26.11.2014 Views

Defining the Himalayan Main Central Thrust in Nepal - Queen's ...

Defining the Himalayan Main Central Thrust in Nepal - Queen's ...

Defining the Himalayan Main Central Thrust in Nepal - Queen's ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Journal of <strong>the</strong> Geological Society, London, Vol. 165, 2008, pp. 523–534. Pr<strong>in</strong>ted <strong>in</strong> Great Brita<strong>in</strong>.<br />

<strong>Def<strong>in</strong><strong>in</strong>g</strong> <strong>the</strong> <strong>Himalayan</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> <strong>in</strong> <strong>Nepal</strong><br />

MICHAEL P. SEARLE 1 ,RICHARDD.LAW 2 , LAURENT GODIN 3 , KYLE P. LARSON 3 , MICHAEL J.<br />

STREULE 1 ,JOHNM.COTTLE 1 & MICAH J. JESSUP 2<br />

1 Department of Earth Sciences, Oxford University, Parks Road, Oxford OX1 3PR, UK (e-mail: mikes@earth.ox.ac.uk)<br />

2 Department of Geological Science, Virg<strong>in</strong>ia Tech, Blacksburg, VA 24061, USA<br />

3 Geological Sciences and Geological Eng<strong>in</strong>eer<strong>in</strong>g, Queen’s University, K<strong>in</strong>gston, Ontario K7L 3N6, Canada<br />

Abstract: An <strong>in</strong>verted metamorphic field gradient associated with a crustal-scale south-vergent thrust fault,<br />

<strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, has been recognized along <strong>the</strong> Himalaya for over 100 years. A major problem <strong>in</strong><br />

<strong>Himalayan</strong> structural geology is that recent workers have mapped <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> with<strong>in</strong> <strong>the</strong> Greater<br />

<strong>Himalayan</strong> Sequence high-grade metamorphic sequence along several different structural levels. Some workers<br />

map <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> as co<strong>in</strong>cid<strong>in</strong>g with a lithological contact, o<strong>the</strong>rs as co<strong>in</strong>cident with <strong>the</strong> kyanite<br />

isograd, up to 1–3 km structurally up-section <strong>in</strong>to <strong>the</strong> Tertiary metamorphic sequence, without support<strong>in</strong>g<br />

structural data. Some workers recognize a <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone of high ductile stra<strong>in</strong> up to 2–3 km thick,<br />

bounded by an upper thrust, MCT-2 (¼ Vaikrita thrust), and a lower thrust, MCT-1 (¼ Munsiari thrust). Some<br />

workers def<strong>in</strong>e an ‘upper Lesser Himalaya’ thrust sheet that shows similar P–T conditions to <strong>the</strong> Greater<br />

<strong>Himalayan</strong> Sequence. O<strong>the</strong>rs def<strong>in</strong>e <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> ei<strong>the</strong>r on isotopic (Nd, Sr) differences,<br />

differences <strong>in</strong> detrital zircon ages, or as be<strong>in</strong>g co<strong>in</strong>cident with a zone of young (,5 Ma) Th–Pb monazite<br />

ages. Very few papers <strong>in</strong>corporate any structural data <strong>in</strong> justify<strong>in</strong>g <strong>the</strong> position of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>.<br />

These studies, comb<strong>in</strong>ed with recent quantitative stra<strong>in</strong> analyses from <strong>the</strong> Everest and Annapurna Greater<br />

<strong>Himalayan</strong> Sequence, show that a wide region of high stra<strong>in</strong> characterizes most of <strong>the</strong> Greater <strong>Himalayan</strong><br />

Sequence with a concentration along <strong>the</strong> bound<strong>in</strong>g marg<strong>in</strong>s of <strong>the</strong> South Tibetan Detachment along <strong>the</strong> top,<br />

and <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> along <strong>the</strong> base. We suggest that <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> has to be def<strong>in</strong>ed and<br />

mapped on stra<strong>in</strong> criteria, not on stratigraphic, lithological, isotopic or geochronological criteria. The most<br />

logical place to map <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> is along <strong>the</strong> high-stra<strong>in</strong> zone that commonly occurs along <strong>the</strong><br />

base of <strong>the</strong> ductile shear zone and <strong>in</strong>verted metamorphic sequence. Above that horizon, all rocks show some<br />

degree of Tertiary <strong>Himalayan</strong> metamorphism, and most of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence metamorphic or<br />

migmatitic rocks show some degree of pure shear and simple shear ductile stra<strong>in</strong> that occurs throughout <strong>the</strong><br />

mid-crustal Greater <strong>Himalayan</strong> Sequence channel. The <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> evolved both <strong>in</strong> time (early–<br />

middle Miocene) and space from a deep-level ductile shear zone to a shallow brittle thrust fault.<br />

The <strong>Himalayan</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, with its zone of <strong>in</strong>verted<br />

metamorphic isograds from sillimanite grade down to biotite<br />

grade, is one of <strong>the</strong> largest ductile shear zones known from any<br />

collision-related mounta<strong>in</strong> belt. The <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> crops<br />

out along c. 2200 km length of <strong>the</strong> Himalaya from western<br />

Zanskar to Bhutan and Arunachal Pradesh (Fig. 1). It dips north<br />

and places high-grade metamorphic rocks of <strong>the</strong> Greater Himalaya<br />

south over unmetamorphosed rocks of <strong>the</strong> Lesser Himalaya.<br />

S<strong>in</strong>ce <strong>the</strong> discovery of an <strong>in</strong>verted metamorphic field gradient<br />

across <strong>the</strong> Darjeel<strong>in</strong>g–Sikkim Himalaya by Mallet (1874) and<br />

von Loczy (1878), and across <strong>the</strong> Indian Himalaya by Oldham<br />

(1883), it has been recognized that metamorphic grade <strong>in</strong>creases<br />

up-structural section towards <strong>the</strong> north from <strong>the</strong> Lesser Himalaya<br />

to <strong>the</strong> Greater Himalaya. In <strong>Nepal</strong>, pioneer<strong>in</strong>g geological studies<br />

by Hagen (1954), Hashimoto (1959, 1973) and Bordet (1961)<br />

also recognized <strong>the</strong> <strong>in</strong>crease of metamorphic grade up-structural<br />

section. The <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone was def<strong>in</strong>ed by Heim &<br />

Gansser (1939) and Gansser (1964) as <strong>the</strong> thrust fault that places<br />

high-grade metamorphic rocks of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence<br />

southward over low-grade rocks of <strong>the</strong> Lesser Himalaya.<br />

Unfortunately, this def<strong>in</strong>ition is not useful, because Greater and<br />

Lesser <strong>Himalayan</strong> rocks refer to thrust-bounded structural<br />

packages. Typically thrusts cut up-stratigraphic section <strong>in</strong> <strong>the</strong><br />

footwall, along ramps plac<strong>in</strong>g older rocks over younger rocks.<br />

<strong>Thrust</strong>s may also follow flats plac<strong>in</strong>g similar age or younger<br />

rocks over older rocks. We emphasize here <strong>the</strong> dist<strong>in</strong>ction<br />

between structural term<strong>in</strong>ology (Greater <strong>Himalayan</strong> Sequence/<br />

Lesser Himalaya Sequence) and stratigraphic term<strong>in</strong>ology. S<strong>in</strong>ce<br />

<strong>the</strong> orig<strong>in</strong>al recognition of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> along <strong>the</strong><br />

Himalaya, <strong>the</strong>re has been a great amount of confusion regard<strong>in</strong>g<br />

<strong>the</strong> structural position and tim<strong>in</strong>g of slip along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong>. This has come about because many different structures<br />

are called <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> by different workers. Clearly<br />

<strong>the</strong>re is an urgent need to f<strong>in</strong>d a common def<strong>in</strong>ition and location<br />

of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, and <strong>in</strong> this paper we attempt to do<br />

that, based on comb<strong>in</strong>ed stra<strong>in</strong> and metamorphic criteria.<br />

Previous attempts to map <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> have used<br />

<strong>in</strong>direct methods such as: (1) a lithological contrast follow<strong>in</strong>g a<br />

dist<strong>in</strong>ctive quartzite unit beneath an orthogneiss unit (e.g.<br />

Gansser 1983; Daniel et al. 2003); (2) follow<strong>in</strong>g <strong>the</strong> kyanite<br />

isograd (e.g. Bordet 1961; LeFort 1975; Colchen et al. 1986); (3)<br />

differences <strong>in</strong> U–Pb detrital zircon ages (e.g. Parrish & Hodges<br />

1996; Ahmad et al. 2000; DeCelles et al. 2000); (4) differences<br />

<strong>in</strong> Nd isotope compositions (e.g. Rob<strong>in</strong>son et al. 2001; Mart<strong>in</strong> et<br />

al. 2005; Richards et al. 2005, 2006); (5) location of young U–<br />

Pb and Th–Pb monazite ages (e.g. Harrison et al. 1997; Catlos<br />

et al. 2001, 2002;). None of <strong>the</strong>se methods <strong>in</strong> <strong>the</strong>mselves can be<br />

used <strong>in</strong>dependently to def<strong>in</strong>e a thrust fault. Lithology, detrital<br />

zircon ages and Nd isotopes give <strong>in</strong>formation on stratigraphy, not<br />

structural relationships. Isograds and monazite ages give <strong>in</strong>formation<br />

on metamorphic reactions, fluids, and tim<strong>in</strong>g of m<strong>in</strong>eral<br />

growth, not structure.<br />

523


524<br />

M. P. SEARLE ET AL.<br />

Fig. 1. Geological map of <strong>the</strong> Himalaya.<br />

Zones of high stra<strong>in</strong> have been documented across <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> zone <strong>in</strong> <strong>the</strong> Arun Valley by Brunel (1986) and<br />

Brunel & Kienast (1986) us<strong>in</strong>g k<strong>in</strong>ematic criteria. Abundant<br />

shear criteria (S–C fabrics, rolled garnets, etc.) and stretch<strong>in</strong>g<br />

l<strong>in</strong>eations show southward transport of <strong>the</strong> Greater <strong>Himalayan</strong><br />

Sequence along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone. Quantitative<br />

vorticity studies were first reported from <strong>the</strong> Sutlej Valley, India<br />

by Grasemann et al. (1999). Law et al. (2004) and Jessup et al.<br />

(2006) showed that mean k<strong>in</strong>ematic vorticity numbers from <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone along <strong>the</strong> Everest profile <strong>in</strong> <strong>Nepal</strong><br />

yielded 58–44% pure shear component <strong>in</strong> addition to <strong>the</strong><br />

dom<strong>in</strong>ant top-to-south simple shear. The fact that metamorphic<br />

isograds across <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> ductile shear zone have<br />

been compressed or telescoped <strong>in</strong>to a 1–2 km thick section<br />

(Searle & Rex 1989; Hubbard 1996) shows that shear<strong>in</strong>g postdates<br />

peak metamorphism.<br />

Here we discuss <strong>the</strong> various previous methods used to map or<br />

def<strong>in</strong>e <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> and <strong>the</strong>n we propose a unify<strong>in</strong>g<br />

def<strong>in</strong>ition and map location, <strong>in</strong> <strong>the</strong> hope that future studies<br />

relat<strong>in</strong>g to <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> will refer to one s<strong>in</strong>gle <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> ductile shear zone and brittle thrust fault.<br />

Metamorphic isograds and <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

The earliest attempt to map <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> was carried<br />

out by Bordet (1961), who mapped it along a prom<strong>in</strong>ent kyanitebear<strong>in</strong>g<br />

pelite band with<strong>in</strong> sillimanite gneisses <strong>in</strong> <strong>the</strong> Arun valley,<br />

but did not describe any structural criteria to support this<br />

placement (Figs 2 and 3). The same location for <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> was subsequently used by Lombardo et al. (1993)<br />

and Pognante & Benna (1993), who mapped <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> fur<strong>the</strong>r north as far as Kharta <strong>in</strong> sou<strong>the</strong>rn Tibet. Above<br />

this kyanite-bear<strong>in</strong>g pelite horizon are some 5–8 km thickness of<br />

sillimanite + garnet + biotite + cordierite orthogneiss (variously<br />

called <strong>the</strong> Barun gneiss, Black gneiss or Jannu–Kangchenjunga<br />

gneiss) with evidence of abundant partial melt<strong>in</strong>g (Brunel &<br />

Kienast 1986; Lombardo et al. 1993; Pognante & Benna 1993;<br />

Searle & Szulc 2005). High-grade calc-silicate gneisses and


HIMALAYAN MAIN CENTRAL THRUST, NEPAL 525<br />

Fig. 2. Map of <strong>the</strong> Everest–Makalu–<br />

Kangchenjunga Himalaya <strong>in</strong> east <strong>Nepal</strong><br />

show<strong>in</strong>g location of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> (MCT) and Greater <strong>Himalayan</strong><br />

Series. Shaded area represents <strong>the</strong> partially<br />

molten channel conta<strong>in</strong><strong>in</strong>g migmatites and<br />

leucogranites. The Bordet (1961) <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> follows a prom<strong>in</strong>ent band of<br />

kyanite gneisses at Tashigaon village,<br />

with<strong>in</strong> <strong>the</strong> sillimanite-grade gneisses. Our<br />

proposed location of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> is <strong>the</strong> ductile shear zone<br />

correspond<strong>in</strong>g to <strong>the</strong> zone of <strong>in</strong>verted<br />

metamorphic isograds above Tuml<strong>in</strong>gtar<br />

village along <strong>the</strong> Arun river, and near<br />

Taplejung <strong>in</strong> <strong>the</strong> Tamur river dra<strong>in</strong>age.<br />

LHS, Lesser Himalaya Series; GHS,<br />

Greater Himalaya Series; TSS, Tethyan<br />

sedimentary series.<br />

marbles conta<strong>in</strong><strong>in</strong>g oliv<strong>in</strong>e, cl<strong>in</strong>opyroxene, wollastonite and<br />

scapolite are <strong>in</strong>tercalated with <strong>the</strong> orthogneiss. P–T conditions<br />

reached upper amphibolite facies and even granulite facies at<br />

800–850 8C and 10–12 kbar, with later decompression follow<strong>in</strong>g<br />

a clockwise P–T–t path to 4–6 kbar dur<strong>in</strong>g sillimanite-grade<br />

metamorphism (Goscombe & Hand 2000; Dasgupta et al. 2004).<br />

Below <strong>the</strong> Bordet (1961) <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> horizon, <strong>the</strong><br />

Num orthogneiss is a 3–4 km thick unit of sillimanite + K-<br />

feldspar orthogneiss with about 15–20% <strong>in</strong> situ partial melt (Fig.<br />

3). Metamorphic grade is similar above and below this kyanite<br />

gneiss horizon, although protolith rocks are probably from different<br />

stratigraphic levels. Internal stra<strong>in</strong> is extremely high across<br />

this entire package with consistent shear criteria <strong>in</strong>dicat<strong>in</strong>g<br />

south-directed simple shear with a significant component of<br />

coaxial pure shear. Structurally beneath <strong>the</strong> Num orthogneiss,<br />

near Tuml<strong>in</strong>gtar, is a telescoped and highly sheared <strong>in</strong>verted<br />

metamorphic isograd sequence from sillimanite through kyanite<br />

and staurolite to biotite grade, where we map our preferred<br />

location for <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. Above this, all rocks have<br />

a Tertiary metamorphic impr<strong>in</strong>t on protoliths that range from<br />

Proterozoic to Mesozoic. Below our <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, rocks<br />

have little or no Tertiary metamorphic overpr<strong>in</strong>t.<br />

Goscombe et al. (2006) recognized that <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> had been <strong>in</strong>correctly mapped as follow<strong>in</strong>g a stratigraphic<br />

boundary (<strong>the</strong>ir ‘<strong>Himalayan</strong> unconformity’) and <strong>the</strong>y mapped <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> beneath <strong>the</strong> Ulleri–Phaplu augen gneiss <strong>in</strong>


526<br />

M. P. SEARLE ET AL.<br />

Fig. 3. Simplified schematic section across<br />

<strong>the</strong> Everest–Makalu Himalaya show<strong>in</strong>g key<br />

features of <strong>the</strong> structure, stratigraphy and<br />

m<strong>in</strong>eral isograds, toge<strong>the</strong>r with our<br />

proposed location of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> <strong>in</strong> eastern <strong>Nepal</strong>. bt, biotite; grt,<br />

garnet; st, staurolite; ky, kyanite; sill,<br />

sillimanite; crd, cordierite; ms, muscovite;<br />

kfs, K-feldspar.<br />

eastern <strong>Nepal</strong>. However, <strong>the</strong>y also def<strong>in</strong>ed a new structure, <strong>the</strong><br />

‘High Himal <strong>Thrust</strong>’, close to, or along <strong>the</strong> kyanite pelite band<br />

and <strong>the</strong> Bordet (1961) <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. Sillimanite +<br />

muscovite + K-feldspar grade gneisses and migmatites occur<br />

both above and below this horizon, and we suggest that both are<br />

part of <strong>the</strong> same Greater <strong>Himalayan</strong> Sequence metamorphic<br />

package. Goscombe et al. (2006) def<strong>in</strong>ed a <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

zone that extends down-section from this horizon south as far as<br />

<strong>the</strong> garnet isograd beneath <strong>the</strong> Ulleri–Phaplu orthogneiss (Fig.<br />

3).<br />

In <strong>the</strong> Annapurna–Manaslu region of central <strong>Nepal</strong> LeFort<br />

(1975), Colchen et al. (1986) and Pêcher (1989) mapped <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> as follow<strong>in</strong>g <strong>the</strong> kyanite isograd (at Dana <strong>in</strong><br />

<strong>the</strong> Kali Gandaki valley and Bahundanda <strong>in</strong> <strong>the</strong> Marsyandi<br />

valley; Figs 4–6). There is no doubt that <strong>the</strong> kyanite gneisses are<br />

highly stra<strong>in</strong>ed, but so too are most of <strong>the</strong> rocks structurally<br />

above and below this horizon. High-stra<strong>in</strong> shear fabrics are<br />

particularly prom<strong>in</strong>ent <strong>in</strong> pelitic schists and K-feldspar augen<br />

gneisses, but less apparent <strong>in</strong> <strong>the</strong> more massive homogeneous<br />

marble horizons. The location of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> as<br />

mapped by LeFort (1975) and Colchen et al. (1986) was followed<br />

by most subsequent workers <strong>in</strong> <strong>the</strong> region (e.g. Harrison et al.<br />

1997; Kohn et al. 2001).<br />

Hodges et al. (1996) recognized several shear zones and<br />

thrusts across a ‘<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone’ and mapped <strong>the</strong><br />

sou<strong>the</strong>rn limit of <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> shear<strong>in</strong>g close to <strong>the</strong><br />

village of Lamdrung <strong>in</strong> <strong>the</strong> Modi khola. Searle & God<strong>in</strong> (2003)<br />

mapped <strong>the</strong> entire <strong>in</strong>verted metamorphic sequence as be<strong>in</strong>g part<br />

of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence and placed <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> brittle fault along <strong>the</strong> base of <strong>the</strong> <strong>in</strong>verted metamorphic<br />

sequence (Fig. 5). This locality marks a sharp break between<br />

highly stra<strong>in</strong>ed rocks affected by Tertiary metamorphism <strong>in</strong> <strong>the</strong><br />

hang<strong>in</strong>g wall and rocks beneath this that are not highly<br />

metamorphosed or highly stra<strong>in</strong>ed. Orthogneiss horizons such as<br />

<strong>the</strong> Proterozoic Ulleri augen gneiss <strong>in</strong> <strong>the</strong> Annapurna region, or<br />

<strong>the</strong> Phaplu augen gneiss <strong>in</strong> <strong>the</strong> Everest region, previously<br />

assigned to <strong>the</strong> ‘upper Lesser Himalaya’ thrust sheet, are now<br />

more logically placed with<strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence<br />

thrust sheet above <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>.<br />

Arita (1983) mapped a <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone of high<br />

ductile stra<strong>in</strong> up to 2–3 km thick, bounded by an upper thrust,<br />

MCT-2 (Vaikrita thrust), and a lower thrust, MCT-1 (Munsiari<br />

thrust). The earlier, upper MCT-2 corresponds to <strong>the</strong> Colchen et<br />

al. (1986) <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> whereas <strong>the</strong> lower, later MCT-1<br />

corresponds to our proposed location of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>.<br />

The rocks between <strong>the</strong>se two thrusts show peak metamorphic<br />

temperatures rang<strong>in</strong>g between 550 8C and less than 330 8C, with<br />

an <strong>in</strong>verted <strong>the</strong>rmal gradient as deduced from Raman spectroscopy<br />

of carbonaceous material by Beyssac et al. (2004) and<br />

Boll<strong>in</strong>ger et al. (2004). Those workers accepted <strong>the</strong> location of<br />

<strong>the</strong> Colchen et al. (1986) <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, and so placed<br />

<strong>the</strong>se rocks <strong>in</strong> <strong>the</strong> Lesser Himalaya. However, we <strong>in</strong>clude all<br />

<strong>the</strong>se metamorphic rocks <strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence<br />

above our proposed <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>.<br />

With<strong>in</strong> <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> ductile shear zone, Kohn et al.<br />

(2001) showed that garnets from structurally lower locations<br />

grew with <strong>in</strong>creas<strong>in</strong>g P and T (load<strong>in</strong>g), whereas garnets from<br />

structurally higher locations grew with <strong>in</strong>creas<strong>in</strong>g T but decreas<strong>in</strong>g<br />

P (exhumation). This records a snapshot <strong>in</strong> time of <strong>the</strong><br />

cont<strong>in</strong>uously evolv<strong>in</strong>g northward burial (prograde metamorphism)<br />

and southward exhumation (decompression and retrograde<br />

metamorphism) particle paths of Greater <strong>Himalayan</strong> Sequence<br />

metamorphic rocks. Searle et al. (2002) suggested that <strong>the</strong><br />

position of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> along <strong>the</strong> Darondi valley<br />

should be 15 km south of where it was mapped by Colchen et al.<br />

(1986) and Kohn et al. (2001), along <strong>the</strong> base of <strong>the</strong> <strong>in</strong>verted<br />

metamorphic sequence (‘Location of <strong>in</strong>ferred structure’ <strong>in</strong> fig. 1<br />

of Kohn et al. 2001).<br />

Dadeldhura and Ramgarh thrusts<br />

DeCelles et al. (2001) and Rob<strong>in</strong>son et al. (2003, 2006) mapped<br />

two thrust sheets structurally beneath <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> <strong>in</strong><br />

western <strong>Nepal</strong>, <strong>the</strong> higher Dadeldhura and lower Ramgarh thrust<br />

sheets. They failed to locate a discrete thrust at <strong>the</strong> position of<br />

<strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> and <strong>in</strong>ferred its presence from extrapolation<br />

along strike <strong>in</strong> <strong>the</strong> Karnali Valley. The Dadeldhura thrust<br />

sheet consists of garnet–muscovite–biotite schists, mylonitic<br />

augen gneiss and Cambrian–Ordovician granites. The Ramgarh<br />

thrust sheet consists of greenschist-facies metasedimentary rocks


HIMALAYAN MAIN CENTRAL THRUST, NEPAL 527<br />

Fig. 4. Map of <strong>the</strong> Annapurna–Manaslu<br />

Himalaya, show<strong>in</strong>g <strong>the</strong> structure of <strong>the</strong><br />

Greater <strong>Himalayan</strong> Sequence and our<br />

proposed location of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong>. The Colchen et al. (1986) <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> is co<strong>in</strong>cident with <strong>the</strong><br />

kyanite isograd runn<strong>in</strong>g through Dana and<br />

Bahundanda villages. Our <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> is located along a high-stra<strong>in</strong> zone<br />

fur<strong>the</strong>r south, south of Gorhka, and<br />

corresponds to <strong>the</strong> sou<strong>the</strong>rn limit of Tertiary<br />

metamorphism. The mapped locations of<br />

<strong>the</strong> South Tibetan Detachment system<br />

(STDS) normal faults are from Searle &<br />

God<strong>in</strong> (2003).<br />

Fig. 5. Simplified, schematic section across<br />

<strong>the</strong> Annapurna Himalaya show<strong>in</strong>g key<br />

features of <strong>the</strong> structure, stratigraphy and<br />

m<strong>in</strong>eral isograds, and our proposed location<br />

of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> <strong>in</strong> central <strong>Nepal</strong>.<br />

Shaded area represents <strong>the</strong> migmatites and<br />

leucogranites with<strong>in</strong> <strong>the</strong> partially molten<br />

channel.<br />

of <strong>the</strong> Kushma and Ranimata Formations. Both <strong>the</strong> Dadeldhura<br />

and Ramgarh thrust sheets occur <strong>in</strong> a synformal klippe that is a<br />

structural equivalent of <strong>the</strong> Almora klippe to <strong>the</strong> west <strong>in</strong> India<br />

and <strong>the</strong> Kathmandu klippe to <strong>the</strong> east. The Ramgarh thrust forms<br />

<strong>the</strong> roof thrust to a series of imbricated thrust slices of<br />

unmetamorphosed Lesser <strong>Himalayan</strong> rocks of Late Archaean,<br />

Proterozoic and Cambrian age (DeCelles et al. 2001; Rob<strong>in</strong>son<br />

et al. 2006).


528<br />

M. P. SEARLE ET AL.<br />

Fig. 6. Simplified, schematic section across<br />

<strong>the</strong> Manaslu Himalaya, show<strong>in</strong>g key<br />

features and our proposed location of <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. Shaded area<br />

represents <strong>the</strong> zone of partial melt<strong>in</strong>g with<br />

migmatites and leucogranites (crosses). The<br />

Manaslu leuocogranite is wholly with<strong>in</strong> <strong>the</strong><br />

Greater <strong>Himalayan</strong> Sequence, follow<strong>in</strong>g<br />

Searle & God<strong>in</strong> (2003), with <strong>the</strong> South<br />

Tibetan detachment (STD) wrapp<strong>in</strong>g around<br />

<strong>the</strong> upper level of <strong>the</strong> granite.<br />

The Ramgarh thrust marks <strong>the</strong> sou<strong>the</strong>rn limit of Tertiary<br />

<strong>Himalayan</strong> metamorphism <strong>in</strong> western <strong>Nepal</strong> and we prefer to l<strong>in</strong>k<br />

this with <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. The reported location of <strong>the</strong><br />

Ramgarh thrust <strong>in</strong> central and eastern <strong>Nepal</strong> (Mart<strong>in</strong> et al. 2005;<br />

Pearson & DeCelles 2005), however, does not co<strong>in</strong>cide with <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. In central and eastern <strong>Nepal</strong> <strong>the</strong> location of<br />

<strong>the</strong> Ramgarh thrust is almost entirely <strong>in</strong>terpreted from lithological<br />

repetition; a possible fault surface has been observed only<br />

<strong>in</strong> <strong>the</strong> Tribeni area of eastern <strong>Nepal</strong>. A more sou<strong>the</strong>rly location<br />

for <strong>the</strong> Ramgarh thrust is supported by pervasive deformation<br />

documented by quartz c-axis fabrics throughout central <strong>Nepal</strong><br />

(Bouchez & Pêcher 1981). Although <strong>the</strong> recrystallization of<br />

quartz under significantly high stra<strong>in</strong> has been recognized <strong>in</strong> <strong>the</strong><br />

Ramgrah thrust sheet, as mapped by Mart<strong>in</strong> et al. (2005) and<br />

Pearson & DeCelles (2005), it is <strong>in</strong>terpreted to be locally<br />

conf<strong>in</strong>ed to <strong>the</strong> immediate hang<strong>in</strong>g wall of <strong>the</strong> <strong>in</strong>ferred thrust<br />

(Pearson & DeCelles 2005). However, Bouchez & Pêcher (1981)<br />

showed that quartz c-axis fabrics are preserved for more than<br />

6 km far<strong>the</strong>r south than <strong>the</strong> mapped position of <strong>the</strong> Ramgarh<br />

thrust. Both lithological and structural data fit better with a<br />

structurally lower, more sou<strong>the</strong>rly, Ramgarh thrust where it is<br />

co<strong>in</strong>cident with <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>.<br />

Restored sections show that <strong>the</strong> Ramgarh, Dadeldhura and<br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> thrust sheets of DeCelles et al. (2001) all have<br />

Proterozoic sedimentary rocks, Ulleri augen gneiss and Cambrian–Ordovician<br />

sedimentary rocks and granites as protoliths.<br />

Hang<strong>in</strong>g-wall–footwall cut-offs can be successfully matched <strong>in</strong><br />

restored sections (Fig. 7). We <strong>the</strong>refore propose that <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

Fig. 7. Generalized restored section across <strong>the</strong> <strong>Nepal</strong> Himalaya show<strong>in</strong>g <strong>the</strong> pre-thrust<strong>in</strong>g trajectories of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> and South Tibetan<br />

Detachment shear zones and faults. The shaded horizon represents <strong>the</strong> Upper Proterozoic sedimentary rocks of <strong>the</strong> Lesser Himalaya, and Greater<br />

Himalaya. With<strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence <strong>the</strong>se <strong>in</strong>clude <strong>the</strong> metamorphosed rocks of <strong>the</strong> Nawakot Group above <strong>the</strong> Ramgarh thrust (<strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong>) and <strong>the</strong> Bhimpedi Group with<strong>in</strong> <strong>the</strong> Kathmandu nappe, above <strong>the</strong> Mahabharat thrust. Sillimanite-grade pelitic gneisses with<strong>in</strong> <strong>the</strong> Greater<br />

<strong>Himalayan</strong> Sequence are <strong>in</strong>terpreted as metamorphosed Upper Proterozoic sedimentary rocks of <strong>the</strong> Haimanta–Vaikrita Group. The base of <strong>the</strong> Tethyan<br />

Himalaya consists of similar Neoproterozoic sedimentary rocks show<strong>in</strong>g that <strong>the</strong> Lesser, Greater and Tethyan Himalaya were all part of one contiguous<br />

Indian plate.


HIMALAYAN MAIN CENTRAL THRUST, NEPAL 529<br />

<strong>Central</strong> <strong>Thrust</strong> should be mapped along <strong>the</strong> Ramgarh thrust, and<br />

all rocks above that should be <strong>in</strong>corporated <strong>in</strong>to <strong>the</strong> Greater<br />

<strong>Himalayan</strong> Sequence. The major thrust systems propagated<br />

southward with time from <strong>the</strong> Early Miocene motion of <strong>the</strong><br />

Greater <strong>Himalayan</strong> Sequence metamorphic core (Hodges et al.<br />

1996; God<strong>in</strong> et al. 2001) to <strong>the</strong> ,15 Ma motion along <strong>the</strong><br />

Ramgarh thrust (DeCelles et al. 2001; Rob<strong>in</strong>son et al. 2006).<br />

Dur<strong>in</strong>g <strong>the</strong> Late Miocene ductile shear<strong>in</strong>g along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong>–Ramgarh thrust ceased, and thrust<strong>in</strong>g propagated downsection<br />

to <strong>the</strong> Lesser <strong>Himalayan</strong> brittle imbricate thrust system.<br />

At least 120 km of southward translation has been estimated<br />

across <strong>the</strong> Ramgarh thrust sheet (Rob<strong>in</strong>son et al. 2006).<br />

Mahabharat thrust<br />

Several klippen or thrust sheets of high-grade metamorphic rocks<br />

and granites (e.g. Almora klippe; Kathmandu complex) overlie<br />

low-grade or unmetamorphosed rocks of <strong>the</strong> Lesser Himalaya to<br />

<strong>the</strong> south of <strong>the</strong> ma<strong>in</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, as mapped <strong>in</strong><br />

Langtang and <strong>the</strong> Ganesh Himal (Fig. 1). The thrust beneath<br />

<strong>the</strong>se klippen has been variously termed <strong>the</strong> Almora or Munsiari<br />

thrusts <strong>in</strong> India (Heim & Gansser 1939; Valdiya 1980), and <strong>the</strong><br />

Mahabharat and Dadeldhura thrusts <strong>in</strong> <strong>Nepal</strong> (Stöckl<strong>in</strong> &<br />

Bhattarai 1980; Upreti & LeFort 1999; Johnson et al. 2001). The<br />

Mahabharat thrust beneath <strong>the</strong> Kathmandu klippe encircles <strong>the</strong><br />

sou<strong>the</strong>rn Kathmandu valley and l<strong>in</strong>ks up with <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> <strong>in</strong> <strong>the</strong> NW and NE (Upreti & LeFort 1999; Johnson et al.<br />

2001). Rocks above <strong>the</strong> Mahabharat thrust <strong>in</strong>clude Proterozoic<br />

Bhimpedi Group and early–middle Palaeozoic Phulchauki Group<br />

sedimentary rocks, which are <strong>in</strong>truded by Ordovician granites<br />

and augen gneisses. Metamorphism reaches kyanite grade at <strong>the</strong><br />

base and isograds are right-way-up from kyanite through garnet<br />

and biotite to chlorite grade (Johnson et al. 2001). Along <strong>the</strong><br />

Mahabharat thrust dynamic metamorphism has locally <strong>in</strong>verted<br />

<strong>the</strong> <strong>the</strong>rmal gradient with formation of garnet–biotite mylonites<br />

and phyllonites. Beneath <strong>the</strong> Mahabharat thrust carbonaceous<br />

pelites of <strong>the</strong> Nawakot Group show an <strong>in</strong>verted <strong>the</strong>rmal gradient<br />

from 468 8C to less than 330 8C, based on Raman spectroscopy<br />

of carbonaceous material (Beyssac et al. 2004; Boll<strong>in</strong>ger et al.<br />

2004).<br />

These crystall<strong>in</strong>e complexes have been termed ‘Lesser <strong>Himalayan</strong><br />

crystall<strong>in</strong>es’ or ‘Outer Lesser <strong>Himalayan</strong> crystall<strong>in</strong>es’, but<br />

<strong>in</strong> reality <strong>the</strong>y are lateral equivalents to <strong>the</strong> Greater <strong>Himalayan</strong><br />

Sequence above <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. They share similar<br />

upper Proterozoic and lower Palaeozoic protoliths, similar Miocene<br />

metamorphism, and similar 21–18 Ma leucogranite pegmatite<br />

dykes as <strong>the</strong> ma<strong>in</strong> Greater <strong>Himalayan</strong> Sequence to <strong>the</strong> north<br />

(Johnson et al. 2001). We concur with <strong>the</strong> conclusions of<br />

Johnson et al. (2001) and Johnson (2005) that <strong>the</strong> Mahabharat<br />

thrust is <strong>the</strong> same structure as <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, but along<br />

a more sou<strong>the</strong>rly, proximal position <strong>in</strong> <strong>the</strong> restoration. The<br />

Mahabharat thrust climbs up-section <strong>in</strong> <strong>the</strong> transport direction,<br />

from be<strong>in</strong>g along <strong>the</strong> base of <strong>the</strong> <strong>in</strong>verted metamorphic sequence<br />

at Langtang <strong>in</strong> <strong>the</strong> north, to along <strong>the</strong> isograd fold h<strong>in</strong>ge at<br />

Kathmandu (Fig. 8). The youngest thrust splays off beneath <strong>the</strong><br />

Kathmandu complex to l<strong>in</strong>k with <strong>the</strong> Ramgarh thrust to <strong>the</strong> south<br />

and west of <strong>the</strong> Kathmandu complex (Fig. 1). All <strong>the</strong>se thrusts<br />

are part of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> but are restored at different<br />

depths, progressively ramp<strong>in</strong>g up-section towards <strong>the</strong> south.<br />

The Darjeel<strong>in</strong>g klippe is ano<strong>the</strong>r structural outlier of Greater<br />

Fig. 8. Geometry of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone <strong>in</strong> <strong>the</strong> Langtang–Kathmandu nappe region of central <strong>Nepal</strong> show<strong>in</strong>g <strong>the</strong> relationship of <strong>the</strong> Mahabharat<br />

and Ramgarh thrusts to <strong>the</strong> metamorphic isograds. This geometry comb<strong>in</strong>es <strong>the</strong> folded isograd model of Searle & Rex (1989) with <strong>the</strong> channel flow model<br />

for <strong>the</strong> Greater <strong>Himalayan</strong> Sequence (Law et al. 2006; Searle et al. 2006) and with <strong>the</strong> Johnson (2005) structural model for <strong>the</strong> Mahabharat thrust and<br />

Kathmandu nappe. It also expla<strong>in</strong>s <strong>the</strong> structural location of <strong>the</strong> greenschist- and amphibolite-facies metamorphic rocks of <strong>the</strong> Ramgarh thrust sheet<br />

(Beyssac et al. 2004; Boll<strong>in</strong>ger et al. 2004) structurally beneath <strong>the</strong> Mahabharat thrust. Shaded area shows <strong>the</strong> zone of partial melt<strong>in</strong>g (sillimanite +<br />

K-feldspar gneisses, migmatites), with <strong>the</strong> ma<strong>in</strong> leucogranites (crosses) concentrated along <strong>the</strong> north. Metamorphic isograds have been sheared (by a<br />

comb<strong>in</strong>ation of pure shear and south-directed simple shear), and flattened along <strong>the</strong> South Tibetan Detachment ductile shear zone above (right-way-up<br />

metamorphic isograds) and along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> ductile shear zone below (<strong>in</strong>verted metamorphic isograds).


530<br />

M. P. SEARLE ET AL.<br />

<strong>Himalayan</strong> Sequence rocks thrust above unmetamorphosed Lesser<br />

<strong>Himalayan</strong> sedimentary rocks <strong>in</strong> far eastern <strong>Nepal</strong> and<br />

Sikkim–West Bengal (Fig. 1). The full <strong>in</strong>verted metamorphic<br />

isograd sequence has been mapped here, but unlike <strong>the</strong> Kathmandu<br />

complex, and like <strong>the</strong> Greater <strong>Himalayan</strong> Sequence, <strong>the</strong><br />

isograds are structurally <strong>in</strong>verted from sillimanite down to<br />

biotite–chlorite (Mohan et al. 1989; Dasgupta et al. 2004). The<br />

structures and metamorphic P–T conditions clearly show that <strong>the</strong><br />

Darjeel<strong>in</strong>g klippe is l<strong>in</strong>ked to <strong>the</strong> ma<strong>in</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> to<br />

<strong>the</strong> north (Searle & Szulc 2005; Fig. 3).<br />

Detrital zircon ages and <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

Detrital zircon U–Pb ages provide maximum depositional age<br />

constra<strong>in</strong>ts of <strong>the</strong> metamorphic protolith. Parrish & Hodges<br />

(1996) orig<strong>in</strong>ally proposed that Greater and Lesser <strong>Himalayan</strong><br />

rocks, divided by <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, had a significant<br />

difference <strong>in</strong> sedimentary provenance. Zircons from Greater<br />

<strong>Himalayan</strong> Sequence rocks have ma<strong>in</strong>ly late Proterozoic and<br />

early Palaeozoic ages, whereas zircons from <strong>the</strong> Lesser Himalaya<br />

have late Archaean–early Proterozoic ages. DeCelles et al.<br />

(2000) showed that metasedimentary rocks from <strong>the</strong> Greater<br />

<strong>Himalayan</strong> Sequence gave zircon ages of 800–1700 Ma, whereas<br />

quartzites of one unit generally mapped with<strong>in</strong> <strong>the</strong> Lesser<br />

Himalaya Sequence, <strong>the</strong> Nawakot Group, yielded zircons<br />

.1.8 Ga old. This reflects <strong>the</strong> 1866–1833 Ma depositional age<br />

of <strong>the</strong>se units. The upper age limit is given by <strong>the</strong> Ulleri augen<br />

gneiss, which was <strong>in</strong>truded <strong>in</strong>to <strong>the</strong> Nawakot quartzite. However,<br />

a major mylonite zone correspond<strong>in</strong>g to <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

was mapped beneath <strong>the</strong> Ulleri augen gneiss <strong>in</strong> <strong>the</strong> Annapurna<br />

region (Searle & God<strong>in</strong> 2003), so <strong>the</strong>se rocks are now <strong>in</strong>cluded<br />

with<strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence. The Phaplu augen<br />

gneiss <strong>in</strong> <strong>the</strong> Everest profile is a similar age, and at a similar<br />

structural position to <strong>the</strong> Ulleri augen gneiss. The Phaplu gneiss<br />

is highly sheared and overla<strong>in</strong> by staurolite- and sillimanite-grade<br />

rocks typical of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence, so it has also<br />

been <strong>in</strong>cluded here with<strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence<br />

(Jessup et al. 2006; Searle et al. 2006).<br />

There seems little doubt that many Lesser <strong>Himalayan</strong> protoliths,<br />

particularly <strong>in</strong> <strong>Nepal</strong>, are older than <strong>the</strong> exposed Greater<br />

<strong>Himalayan</strong> Sequence protoliths. The upper structural levels of<br />

<strong>the</strong> Lesser Himalaya <strong>in</strong> India (Cambrian Krol and Tal Formations,<br />

which overlie Proterozoic rocks) are lateral equivalents to<br />

<strong>the</strong> base of <strong>the</strong> restored Greater <strong>Himalayan</strong> Sequence (Steck<br />

2003) It is also widely accepted that <strong>the</strong> Greater <strong>Himalayan</strong><br />

Sequence protoliths were similar <strong>in</strong> age to <strong>the</strong> lower levels of <strong>the</strong><br />

Tethyan Himalaya, which range <strong>in</strong> age from Neoproterozoic to<br />

Eocene. Indeed, <strong>in</strong> <strong>the</strong> Zanskar Himalaya <strong>in</strong> India, Searle (1986)<br />

and Walker et al. (2001) were able to correlate thick garnet<br />

amphibolite units <strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence with<br />

unmetamorphosed Permian Panjal volcanic rocks <strong>in</strong> Kashmir,<br />

and thick high-grade marbles of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence<br />

with unmetamorphosed Triassic (and possible Jurassic) shelf<br />

carbonate units with<strong>in</strong> <strong>the</strong> Tethyan Himalaya.<br />

Nd isotopes and <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

Several workers (e.g. DeCelles et al. 2000; Rob<strong>in</strong>son et al. 2001;<br />

Richards et al. 2005) have described <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> as<br />

a ‘discrete ductile shear zone separat<strong>in</strong>g isotopically different<br />

protoliths’. Parrish & Hodges (1996) first proposed that <strong>the</strong>re<br />

was no overlap between <strong>the</strong> ranges of 143 Nd/ 144 Nd ratios between<br />

Greater <strong>Himalayan</strong> Sequence and Lesser Himalaya rocks <strong>in</strong> <strong>the</strong><br />

Langtang region. DeCelles et al. (2000) and Rob<strong>in</strong>son et al.<br />

(2001) showed that ENd(0) average values from Lesser <strong>Himalayan</strong><br />

rocks <strong>in</strong> <strong>Nepal</strong> are 21.5, whereas <strong>the</strong> Greater and Tethyan<br />

Himalaya zones <strong>in</strong> <strong>Nepal</strong> have an average ENd(0) value of 16.<br />

They suggested that <strong>the</strong> Greater <strong>Himalayan</strong> Sequence was not<br />

Indian basement, but ra<strong>the</strong>r a terrane that was accreted onto India<br />

dur<strong>in</strong>g <strong>the</strong> Early Palaeozoic, and that <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

had a large amount of pre-Tertiary displacement. However, <strong>the</strong>re<br />

is no evidence of Palaeozoic suture zone rocks (e.g. ophiolites,<br />

deep-sea sediments, etc.) anywhere along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong>, and pal<strong>in</strong>spastic reconstructions across <strong>the</strong> Western<br />

Himalaya (Searle 1986; Steck 2003), <strong>the</strong> Everest profile <strong>in</strong> <strong>Nepal</strong><br />

(Searle et al. 2006) and <strong>the</strong> Sikkim–Bhutan Himalaya (Searle &<br />

Szulc 2005) show a cont<strong>in</strong>uous sedimentary succession from<br />

proximal to distal across <strong>the</strong> Lesser, Greater and Tethyan<br />

Himalaya.<br />

With larger Nd isotope datasets, <strong>the</strong> dist<strong>in</strong>ctive differences<br />

between Greater <strong>Himalayan</strong> Sequence and Lesser Himalaya<br />

protolith start to vanish. Ahmad et al. (2000) and Richards et al.<br />

(2005) recognized a separate zone termed <strong>the</strong> ‘Outer Lesser<br />

Himalaya’ that had relatively young source rocks, similar to <strong>the</strong><br />

Greater <strong>Himalayan</strong> Sequence. They concluded that Greater<br />

<strong>Himalayan</strong> Sequence and ‘Outer Lesser Himalaya’ rocks showed<br />

a Meso-Palaeo-Proterozoic source, whereas <strong>the</strong> rest of <strong>the</strong> Lesser<br />

Himalaya showed Late Archaean to Early Proterozoic source<br />

rocks. However, Myrow et al. (2003) showed that samples from<br />

<strong>the</strong> base of <strong>the</strong> Tethyan Himalaya, north of <strong>the</strong> Greater <strong>Himalayan</strong><br />

Sequence, have similar detrital zircon age spectra and Nd<br />

isotopic data to samples from <strong>the</strong> Kathmandu klippe and Lesser<br />

Himalaya south of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence, thus elim<strong>in</strong>at<strong>in</strong>g<br />

<strong>the</strong> need for separate Greater <strong>Himalayan</strong> Sequence–Lesser<br />

Himalaya ‘terranes’. Lesser, Greater and Tethyan Himalaya<br />

represent a proximal to distal section across a cont<strong>in</strong>uous Indian<br />

plate prior to collision with Asia (Searle 1986; Myrow et al.<br />

2003; Steck 2003; Searle et al. 2006). Mart<strong>in</strong> et al. (2005) also<br />

correctly recognized that <strong>the</strong> Lesser–Greater <strong>Himalayan</strong> dist<strong>in</strong>ction<br />

was a protolith designation, fixed at <strong>the</strong> time of deposition.<br />

Because <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> is a Tertiary thrust fault that<br />

certa<strong>in</strong>ly cuts across stratigraphy, detrital zircon ages or Nd<br />

isotopes cannot be used to def<strong>in</strong>e <strong>the</strong> location of <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong>.<br />

Metamorphism, U–Th–Pb monazite ages and <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

Inverted metamorphism along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone is<br />

almost certa<strong>in</strong>ly related to movement along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong>. With<strong>in</strong> <strong>the</strong> <strong>in</strong>verted metamorphic field gradient, <strong>the</strong> rocks<br />

are highly sheared show<strong>in</strong>g ubiquitous C–S–C9 fabrics and<br />

north-plung<strong>in</strong>g l<strong>in</strong>eations that <strong>in</strong>dicate southward transport. Approximately<br />

5–8 km of thickness has been flattened by pure<br />

shear to a section 1–2 km thick along <strong>the</strong> <strong>in</strong>verted metamorphic<br />

isograd zone (Searle & Rex 1989). There are no major metamorphic<br />

discont<strong>in</strong>uities with<strong>in</strong> <strong>the</strong> <strong>in</strong>verted metamorphic sequence,<br />

suggest<strong>in</strong>g post-metamorphic pure shear flatten<strong>in</strong>g. The<br />

geometry of <strong>the</strong> <strong>in</strong>verted isograds along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

zone is similar along <strong>the</strong> entire <strong>Himalayan</strong> cha<strong>in</strong> between <strong>the</strong><br />

Zanskar–Kishtwar area <strong>in</strong> <strong>the</strong> west (Stephenson et al. 2000,<br />

2001) to <strong>Nepal</strong>, Sikkim and Bhutan <strong>in</strong> <strong>the</strong> east (e.g. Boll<strong>in</strong>ger et<br />

al. 2004; Dasgupta et al. 2004; Jessup et al. 2006). Dat<strong>in</strong>g of<br />

peak metamorphism has relied on Sm–Nd dat<strong>in</strong>g of garnet (e.g.<br />

Vance & Harris 1999), or U–Pb dat<strong>in</strong>g of monazites (eg: Walker<br />

et al. 1999; Simpson et al. 2000; Foster et al. 2002), that grew <strong>in</strong><br />

equilibrium with kyanite or sillimanite. These are <strong>the</strong> only<br />

methods that date m<strong>in</strong>erals with high enough closure tempera-


HIMALAYAN MAIN CENTRAL THRUST, NEPAL 531<br />

tures. 40 Ar/ 39 Ar ages of hornblendes or micas record only a po<strong>in</strong>t<br />

on <strong>the</strong> cool<strong>in</strong>g path after peak metamorphism, dur<strong>in</strong>g exhumation.<br />

With<strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence <strong>in</strong>itiation of garnet<br />

growth and burial metamorphism occurred at 44 Ma, with garnet<br />

rims grow<strong>in</strong>g as late as 29 3 Ma. (Pr<strong>in</strong>ce et al. 2001; Foster et<br />

al. 2002). In <strong>the</strong> Everest region peak kyanite-grade metamorphism<br />

with<strong>in</strong> <strong>the</strong> Greater <strong>Himalayan</strong> Sequence has been dated by<br />

U–Pb monazite at 32.2 0.4 Ma with later HT–LP sillimanitegrade<br />

metamorphism at 22.7 0.2 Ma (Simpson et al. 2000).<br />

Most leucogranite ages, <strong>in</strong>terpreted as dat<strong>in</strong>g peak sillimanitegrade<br />

metamorphism and migmatization, along <strong>the</strong> <strong>Nepal</strong>ese<br />

Himalaya range from c. 24 to 16 Ma (see Searle et al. 1997,<br />

2003, for reviews).<br />

Along <strong>the</strong> base of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence <strong>in</strong> <strong>Nepal</strong>,<br />

Harrison et al. (1997) first reported surpris<strong>in</strong>gly young ages of c.<br />

6 Ma from <strong>in</strong> situ 208 Pb/ 232 Th dat<strong>in</strong>g of monazite <strong>in</strong>clusions <strong>in</strong><br />

garnet. These were <strong>in</strong>terpreted as dat<strong>in</strong>g metamorphic recrystallization.<br />

Catlos et al. (2001) reported ages as young as<br />

3.3 0.1 Ma from <strong>the</strong> Marysandi Valley and Kohn et al. (2001)<br />

reported ages of 9–8 Ma from <strong>the</strong> Lesser Himalaya (re<strong>in</strong>terpreted<br />

here as <strong>the</strong> lower levels of <strong>the</strong> Greater <strong>Himalayan</strong><br />

Sequence above <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>). Th–Pb monazite ages<br />

from <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone <strong>in</strong> Langtang are c. 16Ma<br />

(Kohn et al. 2005), and <strong>the</strong> youngest monazite along <strong>the</strong> Dudh<br />

Kosi transect south of Everest has an age of 10.3 0.8 Ma<br />

(Catlos et al. 2002). In Sikkim, <strong>in</strong> situ Th–Pb monazite ages<br />

from <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone cluster at c. 22, 15–14 and<br />

12–10 Ma (Catlos et al. 2004). These workers all <strong>in</strong>terpreted <strong>the</strong><br />

Th–Pb monazite ages as dat<strong>in</strong>g metamorphism along <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong>, and hence tim<strong>in</strong>g of slip. However, Boll<strong>in</strong>ger &<br />

Janots (2006) cautioned that some young <strong>Himalayan</strong> monazites<br />

were a retrograde growth product from <strong>the</strong> breakdown of allanite<br />

at low temperatures (,370 8C). In this scenario, <strong>the</strong> Th–Pb<br />

monazite ages only date a retrograde event at low temperature<br />

and may have noth<strong>in</strong>g to do with slip along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong>. It seems quite likely that <strong>the</strong> very young monazite ages<br />

along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone may record crystallization<br />

from metasomatic hydro<strong>the</strong>rmal fluids ra<strong>the</strong>r than a ‘metamorphic<br />

event’. The ubiquitous presence of hot spr<strong>in</strong>gs along <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone today testifies to <strong>the</strong> high level of<br />

hydro<strong>the</strong>rmal fluids channelled up along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

zone, and to <strong>the</strong> <strong>in</strong>significance of shear heat<strong>in</strong>g along <strong>the</strong> thrust.<br />

The <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> system evolved through time and<br />

space. Deep crustal levels show a wide ductile shear zone <strong>in</strong><br />

kyanite-grade rocks. Higher-level brittle faults show more discrete<br />

fracture planes (e.g. <strong>the</strong> Ramgarh thrust). Rocks with young<br />

matrix monazite ages would be expected above <strong>the</strong> youngest,<br />

sou<strong>the</strong>rnmost thrust planes of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>.<br />

Restoration of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> system<br />

A generalized restoration of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> system <strong>in</strong><br />

<strong>Nepal</strong> is shown <strong>in</strong> Figure 7. Late Proterozoic rocks extend from<br />

<strong>the</strong> Lesser Himalaya to <strong>the</strong> Greater <strong>Himalayan</strong> Sequence and<br />

across to <strong>the</strong> base of <strong>the</strong> Tethyan Himalaya (Haimanta Group–<br />

Cheka Formation). Unmetamorphosed Nawakot Group sedimentary<br />

rocks <strong>in</strong> <strong>the</strong> Lesser Himalaya pass north <strong>in</strong>to <strong>the</strong> same<br />

protolith age rocks which have been metamorphosed to greenschist–upper<br />

amphibolite facies <strong>in</strong> <strong>the</strong> Ramgarh thrust sheet<br />

(Beyssac et al. 2004), up to kyanite grade <strong>in</strong> <strong>the</strong> Kathmandu<br />

thrust sheet (Johnson et al. 2001), and f<strong>in</strong>ally <strong>in</strong>to high-grade<br />

sillimanite gneisses <strong>in</strong> <strong>the</strong> Dadeldhura and Greater <strong>Himalayan</strong><br />

Sequence thrust sheets <strong>in</strong> <strong>the</strong> high Himalaya. In <strong>the</strong> <strong>in</strong>ternal parts<br />

of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence, <strong>the</strong> sillimanite gneisses<br />

(commonly referred to as Greater <strong>Himalayan</strong> Sequence Formation<br />

1; Colchen et al. 1986) are metamorphosed equivalents of<br />

<strong>the</strong> same late Proterozoic protoliths. As <strong>the</strong>re are no suture zone<br />

rocks along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> and <strong>the</strong> stratigraphy is<br />

cont<strong>in</strong>uous across <strong>the</strong> restored thrusts, <strong>the</strong>re cannot be a Palaeozoic<br />

suture zone along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> (DeCelles et al.<br />

2000; Richards et al. 2005). Instead, <strong>the</strong> restored Lesser–<br />

Greater–Tethyan Himalaya was a contiguous cont<strong>in</strong>ental section<br />

(Searle 1986; Myrow et al. 2003; Steck 2003; Searle et al.<br />

2006).<br />

One major implication of <strong>the</strong> restoration of <strong>the</strong> Himalaya is<br />

that <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> follows a flat for a long distance<br />

across strike. This flat follows a rheologically weak horizon<br />

along <strong>the</strong> Neoproterozoic shales. True Indian basement rocks<br />

(Archaean–Lower Proterozoic) are never exposed <strong>in</strong> <strong>the</strong> Himalaya.<br />

Although impossible to determ<strong>in</strong>e accurately because of <strong>the</strong><br />

high degree of ductile stra<strong>in</strong>, m<strong>in</strong>imum crustal shorten<strong>in</strong>g<br />

estimates from <strong>the</strong> Himalaya are between 500 and 900 km<br />

(Searle 1986; DeCelles et al. 2001; Rob<strong>in</strong>son et al. 2006).<br />

Because <strong>the</strong> across-strike width of <strong>the</strong> basement and cover<br />

(Neoproterozoic–Eocene rocks exposed <strong>in</strong> <strong>the</strong> Himalaya) must<br />

balance on <strong>the</strong> restoration, it must be concluded that Indian<br />

basement has underthrust northwards beneath <strong>the</strong> Himalaya north<br />

of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, and beneath <strong>the</strong> sou<strong>the</strong>rn marg<strong>in</strong> of<br />

Asia (Lhasa block) by a similar distance across strike. Very<br />

large-scale subhorizontal detachments such as <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> <strong>the</strong>refore can play a major role <strong>in</strong> <strong>the</strong> mechanical<br />

decoupl<strong>in</strong>g of <strong>the</strong> crust. In <strong>the</strong> Himalaya, <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong><br />

<strong>Thrust</strong> and <strong>the</strong> South Tibetan Detachment effectively decouple<br />

<strong>the</strong> upper (Tethyan Himalaya), middle (Greater Himalaya) and<br />

lower (Indian basement) crust.<br />

The <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>, like all major thrust faults, changes<br />

<strong>in</strong> style <strong>in</strong> both <strong>the</strong> horizontal and vertical planes (Fig. 8), as well<br />

as through time. In deep crustal profiles it is a 1–3 km thick<br />

ductile shear zone <strong>in</strong> kyanite-grade metamorphic rocks. U–Th–<br />

Pb monazite ages of <strong>the</strong>se rocks range between c. 24 and 18 Ma<br />

(for a review, see God<strong>in</strong> et al. 2006). At higher structural levels<br />

ductile shear<strong>in</strong>g passes up <strong>in</strong>to brittle thrust fault<strong>in</strong>g along a<br />

discrete thrust plane. Many of <strong>the</strong>se shallower levels follow flats<br />

plac<strong>in</strong>g similar age or younger rocks over similar age or older<br />

rocks (Fig. 7). In more sou<strong>the</strong>rly, outboard profiles such as at <strong>the</strong><br />

Kathmandu complex, <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> (Mahabharat<br />

thrust) ramps up to higher levels where metamorphism <strong>in</strong> <strong>the</strong><br />

hang<strong>in</strong>g wall is right-way-up (Fig. 8). Here, a new, younger<br />

brittle thrust (Ramgarh thrust) developed <strong>in</strong> <strong>the</strong> footwall of <strong>the</strong><br />

Mahabharat thrust, plac<strong>in</strong>g greenschist-grade rocks over unmetamorphosed<br />

Lesser <strong>Himalayan</strong> rocks (Fig. 8). In <strong>the</strong> Darjeel<strong>in</strong>g<br />

area, a late out-of-sequence breakback <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

developed beh<strong>in</strong>d <strong>the</strong> Darjeel<strong>in</strong>g klippe (Searle & Szulc 2005).<br />

All <strong>the</strong>se ductile shear zones and brittle thrust faults are part of<br />

<strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> system, which developed over a period<br />

.14 Ma (from c. 24 to 10 Ma), from depths equivalent to at least<br />

10–12 kbar (35–40 km) to <strong>the</strong> surface. Because <strong>the</strong> <strong>in</strong>verted<br />

metamorphic gradient resulted from post-metamorphic shear<strong>in</strong>g,<br />

<strong>the</strong> observed metamorphic gradient should not be directly compared<br />

with <strong>the</strong> geo<strong>the</strong>rm (Searle & Rex 1989; Boll<strong>in</strong>ger et al.<br />

2004).<br />

Conclusions<br />

None of <strong>the</strong> criteria commonly used for mapp<strong>in</strong>g <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> are valid <strong>in</strong> <strong>the</strong>mselves for def<strong>in</strong><strong>in</strong>g <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong>. Lithology, detrital zircon U–Pb ages, and Nd<br />

isotope signatures reveal <strong>in</strong>formation about provenance and


532<br />

M. P. SEARLE ET AL.<br />

stratigraphy, but not structure. Because thrust trajectories cut up<br />

and across stratigraphic section <strong>in</strong> <strong>the</strong> transport direction, <strong>the</strong>se<br />

methods are clearly not useful <strong>in</strong> def<strong>in</strong><strong>in</strong>g <strong>the</strong> position of thrust<br />

faults such as <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. Isograds are metamorphic<br />

reactions that can be mapped (with difficulty) <strong>in</strong> <strong>the</strong> field by <strong>the</strong><br />

first appearance of key <strong>in</strong>dex m<strong>in</strong>erals (sillimanite, kyanite,<br />

staurolite, garnet). Young monazite ages reveal specific <strong>in</strong>formation<br />

on growth of <strong>the</strong> garnet that armours <strong>the</strong>m, or <strong>the</strong> matrix<br />

that conta<strong>in</strong>s <strong>the</strong>m, and fluid <strong>in</strong>filtration, nei<strong>the</strong>r of which are<br />

def<strong>in</strong>itively associated with motion along a thrust fault. Only<br />

structural mapp<strong>in</strong>g and stra<strong>in</strong> <strong>in</strong>dicators can def<strong>in</strong>e <strong>the</strong> position<br />

of <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>. Follow<strong>in</strong>g Hanmer & Passchier<br />

(1991) and Passchier & Trouw (2005), <strong>the</strong> essential criteria to<br />

def<strong>in</strong>e a shear zone are <strong>the</strong> identification of a stra<strong>in</strong> gradient and<br />

<strong>the</strong> clear localization of stra<strong>in</strong>.<br />

As <strong>the</strong> metamorphic isograds are always telescoped along <strong>the</strong><br />

base of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence with up to 50% pure<br />

shear flatten<strong>in</strong>g superimposed on <strong>the</strong> already ‘frozen-<strong>in</strong>’ isograds<br />

(Jessup et al. 2006), <strong>the</strong> position of <strong>the</strong> <strong>in</strong>verted metamorphism<br />

often correlates closely, or precisely, with <strong>the</strong> position of <strong>the</strong><br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> ductile shear zone. In <strong>the</strong> western Himalaya,<br />

<strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone <strong>in</strong>verted metamorphic isograd<br />

sequence along <strong>the</strong> base of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence has<br />

been mapped around a NW-plung<strong>in</strong>g recumbent anticl<strong>in</strong>e, and<br />

has been shown to jo<strong>in</strong> up with right-way-up isograds along <strong>the</strong><br />

footwall of <strong>the</strong> South Tibetan Detachment low-angle normal fault<br />

at <strong>the</strong> top of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence (Searle & Rex<br />

1989). The map relationship and tim<strong>in</strong>g constra<strong>in</strong>ts (Hodges et<br />

al. 1996) show that movement along <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

and South Tibetan Detachment were synchronous, and that <strong>the</strong><br />

Greater <strong>Himalayan</strong> Sequence moved south, bounded by <strong>the</strong>se<br />

shear zones above and below, dur<strong>in</strong>g southward extrusion of <strong>the</strong><br />

ductile partially molten core of <strong>the</strong> Greater <strong>Himalayan</strong> Sequence<br />

(Fig. 9; channel flow model).<br />

We suggest that a common unify<strong>in</strong>g def<strong>in</strong>ition for <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> should be ‘<strong>the</strong> base of <strong>the</strong> large-scale zone of<br />

high stra<strong>in</strong> and ductile deformation, commonly co<strong>in</strong>cid<strong>in</strong>g with<br />

<strong>the</strong> base of <strong>the</strong> zone of <strong>in</strong>verted metamorphic isograds, which<br />

places Tertiary metamorphic rocks of <strong>the</strong> Greater <strong>Himalayan</strong><br />

Sequence over unmetamorphosed or low-grade rocks of <strong>the</strong><br />

Lesser Himalaya’, similar to that suggested for <strong>the</strong> Kishtwar<br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> section by Stephenson et al. (2000, 2001).<br />

Whereas <strong>the</strong> Kishtwar section shows an exhumed, deeper, more<br />

<strong>in</strong>ternal section across <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> zone, <strong>the</strong><br />

Kathmandu nappe and Ramgarh thrust sheets show a shallower,<br />

more external section across <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> (Fig. 8).<br />

The <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> ductile shear zone is commonly<br />

bounded along <strong>the</strong> south (base) by a brittle thrust fault, so a<br />

dist<strong>in</strong>ction could be made between <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong><br />

ductile shear zone (up to 2 km or more thick) and <strong>the</strong> brittle<br />

<strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong> fault (sensu stricto) along its base.<br />

We acknowledge NERC, NSF and NSERC grants respectively to M.P.S.,<br />

R.D.L. and L.G., and UK (M.J.S.), New Zealand (J.M.C.), Canadian<br />

(K.P.L.) and US (M.J.J.) PhD studentships grants. We thank <strong>the</strong> late<br />

Pasang Tamang, and Pradap Tamang and team for excellent trekk<strong>in</strong>g<br />

logistics <strong>in</strong> <strong>the</strong> Annapurnas, Tashi Sherpa and Sonam Wangdu <strong>in</strong> <strong>the</strong><br />

Everest region, and Suka Ghale and team <strong>in</strong> <strong>the</strong> Manaslu region. The<br />

paper benefited greatly from reviews by Paul Myrow and Richard Brown,<br />

and discussions with Mike Johnson, Randy Parrish and Laurent Boll<strong>in</strong>ger.<br />

Fig. 9. Generalized model for <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> ductile shear zone and thrust<br />

fault, and Greater <strong>Himalayan</strong> Sequence<br />

channel flow along <strong>the</strong> Himalaya. The<br />

South Tibetan Detachment and <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> were active simultaneously<br />

dur<strong>in</strong>g <strong>the</strong> early to middle Miocene, and <strong>the</strong><br />

deeper ductile shear zones pass upward and<br />

outward <strong>in</strong>to brittle faults with time. The<br />

mid-crustal channel of partially molten<br />

crust separates <strong>the</strong> brittle deform<strong>in</strong>g<br />

seismogenic upper crust from <strong>the</strong> rigid,<br />

high-pressure granulite lower crust of <strong>the</strong><br />

subducted Indian Shield.


HIMALAYAN MAIN CENTRAL THRUST, NEPAL 533<br />

References<br />

Ahmad, T., Harris, N., Bickle, M., Chapman, H., Bunbury, J. & Pr<strong>in</strong>ce, C.<br />

2000. Isotopic constra<strong>in</strong>ts on <strong>the</strong> structural relationships between Lesser<br />

<strong>Himalayan</strong> series and High <strong>Himalayan</strong> crystall<strong>in</strong>e series, Garhwal Himalaya.<br />

Tectonophysics, 112, 467–477.<br />

Arita, K. 1983. Orig<strong>in</strong> of <strong>the</strong> <strong>in</strong>verted metamorphism of <strong>the</strong> lower Himalaya,<br />

central <strong>Nepal</strong>. Tectonophysics, 95, 43–60.<br />

Beyssac, O., Boll<strong>in</strong>ger, L., Avouac, J.-P. & Goffé, B. 2004. Thermal<br />

metamorphism <strong>in</strong> <strong>the</strong> lesser Himalaya of <strong>Nepal</strong> determ<strong>in</strong>ed from Raman<br />

spectroscopy of carbonaceous material. Earth and Planetary Science Letters,<br />

225, 233–241.<br />

Boll<strong>in</strong>ger, L. & Janots, E. 2006. Evidence for Mio-Pliocene retrograde monazite<br />

<strong>in</strong> <strong>the</strong> lesser Himalaya, far western <strong>Nepal</strong>. European Journal of M<strong>in</strong>eralogy,<br />

18, 289–297.<br />

Boll<strong>in</strong>ger, L., Avouac, J.-P., Beyssac, O., et al. 2004. Thermal structure and<br />

exhumation history of <strong>the</strong> Lesser Himalaya <strong>in</strong> central <strong>Nepal</strong>. Tectonics, 23,<br />

doi:10.1029/2003TC001564.<br />

Bordet, P. 1961. Recherches Géologiques dans l’Himalaya du Népal, Région du<br />

Makalu. CNRS, Paris.<br />

Bouchez, J.-L. & Pêcher, A. 1981. <strong>Himalayan</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> thrust pile and its<br />

quartz-rich tectonites <strong>in</strong> central <strong>Nepal</strong>. Tectonophysics, 78, 23–50.<br />

Brunel, M. 1986. Ductile thrust<strong>in</strong>g <strong>in</strong> <strong>the</strong> Himalayas: Shear sense criteria and<br />

stretch<strong>in</strong>g l<strong>in</strong>eations. Tectonics, 5, 247–265.<br />

Brunel, M. & Kienast, J.-R. 1986. Etude pétro-structurale des chevauchements<br />

ductiles himalayaens sur la transversale de l’Everest–Makalu (<strong>Nepal</strong> oriental).<br />

Canadian Journal of Earth Sciences, 23, 1117–1137.<br />

Catlos, E.J., Harrison, T.M. & Kohn, M.J. et al. 2001. Geochronologic and<br />

<strong>the</strong>rmobarometric constra<strong>in</strong>ts on <strong>the</strong> evolution of <strong>the</strong> <strong>Ma<strong>in</strong></strong> central thrust,<br />

central <strong>Nepal</strong> Himalaya. Journal of Geophysical Research, 106, 16177–16204.<br />

Catlos, E.J., Harrison, T.M., Mann<strong>in</strong>g, C.E., Grove, M., Rai, S.M., Hubbard,<br />

M.S. & Upreti, B.N. 2002. Records of <strong>the</strong> evolution of <strong>the</strong> <strong>Himalayan</strong><br />

orogen from <strong>in</strong> situ Th–Pb ion microprobe dat<strong>in</strong>g of monazite: Eastern <strong>Nepal</strong><br />

and western Garhwal. Journal of Asian Earth Sciences, 20, 459–479.<br />

Catlos, E.J., Dubey, C.S., Harrison, T.M. & Edwards, M.A. 2004. Late<br />

Miocene movement with<strong>in</strong> <strong>the</strong> <strong>Himalayan</strong> <strong>Ma<strong>in</strong></strong> central thrust shear zone,<br />

Sikkim, nor<strong>the</strong>ast India. Journal of Metamorphic Geology, 22, 207–226.<br />

Colchen, M., LeFort, P. & Pêcher, A. 1986. Carte Géologique Annapurna–<br />

Manaslu–Ganesh, Himalaya du Népal. CNRS, Paris.<br />

Daniel, C.G., Hollister, L., Parrish, R.R. & Grujic, D. 2003. Exhumation of<br />

<strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> thrust from lower crustal depths, Eastern Bhutan Himalaya.<br />

Journal of Metamorphic Geology, 21, 317–334.<br />

Dasgupta, S., Ganguly, J. & Neogi, S. 2004. Inverted metamorphic sequence <strong>in</strong><br />

<strong>the</strong> Sikkim Himalayas: crystallization history, P–T gradient and implications.<br />

Journal of Metamorphic Geology, 22, 395–412.<br />

DeCelles, P.G., Gehrels, G.E., Quade, J., Lareau, B. & Spurl<strong>in</strong>, M. 2000.<br />

Tectonic implications of U–Pb zircon ages of <strong>the</strong> <strong>Himalayan</strong> orogenic belt <strong>in</strong><br />

<strong>Nepal</strong>. Science, 288, 497–499.<br />

DeCelles, P.G., Rob<strong>in</strong>son, D.M., Quade, J., Ojha, T.P., Garzione, C.N.,<br />

Copeland, P. & Upreti, B.N. 2001. Stratigraphy, structure and tectonic<br />

evolution of <strong>the</strong> <strong>Himalayan</strong> fold–thrust belt <strong>in</strong> western <strong>Nepal</strong>. Tectonics, 20,<br />

487–509.<br />

Foster, G., Vance, D., Harris, N. & Argles, T. 2002. The Tertiary collisionrelated<br />

<strong>the</strong>rmal history and tectonic evolution of <strong>the</strong> NW Himalaya. Journal<br />

of Metamorphic Geology, 20, 827–844.<br />

Gansser, A. 1964. Geology of <strong>the</strong> Himalaya. Wiley, Chichester.<br />

Gansser, A. 1983. Geology of <strong>the</strong> Bhutan Himalaya. Birkhauser, Basel.<br />

God<strong>in</strong>, L., Parrish, R.R., Brown, R.L. & Hodges, K.V. 2001. Crustal thicken<strong>in</strong>g<br />

lead<strong>in</strong>g to exhumation of <strong>the</strong> <strong>Himalayan</strong> metamorphic core of central <strong>Nepal</strong>:<br />

Insights from U–Pb geochronology and 40 Ar/ 39 Ar <strong>the</strong>rmochronology. Tectonics,<br />

20, 729–747.<br />

God<strong>in</strong>, L., Grujic, D., Law, R.D. & Searle, M.P. 2006. Channel flow, ductle<br />

extrusion and exhumation <strong>in</strong> cont<strong>in</strong>ental collision zones: an <strong>in</strong>troduction. In:<br />

Law, R.D., Searle, M.P. & God<strong>in</strong>, L. (eds) Channel Flow, Ductile<br />

Extrusion and Exhumation <strong>in</strong> Cont<strong>in</strong>ental Collision Zones. Geological<br />

Society, London, Special Publications, 268, 1–23.<br />

Goscombe, B.D. & Hand, M. 2000. Contrast<strong>in</strong>g P–T paths <strong>in</strong> <strong>the</strong> Eastern<br />

Himalaya, <strong>Nepal</strong>: <strong>in</strong>verted isograds <strong>in</strong> a paired metamorphic belt. Journal of<br />

Petrology, 41, 1673–1719.<br />

Goscombe, B., Gray, D. & Hand, M. 2006. Crustal architecture of <strong>the</strong> <strong>Himalayan</strong><br />

metamorphic front <strong>in</strong> eastern <strong>Nepal</strong>. Gondwana Research, 10, 232–255.<br />

Grasemann, B., Fritz, H. & Vannay, J.-C. 1999. Quantitative k<strong>in</strong>ematic flow<br />

analysis from <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> thrust zone (NW Himalaya): implications for<br />

a decelerat<strong>in</strong>g stra<strong>in</strong> path and <strong>the</strong> extrusion of orogenic wedges. Journal of<br />

Structural Geology, 21, 837–843.<br />

Hagen, T. 1954. Uber die raulmliche Verteilung der Intrusionen im <strong>Nepal</strong><br />

Himalaya. Schweizerische M<strong>in</strong>eralogische und Petrographische Mitteilungen,<br />

34, 300–308.<br />

Hanmer, S. & Passchier, C. 1991. Shear Sense Indicators: a Review. Geological<br />

Survey of Canada, Papers, 90-17.<br />

Harrison, T.M., Ryerson, F.J., LeFort, P., Y<strong>in</strong>, A., Lovera, O.M. & Catlos,<br />

E.J. 1997. A Late Miocene–Pliocene orig<strong>in</strong> for <strong>the</strong> central <strong>Himalayan</strong><br />

<strong>in</strong>verted metamorphism. Earth and Planetary Science Letters, 146, E1–E7.<br />

Hashimoto, S. 1959. Some notes on <strong>the</strong> geology and petrography of <strong>the</strong> sou<strong>the</strong>rn<br />

approach to Mt. Manaslu <strong>in</strong> <strong>the</strong> <strong>Nepal</strong> Himalaya. Journal of Science,<br />

Hokkaido University, 10, 1–15.<br />

Hashimoto, S. 1973. Geology of <strong>the</strong> <strong>Nepal</strong> Himalayas. Saikon, Apporo.<br />

Heim, A. & Gansser, A. 1939. <strong>Central</strong> Himalaya: Geological Observations of <strong>the</strong><br />

Swiss Expedition, 1936. Memoires de la Société helvetique des Sciences<br />

Naturelles, 73.<br />

Hodges, K.V., Parrish, R.R. & Searle, M.P. 1996. Tectonic evolution of <strong>the</strong><br />

central Annapurna range, <strong>Nepal</strong>ese Himalaya. Tectonics, 15, 1264–1291.<br />

Hubbard, M.S. 1996. Ductile shear as a cause of <strong>in</strong>verted metamorphism: example<br />

from <strong>the</strong> <strong>Nepal</strong> Himalaya. Journal of Geology, 104, 493–499.<br />

Jessup, M.J., Law, R.D., Searle, M.P. & Hubbard, M.S. 2006. Structural<br />

evolution and vorticity of flow dur<strong>in</strong>g extrusion and exhumation of <strong>the</strong><br />

Greater <strong>Himalayan</strong> Slab, Mount Everest Massif, Tibet/<strong>Nepal</strong>: implications for<br />

orogen-scale flow partition<strong>in</strong>g. In: Law, R.D., Searle, M.P. & God<strong>in</strong>, L.<br />

(eds) Channel Flow, Ductile Extrusion and Exhumation <strong>in</strong> Cont<strong>in</strong>ental<br />

Collision Zones. Geological Society, London, Special Publications, 268, 379–<br />

413.<br />

Johnson, M.R.W. 2005. Structural sett<strong>in</strong>gs for <strong>the</strong> contrary metamorphic zonal<br />

sequences <strong>in</strong> <strong>the</strong> <strong>in</strong>ternal and external zones of <strong>the</strong> Himalaya. Journal of<br />

Asian Earth Sciences, 25, 695–706.<br />

Johnson, M.R.W., Oliver, G.J.H., Parrish, R.R. & Johnson, S.P. 2001. Synthrust<strong>in</strong>g<br />

metamorphism, cool<strong>in</strong>g and erosion of <strong>the</strong> <strong>Himalayan</strong> Katmandu<br />

Complex, <strong>Nepal</strong>. Tectonics, 20, 394–415.<br />

Kohn, M.J., Catlos, E.J., Ryerson, F.J. & Harrison, T.M. 2001. Pressure–<br />

temperature–time discont<strong>in</strong>uity <strong>in</strong> <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> thrust zone, central<br />

<strong>Nepal</strong>. Geology, 29, 571–574.<br />

Kohn, M.J., Wieland, M.S., Park<strong>in</strong>son, C.D. & Upreti, B.N. 2005. Five<br />

generations of monazite <strong>in</strong> Langtang gneisses: implications for chronology of<br />

<strong>the</strong> <strong>Himalayan</strong> metamorphic core. Journal of Metamorphic Geology, 23, 399–<br />

406.<br />

Law, R.D., Searle, M.P. & Simpson, R.L. 2004. Stra<strong>in</strong>, deformation temperatures<br />

and vorticity of flow at <strong>the</strong> top of <strong>the</strong> Greater <strong>Himalayan</strong> slab, Everest massif,<br />

Tibet. Journal of <strong>the</strong> Geological Society, London, 161, 305–320.<br />

Law, R.D., Searle, M.P. & God<strong>in</strong>, L. (eds) 2006. Channel Flow, Ductile<br />

Extrusion and Exhumation <strong>in</strong> Cont<strong>in</strong>ental Collision Zones. Geological<br />

Society, London, Special Publications, 268.<br />

LeFort, P. 1975. Himalayas, <strong>the</strong> collided range: present knowledge of <strong>the</strong><br />

cont<strong>in</strong>ental arc. American Journal of Science, 275, 1–44.<br />

Lombardo, B., Pertusati, P. & Borgi, S. 1993. Geology and tectonomagmatic<br />

evolution of <strong>the</strong> eastern Himalaya along <strong>the</strong> Chomolangma–Makalu transect.<br />

In: Treloar, P.J. & Searle, M.P. (eds) <strong>Himalayan</strong> Tectonics. Geological<br />

Society, London, Special Publications, 74, 341–355.<br />

Mallet, F.R. 1874. On <strong>the</strong> Geology of <strong>the</strong> Darjeel<strong>in</strong>g District and <strong>the</strong> Western<br />

Duars. Memoir of Geological Survey of India, 11.<br />

Mart<strong>in</strong>, A.J., DeCelles, P.G., Gehrels, G.E., Patchett, P.J. & Isachsen, C.<br />

2005. Isotopic and structural constra<strong>in</strong>ts on <strong>the</strong> location of <strong>the</strong> <strong>Ma<strong>in</strong></strong> central<br />

thrust <strong>in</strong> <strong>the</strong> Annapurna Range, central <strong>Nepal</strong> Himalaya. Geological Society<br />

of America Bullet<strong>in</strong>, 117, 926–944.<br />

Mohan, A., W<strong>in</strong>dley, B.F. & Searle, M.P. 1989. Geo<strong>the</strong>rmobarometry and<br />

development of <strong>in</strong>verted metamorphism <strong>in</strong> <strong>the</strong> Darjeel<strong>in</strong>g–Sikkim region of<br />

<strong>the</strong> eastern Himalaya. Journal of Metamorphic Geology, 7, 95–110.<br />

Myrow, P.M., Hughes, N.C. & Paulsen, T.S. et al. 2003. Integrated<br />

tectonostratigraphic analysis of <strong>the</strong> Himalaya and implications for its tectonic<br />

reconstruction. Earth and Planetary Science Letters, 212, 433–441.<br />

Oldham, R.D. 1883. Notes on a traverse between Almora and Mussorree made <strong>in</strong><br />

October 1882. Records of <strong>the</strong> Geological Survey of India, 16, 162–164.<br />

Parrish, R.R. & Hodges, K.V. 1996. Isotopic constra<strong>in</strong>ts on <strong>the</strong> age and<br />

provenance of <strong>the</strong> Lesser and Greater <strong>Himalayan</strong> sequences, <strong>Nepal</strong>ese<br />

Himalaya. Geological Society of America Bullet<strong>in</strong>, 108, 904–911.<br />

Passchier, C.W. & Trouw, R.A.J. 2005. Microtectonics, 2nd. Spr<strong>in</strong>ger, Berl<strong>in</strong>.<br />

Pearson, O.N. & DeCelles, P.G., 2005. Structural geology and regional tectonic<br />

significance of <strong>the</strong> Ramgarh thrust, <strong>Himalayan</strong> fold–thrust belt of <strong>Nepal</strong>.<br />

Tectonics, 14, doi:10.1029/2003TC001617.<br />

Pêcher, A. 1989. The metamorphism of <strong>the</strong> <strong>Central</strong> Himalaya. Journal of<br />

Metamorphic Geology, 7, 31–41.<br />

Pognante, U. & Benna, P. 1993. Metamorphic zonation, migmatization and<br />

leucogranites along <strong>the</strong> Everest transect of Eastern <strong>Nepal</strong> and Tibet: record of<br />

an exhumation history. In: Treloar, P.J. & Searle, M.P. (eds) <strong>Himalayan</strong><br />

Tectonics. Geological Society, London, Special Publications, 74, 323–340.<br />

Pr<strong>in</strong>ce, C., Harris, N. & Vance, D. 2001. Fluid-enhanced melt<strong>in</strong>g dur<strong>in</strong>g<br />

prograde metamorphism. Journal of <strong>the</strong> Geological Society, London, 158,<br />

233–241.


534<br />

M. P. SEARLE ET AL.<br />

Richards, A., Argles, T., Harris, N., Parrish, R., Ahmad, T., Darbyshire, F.<br />

& Draganits, E. 2005. <strong>Himalayan</strong> architecture constra<strong>in</strong>ed by isotopic<br />

tracers from clastic sediments. Earth and Planetary Science Letters, 236,<br />

773–796.<br />

Richards, A., Parrish, R., Harris, N., Argles, T. & Zhang, L. 2006. Correlation of<br />

lithotectonic units across <strong>the</strong> eastern Himalaya, Bhutan. Geology, 34, 341–344.<br />

Rob<strong>in</strong>son, D.M., DeCelles, P.G., Garzione, C.N., Pearson, O.N., Harrison,<br />

T.M. & Catlos, E.J. 2001. The k<strong>in</strong>ematic evolution of <strong>the</strong> <strong>Nepal</strong>ese<br />

Himalaya <strong>in</strong>terpreted from Nd isotopes. Earth and Planetary Science Letters,<br />

192, 507–521.<br />

Rob<strong>in</strong>son, D.M., DeCelles, P.G., Patchett, J. & Garzione, C.N. 2003.<br />

K<strong>in</strong>ematic model for <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> thrust <strong>in</strong> <strong>Nepal</strong>. Geology, 31, 359–362.<br />

Rob<strong>in</strong>son, D.M., DeCelles, P.G. & Copeland, P. 2006. Tectonic evolution of <strong>the</strong><br />

<strong>Himalayan</strong> thrust belt <strong>in</strong> western <strong>Nepal</strong>: Implications for channel flow<br />

models. Geological Society of America Bullet<strong>in</strong>, 118, 865–885.<br />

Searle, M.P. 1986. Structural evolution and sequence of thrust<strong>in</strong>g <strong>in</strong> <strong>the</strong> High<br />

Himalaya, Tibetan–Tethys and Indus suture zones of Zanskar and Ladakh,<br />

Western Himalaya. Journal of Structural Geology, 8, 923–936.<br />

Searle, M.P. & God<strong>in</strong>, L. 2003. The South Tibetan Detachment and <strong>the</strong> Manaslu<br />

Leucogranite: A structural re-<strong>in</strong>terpretation and restoration of <strong>the</strong> Annapurna–Manaslu<br />

Himalaya, <strong>Nepal</strong>. Journal of Geology, 111, 505–523.<br />

Searle, M.P. & Rex, A.J. 1989. Thermal model for <strong>the</strong> Zanskar Himalaya. Journal<br />

of Metamorphic Geology, 7, 127–134.<br />

Searle, M.P. & Szulc, A.G. 2005. Channel flow and ductile extrusion of <strong>the</strong> high<br />

<strong>Himalayan</strong> slab—<strong>the</strong> Kangchenjunga–Darjeel<strong>in</strong>g profile, Sikkim Himalaya.<br />

Journal of Asian Earth Sciences, 25, 173–185.<br />

Searle, M.P., Parrish, R.R., Hodges, K.V., Hurford, A., Ayres, M.W. &<br />

Whitehouse, M.J. 1997. Shisha Pangma leucogranite, South Tibetan<br />

Himalaya: Field relations, geochemistry, age, orig<strong>in</strong> and emplacement.<br />

Journal of Geology, 105, 295–317.<br />

Searle, M.P., Waters, D.J. & Stephenson, B.J. 2002. Pressure–temperature–<br />

time path discont<strong>in</strong>uity <strong>in</strong> <strong>the</strong> <strong>Ma<strong>in</strong></strong> central thrust zone, central <strong>Nepal</strong>:<br />

Comment. Geology, 30, 480–481.<br />

Searle, M.P., Simpson, R.L., Law, R.D., Parrish, R.R. & Waters, D.J. 2003.<br />

The structural geometry, metamorphic and magmatic evolution of <strong>the</strong> Everest<br />

massif, High Himalaya of <strong>Nepal</strong>–south Tibet. Journal of <strong>the</strong> Geological<br />

Society, London, 160, 345–366.<br />

Searle, M.P., Law, R.D. & Jessup, M.J. 2006. Crustal structure, restoration and<br />

evolution of <strong>the</strong> greater Himalaya <strong>in</strong> <strong>Nepal</strong>–South Tibet: implications for<br />

channel flow and ductile extrusion of <strong>the</strong> middle crust. In: Law, R.D.,<br />

Searle, M.P. & God<strong>in</strong>, L. (eds) Channel Flow, Ductile Extrusion and<br />

Exhumation <strong>in</strong> Cont<strong>in</strong>ental Collision Zones. Geological Society, London,<br />

Special Publications, 268, 355–378.<br />

Simpson, R.L., Parrish, R.R., Searle, M.P. & Waters, D.J. 2000. Two<br />

episodes of monazite crystallization dur<strong>in</strong>g metamorphism and crustal<br />

melt<strong>in</strong>g <strong>in</strong> <strong>the</strong> Everest region of <strong>the</strong> <strong>Nepal</strong>ese Himalaya. Geology, 28, 403–<br />

406.<br />

Steck, A. 2003. Geology of <strong>the</strong> NW Indian Himalaya. Eclogae Geologicae<br />

Helvetiae, 96, 147–196.<br />

Stephenson, B.J., Waters, D.J. & Searle, M.P. 2000. Inverted metamorphism<br />

and <strong>the</strong> <strong>Ma<strong>in</strong></strong> <strong>Central</strong> <strong>Thrust</strong>: field relations and <strong>the</strong>rmobarometric constra<strong>in</strong>ts<br />

from <strong>the</strong> Kishtwar W<strong>in</strong>dow, NW Indian Himalaya. Journal of Metamorphic<br />

Geology, 18, 571–590.<br />

Stephenson, B.J., Searle, M.P. & Waters, D.J. 2001. Structure of <strong>the</strong> <strong>Ma<strong>in</strong></strong><br />

<strong>Central</strong> <strong>Thrust</strong> zone and extrusion of <strong>the</strong> High <strong>Himalayan</strong> crustal wedge,<br />

Kishtwar–Zanskar Himalaya. Journal of <strong>the</strong> Geological Society, London, 158,<br />

637–652.<br />

Stöckl<strong>in</strong>, J. & Bhattarai, K.D. 1980. Geological map of <strong>the</strong> Kathmandu area<br />

and <strong>Central</strong> Mahabharat Range 1:250 000. HM Government of <strong>Nepal</strong>,<br />

Department of M<strong>in</strong>es and Geology, Kathmandu.<br />

Upreti, B.N. & LeFort, P. 1999. Lesser <strong>Himalayan</strong> crystall<strong>in</strong>e nappes of <strong>Nepal</strong>:<br />

problems of <strong>the</strong>ir orig<strong>in</strong>. In: Macfarlane, A.M. & Sorkhabi, R.B. (eds)<br />

Himalaya and Tibet: Mounta<strong>in</strong> Roots to Mounta<strong>in</strong> Tops. Geological Society<br />

of America, Special Papers, 328, 225–238.<br />

Valdiya, K.S. 1980. Geology of <strong>the</strong> Kumaun Lesser Himalaya. Himachal Press,<br />

Wadia Institute of <strong>Himalayan</strong> Geology, Dehra Dun.<br />

Vance, D. & Harris, N. 1999. Tim<strong>in</strong>g of prograde metamorphism <strong>in</strong> <strong>the</strong> Zanskar<br />

Himalaya. Geology, 27, 395–398.<br />

von Loczy, L. 1878. Beobachtungen im östichen Himalaya. Földr Közlem, 35, 1–<br />

24.<br />

Walker, C.B., Searle, M.P. & Waters, D.J. 2001. An <strong>in</strong>tegrated tectono<strong>the</strong>rmal<br />

model for <strong>the</strong> evolution of <strong>the</strong> High Himalaya <strong>in</strong> western Zanskar with<br />

constra<strong>in</strong>ts from <strong>the</strong>rmobarometry and metamorphic modell<strong>in</strong>g. Tectonics, 20,<br />

810–833.<br />

Walker, J.D., Mart<strong>in</strong>, M.W., Bowr<strong>in</strong>g, S.A., Searle, M.P., Waters, D.J. &<br />

Hodges, K.V. 1999. Metamorphism, melt<strong>in</strong>g and extension: age constra<strong>in</strong>ts<br />

from <strong>the</strong> High <strong>Himalayan</strong> slab of SE Zanskar and NE Lahoul. Journal of<br />

Geology, 107, 473–495.<br />

Received 14 May 2007; revised typescript accepted 9 August 2007.<br />

Scientific edit<strong>in</strong>g by Rob Strachan

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!