide an evaluation <strong>of</strong> <strong>the</strong> relationship between slip systems, c-<br />

axis opening angles, and temperature <strong>of</strong> deformation (Chapter 2.5.1.3; Law, 1991; Law et al.,<br />

2004). The aforementioned studies by Toy et al. (1998), and Peternell et al. (2010) both conclude<br />

that CPO patterns did not consistently reflect changes in temperature. This study would attempt to<br />

clarify this by comparing deformation temperatures estimated from quartz c-axis opening angles,<br />

and slip systems, with externally derived temperature estimates.<br />

4.2.2 Lateral variations in strain during extrusion <strong>of</strong> <strong>the</strong> Greater Himalayan sequence<br />

The Greater Himalayan Sequence (GHS) is <strong>the</strong> high-metamorphic-grade and anatectic<br />

core <strong>of</strong> <strong>the</strong> Himalaya, commonly interpreted to represent an extruded low-viscosity mid-crustal<br />

channel (Grujic et al. 1996, 2002; Beaumont et al. 2001, 2004; Jamieson et al 2004; Searle and<br />

Szulc 2005; Law et al. 2004; Lee and Whitehouse 2007). A significant kinematic database <strong>of</strong> <strong>the</strong><br />

extruded GHS has been established, largely through analyses <strong>of</strong> quartz c-axis fabrics and vorticity<br />

(Bouchez and Pêcher, 1976; Brunel and Kienast, 1986; Grujic et al., 1996; Grasemann et al.,<br />

1999; Bhattacharya and Weber, 2004; Law et al., 2004; Carosi et al., 2006; Jessup et al., 2006,<br />

2007; Larson and Godin, 2009; Lee and Wagner, 2009; Langille et al., 2010; Larson et al.,<br />

2010a,b). However, this data set commonly assumes plane strain conditions, and does not<br />

sufficiently address lateral strain variations (i.e. deviations from plane strain towards constriction<br />

or flattening).<br />

112


Understanding deviations from plane strain is critical for two reasons: (1) contributions <strong>of</strong><br />

flattening and constriction have a direct effect on extrusion models, and (2) temperature estimates<br />

from quartz c-axis slip system and c-axis opening angles, and vorticity interpretations, rely on <strong>the</strong><br />

assumption that deformation occurs in a plane strain environment. Deviation from plane strain<br />

may negate <strong>the</strong> validity <strong>of</strong> <strong>the</strong>se methods.<br />

Several quartz CPO studies <strong>of</strong> <strong>the</strong> GHS have presented type II cross-girdle c-axis fabrics<br />

(Grujic et al., 1996; Bhattacharya and Weber, 2004; Law et al., 2004; Larson and Godin, 2009;<br />

Larson et al., 2010a,b), indicative <strong>of</strong> constrictional deformation (Fig. 4.1; Bouchez 1978; Schmid<br />

and Casey, 1986). The goal <strong>of</strong> <strong>the</strong> proposed study is to assess deviations from plane strain<br />

towards flattening or constriction throughout <strong>the</strong> GHS. Quartz c- and a-axes across and along<br />

strike will be analyzed and combined with <strong>the</strong> current kinematic database. To date, quartz a-axes<br />

are largely undocumented throughout <strong>the</strong> GHS. Obtaining a-axes may be critical, as <strong>the</strong>y provide<br />

an independent assessment <strong>of</strong> strain conditions (Passchier and Trouw, 2005; Sullivan and Beane,<br />

2010)<br />

Results from this study will reveal lateral strain variations throughout <strong>the</strong> GHS. These<br />

results will affect <strong>the</strong> validity <strong>of</strong> <strong>the</strong> plane strain assumption underlying all quartz CPO and<br />

vorticity interpretations, and influence GHS extrusion models.<br />

4.3 Evaluation <strong>of</strong> <strong>the</strong> rigid grain method for estimating kinematic vorticity<br />

numbers; future research opportunities<br />

Vorticity analyses quantify <strong>the</strong> internal rotational component <strong>of</strong> flow in order to estimate<br />

<strong>the</strong> relative contributions <strong>of</strong> pure and simple shear throughout deformation (Passchier and Trouw,<br />

2005). The concept <strong>of</strong> rigid grain vorticity analysis was initially developed by Jeffery (1922) and<br />

Ghosh and Ramberg (1976), and has since been modified for use in geologic problems (Passchier,<br />

113


1987; Simpson and De Paor, 1993, 1997; Wallis, 1992, 1995; Wallis et al., 1993; Jessup et al.,<br />

2007). It is now commonly used to characterize flow within shear zones, particularly in <strong>the</strong><br />

Himalaya (Law et al., 2004; Carosi et al., 2006; Jessup et al., 2006, 2007; Larson and Godin,<br />

2009; Lee and Wagner, 2009; Langille et al., 2010; Larson et al., 2010a,b).<br />

The basis <strong>of</strong> <strong>the</strong> rigid grain vorticity method relies on a ma<strong>the</strong>matic relationship between<br />

<strong>the</strong> aspect ratios and orientations <strong>of</strong> rigid objects, commonly porphyroclasts, rotating in a flowing<br />

matrix (Jeffery, 1922; Ghosh and Ramberg, 1976). This relationship predicts that during<br />

deformation, porphyroclasts <strong>of</strong> certain aspect ratios will rotate into a stable-sink orientation.<br />

Assuming <strong>the</strong> following assumptions <strong>of</strong> Passchier (1987) are met, this orientation should be a<br />

reflection <strong>of</strong> <strong>the</strong> contribution <strong>of</strong> pure and simple shear active throughout deformation:<br />

(1) Presence <strong>of</strong> abundant rigid pre-deformational porphyroclasts<br />

(2) Porphyroclasts are significantly larger <strong>the</strong>n host matrix grains<br />

(3) No mechanical interaction between porphyroclasts<br />

(4) Significant quantity <strong>of</strong> porphyroclasts with a wide range <strong>of</strong> aspect ratios<br />

(5) Homogenously deforming matrix<br />

(6) High and protracted strains allowing grains to rotate to stable-sink positions<br />

An additional fundamental assumption when performing vorticity analyses, is that all<br />

deformation occurs in two-dimensional flow. This evokes two major limitations. First, flow must<br />

be monoclinic (Fig. 4.4). This implies that <strong>the</strong> vorticity vector must be parallel to one <strong>of</strong> <strong>the</strong> three<br />

instantaneous stretching axes, and both <strong>of</strong> <strong>the</strong>se vectors must be perpendicular to <strong>the</strong> plane<br />

containing <strong>the</strong> best-developed asymmetric structures, as this is <strong>the</strong> plane that will be analyzed for<br />

vorticity (Fig. 4.4; Passchier, 1987). This assumption may be problematic as not all natural shear<br />

zones have monoclinic symmetry (Lin et al., 1998; Jones and Holdsworth, 1998; Jiang and<br />

Williams, 1998; Forte and Bailey, 2007; Iacopini et al., 2007; Fernandez and Diaz-Azpiroz,<br />

114


Figure 4.4 Illustration <strong>of</strong> monoclinic (instantaneous stretching axis, ISA, parallel to vorticity<br />

vector), and triclinic (ISA oblique to vorticity vector) shear zone geometries. Blue ellipses<br />

show orientation <strong>of</strong> finite strain axes, yellow arrows show ISA, green arrows show shear sense<br />

and purple rectangle shows vorticity analysis plane. Modified after Iacopini et al. (2007).<br />

115


2009). Fur<strong>the</strong>rmore, <strong>the</strong>re is rising debate on which type <strong>of</strong> flow, monoclinic or triclinic, is<br />

predominant in <strong>the</strong> majority <strong>of</strong> shear zones (Jiang and Williams, 1998; Jiang et al., 2001).<br />

Second, deformation must be plane strain. This implies that all deformation is parallel to<br />

<strong>the</strong> plane in which vorticity will be analyzed. Deviations from plane strain towards constriction or<br />

flattening are not incorporated into <strong>the</strong> rigid grain vorticity <strong>the</strong>ory. This assumption may be<br />

problematic when assessing transtensional or transpressional shear zones, which may exhibit<br />

constriction or flattening strain (see review by Dewey, 2002).<br />

Past critiques <strong>of</strong> <strong>the</strong> rigid clast vorticity method highlight systematic error in estimating<br />

vorticity due to <strong>the</strong> two-dimensional nature <strong>of</strong> <strong>the</strong> analysis (Tik<strong>of</strong>f and Fossen, 1995; Forte and<br />

Bailey 2007; Iacopini et al., 2008). Building upon <strong>the</strong>se critiques, a recent study by Li and Jiang<br />

(2011) used numerical modeling to test <strong>the</strong> reliability <strong>of</strong> <strong>the</strong> rigid grain method for evaluating<br />

vorticity. Despite respecting all stipulated assumptions, results <strong>of</strong> <strong>the</strong>ir study yield large<br />

uncertainties that call into question <strong>the</strong> applicability <strong>of</strong> this method.<br />

Uncertainties derived from <strong>the</strong> study <strong>of</strong> Li and Jiang (2011) were largely attributed to <strong>the</strong><br />

inability <strong>of</strong> randomly oriented clasts to rotate into a stable-sink position under natural amounts <strong>of</strong><br />

strain. To perform vorticity analyses, it is essential that rigid clasts rotate into a stable-sink<br />

position. Once in this position, <strong>the</strong> vorticity vector should be parallel to <strong>the</strong> intermediate axis <strong>of</strong><br />

<strong>the</strong> clast, and only <strong>the</strong>n will clast aspect ratios and orientations be relatable to vorticity (Fig. 4.4,<br />

Passchier, 1987). Li and Jiang (2011) postulate that this assumption is not valid, as <strong>the</strong>ir model<br />

predicts that <strong>the</strong> alignment <strong>of</strong> <strong>the</strong>se axes requires unnaturally large amounts <strong>of</strong> strain. These<br />

authors conclude that <strong>the</strong> current 2D models <strong>of</strong> analyzing vorticity do not simulate natural 3D<br />

conditions and <strong>the</strong>refore do not provide a valid assessment <strong>of</strong> coaxial versus non-coaxial strain.<br />

116


The following section proposes an alternative study <strong>of</strong> rigid clast vorticity based on<br />

analogue modeling. Results <strong>of</strong> this study would provide an additional evaluation <strong>of</strong> <strong>the</strong> rigid clast<br />

methods used to estimate vorticity.<br />

4.3.1 Analogue modeling <strong>of</strong> rigid clasts in a 3D flowing matrix<br />

The goal <strong>of</strong> this study is to simulate <strong>the</strong> rotation <strong>of</strong> rigid clasts in a flowing matrix under<br />

varied natural geologic conditions. Vorticity numbers estimated from <strong>the</strong> tests will be compared<br />

with known input values and will provide conclusions on <strong>the</strong> applicability <strong>of</strong> <strong>the</strong> rigid grain<br />

vorticity <strong>the</strong>ory to practical geologic problems.<br />

The basic model (Fig. 4.5) would place randomly oriented rigid blocks <strong>of</strong> known aspect<br />

ratios within a material simulating a geologically-realistic viscous rheology. This material will<br />

<strong>the</strong>n be deformed within a shear box (Fig. 4.5), at a predetermined parameter (listed below). At<br />

intervals throughout <strong>the</strong> deformation process, <strong>the</strong> block will be imaged in 2D slices normal to <strong>the</strong><br />

vorticity vector, and in 3D (using a tomodensitometer or “CT-scanner” as per Harris et al.,<br />

2012a,b). A similar study was done by Piazolo et al. (2002), however <strong>the</strong>ir analogue model was<br />

limited to 2D and was only able to simulate plane strain. Our proposed model will operate in 3D<br />

and will simulate plane strain, constriction and flattening.<br />

This experiment will be repeated with variations in <strong>the</strong> following parameters: (1) aspect<br />

ratios <strong>of</strong> rigid clasts, (2) strain rates (low to high), (3) contributions <strong>of</strong> coaxial and non-coaxial<br />

strain, (4) monoclinic and triclinic flow, (5) plane, constrictional and flattening strain, (6) variable<br />

matrix rheology properties, and (7) alternating layers <strong>of</strong> variable rheology.<br />

2D images will be used to estimate vorticity at intervals (or “slices”) across <strong>the</strong> model<br />

using <strong>the</strong> Rigid Grain Net method (Jessup et al., 2007). These results will show under what<br />

geologic conditions <strong>the</strong> 2D rigid grain method can accurately estimate <strong>the</strong> kinematic vorticity<br />

117


Figure 4.5 Simplified set up <strong>of</strong> analogue model showing (a) randomly<br />

oriented rigid clasts (purple) within a material <strong>of</strong> geologically realistic<br />

rheology (pink), and (b), <strong>the</strong> material from (a) within a shear box. Blue<br />

arrows correspond to shear direction. Simultaneous shearing and<br />

widening/narrowing <strong>of</strong> <strong>the</strong> box parallel to <strong>the</strong> black arrows would<br />

simulate constriction/flattening (i.e. transtension/transpression).<br />

118


number. Additionally, comparing 2D slices across <strong>the</strong> model will show <strong>the</strong> consistency <strong>of</strong> repeat<br />

rigid grain analyses within a single sample. Ideally, statistical analyses <strong>of</strong> results will<br />

establishment <strong>of</strong> a minimum threshold number <strong>of</strong> grains required to be measured in order to<br />

accurately estimate vorticity.<br />

2D and 3D images will be used to assess <strong>the</strong> time required at established strain rates to<br />

bring an axis <strong>of</strong> all rigid clasts parallel to <strong>the</strong> vorticity vector (i.e. into stable-sink position). These<br />

values can be compared to geologic time and natural strain rates to verify <strong>the</strong> assumption that<br />

under natural conditions, rigid clasts will orient <strong>the</strong>mselves in a stable-sink position.<br />

Ultimately, <strong>the</strong> study will <strong>of</strong>fer an analogue modeling based evaluation <strong>of</strong> <strong>the</strong><br />

applicability <strong>of</strong> <strong>the</strong> rigid grain method in estimating vorticity. In combination with prior critiques<br />

<strong>of</strong> vorticity analyses (Tik<strong>of</strong>f and Fossen, 1995; Forte and Bailey 2007; Iacopini et al., 2008; Li<br />

and Jiang, 2011), users <strong>of</strong> <strong>the</strong> method can <strong>the</strong>n determine if it is still an appropriate tool to<br />

characterize <strong>the</strong> kinematics <strong>of</strong> flow.<br />

4.4 Conclusions<br />

Quartz CPO and rigid grain vorticity techniques allow for <strong>the</strong> analysis and extraction <strong>of</strong><br />

kinematic data from ductilly deformed rocks. Notably, quartz CPO analyses have <strong>the</strong> potential for<br />

innovative studies and for resolving kinematic data on a microstructural scale.<br />

Despite <strong>the</strong> usefulness <strong>of</strong> microstructural techniques, <strong>the</strong> applicability <strong>of</strong> <strong>the</strong>se<br />

approaches towards geologically realistic scenarios remains poorly understood. In particular,<br />

results derived from numerical modeling <strong>of</strong> rigid grain vorticity analyses question <strong>the</strong> validity <strong>of</strong><br />

this technique. A robust understanding <strong>of</strong> <strong>the</strong> true significance <strong>of</strong> microstructural results is<br />

required for a proper geologic interpretation.<br />

119


Appendix A<br />

Station locations and structural data<br />

* Latitude and Longitude based on <strong>the</strong> WGS84 datum<br />

** (2) refers to secondary structures<br />

Station Latitude Longitude<br />

Elevation<br />

(m)<br />

120<br />

Structure<br />

Azimuth<br />

Dip/<br />

plunge<br />

HK100 29 58 21.3 81 48 37.7 3142 foliation (dominant) 270 15<br />

fold axis 260 8<br />

axial plane 260 66<br />

HK101 29 58 21.1 81 47 59.8 3134 foliation (dominant) 281 27<br />

mineral lineation 325 3<br />

HK102 30 01 05.9 81 45 18.8 2432 foliation (dominant) 315 36<br />

crenulation lineation 115 10<br />

HK103 30 01 59.3 81 44 49.0 2431 foliation (dominant) 81 23<br />

mineral lineation 310 7<br />

HK104 30 02 14.5 81 44 34.8 2565 foliation (dominant) 87 50<br />

mineral lineation 135 40<br />

joint 191 65<br />

joint 215 67<br />

joint 210 35<br />

joint 230 50<br />

HK105 30 02 52.0 81 42 22.8 2736 foliation (dominant) 100 34<br />

mineral lineation 270 6<br />

fold axis isoclinal) 100 3<br />

HK106 30 02 56.2 81 41 35.2 2837 foliation (dominant) 120 25<br />

mineral lineation 287 4<br />

fold axis (isoclinal) 290 3<br />

HK107 30 02 52.6 81 04 01.5 3087 foliation (dominant) 272 38<br />

Comments<br />

mineral lineation 295 5 L>S tectonite<br />

minor open<br />

HK108 30 02 32.2 81 38 08.2 2931 foliation (dominant) 120 30 folds<br />

mineral lineation 284 2<br />

HK109 30 03 00.1 81 37 51.9 2941 foliation (dominant) 111 52<br />

mineral lineation 290 5<br />

HK110A 30 03 50.6 81 33 19.9 2950 foliation (dominant) 50 29<br />

HK110B<br />

30 03<br />

48.67<br />

mineral lineation 120 27<br />

81 33<br />

24.11 3001 foliation (dominant) 91 26<br />

mineral lineation 100 1<br />

HK111 30 03 41.7 81 33 47.1 3157 foliation (dominant) 74 30<br />

mineral lineation 128 5<br />

HK112 30 03 45.1 81 33 39.5 3077 foliation (dominant) 89 30<br />

HK113 30 03 39.1 81 34 29.4 2978<br />

mineral lineation 121 15<br />

HK114 30 03 26.1 81 35 20.8 2979 joint 335 84<br />

joint 340 88


Station Latitude Longitude<br />

Elevation<br />

(m)<br />

121<br />

Structure<br />

Azimuth<br />

Dip/<br />

plunge<br />

joint 95 39<br />

joint 98 44<br />

HK115 30 04 39.1 81 36 05.2 3450 foliation (dominant) 355 20<br />

mineral lineation 110 14<br />

HK116 30 05 20.3 81 36 09.3 3575 foliation (dominant) 58 33<br />

mineral lineation 80 5<br />

HK117 30 06 15.5 81 36 19.6 3950 foliation (dominant) 255 88<br />

mineral lineation 70 30<br />

HK118 30 07 42.1 81 37 48.4 4320 foliation (dominant) 90 39<br />

mineral lineation 104 36<br />

mineral lineation (2) 175 57<br />

HK119 30 07 00.2 81 37 09.0 4227 foliation (dominant) 264 49<br />

mineral lineation 40 20<br />

HK120 30 04 24.5 81 36 17.1 3438 foliation (dominant) 74 46<br />

mineral lineation 103 17<br />

HK121 30 03 04.8 81 36 54.9 3145 foliation (dominant) 54 30<br />

mineral lineation 116 28<br />

Comments<br />

HK122 30 03 07.9 81 36 59.9 3074 foliation (dominant) 160 34 cataclasite<br />

fold axis (dominant) 295 41<br />

fold axis (2) 320 64<br />

fold axis (2) 350 25<br />

joint 315 39<br />

striations 295 17<br />

HK123 30 02 35.8 81 36 39.7 3240 foliation (dominant) 145 36<br />

mineral lineation 295 7<br />

HK124 30 02 14.1 81 35 52.5 3600 foliation (dominant) 107 65<br />

foliation (dominant) 115 31<br />

mineral lineation 270 12<br />

mineral lineation 270 45<br />

fold axis (isoclinal) 270 12<br />

axial plane 107 65<br />

HK125 30 02 19.1 81 35 56.7 3589 foliation (dominant) 122 51<br />

mineral lineation 282 13<br />

HK126 30 02 17.3 81 35 45.2 3745 foliation (dominant) 122 28<br />

crenulation lineation 260 23<br />

mineral lineation (2) 250 24<br />

HK127 30 02 17.9 81 35 46.2 3744 foliation (dominant) 105 23<br />

crenulation lineation 250 18<br />

HK128 30 02 17.2 81 35 39.9 3763 foliation (dominant) 113 31<br />

crenulation lineation 138 21<br />

minearl lineation (2) 245 25<br />

HK129 30 02 09.2 81 35 08.9 3975 foliation (dominant) 115 26<br />

cross-cutting<br />

mylonite<br />

cataclastic<br />

carbonate<br />

bedding 140 40 not consistent<br />

mineral lineation 122 11<br />

HK130 30 01 56.9 81 34 55.8 4033 foliation (dominant) 94 32<br />

spaced<br />

cleavage


Station Latitude Longitude<br />

Elevation<br />

(m)<br />

122<br />

Structure<br />

Azimuth<br />

Dip/<br />

plunge<br />

fold axis 265 26<br />

axial plane 94 32<br />

HK131 30 02 22.5 81 36 12.2 3425 foliation (dominant) 117 41<br />

HK132 29 59 26.3<br />

HK133<br />

30 00<br />

05.64<br />

foliation (dominant) 266 14<br />

bedding 146 28<br />

Comments<br />

mineral lineation 287 9<br />

fold axis 282 17<br />

broad open<br />

fold<br />

81 36<br />

49.23 3916 foliation (dominant) 153 28 no GPS lock<br />

fold axis 307 15<br />

81 37<br />

00.99 3547 foliation (dominant) 146 31 no GPS lock<br />

mineral lineation 285 18<br />

fold axis 315 20<br />

HK134 30 00 29.6 81 36 54.6 3768 foliation (dominant) 242 16<br />

HK135 30 00 32.4 81 37 22.3 3713 foliation (dominant) 192 19<br />

foliation (dominant) 101 76<br />

mineral lineation 184 76<br />

HK136 30 00 35.8 81 37 27.4 3685 foliation (dominant) 84 33<br />

HK 137<br />

HK138<br />

HK139<br />

30 00<br />

37.79<br />

30 00<br />

39.53<br />

mineral lineation 120 23<br />

81 37<br />

29.13 3671 foliation (dominant) 142 44<br />

mineral lineation 305 17<br />

fold axis 140 18<br />

81 37<br />

30.87 3655 foliation (dominant) 70 45<br />

mineral lineation 120 20<br />

30 01 03.<br />

0 81 37 42.0 3585 foliation (dominant) 102 35<br />

mineral lineation 130 12<br />

HK140 30 01 19.4 81 37 59.8 3501 foliation (dominant) 140 28<br />

mineral lineation 305 5<br />

HK141 30 01 32.3 81 38 14.1 3924 foliation (dominant) 90 47<br />

HK142 30 01 36.0 81 38 21.3 3421 foliation (dominant) 110 32<br />

TSS - GHS<br />

contact<br />

mineral lineation 204 ~30 no GPS lock<br />

HK143 30 01 52.5 81 38 39.0 3330 foliation (dominant) 89 31<br />

HK144 30 01 56.5 81 38 41.8 3160 foliation (dominant) 116 5<br />

HK145<br />

30 01<br />

59.25<br />

mineral lineation 114 2<br />

crenulation lineation 240 11<br />

fold axis (isoclinal) 300 11<br />

81 38<br />

43.48 3147 foliation (dominant) 110 2<br />

HK146 30 02 18.8 81 44 15.4 2537 foliation (dominant) 8 25<br />

KH001<br />

29 58<br />

21.12<br />

mineral lineation 278 23<br />

foliation (2) 210 26<br />

minearl lineation (2) 260 15<br />

081 48<br />

38.04 foliation (dominant) 350 10<br />

cross-cutting<br />

mylonite


Station Latitude Longitude<br />

Elevation<br />

(m)<br />

123<br />

Structure<br />

Azimuth<br />

Dip/<br />

plunge<br />

KH002 29 59 12.3 81 47 45.9 foliation (dominant) 180 25<br />

KH003<br />

30 00<br />

01.0200<br />

81 47<br />

23.760 foliation (dominant) 335 13<br />

KH004<br />

30 01<br />

03.720<br />

81 45<br />

22.560 foliation (dominant) 326 28<br />

KH005<br />

KH006<br />

KH007<br />

KH008<br />

KH009<br />

30 01<br />

57.30<br />

mineral lineation 120 25<br />

joint 127 37<br />

joint 110 45<br />

Dyke 240 70<br />

81 44<br />

51.06 foliation (dominant) 138 45<br />

mineral lineation 295 20<br />

joint 40 90<br />

30 02<br />

16.62 81 44 23.7 foliation (dominant) 60 35<br />

30 02<br />

20.58<br />

30 02<br />

37.86<br />

30 03<br />

06.66<br />

KH010 30 03 55.5<br />

KH011<br />

KH012<br />

KH013<br />

KH014<br />

KH015<br />

KH016<br />

KH017<br />

KH018<br />

30 03<br />

22.98<br />

30 03<br />

35.34<br />

30 03<br />

49.44<br />

30 05<br />

19.76<br />

30 05<br />

13.80<br />

30 05<br />

07.88<br />

30 05<br />

02.29<br />

30 04<br />

58.35<br />

Comments<br />

Also named<br />

HK004<br />

mineral lineation 108 30<br />

81 38<br />

51.18 foliation (dominant) 255 12<br />

mineral lineation 285 9<br />

81 38<br />

16.25 foliation (dominant) 315 24<br />

mineral lineation 117 18<br />

81 37<br />

German<br />

45.48 foliation (dominant) 110 74 granite<br />

mineral lineation 290 13<br />

81 36<br />

06.000 foliation (dominant) 59 27<br />

mineral lineation 108 22<br />

81 35<br />

13.62 foliation (dominant) 106 54<br />

mineral lineation 125 20<br />

81 34<br />

30.12 foliation (dominant) 65 70<br />

mineral lineation 105 44<br />

81 33<br />

20.04 foliation (dominant) 25 25<br />

mineral lineation 115 20<br />

81 31<br />

59.61 foliation (dominant) 140 55 no GPS lock<br />

mineral lineation 290 32<br />

joint 0 30<br />

81 31<br />

44.27 foliation (dominant) 140 25 no GPS lock<br />

81 31<br />

31.54 foliation (dominant) 100 43 no GPS lock<br />

mineral lineation 110 25<br />

joint 165 90<br />

81 31<br />

30.75 foliation (dominant) 107 55 pseudotachylite<br />

minor open<br />

folds<br />

fold axis 145 40<br />

81 31<br />

30.79 3058 foliation (dominant) 90 52<br />

fold axis 52 63<br />

fold axis 40 45


Station Latitude Longitude<br />

KH019<br />

KH020<br />

KH021<br />

KH022<br />

KH023<br />

KH024<br />

30 04<br />

29.16<br />

30 04<br />

29.57<br />

30 04<br />

32.40<br />

Elevation<br />

(m)<br />

124<br />

Structure<br />

Azimuth<br />

Dip/<br />

plunge<br />

minearl lineation (2) 240 10<br />

81 30<br />

00.36 3338<br />

81 29<br />

59.51 foliation (dominant) 132 37<br />

joint 350 55<br />

81 30<br />

03.96 3330 foliation (dominant) 127 38<br />

bedding 128 15<br />

fold axis 133 1<br />

axial plane 127 38<br />

joint 15 60<br />

30 04 81 30<br />

40.920 11.580 foliation (dominant) 116 33<br />

30 04 81 30<br />

46.82 14.93 3282 foliation (dominant) 110 32<br />

30 04<br />

52.92 81 30 30.3<br />

30 05 81 30<br />

01.86 19.98 foliation (dominant) 115 34<br />

Comments<br />

broad NE<br />

verging fold<br />

N verging<br />

synform<br />

N verging<br />

synform<br />

N verging<br />

synform<br />

transposed<br />

bedding<br />

transposed<br />

bedding<br />

KH025<br />

axial plane 115 34 isoclinal folds<br />

30 04 81 29<br />

KH026 53.69 45.17 foliation (dominant) 130 35<br />

30 04 81 29<br />

KH027 46.980 39.24 4015 foliation (dominant) 120 25<br />

30 04 81 29<br />

KH028 42.84 30.000 4126 foliation (dominant) 280 32<br />

30 04 81 28<br />

KH029 41.28 51.54 4300 fold axis 150 3<br />

axial plane 135 34<br />

30 04 81 28<br />

KH030 29.82 52.56 foliation (dominant) 110 35<br />

30 04 81 28<br />

transposed<br />

KH031 28.980 17.94 4555 foliation (dominant) 110 13 bedding<br />

30 04 81 28<br />

KH032 27.54 13.62 4640 foliation (dominant) 125 26<br />

mineral lineation 280 16<br />

KH033 30 04 28.8 81 28 09.0 4707 foliation (dominant) 150 36<br />

bedding 147 24<br />

mineral lineation 275 25<br />

30 04<br />

KH034 29.04 81 28 06.9 4765 foliation (dominant) 155 40<br />

30 03 81 33<br />

KH035 51.12 09.12 3004 foliation (dominant) 87 45<br />

KH036<br />

KH037<br />

KH039<br />

30 03<br />

56.52<br />

30 03<br />

06.72<br />

30 03<br />

21.03<br />

mineral lineation 115 33<br />

81 33<br />

47.76 3038<br />

81 36<br />

58.98 3077 fold axis 70 30 cataclasite<br />

fold axis 165 24<br />

foliation (dominant) 80 30<br />

81 37<br />

06.75 2928 foliation (dominant) 98 56<br />

mineral lineation 110 20


Appendix B<br />

Quartz CPO data<br />

*For EBSD data, non-processed refers to <strong>the</strong> raw data, and processed refers to data that has been<br />

reduced whereby each data point represents an individual grain (see Chapter 2.5.1.1).<br />

** Additionally for EBSD data, represents <strong>the</strong> axis, and , and {11-20}<br />

represent <strong>the</strong> axis.<br />

HK 140 – U-Stage data<br />

125


HK 140 – Non-processed EBSD data<br />

126


HK 140 – Processed EBSD data (e.g. 1 point per grain)<br />

127


HK 131A – U-Stage data<br />

128


HK 131A (continued)<br />

129


HK 131A – Non-processed EBSD data<br />

130


131


HK 139D – Non-processed EBSD data<br />

132


HK 139D – Non-processed EBSD data<br />

133


HK 105A – U-Stage data<br />

134


HK 105A – Non-processed EBSD data<br />

135


HK 105A – Processed EBSD data (e.g. 1 point per grain)<br />

136


HK 117A – Non-processed EBSD data<br />

137


HK 117A – Processed EBSD data (e.g. 1 point per grain)<br />

138


HK 120A – Non-processed EBSD data<br />

139


HK 120A – Processed EBSD data (e.g. 1 point per grain)<br />

140


HK 125 – Non-processed EBSD data<br />

141


HK 125 – Processed EBSD data (e.g. 1 point per grain)<br />

142


HK 145 – Non-processed EBSD data<br />

143


HK 145 – Processed EBSD data (e.g. 1 point per grain)<br />

144


HK 123 – Non-processed EBSD data<br />

145


HK 123 – Processed EBSD data (e.g. 1 point per grain)<br />

146


HK 123 – Processed EBSD data (e.g. 1 point per grain)<br />

147


HK 124B2 – Processed EBSD data (e.g. 1 point per grain)<br />

148


Appendix C<br />

Monazite images and elemental maps<br />

* Location <strong>of</strong> monazite grains for each thin section seen on thin section image.<br />

** Numbers on x-ray maps refer to U-Th-Pb spot analyses.<br />

***SL and CP (or COMPO) refer to secondary electron and backscatter electron images,<br />

respectively.<br />

KH18a – Thin section<br />

149


KH18a – Monazite 1<br />

150


151


KH18a – Monazite 2<br />

152


5<br />

6 18<br />

4<br />

19<br />

3<br />

13 16<br />

2<br />

1<br />

12<br />

15<br />

11 14 10<br />

7 8<br />

9<br />

17<br />

20<br />

153


KH18a – Monazite 3<br />

154


155


KH18a – Monazite 4<br />

156


5<br />

10<br />

4<br />

3<br />

9<br />

11 12<br />

2<br />

6 7 8<br />

1<br />

13 15<br />

14<br />

157


KH18a – Monazite 5<br />

158


1 2 5<br />

3 4<br />

6<br />

7<br />

10<br />

9<br />

8<br />

159


KH18a – Monazite 6<br />

160


1<br />

13 14 8 7<br />

12<br />

9 6<br />

11 15 10<br />

5<br />

4<br />

2 3<br />

161


KH18a – Monazite 7<br />

162


10<br />

9<br />

8<br />

1 7<br />

2<br />

3 4 5 6<br />

13<br />

11<br />

12<br />

15<br />

14<br />

163


KH18a – Monazite 8<br />

164


9 8<br />

20 19 7 5 3 1<br />

2<br />

10 6 4<br />

16<br />

15 17<br />

14<br />

12<br />

13 18<br />

11<br />

165


HK109 – Thin section<br />

* For HK109_monazites, scale bars for all x-ray maps are 10 um (seen in top left or<br />

bottom left <strong>of</strong> image).<br />

166


HK109 – Monazite 2<br />

167


U<br />

Th<br />

4<br />

3<br />

2<br />

1<br />

5<br />

10<br />

9<br />

11<br />

7<br />

8<br />

6<br />

Nd<br />

Y<br />

168


CP<br />

SL<br />

169


HK109 – Monazite 3<br />

170


1<br />

2<br />

3<br />

4<br />

5 6<br />

Y<br />

Th<br />

Nd<br />

171


U<br />

SL<br />

CP<br />

172


HK109 – Monazite 4<br />

173


U<br />

Th<br />

SL<br />

174


Nd<br />

CP<br />

5<br />

4<br />

6<br />

7<br />

8<br />

9<br />

3<br />

2<br />

10<br />

1<br />

Y<br />

175


HK109 – Monazite 5<br />

176


5<br />

3 4<br />

7<br />

8<br />

9<br />

6<br />

10<br />

1 2<br />

Y<br />

Th<br />

SL<br />

177


U<br />

Nd<br />

CP<br />

178


HK109 – Monazite 6<br />

179


U<br />

Nd<br />

Th<br />

180


SL<br />

CP<br />

1<br />

2<br />

3 4<br />

5<br />

9<br />

6 7 8<br />

10<br />

Y<br />

181


HK109 – Monazite 7<br />

182


CP<br />

SL<br />

183


Y<br />

Nd<br />

U<br />

Th<br />

184


HK109 – Monazite 8<br />

185


U<br />

Th<br />

SL<br />

186


P<br />

Nd<br />

Y<br />

CP<br />

187


HK109 – Monazite 9<br />

188


U<br />

Th<br />

SL<br />

189


CP<br />

Nd<br />

2<br />

3 4<br />

6<br />

5<br />

1<br />

Y<br />

P<br />

190


HK109 – Monazite 10<br />

191


U<br />

Th<br />

SL<br />

192


CP<br />

Nd<br />

4<br />

6<br />

5<br />

7<br />

2<br />

3<br />

1<br />

Y<br />

P<br />

193


HK109 – Monazite 11<br />

194


U<br />

Th<br />

SL<br />

195


CP<br />

Nd<br />

Y<br />

P<br />

196


HK109 – Monazite 12<br />

197


U<br />

Th<br />

SL<br />

198


CP<br />

P<br />

Y<br />

Nd<br />

199


HK117a – Thin section<br />

200


HK117a – Monazite 1<br />

201


11<br />

12<br />

10<br />

8<br />

9<br />

13 14 15<br />

7<br />

6<br />

4<br />

2<br />

5<br />

3<br />

1<br />

202


HK117a – Monazite 2<br />

203


4<br />

1 5 9<br />

10<br />

2<br />

11<br />

6 16<br />

3<br />

7 17<br />

14<br />

8<br />

18<br />

19<br />

12<br />

13<br />

15<br />

204


HK117a – Monazite 3<br />

205


9<br />

8<br />

7<br />

10<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

206


HK117a – Monazite 4<br />

207


7<br />

6<br />

8<br />

9<br />

13<br />

14<br />

15<br />

4<br />

5<br />

1<br />

2<br />

3<br />

10<br />

11 12<br />

208


HK117a – Monazite 5<br />

209


210


HK117a – Monazite 6<br />

211


5<br />

4<br />

9<br />

6<br />

7<br />

8<br />

3<br />

2<br />

1<br />

10<br />

212


HK117a – Monazite 7<br />

213


2<br />

3<br />

4<br />

1<br />

6<br />

7<br />

8 9<br />

10<br />

5<br />

214


HK117a – Monazite 8<br />

215


216


HK117a – Monazite 9<br />

217


6<br />

5<br />

12<br />

11 10<br />

9<br />

8<br />

7<br />

1<br />

2 3 4<br />

218


Appendix D<br />

Muscovite 40 Ar/ 39 Ar <strong>the</strong>rmochronology<br />

219


220


221


222


223


224


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


References<br />

Aikman, A.B., Harrison, T.M., and Lin, D. 2008. Evidence for Early (>44 Ma)<br />

Himalayan Crustal Thickening, Tethyan Himalaya, sou<strong>the</strong>astern Tibet. Earth and<br />

Planetary Science Letters, 274: 14-23.<br />

Aleinik<strong>of</strong>f, J.N., Schenck, W.S., Plank, M.O., Srogi, L., Fanning, C.M., Kamo, S.L., and<br />

Bosbyshell, H. 2006. Deciphering igneous and metamorphic events in high-grade rocks<br />

<strong>of</strong> <strong>the</strong> Wilmington Complex, Delaware: Morphology, cathodoluminescence and<br />

backscattered electron zoning, and SHRIMP U-Pb geochronology <strong>of</strong> zircon and<br />

monazite. Geological Society <strong>of</strong> America Bulletin. 118: 39–64.<br />

Allègre, C. J., Courtillot, V., Tapponnier, P., Hirn, A., Mattauer, M., Coulon, C., Jaeger, J. J.,<br />

Achache, J., Scharer, U., Marcoux, J., Burg, J. P., Girardeau, J., Armijo, R., Gariepy, C.,<br />

Gopel, C., Tindong, Li, Xuchang, Xiao, Chenfa, Chang, Guangqin, Li, Baoyu, Lin,<br />

Jiwen, Teng, Naiwen, Wang, Guoming, Chen, Tonglin, Han, Xibin, Wang, Wanming,<br />

Den, Huaibin, Sheng, Yougong, Cao, Ji, Zhou, Hongrong, Qiu, Peisheng, Bao, Songchan,<br />

Wang, Bixiang, Wang, Yaoxiu, Zhou, and Xu, Ronghua. 1984. Structure and evolution <strong>of</strong><br />

<strong>the</strong> Himalaya-Tibet orogenic belt. Nature, 307: 17-22.<br />

Allmendinger, R.W., Reilinger, R., and Loveless J. 2007. Strain and rotation rate from GPS in<br />

Tibet, Anatolia, and <strong>the</strong> Altiplano. Tectonics, 26: TC3013.<br />

Amelin, Y. and Zaitsev, A. 2002. Precise geochronology <strong>of</strong> phoscorites and carbonatites: <strong>the</strong><br />

critical role <strong>of</strong> U-series disequilibrium in age interpretations. Geochimica et<br />

Cosmochimica Acta, 66: 2399–2419.<br />

Antolín, B., Schill, E., Grujic, D., Baule, S., Quidelleur, X., Appel, E., and Waldhör, M. 2012.<br />

E–W extension and block rotation <strong>of</strong> <strong>the</strong> sou<strong>the</strong>astern Tibet: Unravelling late deformation<br />

stages in <strong>the</strong> eastern Himalayas (NW Bhutan) by means <strong>of</strong> pyrrhotite remanences,<br />

Journal <strong>of</strong> Structural <strong>Geology</strong>, 42: 19–33.<br />

Armijo, R., Tapponnier, P., Mercier, J., and Tong-Lin, H. 1986. Quaternary extension in sou<strong>the</strong>rn<br />

Tibet: Field observations and tectonic implications. Journal <strong>of</strong> Geophysical Research. 91:<br />

13,803–13,872.<br />

Armijo, R., Tapponnier, P., and Han, T. 1989. Late Cenozoic right-lateral strike-slip faulting in<br />

sou<strong>the</strong>rn Tibet. Journal <strong>of</strong> Geophysical Research, 94: 2787–2838.<br />

Bailey, C.M., Francis, B.E. and Fahrney, E.E. 2004. Strain and vorticity analysis <strong>of</strong><br />

transpressional high-strain zones from <strong>the</strong> Virginia Piedmont, USA. In: Alsop, G.I.,<br />

263


Holdsworth, R.E., McCaffrey, K.J.W. & Hand, M. (eds). Flow Processes in Faults and<br />

Shear Zones. Geological Society, London, Special Publications, 224: 249–264.<br />

Beaumont, C., Jamieson, R.A., Nguyen, M.H., and Lee, B. 2001. Himalayan tectonics explained<br />

by extrusion <strong>of</strong> a low-viscosity crustal channel coupled to focused surface denudation.<br />

Nature, 414: 738-742.<br />

Beaumont, C., Jamieson, R.A., Nguyen, M.H., and Medvedev, S. 2004. Crustal channel flows: 1.<br />

Numerical models with applications to <strong>the</strong> tectonics <strong>of</strong> <strong>the</strong> Himalayan-Tibetan orogen.<br />

Journal <strong>of</strong> Geophysical Research, 109: B06406.<br />

Beaumont, C., Nguyen, M.H., Jamieson, R.A. and Ellis, S. 2006. Crustal flow modes in large hot<br />

orogens. In: Law, R.D., Searle, M.P. & Godin, L. (eds). Channel Flow, Ductile Extrusion<br />

and Exhumation in Continental Collisional Zones. Geological Society, London, Special<br />

Publications, 268: 91–146.<br />

Bendick, R., and Bilham, R. 2001. How perfect is <strong>the</strong> Himalayan arc <strong>Geology</strong>, 29: 791–794.<br />

Bhattacharya, A.R., and Weber, K. 2004. Fabric development during shear deformation in <strong>the</strong><br />

Main Central Thrust Zone, NW-Himalaya, India. Tectonophysics, 387: 23-46.<br />

Blumenfeld, P., Mainprice, D., and Bouchez, J.L. 1986. c-slip in quartz from subsolidus<br />

deformed granite. Tectonophysics, 127: 97-115.<br />

Boucher, J.L., 1977. Le quartz et la cinématique des zones ductiles. Ph.D. <strong>the</strong>sis, A l'U. E. R, des<br />

Sciences de la Nature de l' Universite de Nantes.<br />

Bouchez, J.L. 1978. Preferred orientations <strong>of</strong> quartz axes in some tectonites; kinematic<br />

inferences. Tectonophysics, 49: T25–T30.<br />

Bouchez, J.L., and Pêcher, A. 1976. Microstructures and quartz preferred orientations in<br />

quartzites <strong>of</strong> <strong>the</strong> Annapurna area (central Nepal), in <strong>the</strong> proximity <strong>of</strong> <strong>the</strong> Main Central<br />

thrust. Himalayan <strong>Geology</strong>, 6: 118-131.<br />

Brunel, M. and Kienast, J.R. 1986. Étude pêtro-structurale des chevauchements ductiles<br />

Himalayens sur la transversale de l’Everest-Makalu [Népal oriental]. Canadian Journal <strong>of</strong><br />

Earth Sciences, 23: 1117-1137.<br />

Burbank, D.W., Derry, L.A., and France-Lanord, C. 1993. Reduced Himalayan sediment<br />

production 8 Myr ago despite an intensified monsoon. Nature, 364: 48–50.<br />

Burchfiel, B.C., and Royden, L.H. 1985. North–south extension within <strong>the</strong> convergent Himalayan<br />

region. <strong>Geology</strong>, 13: 679–682.<br />

264


Burchfiel, B.C., Z. Chen, K. V. Hodges, Y. Liu, L. H. Royden, C. Deng, and J. Xu. 1992. The<br />

South Tibetan Detachment System, Himalayan orogen: Extension contemporaneous with<br />

and parallel to shortening in a collisional mountain belt. Special Paper. Geological<br />

Society <strong>of</strong> America, 269: 41 pp.<br />

Burchfiel, B. C., Chen, Z., Liu,Y., and Royden, L. H. 1995. Tectonics <strong>of</strong> <strong>the</strong> Longmen Shan and<br />

adjacent regions: International Geological Review, 8: 661–735.<br />

Burkhard, M. 1993. Calcite twins, <strong>the</strong>ir geometry, appearance and significance as stress-strain<br />

markers and indicators <strong>of</strong> tectonics regime: a review. Journal <strong>of</strong> Structural <strong>Geology</strong>, 15:<br />

351–368.<br />

Carosi, R., Montomoli, C., Rubatto, D., and Visonà, D. 2006. Normal-sense shear zones in <strong>the</strong><br />

core <strong>of</strong> <strong>the</strong> Higher Himalayan Crystallines (Bhutan Himalaya): evidence for extrusion.<br />

In: Law, R.D., Searle, M.P., and Godin, L. (eds). Channel Flow, Ductile Extrusion and<br />

Exhumation in Continental Collision Zones. Geological Society <strong>of</strong> London Special<br />

Publications, 268: 425-444.<br />

Carosi, R., Montomoli, C., Rubatto, D., and Visonà, D. 2010. Late-Oligocene high temperature<br />

shear zones in <strong>the</strong> core <strong>of</strong> <strong>the</strong> Higher Himalayan Crystallines (Lower Dolpo, Western<br />

Nepal). Tectonics: 29, TC4029.<br />

Cheng, J., and G. Xu. 1987. Geologic map <strong>of</strong> <strong>the</strong> Gerdake region at a scale <strong>of</strong> 1:1000000 and<br />

geologic report (in Chinese). Xizang Bureau <strong>of</strong> <strong>Geology</strong> and Mineral Resources, Lhasa,<br />

China: 363 pp.<br />

Clark, M.K., and Royden, L.H. 2000. Topographic ooze; building <strong>the</strong> eastern margin <strong>of</strong> Tibet by<br />

lower crustal flow. <strong>Geology</strong>, 28: 703–706.<br />

Clark, M.K., House, M.A., Royden, L.H., Whipple, K.X., Burchfiel, B.C., Zhang, X., and Tang,<br />

W. 2005. Late Cenozoic uplift <strong>of</strong> sou<strong>the</strong>astern Tibet. <strong>Geology</strong>, 33: 525–528.<br />

Coleman, M., and Hodges, K. 1995. Evidence for Tibetan Plateau uplift before 14-Myr Ago from<br />

a new minimum age for east-west extension. Nature, 374: 49–52.<br />

Cook, K.L., and Royden, L.H. 2008. The role <strong>of</strong> crustal strength variations in shaping orogenic<br />

plateaus, with application to Tibet. Journal <strong>of</strong> Geophysical Research, 113: B08407<br />

Copley, A. 2008, Kinematics and dynamics <strong>of</strong> <strong>the</strong> sou<strong>the</strong>astern margin <strong>of</strong> <strong>the</strong> Tibetan Plateau.<br />

Geophysical Journal International, 174: 1081–1100.<br />

Copley, A., and McKenzie, D. 2007. Models <strong>of</strong> crustal flow in <strong>the</strong> India-Asia collision zone.<br />

Geophysical Journal International, 169: 683–698.<br />

265


Cottle, J. M., Jessup, M. J., Newell, D. L., Horstwood, M.S. A., Noble, S. R., Parrish, R. R.,<br />

Waters, D. J., and Searle, M. P. 2009a. Geochronology <strong>of</strong> granulitized eclogites from <strong>the</strong><br />

Ama Drime Massif, implications for <strong>the</strong> tectonic evolution <strong>of</strong> <strong>the</strong> South Tibetan<br />

Himalaya. Tectonics, 28: TC1002.<br />

Cottle, J. M., Searle, M. P., Horstwood, M. S. A., and Waters, D. 2009b. Timing <strong>of</strong> midcrustal<br />

metamorphism,melting, and deformation in <strong>the</strong> Mount Everest region <strong>of</strong> sou<strong>the</strong>rn Tibet<br />

revealed by U(-Th)-Pb geochronology. Journal <strong>of</strong> <strong>Geology</strong>, 117: 643–664.<br />

Cottle, J.M., Waters, D.J., Riley, D., Beyssac, O., and Jessup, M.J. 2011. Metamorphic history <strong>of</strong><br />

<strong>the</strong> South Tibetan detachment system, Mt. Everest region, revealed by RSCM<br />

<strong>the</strong>rmometry and phase equilibria modeling. Journal <strong>of</strong> Metamorphic <strong>Geology</strong>, 29: 561–<br />

582.<br />

Curray, 2005. Tectonics and history <strong>of</strong> <strong>the</strong> Andaman Sea region. Journal <strong>of</strong> Asian Earth Sciences,<br />

25: 187-232.<br />

Dettman, D.L., Fang, Xiaomin, Garzione, C.N., Li, Jijun, 2003. Uplift driven climate change at<br />

12 Ma: a long record from <strong>the</strong> NE margin <strong>of</strong> <strong>the</strong> Tibetan Plateau. Earth and Planetary<br />

Science Letters, 214: 267-277.<br />

Derry, L.A., and France-Lanord, C. 1996. Neogene Himalayan wea<strong>the</strong>ring history and river<br />

87 Sr/ 86 Sr: Impact on <strong>the</strong> marine Sr record. Earth and Planetary Science Letters, 142: 59–<br />

74.<br />

Dewane, T., Stockli, D., Hager, C., Taylor, M., Ding, L., Lee, J., and Wallis, S. 2006. Timing <strong>of</strong><br />

Cenozoic E-W extension in <strong>the</strong> Tangra Yum Co-Kung Co Rift, southcentral Tibet. Eos<br />

(Transactions, American Geophysical Union), 87, no. 52, fall meeting supplement,<br />

abstract: T34C–04.<br />

Dewey, J.F. 2002. Transtension in arcs and orogens. International Geological Reviews, 44: 402–<br />

439.<br />

Eskola, P. 1949. The problem <strong>of</strong> mantled gneiss domes. Journal <strong>of</strong> <strong>the</strong> Geological Society <strong>of</strong><br />

London, 104: 461-478.<br />

England, P.C., and Houseman, G.A. 1988. The mechanics <strong>of</strong> <strong>the</strong> Tibetan Plateau. Royal Society<br />

<strong>of</strong> London Philosophical Transactions, ser. A, 326: 301–319.<br />

ETOPO2v2 Global Gridded 2-minute Database. 2006. National Geophysical Data Center,<br />

National Oceanic and Atmospheric Administration, U.S. Dept. <strong>of</strong> Commerce,<br />

http://www.ngdc.noaa.gov/mgg/global/etopo2.html.<br />

266


Everett, J.R., Russell, O.R., Staskowski, R.J., Loyd, S.P., and Malhotra, V. 1990. Regional<br />

tectonics <strong>of</strong> Myanmar (Burma) and adjacent areas (abstract) AAPG Bulletin, 74, p. 651.<br />

Fernández, C., and Díaz-Azpiroz, M. 2009. Triclinic transpression zones with inclined extrusion.<br />

Journal <strong>of</strong> Structural <strong>Geology</strong>, 31: 1255-1269.<br />

Ferrill, D. A., Morris, A. P., Evans, M. A., Burkhard, M., Groshong, R.H., and Onasch,C.M.<br />

2004. Calcite twin morphology: a low-temperature geo<strong>the</strong>rmometer. Journal <strong>of</strong> Structural<br />

<strong>Geology</strong>, 26: 1521–1530.<br />

Fielding, E., Isacks, B., Barazangi, M., and Duncan, C. 1994. How flat is Tibet <strong>Geology</strong>, 22:<br />

163–167.<br />

Forte, A.M., and Bailey, C.M. 2007. Testing <strong>the</strong> utility <strong>of</strong> <strong>the</strong> porphyroclast hyperbolic<br />

distribution method <strong>of</strong> kinematic vorticity analysis. Journal <strong>of</strong> Structural <strong>Geology</strong>, 29:<br />

983-1001.<br />

Fossen, H., and Tik<strong>of</strong>f, B. 1997. Forwardmodeling <strong>of</strong> nonsteady-state deformations and <strong>the</strong><br />

“minimum strain path.” Journal <strong>of</strong> Structural <strong>Geology</strong>, 19: 987–996.<br />

Fossen, H., and Tik<strong>of</strong>f, B. 1998. Forward modeling <strong>of</strong> non-steady-state deformations and <strong>the</strong><br />

“minimum strain path”: reply. Journal <strong>of</strong> Structural <strong>Geology</strong>, 20: 979–981.<br />

Foster, G., Vance, D., Argles, T., and Harris, N. 2002. The Tertiary collision-related <strong>the</strong>rmal<br />

history <strong>of</strong> <strong>the</strong> NW Himalaya. Journal <strong>of</strong> Metamorphic <strong>Geology</strong>, 20: 827-843.<br />

Fuchs, G. 1977. On <strong>the</strong> <strong>Geology</strong> <strong>of</strong> <strong>the</strong> Karnali and Dolpo Regions, West Nepal. Mitteilungen der<br />

Geologischen Gesellschaft in Wien, 66-67: 22- 32.<br />

Gan, W., Zhang, P., Shen, Z., Niu, Z., Wang, M., Wan, Y., -Zhou, D., and Cheng, J. 2007.<br />

Present-day crustal motion within <strong>the</strong> Tibetan Plateau inferred from GPS measurements.<br />

Journal <strong>of</strong> Geophysical Research, 112: B08416.<br />

Gansser, A. 1964. <strong>Geology</strong> <strong>of</strong> <strong>the</strong> Himalayas. Wiley, New York.<br />

Garzanti, E., 1999. Stratigraphy and sedimentary history <strong>of</strong> <strong>the</strong> Nepal Tethys Himalaya passive<br />

margin. Journal <strong>of</strong> Asian Earth Sciences 17: 805–827.<br />

Garzione, C.N., DeCelles, P.G., Hodkinson, D.G., Ojha, T.P., and Upreti, B.N. 2003. East-west<br />

extension and Miocene environmental change in <strong>the</strong> sou<strong>the</strong>rn Tibetan plateau: Thakkhola<br />

graben, central Nepal. Geological Society <strong>of</strong> America Bulletin, 115: 3–20.<br />

267


Ghosh, S.K., and Ramberg, H. 1976. Reorientation <strong>of</strong> inclusions by combination <strong>of</strong> pure shear<br />

and simple shear: Tectonophysics, 24: 1-70.<br />

Godin, L. 2003. Structural evolution <strong>of</strong> <strong>the</strong> Tethyan sedimentary sequence in <strong>the</strong> Annapurna area,<br />

central Nepal Himalaya. Journal <strong>of</strong> Asian Earth Sciences, 22: 307–328.<br />

Godin, L., Brown, R.L., Hanmer, S., and Parrish, R. 1999. Back folds in <strong>the</strong> core <strong>of</strong> <strong>the</strong><br />

Himalayan orogen: An alternative interpretation. <strong>Geology</strong>, 27: 151-154.<br />

Godin, L., Parrish, R.R., Brown, R.L., and Hodges, K.V. 2001. Crustal thickening leading to<br />

exhumation <strong>of</strong> <strong>the</strong> Himalayan Metamorphic core <strong>of</strong> central Nepal: Insight from U-Pb<br />

Geochronology and 40 Ar/ 39 Ar Thermochronology. Tectonics, 20: 729-747.<br />

Godin, L., Grujic, D., Law, R.D., and Searle, M.P. 2006a. Channel flow, ductile extrusion and<br />

exhumation in continental collision zones: an introduction. In: Law, R.D., Searle, M.P.,<br />

and Godin, L. (eds). Channel Flow, Ductile Extrusion and Exhumation in Continental<br />

Collision Zones. Geological Society <strong>of</strong> London Special Publications, 268: 1-23.<br />

Godin, L., Gleeson, T.P., Searle, M.P., Ullrich, T.D., and Parrish, R.R. 2006b. Locking <strong>of</strong> <strong>the</strong><br />

southward extrusion in favour <strong>of</strong> rapid crustal-scale buckling <strong>of</strong> <strong>the</strong> Greater Himalayan<br />

sequence, Nar valley, central Nepal. In: Law, R.D., Searle, M.P., and Godin, L. (eds).<br />

Channel Flow, Ductile Extrusion and Exhumation in Continental Collision Zones.<br />

Geological Society <strong>of</strong> London Special Publications, 268: 269-292.<br />

Grasemann, B., Fritz, H., and Vannay, J. 1999. Quantitative kinematic flow analysis from <strong>the</strong><br />

Main Central Thrust Zone (NW-Himalaya, India): implications for a decelerating strain<br />

path and <strong>the</strong> extrusion <strong>of</strong> orogenic wedges. Journal <strong>of</strong> Structural <strong>Geology</strong> 21: 837-853.<br />

Green, O. R., Searle, M.P., Corfield, R.I., and Corfield, R.M. 2008. Cretaceous-Tertiary<br />

carbonate platform evolution and <strong>the</strong> age <strong>of</strong> <strong>the</strong> India-Asia collision along <strong>the</strong> Ladakh<br />

Himalaya (Northwest India). <strong>Geology</strong>: 116: 331–353.<br />

Grujic, D. 2006. Channel flow and continental collision tectonics: an overview. In: Law, R.D.,<br />

Searle, M.P., and Godin, L. (eds). Channel Flow, Ductile Extrusion and Exhumation in<br />

Continental Collision Zones. Geological Society <strong>of</strong> London Special Publications, 268:<br />

25-37.<br />

Grujic, D., Casey, M., Davidson, C., Hollister, L.S., Kündig, R., Pavlis, T., and Schmid, S. 1996.<br />

Ductile extrusion <strong>of</strong> <strong>the</strong> Higher Himalayan Crystalline in Bhutan: evidence from quartz<br />

micr<strong>of</strong>abrics. Tectonophysics, 260: 21-43.<br />

Grujic, D., Hollister, L.S. and Parrish, R.R. 2002. Himalayan metamorphic sequence as an<br />

orogenic channel: insight from Bhutan. Earth and Planetary Science Letters, 198: 177–<br />

191.<br />

268


Guo, Z.T., Ruddiman,W.F., Hao, Q.Z.,Wu, H.B., Qiao, Y.S., Zhu, R.X., Peng, S.Z., Wei, J.J.,<br />

Yuan, B.Y., Liu, T.S. 2002. Onset <strong>of</strong> Asian desertification by 22 Myr ago inferred from<br />

loess deposits in China. Nature, 416: 159–163.<br />

Hall, R. 2002. Cenozoic geological and plate tectonic evolution <strong>of</strong> SE Asia and <strong>the</strong> SW Pacific:<br />

computer-based reconstructions, model and animations. Journal <strong>of</strong> Asian Earth Sciences,<br />

20: 353–431.<br />

Harris, L.B., Yakymchuk, C., and Godin, L. 2012a. Implications <strong>of</strong> centrifuge simulations <strong>of</strong><br />

channel flow for opening out or destruction <strong>of</strong> folds. Tectonophysics, 526–529: 67–87.<br />

Harris, L.B., Godin, L., and Yakymchuk, C. 2012b. Regional shortening followed by channel<br />

flow induced collapse: a new mechanism for “dome and keel” geometries in Neoarchaean<br />

granite-greenstone terrains. Precambrian Research, 212-213: 139–154.<br />

Harris, N. 2006. The elevation history <strong>of</strong> <strong>the</strong> Tibetan Plateau and its implications for <strong>the</strong> Asian<br />

monsoon. Palaeogeography, Palaeoclimatology, Palaeoecology, 24: 4–15.<br />

Harris, N. 2007. Channel flow and <strong>the</strong> Himalayan-Tibetan orogen: a critical review. Journal <strong>of</strong><br />

<strong>the</strong> Geological Society <strong>of</strong> London 164: 511-523.<br />

Harris, N.B.W., Caddick, M., Kosler, J., Goswami, S., Vance, D., and Tindle, A.G. 2004. The<br />

pressure-temperature-time path <strong>of</strong> migmatites from <strong>the</strong> Sikkim Himalaya. Journal <strong>of</strong><br />

Metamorphic <strong>Geology</strong>, 22: 249-264.<br />

Harrison, T.M. 2006. Did <strong>the</strong> Himalayan Crystallines extrude partially molten from beneath <strong>the</strong><br />

Tibetan Plateau In: Law, R.D., Searle, M.P., and Godin, L. (eds). Channel Flow, Ductile<br />

Extrusion and Exhumation in Continental Collision Zones. Geological Society <strong>of</strong> London<br />

Special Publications, 268: 237-254.<br />

Harrison, T.M., Célérier, J., Aikman, A.B., Hermann, J., and Heizler, M.T. 2009. Diffusion <strong>of</strong><br />

40 Ar in muscovite. Geochimica and Cosmochimica Acta, 73: 1039-1051.<br />

Hintersberger, E., Thiede, R.C., Strecker, M.R., and Hacker, B. 2010. E–W extension in <strong>the</strong><br />

Northwestern Himalaya, NW India. Geological Society <strong>of</strong> America Bulletin, 122: 1499–<br />

1515.<br />

Hodges, K.V. 2000. Tectonics <strong>of</strong> <strong>the</strong> Himalaya and sou<strong>the</strong>rn Tibet from two perspectives.<br />

Geological Society <strong>of</strong> America Bulletin, 112: 324-350.<br />

Hodges, K.V. 1991. Pressure-temperature-time paths. Annual Review <strong>of</strong> Earth and Planetary<br />

Sciences, 19: 207–236.<br />

269


Hodges, K.V., and Silverberg, D.S. 1988. Thermal evolution <strong>of</strong> <strong>the</strong> Greater Himalaya, Garhwal,<br />

India. Tectonics, 7: 583-600.<br />

Hodges, K. V., J. M. Hurtado, and K. X. Whipple. 2001. Southward extrusion <strong>of</strong> Tibetan crust<br />

and its effect on Himalayan tectonics. Tectonics, 20: 799 – 809.<br />

Horstwood, M.S.A., Foster, G.L., Parrish, R.R., Noble, S.R., and Nowell, G.M. 2003. Common-<br />

Pb corrected in situ U-Pb accessory mineral geochronology by LA-MC-ICPMS. Journal<br />

<strong>of</strong> Analytical Atomic Spectrometry, 18: 837–846.<br />

Huyghe, P., Galy, A., Mugnier, J.-L., and C. France-Lanord. 2001. Propagation <strong>of</strong> <strong>the</strong> thrust<br />

system and erosion in <strong>the</strong> Lesser Himalaya: Geochemical and sedimentological<br />

evidence. <strong>Geology</strong>, 29: 1007–1010.<br />

Iacopini, D., Passchier, C.W., Koehn, D., and Carosi, R. 2007. Fabric attractors in general<br />

triclinic flow systems and <strong>the</strong>ir application to high strain shear zones: a dynamical system<br />

approach. Journal <strong>of</strong> Structural <strong>Geology</strong>, 29: 298-317.<br />

Iacopini, D., Carosi, R., Montomoli, C., and Passchier, C. W. 2008. Strain analysis and vorticity<br />

<strong>of</strong> flow in <strong>the</strong> nor<strong>the</strong>rn Sardinian Variscan belt: recognition <strong>of</strong> a partitioned oblique<br />

deformation event. Tectonophysics, 446: 77–96.<br />

Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., and Essling, A.M. 1971. Precision<br />

measurement <strong>of</strong> half-lives and specific activities <strong>of</strong> 235 U and 238 U. Physical Review, 4:<br />

1889–1906.<br />

Jamieson, R.A., Beaumont, C., Medvedev, S., and Nguyen, M.H. 2004. Crustal channel flows: 2.<br />

Numerical models with implications for metamorphism in <strong>the</strong> Himalayan-Tibetan<br />

orogen. Journal <strong>of</strong> Geophysical Research, 109: B06407.<br />

Jeffrey, G. B. 1922. The motion <strong>of</strong> ellipsoidal particles immersed in a viscous fluid. Proceeding<br />

<strong>of</strong> <strong>the</strong> Royal Society <strong>of</strong> London, 102: 161–179.<br />

Jessup M.J., and Cottle J.M. 2010. Progression from South-Directed Extrusion to Orogen-Parallel<br />

Extension in <strong>the</strong> Sou<strong>the</strong>rn Margin <strong>of</strong> <strong>the</strong> Tibetan Plateau, Mount Everest Region, Tibet:<br />

Journal <strong>of</strong> <strong>Geology</strong>, 118: 467–486.<br />

Jessup, M.J., Law, R.D., Searle, M.P., and Hubbard, M.S. 2006. Structural evolution and vorticity<br />

<strong>of</strong> flow during extrusion and exhumation <strong>of</strong> <strong>the</strong> Greater Himalayan Slab, Mount Everest<br />

Massif, Tibet/Nepal: implications for orogen-scale flow partitioning. In: Law, R.D.,<br />

Searle, M.P., and Godin, L. (eds). Channel Flow, Ductile Extrusion and Exhumation in<br />

Continental Collision Zones. Geological Society <strong>of</strong> London Special Publications, 268:<br />

379-413.<br />

270


Jessup, M.J., Law, R.D., and Frassi, C. 2007. The Rigid Grain Net (RGN): An alternative method<br />

for estimating mean kinematic vorticity number (Wm). Journal <strong>of</strong> Structural <strong>Geology</strong> 29,<br />

411-421.<br />

Jessup, M., Newell, D., Cottle, J., Berger, A., and Spotila, J. 2008. Orogen-parallel extension and<br />

exhumation enhanced by denudation in <strong>the</strong> trans-Himalayan Arun River gorge, Ama<br />

Drime Massif, Tibet-Nepal: <strong>Geology</strong>, 36: 587–590.<br />

Jiang, D. 1998. Forward modeling <strong>of</strong> non-steady-state deformations and <strong>the</strong> “minimum strain<br />

path”: discussion. Journal <strong>of</strong> Structural <strong>Geology</strong>, 20: 975–977.<br />

Jiang, D., and Williams, P.F. 1998. High-strain zones: a unified model. Journal <strong>of</strong> Structural<br />

<strong>Geology</strong>, 20: 1105-1120.<br />

Jiang, D., Lin, S., and Williams, P.F. 2001. Deformation paths in high-strain zones, with<br />

reference to slip partitioning in transpressional plate-boundary regions. Journal <strong>of</strong><br />

Structural <strong>Geology</strong>, 23: 991-1005.<br />

Jones, R.R., and Holdsworth, R.E. 1998. Oblique simple shear in transpression zones. In:<br />

Holdsworth, R.E., Strachan, R.A., Dewey, J.F. (eds). Continental Transpressional and<br />

Transtensional Tectonics. Geological Society <strong>of</strong> London Special Publications, 135: 35-<br />

40.<br />

Kali, E., Leloup, P.H., Arnaud, N., Mahéo, G., Liu, D., Boutonnet , E., Van der Woerd, J., Liu,<br />

X., Liu-Zeng, J., and Li, H. 2010. Exhumation history <strong>of</strong> <strong>the</strong> deepest central Himalayan<br />

rocks, Ama Drime range: Key pressure-temperature-deformation-time constraints on<br />

orogenic models. Tectonics, 29: TC2014.<br />

Kellett, D.A. and Godin, L. 2009. Pre-Miocene deformation <strong>of</strong> <strong>the</strong> Himalayan superstructure,<br />

Hidden valley, central Nepal. Journal <strong>of</strong> <strong>the</strong> Geological Society <strong>of</strong> London, 166: 261-275.<br />

Kellett, D.A., Grujic, D., Warren, C., Cottle, J.M., Jamieson, R.A., and Tenzin, T. 2010.<br />

Metamorphic history <strong>of</strong> a synconvergent orogen parallel detachment: The South Tibetan<br />

detachment system, Bhutan Himalaya. Journal <strong>of</strong> Metamorphic <strong>Geology</strong>, 28: 785-808.<br />

Kellett, D.A., and Grujic, D. 2012. New insight into <strong>the</strong> South Tibetan detachment system: not a<br />

single progressive deformation. Tectonics, 31: TC2007.<br />

Klootwijk, C.T., Conaghan, P.J., and Powell, C.M. 1985. The Himalayan arc—Large-scale<br />

continental subduction, oroclinal bending and back-arc spreading. Earth and Planetary<br />

Science Letters, 75: 167–183.<br />

271


Kohn, M.J. 2008. P-T-t data from central Nepal support critical taper and repudiate large-scale<br />

channel flow <strong>of</strong> <strong>the</strong> Greater Himalayan Sequence. Geological Society <strong>of</strong> America<br />

Bulletin, 120: 259-273.<br />

Kroon, D., Steens, T., and Troelstra, S.R. 1991. Onset <strong>of</strong> monsoonal related uplift in <strong>the</strong> western<br />

Arabian sea as revealed by planktonic foraminifera. In: Prell, W.L., Niitsuma, N. et al.,<br />

(eds). Proceedings <strong>of</strong> <strong>the</strong> Ocean Drilling Program, Scientific results, College Station,<br />

Texas, Ocean Drilling Program, 117: 257–263.<br />

Kruhl, J.H. 1998. Reply: Prism- and basal-plane parallel subgrain boundaries in quartz: a<br />

microstructural geo<strong>the</strong>rmobarometer. Journal <strong>of</strong> Metamorphic <strong>Geology</strong>, 16: 142-146.<br />

Langille, J., Lee, J., Hacker, B., and Seward, G. 2010. Middle crustal ductile deformation patterns<br />

in sou<strong>the</strong>rn Tibet: Insights from vorticity studies in <strong>the</strong> Majba Dome. Journal <strong>of</strong><br />

Structural <strong>Geology</strong>, 32: 70-85.<br />

Larson, K.P. and Godin, L. 2009. Kinematics <strong>of</strong> <strong>the</strong> Greater Himalayan sequence, Dhaulagiri<br />

Himal: implications for <strong>the</strong> structural framework <strong>of</strong> central Nepal. Journal <strong>of</strong> <strong>the</strong><br />

Geological Society, London, 166: 25-43.<br />

Larson, K.P., Godin, L., and Price, R.A. 2010a. Relationships between displacement and<br />

distortion in orogens: Linking <strong>the</strong> Himalayan foreland and hinterland in central Nepal.<br />

GSA Bulletin, 122: 1116-1134.<br />

Larson, K.P., Godin, L., Davis, W.J., and Davis, D.W. 2010b. Out-<strong>of</strong>-sequence deformation and<br />

expansion <strong>of</strong> <strong>the</strong> Himalayan orogenic wedge: insight from <strong>the</strong> Changgo culmination,<br />

south-central Tibet. Tectonics, 29: TC4013.<br />

Law, R.D. 1991. Crystallographic fabrics: a selective review <strong>of</strong> <strong>the</strong>ir applications to research in<br />

structural geology. In: Knipe, R.J., and Rutter, E.H. (eds). Deformation Mechanisms,<br />

Rheology, and Tectonics. Geological Society <strong>of</strong> London Special Publications, 54: 335-<br />

352.<br />

Law, R., Searle, M.P., and Simpson, R.L. 2004. Strain, deformation temperatures and vorticity <strong>of</strong><br />

flow at <strong>the</strong> top <strong>of</strong> <strong>the</strong> Greater Himalayan Slab, Everest Massif, Tibet. Journal <strong>of</strong> <strong>the</strong><br />

Geological Society, London, 161: 305-320.<br />

Law, R.D., Jessup, M.J., Searle, M.P., Francsis, M.K., Waters, D.J. and Cottle,<br />

J.M., 2011. Telescoping <strong>of</strong> iso<strong>the</strong>rms beneath <strong>the</strong> South Tibetan Detachment System,<br />

Mount Everest Massif. Journal <strong>of</strong> Structural <strong>Geology</strong>, 33: 1569–1594.<br />

Lee, J., and Whitehouse, M.J. 2007. Onset <strong>of</strong> mid-crustal extensional flow in sou<strong>the</strong>rn Tibet:<br />

Evidence from U/Pb zircon ages. <strong>Geology</strong>, 35: 45–48.<br />

272


Lee, J., and Wagner, T. 2009. Kinematics and vorticity in Kangmar Dome: Testing patterns <strong>of</strong><br />

mid-crustal ductile deformation during <strong>the</strong> Himalayan orogeny. AGU Fall Meeting 2009,<br />

abstract #TC43C-2124.<br />

Lee, J., Hager, C., Wallis, S.R., Stockli, D.F., Whitehouse, M.J., Aoya, M., and Wang, Y. 2011.<br />

Middle to late Miocene extremely rapid exhumation and <strong>the</strong>rmal equilibriation in <strong>the</strong><br />

Kung Co rift, sou<strong>the</strong>rn Tibet. Tectonics, 30: TC002745.\<br />

Leech, M.L. 2008. Does <strong>the</strong> Karakoram fault interrupt mid-crustal channel flow in <strong>the</strong> western<br />

Himalaya Earth and Planetary Science Letters, 276: 314–322.<br />

Li, D., and Yin, A. 2008. Orogen-parallel, active left-slip faults in <strong>the</strong> Eastern Himalaya:<br />

Implications for <strong>the</strong> growth mechanism <strong>of</strong> <strong>the</strong> Himalayan Arc. Earth and Planetary<br />

Science Letters, 274: 258–267.<br />

Li, C., and Jiang, D. 2011. A critique <strong>of</strong> vorticity analysis using rigid clasts. Journal <strong>of</strong> Structural<br />

<strong>Geology</strong>, 33: 203–219.<br />

Lin, S., Jiang, D., and Williams, P.F. 1998. Transpression (or transtension) zones <strong>of</strong> triclinic<br />

symmetry: natural example and <strong>the</strong>oretical modelling. In: Holdsworth, R.E., Strachan,<br />

R.A., Dewey, J.F. (eds). Continental Transpressional and Transtensional Tectonics.<br />

Geological Society <strong>of</strong> London Special Publications, 135: 41-58.<br />

Qianyu Li, Pinxian Wang, Quanhong Zhao, Lei Shao, Guangfa Zhong, Jun Tian, Xinrong Cheng,<br />

Zhimin Jian, and Xin Su. 2006. A 33 Ma lithostratigraphic record <strong>of</strong> tectonic and<br />

paleoceanographic evolution <strong>of</strong> <strong>the</strong> South China Sea. Marine <strong>Geology</strong>, 230: 217-235.<br />

Lister, G. 1977. Crossed-girdle c-axis fabrics in quartzites plastically deformed by plane strain<br />

and progressive simple shear. Tectonophysics, 30: 51-54.<br />

Ludwig, K. 2012. User’s manual for Isoplot v.3.75: a geochronological toolkit for Micros<strong>of</strong>t<br />

Excel, Berkeley, CA, Berkerly Geochronology Center Special Publication, 5: revised<br />

January 30, 2012, 75p.<br />

Mainprice, D., Bouchez, J.-L., Blumenfeld, P., and Tubia, J.M. 1986. Dominant c slip in naturally<br />

deformed quartz: Implications for drastic plastic s<strong>of</strong>tening at high temperature. <strong>Geology</strong>,<br />

14: 819-822.<br />

McCaffrey, R., and Nábelek, J. 1998. Role <strong>of</strong> oblique convergence in <strong>the</strong> active deformation <strong>of</strong><br />

<strong>the</strong> Himalayas and sou<strong>the</strong>rn Tibet plateau. <strong>Geology</strong>, 26: 691–694.<br />

McDougall, I., and Harrison, T.M. 1988. Geochronology and <strong>the</strong>rmochronology by <strong>the</strong> 40 Ar/ 39 Ar<br />

Method. Oxford <strong>University</strong> Press, Oxfordshire.<br />

273


Means, W. D., Hobbs, B. E., Lister, G. S., and Williams, P. F. 1980. Vorticity and non-coaxiality<br />

in progressive deformations. Journal <strong>of</strong> Structural <strong>Geology</strong>, 2: 371–378.<br />

Meigs, A., Burbank, D., and Beck R. 1995. Middle-late Miocene (>10 Ma) formation <strong>of</strong> <strong>the</strong> Main<br />

Boundary thrust in <strong>the</strong> western Himalaya. <strong>Geology</strong>, 23: 423–426.<br />

Mercier, J.-L., Armijo, R., Tapponier, P., Carey-Gailhardis, E. and Lin, H.L. 1987. Change from<br />

Late Tertiary compression to Quaternary extension in sou<strong>the</strong>rn Tibet during <strong>the</strong> India-<br />

Asia collision: Tectonics, 6: 275-304.<br />

Molnar, P., and Tapponnier, P. 1978, Active tectonics <strong>of</strong> Tibet. Journal <strong>of</strong> Geophysical Research,<br />

83: 5361–5375.<br />

Molnar, P., and Chen, W.P. 1983. Focal depths and fault plane solutions <strong>of</strong> earthquakes under <strong>the</strong><br />

Tibetan Plateau. Journal <strong>of</strong> Geophysical Research, 88: 1180–1196.<br />

Murphy, M. A., and A. Yin. 2003. Structural evolution and sequence <strong>of</strong> thrusting in <strong>the</strong> Tethyan<br />

fold-thrust belt. Geological Society <strong>of</strong> America Bulletin, 115: 21 – 34.<br />

Murphy, M.A. and Copeland, P. 2005. Transtensional deformation in <strong>the</strong> central Himalaya and its<br />

role in accommodating growth <strong>of</strong> <strong>the</strong> Himalayan orogen. Tectonics, 24: TC4012.<br />

Murphy, M.A., Yin, A., Kapp, P., Harrison, T.M., Manning, C.E., Ryerson, F.J., Lin, D.,and<br />

Jinghui, G. 2002. Structural evolution <strong>of</strong> <strong>the</strong> Gurla Mandhata detachment system,<br />

southwest Tibet: Implications for <strong>the</strong> eastward extent <strong>of</strong> <strong>the</strong> Karakoram fault system.<br />

Bulletin <strong>of</strong> <strong>the</strong> Geological Society <strong>of</strong> America, 114: 428-447.<br />

Murphy, M.A., Saylor, J.E., and Ling, D. 2009. Landscape evolution <strong>of</strong> southwest Tibet based on<br />

integrated paleoelevation reconstructions and structural history. Earth and Planetary<br />

Science Letters, 282: 1–9.<br />

Murphy, M.A., Sanchez, V., and Taylor, M.H. 2010. Syncollisional extension along <strong>the</strong> India-<br />

Asia suture zone, south-central Tibet: Implications for crustal deformation <strong>of</strong> Tibet. Earth<br />

and Planetary Science Letters, 290: 233–243.<br />

Najman, Y., Appel, E., Boudagher-Fadel, M., Bown, P., Carter, A., Garzanti, E., Godin, L., Han,<br />

J., Liebke, U., Oliver, G., Parrish, R., and Vezzoli, G. 2010. The timing <strong>of</strong> India-Asia<br />

collision: geological, biostratigraphic and palaeomagnetic constraints. Journal <strong>of</strong><br />

Geophysical Research, 115: B12416.<br />

Nakata, T. 1989. Active faults <strong>of</strong> <strong>the</strong> Himalaya <strong>of</strong> India and Nepal: Tectonics <strong>of</strong> <strong>the</strong> western<br />

Himalayas. In: Malinconico, L.L., Jr., and Lillie, R. (eds). Tectonics <strong>of</strong> <strong>the</strong> Western<br />

Himalayas. Geological Society <strong>of</strong> America Special Paper 232: 243–264.<br />

274


Nelson, K. D., Zhao, W., Brown, L. D., Kuo, J., Che, J., Liu, X., Klemperer, S. L., Makovsky, Y.,<br />

Meissner, R., Mechie, J., Kind, R., Wenzel, F., Ni, J., Nabelek, J., Leshou, C., Tan, H.,<br />

Wei, W., Jones, A. G., Booker, J., Unsworth, M., Kidd, W.S.F., Hauck, M., Alsdorf, D.,<br />

Ross, A., Cogan, M., Wu, C., Sandvol, E., and Edwards M. 1996. Partially molten middle<br />

crust beneath sou<strong>the</strong>rn Tibet: Syn<strong>the</strong>sis <strong>of</strong> project INDEPTH results. Science, 274: 1684–<br />

1688.<br />

Paquette, J.L., Nédélec, A., Moine, B., and Rakotondrazafy, M. 1994. U-Pb, single zircon Pbevaporation,<br />

and Sm-Nd isotopic study <strong>of</strong> a granulite domain in SE Madagascar. The<br />

Journal <strong>of</strong> <strong>Geology</strong>, 102: 523–538.<br />

Paton, C., Woodhead, J., Hellstrom, J., Hergt, J., Greig, A., and Maas, R. 2010. Improved laser<br />

ablation U–Pb zircon geochronology through robust downhole fractionation correction.<br />

Geochemistry, Geophysics and Geosystems. 11: Q0AA06.<br />

Passchier, C.W. 1987. Stable positions <strong>of</strong> rigid objects in non-coaxial flow; a study in vorticity<br />

analysis; Shear criteria in rocks. Journal <strong>of</strong> Structural <strong>Geology</strong> 9: 679-690.<br />

Passchier, C.S. and Trouw, R.A. 2005. Microtectonics, Springer-Verlag. 2nd edition. Berlin.<br />

Peternell, M., Hasalová, P., Wilson, C.J.L., Piazolo, S. and Schulmann, K. 2010. Evaluating<br />

quartz crystallographic preferred orientations and <strong>the</strong> role <strong>of</strong> deformation partitioning<br />

using EBSD and fabric analyser techniques. Journal <strong>of</strong> Structural <strong>Geology</strong>. 32, 803–817.<br />

Phillips, R.J., Parrish, R.R., and Searle, M.P. 2004. Age constraints on ductile deformation and<br />

long-term slip rates along <strong>the</strong> Karakoram fault zone, Ladakh. Earth and Planetary Science<br />

Letters, 226: 305–319.<br />

Piazolo, S., Bons, P.D., and Passchier, C.W. 2002. The influence <strong>of</strong> matrix rheology and vorticity<br />

on fabric development <strong>of</strong> populations <strong>of</strong> rigid objects during plane strain deformation.<br />

Tectonophysics, 351: 315–329.<br />

Prell, W.L., Murray, D.W., Clemens, S.C., and Anderson, D.M. 1992. Evolution and variability<br />

<strong>of</strong> <strong>the</strong> Indian Ocean summer monsoon: Evidence from <strong>the</strong> western Arabian Sea drilling<br />

program. In: Duncan, R.A. et al. (eds). Syn<strong>the</strong>sis <strong>of</strong> results from scientific drilling in <strong>the</strong><br />

Indian Ocean. American Geophysical Union Geophysical Monograph, 70: 447–469.<br />

Prince, C., Harris, N. and Vance, D. 2001. Fluid-enhanced melting during prograde<br />

metamorphism. Journal <strong>of</strong> <strong>the</strong> Geological Society, London. 158: 233–241.<br />

Prior, D.J., Boyle, A.P., Brenker, F., Cheadle, M., Day, A., Lopez, G., Peruzzo, L., Potts, G.J.,<br />

Reddy, S., Spiess, R., Timms, N.E., Trimby, P., Wheeler, J., and Zetterström, L.1999.<br />

The application <strong>of</strong> electron backscatter diffraction and orientation contrast imaging in <strong>the</strong><br />

SEM to textural problems in rocks. American Mineralogist, 84: 1741-1759.<br />

275


Prior, D.J. 2009. EBSD in <strong>the</strong> earth sciences: application, common practice and challenges. In:<br />

Schwartz, A.J., Kumar, M., Adams, B.L., Field, D.P. (eds.). Electron Backscatter<br />

Diffraction in Materials Science. Springer Science and Business Media, New York. 345-<br />

360.<br />

Quade, J., Cerling, T.E., and Bowman, J.R. 1989. Development <strong>of</strong> Asian monsoon revealed by<br />

marked ecological shift during <strong>the</strong> latest Miocene in nor<strong>the</strong>rn Pakistan. Nature, 342: 163–<br />

166.<br />

Quade, J., Cater, J.M.L., Ojha, T.P., Adam, J., and Harrison, T.M.. 1995. Late Miocene<br />

environmental change in Nepal and <strong>the</strong> nor<strong>the</strong>rn Indian subcontinent: Stable isotopic<br />

evidence from paleosols. Geological Society <strong>of</strong> America Bulletin, 107: 1381-1397.<br />

Renne, P.R., Swisher, C.C., Deino, A.L., Karner, D.B., Owens, T.L., and DePaolo, D.J. 1998.<br />

Intercalibration <strong>of</strong> standards, absolute ages and uncertainties in 40Ar/39Ar dating.<br />

Chemical <strong>Geology</strong>, 145: 117-152.<br />

Replumaz, A., and P. Tapponnier. 2003. Reconstruction <strong>of</strong> <strong>the</strong> deformed collision zone Between<br />

India and Asia by backward motion <strong>of</strong> lithospheric blocks. Journal <strong>of</strong> Geophysical<br />

Research, 108: 2285.<br />

Ratschbacher, L., Frisch, W., Liu, G.H., and Chen, C.S. 1994. Distributed deformation in<br />

sou<strong>the</strong>rn and western Tibet during and after <strong>the</strong> India-Asia collision. Journal <strong>of</strong><br />

Geophysical Research—Solid Earth, 99: 19,917–19,945.<br />

Robinson, D.M., DeCelles, P.G., and Copeland, P. 2006. Tectonic evolution <strong>of</strong> <strong>the</strong> Himalayan<br />

thrust belt in western Nepal: Implications for channel flow models. Geological Society<br />

<strong>of</strong> America Bulletin, 118: 865-885.<br />

Roddick, J.C. 1983. High precision intercalibration <strong>of</strong> 40 Ar/ 39 Ar standards. Geochimica et<br />

Cosmochimica Acta, 47: 887-898.<br />

Ro<strong>the</strong>ry, D.A., and Drury, S.A. 1984. The neotectonics <strong>of</strong> <strong>the</strong> Tibetan Plateau. Tectonics, 3: 19-<br />

26.<br />

Rowley, D.B. 1996. Age <strong>of</strong> initiation <strong>of</strong> collision between India and Asia: A review <strong>of</strong><br />

stratigraphic data. Earth and Planetary Science Letters, 145: 1-13.<br />

Rowley, D.B. 1998. Minimum age <strong>of</strong> initiation <strong>of</strong> collision between India and Asia north <strong>of</strong><br />

Everest based on <strong>the</strong> subsidence history <strong>of</strong> <strong>the</strong> Zhepure Mountain section. Journal <strong>of</strong><br />

<strong>Geology</strong>, 106, 1-13.<br />

Rowley, D.B., and Currie, B.S. 2006. Palaeo-altimetry <strong>of</strong> <strong>the</strong> late Eocene to Miocene Lunpola<br />

basin, central Tibet. Nature, 439: 677-681.<br />

276


Royden, L.H., Burchfiel, B.C., King, R.W., Wang, E., Chen, Z., Shen, F., and Liu, Y. 1997.<br />

Surface deformation and lower crustal flow in eastern Tibet. Science, 276: 788-790.<br />

Sandeman, H.A., Archibald, D.A., Grant, J.W., Villeneuve, M.E., and Ford, F.D. 1999.<br />

Characterization <strong>of</strong> <strong>the</strong> chemical composition and 40 Ar- 39 Ar systematics <strong>of</strong> <strong>the</strong><br />

interlaboratory standard MAC-83 biotite. In: Age and isotopic studies, Report 12,<br />

Geological Survey <strong>of</strong> Canada, Current Research 1999-F:13-26.<br />

Schelling, D.D. 1992. The tectonostratigraphy and structure <strong>of</strong> <strong>the</strong> eastern Nepal Himalaya.<br />

Tectonics, 11: 925–943.<br />

Schmid, S.M. 1994. Textures <strong>of</strong> geological materials: computer model predictions versus<br />

empirical interpretations based on rock deformation experiments and field studies. In:<br />

Bunge, H.J., Siegesmund, S., Skrotzki, W., Weber, K. (eds). Textures <strong>of</strong> Geological<br />

Materials. DGM Informations-gesellschaft, Oberursel, 279–302.<br />

Schmid, S.M. and Casey, M. 1986. Complete fabric analysis <strong>of</strong> some commonly observed quartz<br />

c-axis pattterns. In: Hobbs, B.E. & Heard, H.C. (eds). Mineral and Rock Deformation:<br />

Laboratory Studies—The Paterson Volume. Geophysical Monograph, American<br />

Geophysical Union, 36: 263–386.<br />

Searle, M.P., and Szulc, A.G. 2005. Channel flow and ductile extrusion <strong>of</strong> <strong>the</strong> high Himalayan<br />

slab—<strong>the</strong> Kangchenjunga–Darjeeling pr<strong>of</strong>ile, Sikkim Himalaya. Journal <strong>of</strong> Asian Earth<br />

Sciences, 25: 173–185.<br />

Seeber, L., and Pêcher, A. 1998. Strain partitioning along <strong>the</strong> Himalayan arc and <strong>the</strong> Nanga<br />

Parbat antiform. <strong>Geology</strong>, 26: 791–794.<br />

Shen, Z.-K., Zhao, C., Yin, A., Li, Y., Jackson, D., Fang, P., and Dong, D. 2000. Contemporary<br />

crustal deformation in East Asia constrained by Global Positioning System measurement.<br />

Journal <strong>of</strong> Geophysical Research, 105: 5721-5734.<br />

Shen, Z.K., Wang, M., Li, Y., Jackson, D., Yin, A., Dong, D., and Peng, F. 2001. Crustal<br />

deformation associated with <strong>the</strong> Altyn Tagh fault system, western China, from GPS.<br />

Journal <strong>of</strong> Geophysical Research, 106: 30607-30621.<br />

Shrestha, S.B. 1987. Geological map <strong>of</strong> mid-western Nepal: Kathmandhu, Royal Nepali<br />

<strong>Department</strong> <strong>of</strong> Mines and geology, scale 1:250,000.<br />

Simpson, C., and DePaor, D.G. 1993. Strain and kinematic analysis in general shear zones.<br />

Journal <strong>of</strong> Structural <strong>Geology</strong>, 15: 1–20<br />

Simpson, C., and DePaor D.G. 1997. Practical analysis <strong>of</strong> general shear zones using <strong>the</strong><br />

porphyroclast hyperbolic distribution method: An example from <strong>the</strong> Scandinavian<br />

277


Caledonides. In, S. Sengupta (Ed.), Evolution <strong>of</strong> Geological Structures in Micro- to<br />

Macro-scales, Chapman and Hall, London: 169–184.<br />

Steiger, R.H. and Jäger, E. 1977. Subcommission on geochronology: Convention on <strong>the</strong> use <strong>of</strong><br />

decay constants in geo- and cosmo-chronology. Earth and Planetary Science Letters, 36:<br />

359-362.<br />

Stipp, M., Stünitz, H., Heilbronner, R., and Schmid, S.M. 2002. Dynamic recrystallization <strong>of</strong><br />

quartz: correlation between natural and experimental conditions. In: De Meer, S., Drury,<br />

De Bresser, J.H.P., and Pennock, G.M. (eds). Deformation Mechanisms and Tectonics:<br />

Current Status and Future Perspectives. Geological Society <strong>of</strong> London Special<br />

Publications, 200: 171-190.<br />

Styron, R.H., Taylor, M.H., and Murphy, M.A. 2011. Oblique convergence, arc-parallel<br />

extension, and <strong>the</strong> role <strong>of</strong> strike-slip faulting in <strong>the</strong> High Himalaya. Geosphere, 7: 582–<br />

596.<br />

Sullivan, W.A., and Beane, R.J. 2010. Asymmetrical quartz crystallographic fabrics produced<br />

during constrictional deformation. Journal <strong>of</strong> Structural <strong>Geology</strong>, 32: 1,430–1,443.<br />

Tapponnier, P., G. Peltzer, A. Y. Le Dain, and R. Armijo. 1982. Propagating extrusion tectonics<br />

in Asia: New insights from simple experiments with plasticine. <strong>Geology</strong>, 10: 611–616.<br />

Taylor, M., and Yin, A. 2009. Active structures <strong>of</strong> <strong>the</strong> Himalayan-Tibetan orogen and <strong>the</strong>ir<br />

relationships to earthquake distribution, contemporary strain field, and Cenozoic<br />

volcanism. Geosphere, 5: 199–214.<br />

Taylor, M., Yin, A., Ryerson, F., Kapp, P., and Ding, L. 2003. Conjugate strike-slip faulting<br />

along <strong>the</strong> Bangong-Nujiang suture zone accommodates coeval east-west extension and<br />

north-south shortening in <strong>the</strong> interior <strong>of</strong> <strong>the</strong> Tibetan Plateau. Tectonics, 22: 1044.<br />

Thiede, R.C., Arrowsmith, J.R., Bookhagen, B., McWilliams, M., Sobel, E.R., and Strecker, M.R.<br />

2006. Dome formation and extension in <strong>the</strong> Tethyan Himalaya, Leo Pargil, northwest<br />

India. Geological Society <strong>of</strong> America Bulletin, 118: 635–650.<br />

Tik<strong>of</strong>f, B., and Fossen, H. 1995. The limitations <strong>of</strong> three-dimensional kinematic vorticity<br />

analysis. Journal <strong>of</strong> Structural <strong>Geology</strong>, 17: 1771–1784.<br />

Toy, V.G., Prior, D.J., and Norris, R.J. 2008. Quartz fabrics in <strong>the</strong> Alpine Fault mylonites:<br />

Influence <strong>of</strong> pre-existing preferred orientations on fabric development during progressive<br />

uplift. Journal <strong>of</strong> Structural <strong>Geology</strong>, 30: 602-621.<br />

Tullis, J., and R. A. Yund. 1991. Diffusion creep in feldspar aggregates: Experimental evidence.<br />

Journal <strong>of</strong> Structural <strong>Geology</strong>, 13: 987 – 1000.<br />

278


Vannay, J.C. and Hodges, K.V. 1996. Tectonometamorphic evolution <strong>of</strong> <strong>the</strong> Himalayan<br />

metamorphic core between <strong>the</strong> Annapurna and Dhaulagiri, central Nepal. Journal <strong>of</strong><br />

Metamorphic <strong>Geology</strong>, 14: 635-656.<br />

Villa, I.M.. 2004. Geochronology <strong>of</strong> metamorphic rocks. Periodico di Mineralogia, 73: 259–271.<br />

Wallis, S.R., 1992. Vorticity analysis in a metachert from <strong>the</strong> Sanbagawa Belt, SW Japan. Journal<br />

<strong>of</strong> Structural <strong>Geology</strong>, 14: 271–280.<br />

Wallis, S.R., 1995. Vorticity analysis and recognition <strong>of</strong> ductile extension in <strong>the</strong> Sanbagawa belt,<br />

SW Japan. Journal <strong>of</strong> Structural <strong>Geology</strong>, 17: 1077–1093.<br />

Wallis, S.R., Platt, J.P. and Knott, S.D. 1993. Recognition <strong>of</strong> syn-convergence extension in<br />

accretionary wedges with examples from <strong>the</strong> Calabrian Arc and <strong>the</strong> Eastern Alps.<br />

American Journal <strong>of</strong> Science, 293: 463–495.<br />

Warren, C. J., Grujic, D., Cottle, J. M. and Rogers, N. W. 2012. Constraining cooling histories:<br />

rutile and titanite chronology and diffusion modelling in NW Bhutan. Journal <strong>of</strong><br />

Metamorphic <strong>Geology</strong>, 30: 113–130.<br />

Webb, A.A.G., Yin, A., Harrison, T.M., Celerier, J. and Burgess, W.P. 2007. The leading edge <strong>of</strong><br />

<strong>the</strong> Greater Himalayan Crystalline complex revealed in <strong>the</strong> NW Indian Himalaya:<br />

Implications for <strong>the</strong> evolution <strong>of</strong> <strong>the</strong> Himalayan orogen. <strong>Geology</strong>, 35: 995–958.<br />

Webb, A.A.G., Yin, A., Harrison, T.M., Celerier, J., Gehrels, G.E., Manning, C.E., and Grove,<br />

M. 2011. Cenozoic tectonic history <strong>of</strong> <strong>the</strong> Himachal Himalaya (NW India) and its<br />

constraints on <strong>the</strong> formation mechanism <strong>of</strong> <strong>the</strong> Himalayan orogen. Geosphere, 7: 1013-<br />

1061.<br />

Williams, P.F. & Jiang, D. 2005. An investigation <strong>of</strong> lower crustal deformation: Evidence for<br />

channel flow and its implications for tectonics and structural studies. Journal <strong>of</strong><br />

Structural <strong>Geology</strong>, 27: 1486–1504.<br />

Williams, H., Turner, S., Kelley, S., and Harris, N. 2001. Age and composition <strong>of</strong> dikes in<br />

Sou<strong>the</strong>rn Tibet: New constraints on <strong>the</strong> timing <strong>of</strong> east-west extension and its relationship<br />

to postcollisional volcanism. <strong>Geology</strong>, 29: 339–342.<br />

Wilson, C.J.L., Russell-Head, D.S., and Sim, H.M., 2003. The application <strong>of</strong> an automated FA<br />

system to <strong>the</strong> textural evolution <strong>of</strong> folded ice layers in shear zones. Annals <strong>of</strong> Glaciology,<br />

37: 7e17.<br />

Wilson, C.J.L., Russell-Head, D.S., Kunze, K., and Viola, G. 2007. The analysis <strong>of</strong> quartz c-axis<br />

fabrics using a modified optical microscope. Journal <strong>of</strong> Microscopy, 227: 30-41.<br />

279


Yakymchuck, C., 2010. Tectonometamorphic evolution <strong>of</strong> <strong>the</strong> Greater Himalayan sequence,<br />

Karnali valley, northwestern Nepal. M.Sc. <strong>the</strong>sis, <strong>Queen's</strong> <strong>University</strong>.<br />

Yakymchuk, C., and Godin, L. 2012. Coupled role <strong>of</strong> deformation and metamorphism in <strong>the</strong><br />

construction <strong>of</strong> inverted metamorphic sequences: an example from far-northwest Nepal.<br />

Journal <strong>of</strong> Metamorphic <strong>Geology</strong>, 30: 513–535.<br />

Yang, Y., Liu, M. 2002. Deformation <strong>of</strong> convergent plates: evidence from discrepancies between<br />

GPS velocities and rigid-plate motions. Geophysical Research Letters, 29: GL13391.<br />

Yin, A. 2006. Cenozoic tectonic evolution <strong>of</strong> <strong>the</strong> Himalayan orogen as constrained by alongstrike<br />

variation <strong>of</strong> structural geometry, exhumation history, and foreland sedimentation.<br />

Earth Science Reviews, 76: 1–131.<br />

Yin, A. and Harrison, T.M. 2000. Geologic Evolution <strong>of</strong> <strong>the</strong> Himalayan-Tibetan Orogen. Annual<br />

Reviews in Earth and Planetary Science, 28: 211-280.<br />

Zhang, P., Shen, Z., Wang, M., Gan, W., Bürgmann, R., Molnar, P., Wang, Q., Niu, Z., Sun, J.,<br />

and Wu, J. 2004. Continuous deformation <strong>of</strong> <strong>the</strong> Tibetan Plateau from global positioning<br />

system data. <strong>Geology</strong>, 32: 809–812.<br />

Zhang, R., Murphy, M.A., Lapen, T.J., Sanchez, V., and Heizler, M. 2010. Late Eocene crustal<br />

thickening followed by Early-Late Oligocene extension along <strong>the</strong> India-Asia suture zone:<br />

Evidence for cyclicity in <strong>the</strong> Himalayan orogeny. In: Leech, M.L., et al., (eds).<br />

Proceedings <strong>of</strong> <strong>the</strong> 25th Himalaya-Karakoram-Tibet Workshop: U.S. Geological Survey<br />

Open-File Report 2010–1099, 2 p., http://pubs.usgs.gov/<strong>of</strong>/2010/1099/zhangmurphy/.<br />

Zhisheng, A., Kutzbach, J.E., Prell, W.L., and Porter, S. 2001. Evolution <strong>of</strong> Asian monsoons and<br />

phased uplift <strong>of</strong> <strong>the</strong> Himalaya-Tibetan Plateau since late Miocene times. Nature, 411: 62-<br />

66.<br />

280

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!