U-Pb dating of Detrital Zircon Grains in the Paleocene Stumpata Formation, Tethyan Himalaya,...

10
U–Pb dating of detrital zircon grains in the Paleocene Stumpata Formation, Tethyan Himalaya, Zanskar, India Peter D. Clift a,b,, Andrew Carter c , Tara N. Jonell a a Department of Geology and Geophysics, E235 Howe-Russell, Louisiana State University, Baton Rouge, LA 70803, USA b South China Sea Institute of Oceanology, Chinese Academy of Sciences, No. 164 Xingang Road, Guangzhou 510301, China c Department of Earth and Planetary Sciences, Birkbeck, University of London, Malet Street, London WC1E 7HX, UK article info Article history: Received 5 September 2013 Received in revised form 29 November 2013 Accepted 16 December 2013 Available online 25 December 2013 Keywords: Provenance Zircon Himalaya Volcanism Ophiolite Collision abstract The sediments deposited on the northern margin of Greater India during the Paleocene allow the timing of collision with the Spontang Ophiolite, the oceanic Kohistan–Dras Arc and Eurasia to be constrained. U–Pb dating of detrital zircon grains from the Danian (61–65 Ma) Stumpata Formation shows a prove- nance that is typical of the Tethyan Himalaya, but with a significant population of grains from 129 ± 7 Ma also accounting for 15% of the total, similar to the synchronous Jidula Formation of south central Tibet. Derivation of these grains from north of the Indus Suture can be ruled out, precluding India’s collision with either Eurasia or the Kohistan–Dras before 61 Ma. Despite the immediate superpo- sition of the Spontang Ophiolite, there are no grains in the Stumpata Formation consistent with erosion from this unit. Either Spontang obduction is younger than previously proposed, or the ophiolite remained submerged and/or uneroded until into the Eocene. The Mesozoic grains correlate well with the timing of 130 Ma volcanism in central Tibet, suggesting that this phase of activity is linked to extension across the whole margin of northern India linked to the separation of India from Australia and Antarctica at that time. Mesozoic zircons in younger sedimentary rocks in Tibet suggest a rapid change in provenance, with strong erosion from within or north of the suture zone starting in the Early Eocene following collision. We find no evidence for strongly diachronous collision from central Tibet to the western Himalaya. Ó 2013 Elsevier Ltd. All rights reserved. 1. Introduction The collision between India, the oceanic Kohistan–Dras Arc and the mainland of Eurasia is the type example of continental collision known on the present Earth. The sedimentary sequences that form the northern margin of the Indian plate and which are preserved within the Tethyan Himalaya (Fig. 1) record the rifting, tectonic evolution and final collision of the northern passive margin of India with the Asian mainland (Chatterjee et al., 2013; Gaetani et al., 1986; Garzanti et al., 1987). As a result they form an important archive of how the margin has developed through time. Despite a large amount of work addressing the nature of the India–Eurasia collision, the timing of the different phases of this process remains controversial. While some workers propose that final collision between India and Eurasia occurred as late as 35 Ma (Aitchison et al., 2007), others argue that it was much earlier, potentially as early as 65 Ma (Cai et al., 2011; Hu et al., 2012; Jaeger et al., 1989), although a consensus of the community would place the closure of the Indus Suture Zone close to 50 Ma based on present evidence (Bouilhol et al., 2013; Clift et al., 2002a; Najman et al., 2010). Collision itself is a term that has been applied to various stages of the ocean closure process and means different things to different people. While some use this term to mean the elimination of oceanic lithosphere from between two continents (de Sigoyer et al., 2000; Leech et al., 2005), others refer to the time when the last marine sedimentation was occurring in the suture zone (Green et al., 2008). In this paper we use collision to mean the time at which all bathymetric or topographic barriers had been eliminated between India and Eurasia, thus allowing interchange of clastic sediment between the two colliding blocks. This is relatively late in the overall process and could significantly postdate the initial subduction of the distal passive margin. The section that we address in this study forms part of the old passive margin of northern (sometimes called ‘‘Greater’’) India formed during its separation from western Australia starting after 132 Ma (Ali and Aitchison, 2005; Gaina et al., 2007). The timing of this separation is relatively well-known because of the well-defined seafloor spreading anomalies found in this part of the Indian Ocean, as well as on study of the sedimentary rocks of what is now called the Tethyan Himalaya. What is less well-known 1367-9120/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.jseaes.2013.12.014 Corresponding author at: Department of Geology and Geophysics, E235 Howe- Russell, Louisiana State University, Baton Rouge, LA 70803, USA. Tel.: +1 225 578 2153. E-mail addresses: [email protected], [email protected] (P.D. Clift). Journal of Asian Earth Sciences 82 (2014) 80–89 Contents lists available at ScienceDirect Journal of Asian Earth Sciences journal homepage: www.elsevier.com/locate/jseaes

Transcript of U-Pb dating of Detrital Zircon Grains in the Paleocene Stumpata Formation, Tethyan Himalaya,...

Journal of Asian Earth Sciences 82 (2014) 80–89

Contents lists available at ScienceDirect

Journal of Asian Earth Sciences

journal homepage: www.elsevier .com/locate / jseaes

U–Pb dating of detrital zircon grains in the Paleocene StumpataFormation, Tethyan Himalaya, Zanskar, India

1367-9120/$ - see front matter � 2013 Elsevier Ltd. All rights reserved.http://dx.doi.org/10.1016/j.jseaes.2013.12.014

⇑ Corresponding author at: Department of Geology and Geophysics, E235 Howe-Russell, Louisiana State University, Baton Rouge, LA 70803, USA. Tel.: +1 225 5782153.

E-mail addresses: [email protected], [email protected] (P.D. Clift).

Peter D. Clift a,b,⇑, Andrew Carter c, Tara N. Jonell a

a Department of Geology and Geophysics, E235 Howe-Russell, Louisiana State University, Baton Rouge, LA 70803, USAb South China Sea Institute of Oceanology, Chinese Academy of Sciences, No. 164 Xingang Road, Guangzhou 510301, Chinac Department of Earth and Planetary Sciences, Birkbeck, University of London, Malet Street, London WC1E 7HX, UK

a r t i c l e i n f o

Article history:Received 5 September 2013Received in revised form 29 November 2013Accepted 16 December 2013Available online 25 December 2013

Keywords:ProvenanceZirconHimalayaVolcanismOphioliteCollision

a b s t r a c t

The sediments deposited on the northern margin of Greater India during the Paleocene allow the timingof collision with the Spontang Ophiolite, the oceanic Kohistan–Dras Arc and Eurasia to be constrained.U–Pb dating of detrital zircon grains from the Danian (61–65 Ma) Stumpata Formation shows a prove-nance that is typical of the Tethyan Himalaya, but with a significant population of grains from129 ± 7 Ma also accounting for �15% of the total, similar to the synchronous Jidula Formation of southcentral Tibet. Derivation of these grains from north of the Indus Suture can be ruled out, precludingIndia’s collision with either Eurasia or the Kohistan–Dras before 61 Ma. Despite the immediate superpo-sition of the Spontang Ophiolite, there are no grains in the Stumpata Formation consistent with erosionfrom this unit. Either Spontang obduction is younger than previously proposed, or the ophiolite remainedsubmerged and/or uneroded until into the Eocene. The Mesozoic grains correlate well with the timing of�130 Ma volcanism in central Tibet, suggesting that this phase of activity is linked to extension across thewhole margin of northern India linked to the separation of India from Australia and Antarctica at thattime. Mesozoic zircons in younger sedimentary rocks in Tibet suggest a rapid change in provenance, withstrong erosion from within or north of the suture zone starting in the Early Eocene following collision. Wefind no evidence for strongly diachronous collision from central Tibet to the western Himalaya.

� 2013 Elsevier Ltd. All rights reserved.

1. Introduction

The collision between India, the oceanic Kohistan–Dras Arc andthe mainland of Eurasia is the type example of continental collisionknown on the present Earth. The sedimentary sequences that formthe northern margin of the Indian plate and which are preservedwithin the Tethyan Himalaya (Fig. 1) record the rifting, tectonicevolution and final collision of the northern passive margin of Indiawith the Asian mainland (Chatterjee et al., 2013; Gaetani et al.,1986; Garzanti et al., 1987). As a result they form an importantarchive of how the margin has developed through time. Despite alarge amount of work addressing the nature of the India–Eurasiacollision, the timing of the different phases of this process remainscontroversial. While some workers propose that final collisionbetween India and Eurasia occurred as late as 35 Ma (Aitchisonet al., 2007), others argue that it was much earlier, potentially asearly as 65 Ma (Cai et al., 2011; Hu et al., 2012; Jaeger et al.,

1989), although a consensus of the community would place theclosure of the Indus Suture Zone close to 50 Ma based on presentevidence (Bouilhol et al., 2013; Clift et al., 2002a; Najman et al.,2010). Collision itself is a term that has been applied to variousstages of the ocean closure process and means different things todifferent people. While some use this term to mean the eliminationof oceanic lithosphere from between two continents (de Sigoyeret al., 2000; Leech et al., 2005), others refer to the time when thelast marine sedimentation was occurring in the suture zone (Greenet al., 2008). In this paper we use collision to mean the time atwhich all bathymetric or topographic barriers had been eliminatedbetween India and Eurasia, thus allowing interchange of clasticsediment between the two colliding blocks. This is relatively latein the overall process and could significantly postdate the initialsubduction of the distal passive margin.

The section that we address in this study forms part of the oldpassive margin of northern (sometimes called ‘‘Greater’’) Indiaformed during its separation from western Australia starting after132 Ma (Ali and Aitchison, 2005; Gaina et al., 2007). The timing ofthis separation is relatively well-known because of thewell-defined seafloor spreading anomalies found in this part ofthe Indian Ocean, as well as on study of the sedimentary rocks ofwhat is now called the Tethyan Himalaya. What is less well-known

Fig. 1. (A) Simplified geologic map of the Himalaya and Tibetan Plateau showing the major units referred to in the text and the location of the study area within the westernHimalaya. TM = Tso Morari, ISZ = Indus-Tsangpo Suture Zone, SSZ = Shyok Suture Zone, MCT = Main Central Thrust, MBT = Main Boundary Thrust, NP = Nanga Parbat. Map ismodified after Garzanti et al. (2005). (B) Detailed geological map of the study area modified from Fuchs (1987). Fm. = Formation, Lst. = Limestone, Sst = Sandstone,Qtz. = Quartzite.

P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89 81

is when this passive margin sedimentary package first started tocollide with the rest of Eurasia.

Controversies concerning the closure of the Neotethys north ofIndia continue. It has recently been suggested that the oceanic is-land arc of Kohistan–Dras collided first with India either at�61 Ma (Khan et al., 2009) or at �50 Ma (Bouilhol et al., 2013),rather than initially with Eurasia at 85–90 Ma, as has been moregenerally accepted (Burg, 2011; Rehman et al., 2011). Closure ofthe northern Shyok Suture between the Kohistan–Dras Arc andEurasia has traditionally been dated as being older than collisionbetween India and this same arc along the Indus Suture. However,recent geochemical evidence from the Shyok Suture raises the pos-sibility that this basin remained open until the Oligocene (�40 Ma)(Bouilhol et al., 2013).

The Indian passive margin has also been affected by the obduc-tion of ophiolites, which are preserved in the western Himalayas inthe form of the Spontang thrust sheet (Fig. 1), but its obduction his-tory has been controversial. The age of seafloor spreading of Spon-tang is dated at 177 Ma (Pedersen et al., 2001), but an andesiticisland arc, the Spong Arc, built on the Spontang basement wasformed during the Late Cretaceous (88 ± 5 Ma) and this has beenconsidered contemporaneous with the initiation of obduction ontothe passive Indian continental margin (Pedersen et al., 2001). TheSpontang Ophiolite and Spong Arc are often considered to be sep-arate tectonic entities from the oceanic Kohistan–Dras Arc, despiteboth being oceanic, Cretaceous subducted-related units. Alterna-tively if the Spontang Ophiolite is actually the forearc to the Koh-istan–Dras Arc then the Spong Arc would be a part of that arc’sproduction. Radiolarian biostratigraphic data now suggest thatthe Spong arc/Spontang ophiolite might have been active for a rel-atively long period of time from 140 Ma to �118 Ma (Baxter et al.,2010). If so then it overlaps with the ages of magmatism in theKohistan–Dras Arc that date back to 121 Ma to as young as27 Ma, but with peak activity at 80–110 Ma (Bouilhol et al.,2010; Heuberger et al.; Jagoutz et al.; Krol et al., 1996; Ravikantet al., 2009; Zeilinger et al., 2001). The obduction history itselfhas been controversial, with some workers suggesting emplace-ment of Spontang during the late Cretaceous (Corfield et al.,1999; Searle, 1986), while others argue for it being thrust overthe northern Indian margin at the time of final closure along the In-dus Suture during the Eocene (Fuchs, 1977; Garzanti et al., 1987).

In this study we focus on the origin of the sediment depositedon the northern passive margin of India during the early Paleocene.The Stumpata Quartzite is one of the youngest units that formspart of the Tethyan Sedimentary Series and is the youngest coarseclastic sedimentary rock to be found lying structurally below theSpontang Ophiolite (Fig. 2). Biostratigraphic data indicate that thissedimentary rock was deposited between 65 and 61 Ma (Greenet al., 2008). In this study we applied the U–Pb dating method tozircon sand grains extracted from the Stumpata Quartzite in orderto understand their origin and constrain the paleogeography of themargin at the time of deposition. We consider this method to beappropriate because of the unique age characteristics of the possi-ble sources on either side of the Indus Suture Zone and the effec-tiveness of this method as applied to sediment provenancewithin the Indus Suture Zone in the past (Cai et al., 2012; Hender-son et al., 2010; Hu et al., 2012; Wang et al., 2011; Wu et al., 2007).

Although zircon is not ubiquitous in crustal rocks and is espe-cially less abundant in mafic rocks, such as those composing theSpontang Ophiolite, it is sufficiently widespread to be a suitablemineral for sedimentary provenance studies. Zircon is amongstthe most stable of heavy minerals and as a result is common inmost clastic sedimentary rocks that derive from weathering ofigneous and metamorphic rocks. The inherent durability of zirconcombined with a high closure temperature (>700 �C, Hodges,2003) for the U–Pb system means that it is ideal for obtainingage information of the original source rock, although reworkingvia sedimentary deposits can be an issue when matching detritalgrain and bedrock ages (Carter and Bristow, 2001).

2. Regional geology

The Stumpata Quartzite lies towards the top of the stratigraphyof the low-grade metasedimentary rocks of the Tethyan Himalaya(Chatterjee et al., 2013) and is positioned structurally below theSpontang Ophiolite, around 30 km to the south of the Indus SutureZone (Fig. 1) (Fuchs, 1979; Reuber, 1986; Searle, 1986). The Stum-pata Quartzite is sandwiched between two carbonate sequences,sharply overlying the older Marpo/Spanboth Limestone and under-lying the Lingshed/Dibling Limestone (Gaetani et al., 1986) (Fig. 2).The Stumpata Quartzite represents the youngest siliciclastic

46474849505152535455565758596061626364656667686970

Age

(Ma)

Eoce

nePa

leoc

ene

Cre

tace

ous

Ypre

sian

Than

etia

nSe

land

ian

Dan

ian

Maa

stric

htia

n

Stumpata

Quartzite

Dib

ling

Lim

esto

ne F

orm

atio

nC

hulu

ng L

aFo

rmat

ion

Tethyan/ZanskarHimalaya

Indus SutureZone - Ladakh

Nur

laFo

rmat

ion

Juru

tze

Form

atio

n

Mar

po/S

panb

oth

Form

atio

n

Num

milit

icLi

mes

toneKong

Formation

ChogdoFormation

Sumda Fmn.

SpontangOphiolite obduction?

Asian zircons in Tethyan Himalaya

Slowing Indianconvergence

India-Kohistancollision?

India-Kohistancollision?

Tso Moririeclogite

Tethyan HimalayaCentral Tibet

Zong

zhuo

Fm

n/Zh

epur

e Sh

an F

mn

Jidu

laFo

rmat

ion

Zong

pu F

orm

aito

nEn

ba F

orm

atio

n

?

?

33343536

ThrustingNo deposition C

hoks

tiFo

rmat

ion

Olig

Sang

danl

in F

orm

atio

nZo

ngzh

uo F

mn

Zhey

a Fo

rmat

ion

?

SagaZhepure

Fig. 2. Stratigraphic time chart of the Zanskar Indian plate margin and Indus suture zone with biostratigraphic data from Green et al. (2008). Tibetan stratigraphy from theSaga area is simplified from Wang et al.(2011), while that from Zhepure is from Hu et al. (2012). Age of India–Kohistan collision from Khan et al. (2009). Age of Tso Moririeclogites from Donaldson et al. (2013). Phases of Ladakh Batholith plutonism from White et al. (2011).

82 P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89

deposit within the Tethyan Himalaya, except for the shaley KongFormation that lies immediately under the Spontang Ophiolite.The Stumpata Quartzite is a massive, thick-bedded texturally andcompositionally mature sandstone, with a high proportion ofquartz. Locally cross-bedding is visible, so that the unit is inter-preted as being a current-deposited, shallow marine sandstone,which weathers out into prominent slightly rusty-colored cliffs atthe base of the Dibling Limestone Formation, known locally asthe Lingshed Limestone (Fuchs, 1987). In the study region, thewhole Tethyan Sedimentary Sequence is deformed into a seriesof SW-vergent folds below the Spontang Ophiolite (Searle et al.,1997; Searle, 1986). The Stumpata Quartzite is dated as Danian(Early Paleocene; Fig. 2) on the basis of its assignment to theSBZ1 zone of larger benthic foraminifera by Green et al. (2008),equivalent to 60.7–65.0 Ma using the timescale of Berggren et al.(1995) or 61.7–65.5 Ma according to Gradstein et al. (2004).

3. Methods

Two large samples of the Stumpata Quartzite were taken at33�54.1100N, 76�50.6930E, approximately 2 km east of LingshedMonastery (Fig. 1B). The rock was jaw-crushed and zircons wereconcentrated using standard heavy liquid methods. Statisticallyadequate datasets of detrital zircon U–Pb ages were measured

using the London Geochronology Centre at University College Lon-don based on a New Wave Nd:YAG 213 nm laser ablation system,coupled to an Agilent 7700 quadrupole ICP-MS. Around 120 grainsare considered generally sufficient for characterizing sand erodedfrom a geologically complicated drainage basin (Vermeesch,2004). Real time data were processed using GLITTER 4.4 datareduction software. Repeated measurements of external zirconstandard Plesovice (TIMS reference age 337.13 ± 0.37 Ma) (Slámaet al., 2008) and NIST 612 silicate glass (Pearce et al., 1997) wereused to correct for instrumental mass bias and depth-dependentinter-element fractionation of Pb, Th and U. For this study206Pb/238U ages are used for those grains <1000 Ma, and for zircongrains >1000 Ma we used the 207Pb/206Pb ages to determine thecrystallization age.

In order to avoid consideration of analyses that may have sam-pled both core and rims of complex grains we chose to only plotgrains <1000 Ma whose discordance were less than 20%. For grains>1000 Ma we chose a discordance threshold of 20%. Although anal-ysis without use of cathodoluminesence imaging can result in dat-ing of old cores and younger rims together, these dates aretypically removed because they are strongly discordant. Further-more, the crucial ages in the study are the younger ones, whichcannot be contaminated with an older core. Cathodoluminesenceimaging of selected grains (Fig. 3) confirms that the Mesozoicgrains have simple structures, with either no or limited zoning

Fig. 3. Example cathodoluminesence images of Mesozoic zircons analyzed in thisstudy demonstrating their simple structure that shows their growth in a singlephase and that the young ages do not represent mixing of an older core and ayounger rim. In the case of the older grains shown the analytical spot can be seen tolie within a single growth phase.

P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89 83

and no older core. Imaging also shows that in complex grains theconcordant analyses lie clearly within a single sector of the crystal,confirming that these ages correspond to real magmatic episodesand are not simply mixtures between two end members. The ana-lytical data are provided in Appendix A.

4. Results

The spectrum of grain cooling ages from each of the consideredsamples is displayed using the Kernel Density Estimation (KDE)method of Vermeesch (2004), which plots the detrital ages as aset of Gaussian distributions, but does not explicitly take into ac-count the analytical uncertainties. This method allows the ageranges and abundances of the different age populations to begraphically assessed and different samples compared. Fig. 4 pre-sents age populations for both samples from the Stumpata Forma-tion. Both samples have similar age spectra with prominent peaksat 2.4–2.6 Ga, 1.6–2.0 Ga, 850–1050 Ma, 450–650 Ma and

Age (Ga)

Stumpata FormationSample 12072102 (N = 110)

0 1 2 3

Stumpata FormationSample 12072103 (N = 157)

Freq

uenc

y

Tethyan Himalaya (N = 3912)

Fig. 4. KDE plots for U–Pb zircon dates from the Stumpata Formation at Lingshedcompared with the Tethyan Himalayan Series compiled by Gehrels et al. (2011).

�129 ± 7 Ma. The age and relative abundance of each agepopulation are quite similar to what has been measured in otherparts of the Tethyan Himalaya (Gehrels et al., 2011), except thatthe youngest population is more abundant than typical for theTethyan Himalaya. This is unsurprising because many of the rocksin the Tethyan Himalaya were deposited before it was possible forthem to contain zircon grains as young as 130 Ma.

5. Discussion

The location of the Stumpata Formation between two marinecarbonate units and just pre-dating the onset of collision suggeststhat the sand may represent erosion from the exposed flexuralforebulge of the colliding margin, much as Hu et al. (2012)proposed for the Jidula Formation. There is no evidence that themargin would have been exposed as a result of falling sealevel at65 Ma (Haq et al., 1987) requiring some tectonically-trigger todrive uplift. In this scenario the limestones of the Marpo/SpanbothFormations and those of the Dibling Formation would representback-bulge and early foredeep deposits respectively (DeCellesand Giles, 1996). The compositional maturity, shown by thequartz-rich bulk composition is consistent with a dominantreworking of material from existing clastic sedimentary rockswithin the Greater Himalaya, and the similar zircon population.However, if the section was being exposed as a forebulge after65 Ma this implies an unrealistically high flexural rigidity to theplate if initial collision between Eurasia and India was <51 Ma asoften estimated in Ladakh (Donaldson et al., 2013). The 14 m.y.duration implies >1500 km of convergence at the rate of109 km/m.y. reconstructed by Molnar and Stock (2009). Theflexural rigidity of modern foreland is a rather high effective elasticthickness (Te) of 50–70 km (Jordan and Watts, 2005) but stillonly yield a basin 400–500 km across. A width of >1500 km isunrealistic, especially considering the likely lower Te associatedwith the rifted margin Tethyan Himalayan crust but could becaused by earlier collision with ophiolites onto the margin at thetime of Stumpata/Jidula Formation sedimentation, much as sug-gested by Ding et al. (2005).

5.1. Provenance of the Stumpata Formation

The similarity between our samples and the average for theTethyan Himalaya indicates that the Stumpata Formation wasderiving most of its sediment from the same sources as other unitswithin this tectonic block, i.e., from within the margin of northernGreater India, which is consistent with its tectonic setting.Although 280–300 Ma granitoids have been dated in the TethyanHimalaya (Horton, 2011) these are not voluminous and were likelystill buried at the time of deposition of the Stumpata Formation. Ifthe 280–300 Ma granites had been linked to volcanoes that couldhave provided material to the shelf then these were not erodedby the rivers feeding material to our study site, which may implylimited area for the drainages supplying this part of the shelf.

Here we consider the source of the youngest Mesozoic zirconsin the Stumpata Formation by comparing these with the potentialsources that could have been available since these are especiallydiagnostic of sources across the suture zone. As well as using theU–Pb ages the Th/U ratio of zircons has been commonly used toestimate their origins (Maas et al., 1992), although recent studiessuggest that Th/U values are largely reflections of protolith charac-teristics and that discrimination of igneous from metamorphic zir-cons is best done when Th/U values are combined withcathodoluminesence images (Harley et al., 2007; Schulz et al.,2006). The Mesozoic zircons analyzed here have Th/U values thataverage 0.75, although they range from a low of 0.54 to as high

Ladakh Batholith

(N = 226)

Spong Arc

Stumpata Formation

(N = 42)

Fre

quency

Indus Molasse

(N = 290)

Karakoram

(N = 259)

Tethyan Himalayan volcanic rocks - Tibet

(N = 48)

Age (Ma)0 100

Kohistan

(N = 44)Jijal Complex

Yasin Group

Suru/Nindam Fmn

Pumunong Mélange, Suture Zone

Jurassic-E. Cretaceous

(N = 65)

Rongmawa Formation, Suture Zone

L. Cretaceous

(N = 66)

Lhasa Block, Magmatic rocks

(N = 219)

Fig. 5. KDE plots for U–Pb zircon dates since 150 Ma comparing grains in theStumpata Formation with potential source regions. Tethyan volcanic rock ages arefrom Zhu et al. (2005), Liu et al. (2013) and Hu et al. (2010). Indus Molasse data isfrom Wu et al. (2007) and Henderson et al. (2010). Ladakh Batholith data is fromWhite et al. (2011) and references compiled therein. Ages from the Karakoram arefrom Fraser et al. (2001), Schärer et al. (1990), Parrish and Tirrul (1989), Ravikantet al. (2009). Lhasa Block ages are compiled by Hu et al. (2012). Kohistan data isfrom Zeilinger et al. (2001), Krol et al. (1996), Heuberger et al. (2007), Ravikant et al.(2009), Jagoutz et al. (2009) and Bouilhol et al. (2010). Age of the Jijal Complex isfrom Yamamoto and Nakamura (2000). Age of Yasin Group is from Pudsey (1986).Age of the Nindam and Suru Formations is from Reuber (1989). Age of Spong arc isfrom Pedersen et al. (2001). Ages for the Rongmawa Formation and PumunongMélange are from Cai et al. (2012).

84 P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89

as 1.22, supportive of our interpretation that the analyzed zirconshaving a single stage igneous origin. In contrast, typical metamor-phic zircons with Th/U ratios (<0.1) (Kinny et al., 1990) are absentfrom the Mesozoic population group.

The Mesozoic zircons in the Stumpata Formation, which clusteraround 130 Ma, are too young to have been derived from erosion ofthe preserved remnants of the Spontang Ophiolite (�177 Ma), andthey are too old to provide a good match with the �88 Ma andesite

known as the Spong Arc, which overlies the earlier Spontangoceanic crust (Pedersen et al., 2001) (Fig. 5). However, erosion fromthe Spongtang Ophiolite cannot be ruled out because of radiolarianbiostratigraphic dates that now argues for the magmatism beingactive from 142 to 118 Ma (Baxter et al., 2010). Nonetheless, thewell-defined age peaks seen in the Stumpata Formation is not anespecially close match to the known ages in the Spontang. Thisobservation is by itself interesting because Pedersen et al. (2001)proposed that the Spontang was emplaced onto the margin of Indiano later than the Maastrichtian (>65 Ma), that is before the sedi-mentation of the Stumpata Formation. Our data require that eitherSpontang obduction was later than proposed by this study, or thatthe ophiolite remained unexposed and uneroded immediately afterits emplacement, possibly because of a submarine location on theleading distal edge of the Indian continental margin (Fig. 6B). Inthis case it would have been thrust farther south during the finalEocene collision (Corfield et al., 1999; Searle, 1986). Whether theSpontang Ophiolite had equivalents further east is questionable,although ophiolites of similar age (Zedong and Bainang) are knownfrom central Tibet (McDermid et al., 2002; Ziabrev et al., 2004) andmight be part of a larger supra-subduction zone complex. Dinget al. (2005) dated hornblende from mafic schists from withinthe Yarlung Suture at �63 Ma, which they used to propose ophio-lite obduction around that time. This is younger than Pedersenet al.’s (2001) age for Spontang obduction, but is also too old tobe consistent with the zircon population in the Stumpata or JidulaFormations.

Most tectonic models for the closure of the Neotethys indicatethat the Stumpata Formation is too old to have received materialfrom north of the Indus Suture Zone. Our data is consistent withthis view. It is clear that the zircons in the Stumpata Formationare consistently older than those found in the Indus Molasse (Hen-derson et al., 2010; Wu et al., 2007), as well as being older than theLadakh Batholith, part of the continental Transhimalaya Arc, whichis generally considered to be the source of the Molasse (Fig. 5). Ifthe Neotethys Ocean, now represented by the Indus Suture, wasnot yet closed during Stumpata sedimentation then the absenceof these grains requires little explanation. Suture zone mélangesand sedimentary rocks along strike in central Tibet and that havehad their zircons dated (Cai et al., 2012) show a quite different ar-ray of ages, younger than 150 Ma, with many grains dated at lessthan 100 Ma, whereas there are no grains of that age in the Stum-pata or Jidula Formations.

We note that the zircons in our samples are consistently olderthan those typically found in either the Karakorum or the LhasaBlock (Fig. 5). Although Heuberger et al. (2007) have revealed zir-cons spanning 130–104 Ma in the Karakoram, ages of �130 Ma arerare in this range and our sandstone does not contain the youngerzircons that make up the bulk of the known Karakoram ages, andwhich would have been available for erosion at the time of sedi-mentation if the Karakoram had been supplying material to theTethyan Himalaya. Given that most reconstructions require theKarakoram to have been separated from the Zanskar Himalaya bytwo oceans basins (Shyok and Tethys; Fig. 7) at the time of sedi-mentation, the lack of a good match is perhaps no great surprise.

If the Mesozoic zircons are not from north of the Indus Suture orfrom the Spongtang Ophiolite, then the material could conceivablyhave been derived from the Kohistan–Dras Arc. It is noteworthythat the oldest volcanic and volcaniclastic rocks that comprisethe upper parts of the Kohistan–Dras Arc are generally older thanthe plutonic rocks that are dated in the core of the arc. The Suruand Nindam Formations of the Dras Arc (i.e., eastern part ofKohistan) are only constrained as being <144 Ma on the basis ofradiolarian ages (Honegger, 1983), although foraminifera in theNindam Formation indicate sedimentation from 137 to 93 Ma(Sutre, 1991).

Fig. 6. Schematic representation of the two possible tectonic evolutionary paths that are compatible with our new zircons data. The closure of the northern Shyok Suture iscompletely uncontrolled by our study and could have remained open until the Oligocene. K–D = Kohistan–Dras.

P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89 85

Although these ages suggest that a volcanic source of this agecould have been the source of Mesozoic zircons into the StumpataFormation, we do not favor the Kohistan–Dras Arc as a source be-cause this would require closure of the Indus suture prior to 61 Ma,when most other indicators would argue for an open Tethys Ocean(Fig. 7). In addition and more importantly, it is hard to understandwhy only the older volcanic rocks would be eroded into the Stum-pata Formation without any influence from the younger units thatdominate the preserved arc sections. The ages in the Stumpata For-mation form a relatively coherent single population that suggests asimple source rather than a mixture from a long-lived feature suchas the Kohistan–Dras Arc (Fig. 5).

5.2. Rift magmatism

The best match for our measured zircon ages is provided by vol-canic rocks and hypabyssal dikes that have been observed withinthe Tethyan Himalaya in central Tibet (Hu et al., 2010; Liu et al.,2013; Wan et al., 2011; Zhu et al., 2005). Although these are lo-cated �1350 km to the east they do provide an excellent agematch. Moreover, there is evidence for tectonic and volcanic dis-ruption of the Tethyan Himalayan shelf, as preserved in Zanskarduring Early Cretaceous time. Plate reconstructions indicate thatIndia began separating along its eastern and northern boundariesfrom western Australia starting after 132 Ma (Ali and Aitchison,2005; Gaina et al., 2007), as it rotated to the west along the Walla-by–Zenith Fracture Zone (Gibbons et al., 2012). Extension and mag-matism related to this process is the obvious cause of the �130 Mamagmatic zircons that we document with no need to invoke theinfluence of a plume (Zhu et al., 2007).

Garzanti (1993) provides evidence for uplift and tectonic reju-venation of the Tethyan Platform during the Neocomian and Barre-mian (�120–130 Ma) caused by break-up. Our evidence for EarlyCretaceous magmatism in Zanskar is also consistent with evidencefor volcanism in the deep-water Lamayuru Complex that forms thedistal edge of the continental slope and which is poorly dated

within the Cretaceous (Robertson and Degnan, 1993). We do nothowever see any evidence of erosion from a later stage of volca-nism described as trachytic and dated at 107–110 Ma (Albian)(Sutre, 1991).

5.3. Timing of collision

Our results from Zanskar now provide us the opportunity to as-sess the timing of continental collision because other data sets, lar-gely from central Tibet, have been used to argue for an earlycollision between India and Eurasia (Cai et al., 2011; Hu et al.,2012) in contrast to recent models for an Eocene closure of bothfhe Tethys and Shyok Oceans (Bouilhol et al., 2013). The ZongzhuoFormation (Fig. 2), which was deposited during the Maastrichtian(Late Cretaceous, 65–71 Ma) is documented to contain a numberof Cretaceous zircon grains with juvenile Hf isotopic compositionsand positive eNd(0) sediment. Cai et al. (2011) suggested that thisrepresented erosion from an arc or suture-zone and therefore thatcollision had already occurred at 65–71 Ma. However, we note thatany primitive magmatic source could supply a similar isotopic sig-nature, including rift volcanic rocks within the Tethyan Himalaya.Such an early age of collision is inconsistent both with interpreta-tions from the Western Himalaya (Bouilhol et al., 2013; Najmanet al., 2010) (implying very large-scale diachroneity) and withplate reconstructions that would require a prohibitively largeGreater India that is impossible to reconstruct into a pre-breakupGondwana (Ali and Aitchison, 2005).

Fig. 8 shows the development of detrital zircon U–Pb popula-tions through time over the initial collision period, mainly basedon data from central southern Tibet, but also including our newdata from Zanskar. When looking at the entire age spectrum wenote that the proportion of grains older than 400 Ma (i.e., notarc-related) is relatively high in the Zongzhuo Formation (61%),as well as in the Stumpata Formation (84%) and its along strikeequivalent, the Jidula Formation (73%), both dated as Paleocene(Danian) (Hu et al., 2012). The Jidula and the Stumpata Formations

Eurasia

GangdeseKarakoram

Shyok Sea

KDA

India

Eurasia

GangdeseKarakoram

Shyok Sea

KDA

Tethys

KLA

Eurasia

GangdeseKarakoram KKF(?)

KDA

India

Spontang

Eurasia

GangdeseKarakoram

Shyok Sea

KDA

Tethys

Spontang

Spontang

Spontang

Eurasia

GangdeseKarakoram

Shyok Sea

KDA

India

50 Ma

Eurasia

GangdeseKarakoram

Shyok Sea

KDA

Tethys

Scenario 1 Scenario 2

KLA

Eurasia

GangdeseKarakoram KKF(?)

KDA

India

40 Ma

100 Ma

Spontang

Eurasia

GangdeseKarakoram

Shyok Sea

KDA

Tethys

63 Ma

Spontang

Spontang

Spontang

Eurasia

GangdeseKarakoram

Eurasia

GangdeseKarakoram

Shyok Sea

KDA

Tethys

KLA

Eurasia

GangdeseKarakoram KKF(?)

KDA

India

Eurasia

GangdeseKarakoram

Tethys

Spontang

Spontang

40 Ma

63 Ma

100 Ma

Scenario 3

Spontang

KDA

India

Spontang

50 Ma

KDA

IndiaIndia India

IndiaIndia India

St St St

St

J

J

J

JJJ

St

St St St

Fig. 7. Cartoon representations of three possible tectonic models for closure of oceans since the Early Cretaceous in South Asia, as permitted by the data synthesized in thispaper. Modified from Bouilhol et al. (2013). KDA = Kohistan–Dras Arc. KKF = Karakoram Fault. St = Stumpata Formation outcrop. J = Jidula Formation outcrop. Closure of theShyok Ocean is not constrained by this study.

86 P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89

do share some age populations, but also show significant differ-ences. In particular, the Stumpata has significantly more grainsdating at �1.5–2.0 Ga (23% compared to 8%) and at 2.3–2.6 Ga(9% compared to 5%) and a lower proportion of the Mesozoic(16% compared to 27%) suggesting that the Jidula depositional areanear Zhepure was receiving more sediment from volcanic sourcesthan from other parts of the Tethyan Himalaya compared to the ba-sin in Zanskar. As time progressed into the Eocene the dominanceof young (i.e., arc-like, whether that be Kohistan–Dras or the La-dakh–Gangdese continental margin) grains becomes greater andgreater (47% <400 Ma in the Sangdanlin, 62% in the Enba and62% in the Zhaguo) as the focus of erosion moved from south ofthe suture zone towards the north. Looking at this larger scalewe see a major change in provenance between the Paleocene andthe Eocene, suggestive of this being a time of significant tectonictransition, as might be expected given the wide, if not universal,acceptance of suturing along the Indus Suture at �50 Ma (Fig. 7).

Fig. 9 demonstrates more clearly how the source of Mesozoicerosion changes during the collisional period. This figure in partic-ular shows how extremely similar the <150 Ma age populations arein the Jidula and Stumpata Formations and that they are likelyeroding very similar sources, which we inferred to be the rift-re-lated volcanic rocks within the Tethyan Himalaya. Consequentlythere is no evidence for India–Eurasia collision prior to �61 Ma(Fig. 7), although we cannot say anything about an earlier suturingof the Shyok Ocean between the Kohistan–Dras Arc and mainlandAsia which has traditionally been considered to have been Creta-ceous (Clift et al., 2002b; Robertson and Collins, 2002) but has re-cently been argued to have been young �40 Ma (Bouilhol et al.,2013). We infer that the juvenile grains reported by Cai et al.(2011) in the underlying Zongzhuo Formation are likely rift volca-nic-derived or are from subduction-related ophiolites or arc frag-ments obducted onto the Indian passive margin during theCretaceous. Such arc ophiolites are the along-strike equivalent to

Age (Ga)0 1 2 3

Stumpata FormationDanian/Paleocene

(N = 267)Fr

eque

ncy

Zongzhuo FormationMaastrichtian-L. Paleocene

(N = 499)

Cai et al. (2011)Wang et al. (2011)

Jidula FormationDanian/Paleocene

(N = 74)

Hu et al. (2012)

Enba FormationEocene

(N = 231)

Zhaguo FormationOligocene(N = 266)

Hu et al. (2012)Najman et al. (2010)

Hu et al. (2012)Najman et al. (2010)

Wang et al. (2011) Sangdanlin FormationPaleocene-Eocene

(N = 100)

Youn

ger d

epos

ition

al a

ges

Fig. 8. KDE plots for U–Pb zircon dates from the Stumpata Formation since 3.5 Gacompared with U–Pb ages from other sedimentary rocks deposited at thestratigraphic top of the Tethyan Himalaya in central Tibet. Data from the ZongzhuFormation is from Cai et al. (2011) and Wang et al.(2011). Data from the JidulaFormation is from Hu et al. (2012). Data from the Sangdanlin Formation is fromWang et al.(2011). Data from the Zhaguo and Enba Formations is compiled fromNajman et al. (2010) and Hu et al. (2012).

Zongzhuo FormationMaastrichtian-L. Paleocene(N = 176)

Stumpata FormationDanian/Paleocene(N = 42)

Age (Ma)0 100

Jidula FormationDanian/Paleocene(N = 74)

Enba FormationEocene(N = 100)

Zhaguo FormationOligocene(N = 128)

Sangdanlin FormationPaleocene-Eocene(N = 100)

Zheya FormationEocene(N = 72)

Youn

ger d

epos

ition

al a

ges

Fig. 9. KDE plots for U–Pb zircon dates since 150 Ma comparing grains in theStumpata Formation with those from sedimentary units in the Tethyan Himalaya inCentral Tibet. Data from the Zongazhuo Formation is from Cai et al. (2011). Datafrom Jidula Formation is from Hu et al.(2012). Data from the Sangdanlin and ZheyaFormations are from Wang et al.(2011). Data from the Zhaguo and Enba Formationsis compiled from Hu et al.(2012) and Najman et al. (2010). Grey-coloured barsrepresent time of sedimentation.

P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89 87

the Kohistan–Dras Arc, such as the Zedong Arc or DazhuquOphiolite (Aitchison et al., 2000; Ziabrev et al., 2004) which wereactive during the Barremian–Aptian (�112–127 Ma) and whichmay have been obducted onto the margin of India during the lateCretaceous (Fig. 7; Scenarios 1 and 3). It is not presently knownif grains of this provenance are also found in the western Zanskarregions.

During the transition from the Paleocene to the Eocene therewas a sharp increase in the proportion of grains dated around90 Ma and an almost complete disappearance of the �130 Mazircons that characterize the Jidula and Stumpata Formations. Thisprobably reflects increased erosion from new sources into thebasins, rather than the end of erosion of the volcanic TethyanHimalayan sources. The Sangdanlin Formation population is dom-inated by 90 Ma grains that match closely to those found withinthe Jurassic-Cretaceous Pumunong Mélange within the suture zone(Cai et al., 2012). Potential source rocks of this age also found with-in the Lhasa Block, although we do note that this age also corre-sponds to the Spong Arc in Zanskar (Pedersen et al., 2001), whichmay have had equivalents in the central part of Tibet, even if thiscannot be proven from the modern outcrop. If the grains are fromthe Lhasa Block then this would require closure of both the Shyokand Tethys Ocean by the Paleocene–Eocene (Fig. 7, Scenario 3). TheMesozoic grains within the Sangdanlin Formation are the firststrong indication of erosion from north of the suture zone and

are supportive of, but not conclusive of, collision close to thePaleocene–Eocene boundary in south central Tibet. This conclusionis consistent with the cessation of marine sedimentation in thesuture zone at 50.6–52.8 Ma estimated by Najman et al. (2010).Uncertainties in the depositional age of the Sangdanlin Formationmake it difficult to assign a precise time for the start of collisionin south central Tibet, but we generally concur with Wang et al.(2011) in favoring a collision before �50 Ma and rule out the62 Ma collision favored by Hu et al. (2012) or the 65–70 Maestimate of Cai et al. (2011) for central southern Tibet, whichwould have implied large scale diachroneity between collision incentral Tibet and that in NW India. Any collision experienced byIndia prior to 61–65 Ma must have been with an oceanic arc orophiolite complex and not with the active margin of Eurasia. Inthe event that the depositional age of the Sangdanlin Formationare revised this conclusion would have to be re-evaluated.

It is noteworthy that this timing of collision in south centralTibet is close to that estimated from Zanskar and the western

88 P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89

Himalaya (Green et al., 2008), so arguing that the degree ofdiachroneity of initial collision between the central Himalayas inthe western Himalayas is much smaller than has been suggestedin some recent studies (Hu et al., 2012), but is critically dependenton the depositional age of the Sangdanlin Formation. At present theuncertainties mean that there is no need to propose significantdiachroneity.

Finally we note that during in the Eocene and into the Oligocenesignificant numbers of grains with �60 Ma ages are deposited ontothe colliding Indian margin and preserved in the Tethyan Hima-laya. Comparison with the possible source regions (Fig. 5) showsa good match with the Ladakh/Transhimalayan Batholith, as wellas with other detrital grains in the Indus Molasse, which are them-selves inferred to have been eroded from the Ladakh Batholith(Henderson et al., 2010; Wu et al., 2007) and which were depositedat the same time in the Western Himalaya adjacent to Zanskar.These data suggest that the magmatic arc, which formed alongthe pre-collisional active margin of Eurasia was uplifted anderoded shortly after the collision starting in the Early Eocene alongthe entire suture zone. This also implies closure of the Indus Suture(Tethys Ocean) by the time of sedimentation if material is to be de-rived from these arcs onto the Indian margin. Thermochronologydata from the Ladakh Batholith are consistent with rapid erosionfrom this source at that time (Clift et al., 2002a) but does not re-quire suturing of the Shyok Ocean (Fig. 7, Scenario 2).

6. Conclusions

Dating of detrital zircons grains from the Stumpata Formationlocated under the Spontang Ophiolite of the Zanskar Himalaya inLadakh (India) indicates that this Paleocene sandstone is typicalof other sedimentary rocks from the Tethyan Himalaya, but thatit contains a significant proportion (�15%) of Mesozoic grainsmostly dating 129 ± 7 Ma. Comparison with possible sources sug-gests that these grains were probably not eroded from the nearbySpontang Ophiolite or Spong Arc, indicating that if these ophioliteand arc volcanic rocks were emplaced onto the Indian margin by61–65 Ma, they must have remained submarine and/or unerodeduntil later rethrusting after final suturing. Alternatively, the dataare consistent with ophiolite obduction after 61 Ma (Fig. 6). Ero-sion from north of the Indus Suture from either the Ladakh Batho-lith or the Kohistan–Dras Arc is considered unlikely, requiringclosure along the Indus Suture to be no earlier than 61 Ma inZanskar.

We favor the Mesozoic zircons in the Stumpata (and alongstrike equivalent Jidula) Formation being reworked from Early Cre-taceous volcanic rocks erupted as part of a phase of magmatismand extension during rifting of India from Australia and Antarctica.Synchroneity of volcanism in Zanskar and central Tibet �1350 kmapart indicates that magmatism in Tibet is part of a wider tectono-magmatic province driven by plate tectonic stresses affectingmuch of Greater India during the Early Cretaceous. The sharpreduction in the proportion of grains of this age in Eocene and Oli-gocene sedimentary rocks in the overlying units in south centralTibet (Sangdanlin, Zheya, Enba and Zhaguo Formations) is causedby enhanced erosion and sediment supply from north of and with-in the suture zone following the start of India–Eurasia collision inthe early Eocene, consistent with constraints on the collision age inLadakh. Mesozoic zircon grains within these young the units areconsistently younger than those in the Stumpata and Jidula Forma-tions and require significant erosion that cannot be accounted forby sources now known within the Tethyan Himalaya. Our new datafrom Zanskar are consistent with a broadly synchronous collisionbetween India and at least the Ladakh Batholith/Gangdese Arc inthe Western Himalaya and in south central Tibet during the EarlyEocene. Our data do not allow the age of closure along the ShyokSuture to be constrained.

Acknowledgements

We thank Pierre Bouilhol and an anonymous reviewer for theircomments in improving the original manuscript. PC and TJ thankFida Hussein Mittoo and his colleagues in Leh for their logisticalhelp during the fieldwork. The work was supported by the CharlesT. McCord Chair at Louisiana State University.

Appendix A. Supplementary material

Supplementary data associated with this article can be found,in the online version, at http://dx.doi.org/10.1016/j.jseaes.2013.12.014.

References

Aitchison, J.C., Badengzhu, D., Davis, A.M., Liu, J., Luo, H., Malpas, J., McDermid,I.M.C., Wu, H., Ziabrev, S., Zhou, M.F., 2000. Remnants of a Cretaceous intra-oceanic subduction system within the Yarlung–Zangbo suture (southern Tibet).Earth Planet. Sci. Lett. 183, 231–244.

Aitchison, J.C., Ali, J.R., Davis, A.S., 2007. When and where did India and Asia collide?J. Geophys. Res., 112.

Ali, J.R., Aitchison, J.C., 2005. Greater India. Earth-Sci. Rev. 72, 169–188.Baxter, A.T., Aitchison, J.C., Ali, J.R., Zyabrev, S.V., 2010. Early Cretaceous radiolarians

from the Spongtang massif, Ladakh, NW India: implications for Neo-Tethyanevolution. J. Geol. Soc. 167, 511–517.

Berggren, W.A., Kent, D.V., Swisher, C.C., Aubry, M.P., 1995. Revised Cenozoicgeochronology and chronostratigraphy. In: Berggren, W.A., Kent, D.V., Aubry,M.P., Hardenbol, J. (Eds.), Geochronology, Time Scales and Global StratigraphicCorrelation. Society of Economic Paleontologists and Mineralogists, pp. 129–212.

Bouilhol, P., Schaltegger, U., Chiaradia, M., Ovtcharova, M., Stracke, A., Burg, J.-P.,Dawood, H., 2010. Timing of juvenile arc crust formation and evolution in theSapat Complex (Kohistan–Pakistan). Chem. Geol. 280, 243–256.

Bouilhol, P., Jagoutz, O., Hanchar, J.M., Dudas, F.O., 2013. Dating the India–Eurasiacollision through arc magmatic records. Earth Planet. Sci. Lett. 366, 163–175.

Burg, J.P., 2011. The Asia–Kohistan–India collision: review and discussion. In: Brown,D., Ryan, P.D. (Eds.), Arc-continent Collision. Springer-Verlag, Berlin, pp. 279–309.

Cai, F., Ding, L., Yue, Y., 2011. Provenance analysis of upper Cretaceous strata in theTethys Himalaya, southern Tibet: implications for timing of India–Asia collision.Earth Planet. Sci. Lett. 305, 195–206.

Cai, F., Ding, L., Leary, R.J., Wang, H., Xu, Q., Zhang, L., Yue, Y., 2012.Tectonostratigraphy and provenance of an accretionary complex within theYarlung-Zangpo suture zone, southern Tibet: insights into subduction–accretion processes in the Neo-Tethys. Tectonophysics 574–575, 181–192.

Carter, A., Bristow, C.S., 2001. Detrital zircon geochronology: enhancing the qualityof sedimentary source information through improved methodology andcombined U–Pb and fission-track techniques. Basin Res. 12, 47–57.

Chatterjee, S., Goswami, A., Scotese, C.R., 2013. The longest voyage: tectonic,magmatic, and paleoclimatic evolution of the Indian plate during its northwardflight from Gondwana to Asia. Gondwana Res. 23, 238–267.

Clift, P.D., Carter, A., Krol, M., Kirby, E., 2002a. Constraints on India; Eurasiacollision in the Arabian Sea region taken from the Indus Group, LadakhHimalaya, India. In: Clift, P.D., Kroon, D., Gaedicke, C., Craig, J. (Eds.), Thetectonic and climatic evolution of the Arabian Sea region. Geological Society,London, pp. 97–116.

Clift, P.D., Hannigan, R., Blusztajn, J., Draut, A.E., 2002b. Geochemical evolution ofthe Dras–Kohistan Arc during collision with Eurasia; evidence from the LadakhHimalaya, India. Island Arc 11, 255–273.

Corfield, R.I., Searle, M.P., Green, O.R., 1999. Photang thrust sheet: an accretionarycomplex structurally below the Spontang ophiolite constraining timing andtectonic environment of ophiolite obduction, Ladakh Himalaya, NW India. J.Geol. Soc. 156, 1031–1044.

de Sigoyer, J., Chavagnac, V., Blichert-Toft, J., Villa, I.M., Luais, B., Guillot, S., Cosca,M., Mascle, G., 2000. Dating the Indian continental subduction and collisionalthickening in the Northwest Himalaya; multichronology of the Tso Morarieclogites. Geology 28, 487–490.

DeCelles, P.G., Giles, K.A., 1996. Foreland basin systems. Basin Res. 8, 105–123.Ding, L., Kapp, P., Wan, X.Q., 2005. Paleocene–Eocene record of ophiolite obduction

and initial India–Asia collision, South central Tibet. Tectonics, 24.Donaldson, D.G., Webb, A.A.G., Menold, C.A., Kylander-Clark, A.R.C., Hacker, B.R.,

2013. Petrochronology of Himalayan ultrahigh-pressure eclogite. Geology 41,835–838.

Fraser, J.E., Searle, M.P., Parrish, R.R., Noble, S.R., 2001. Chronology of deformation,metamorphism, and magmatism in the southern Karakoram Mountains. Geol.Soc. Am. Bull. 113, 1443–1455.

Fuchs, G., 1977. Traverse of Zanskar from the Indus to the valley of Kashmir–apreliminary note. Jahrbuch Geol. Bund. 120, 165–217.

Fuchs, G., 1979. On the geology of Western Ladakh. Jahrbuch der GeologischenWien, 513–540.

Fuchs, G., 1987. The geology of Southern Zanskar (Ladakh) – evidence for theAutochthony of the Tethys Zone of the Himalaya. Jahrbuch Geol. Bund. 130,465–491.

P.D. Clift et al. / Journal of Asian Earth Sciences 82 (2014) 80–89 89

Gaetani, M., Casnedi, P., Fois, E., Garzanti, E., Jadoul, F., Nicora, A., Tintori, A., 1986.Stratigraphy of the Tethys Himalaya in Zanskar, Ladakh. Riv. Ital. Paleontol.Stratigr. 91, 443–478.

Gaina, C., Müller, R.D., Brown, B., Ishihara, T., Ivanov, S., 2007. Breakup and earlyseafloor spreading between India and Antarctica. Geophys. J. Int. 170, 151–169.

Garzanti, E., 1993. Sedimentary evolution and drowning of a passive margin shelf(Giumal Group; Zanskar Tethys Himalaya, India): palaeoenvironmental changesduring final break-up of Gondwanaland. In: Treloar, P.J., Searle, M.P. (Eds.),Himalayan Tectonics. Geological Society, London, pp. 277–298.

Garzanti, E., Baud, A., Mascle, G., 1987. Sedimentary record of the northward flightof India and its collision with Eurasia (Ladakh Himalaya, India). Geodin. Acta 1,297–312.

Garzanti, E., Sciunnach, D., Gaetani, M., Corfield, R.I., Watts, A.B., Searle, M.P., 2005.Discussion on subsidence history of the north Indian continental margin,Zanskar-Ladakh Himalaya, NW India. J. Geol. Soc. 162, 889–892.

Gehrels, G.E., Kapp, P., DeCelles, P., Pullen, A., Blakely, R., Weislgel, A., Ding, L.,Guynn, J., Marin, A., McQuarrie, N., Yin, A., 2011. Detrital zircon geochronologyof pre-Tertiary strata in the Tibetan–Himalayan orogen. Tectonics, 30.

Gibbons, A.D., Barckhausen, U., Bogaard, P.v.d., Hoernle, K., Werner, R., Whittaker, J.M.,Müller, R.D., 2012. Constraining the Jurassic extent of Greater India: tectonicevolution of the West Australian margin. Geochem. Geophys. Geosyst., 13.

Gradstein, F.M., Ogg, J.G., Smith, A.G., 2004. A Geologic Time Scale. CambridgeUniversity Press, Cambridge.

Green, O.R., Searle, M.P., Corfield, R.I., Corfield, R.M., 2008. Cretaceous-Tertiarycarbonate platform evolution and the age of the India–Asia collision along theLadakh Himalaya (Northwest India). J. Geol. 116, 331–353.

Haq, B.U., Hardenbol, J., Vail, P.R., 1987. Chronology of fluctuating sea levels sincethe Triassic. Science 235, 1156–1167.

Harley, S.L., Kelly, N.M., Möller, A., 2007. Zircon behaviour and the thermal historiesof mountain chains. Elements 3, 25–30.

Henderson, A.L., Najman, Y., Parrish, R., BouDagher-Fadel, M., Barford, D., Garzanti,E., Andò, S., 2010. Geology of the Cenozoic Indus Basin sedimentary rocks:paleoenvironmental interpretation of sedimentation from the westernHimalaya during the early phases of India–Eurasia collision. Tectonics, 29.

Heuberger, S., Schaltegger, U., Burg, J.-P., Villa, I.M., Frank, M., Dawood, H., Hussain,S., Zanchi, A., 2007. Age and isotopic constraints on magmatism along theKarakoram–Kohistan suture zone, NW Pakistan; evidence for subduction andcontinued convergence after India–Asia collision. Swiss J. Geosci. 100, 85–107.

Honegger, K.H., 1983. Strukturen und metamorphose im Zanskarkristallin.Eidgenössiche Technische Hochschule (E.T.H.), Zürich, Switzerland, p. 117.

Horton, F.M., 2011. Geochronology and zircon geochemistry of Greater Himalayaand leucogranites in Zanskar. San Francisco State University, Northwest India,Geosciences, p. 112.

Hu, X., Jansa, L., Chen, L., Griffin, W.L., O’Reilly, S.Y., Wang, J., 2010. Provenance ofLower Cretaceous Wolong Volcaniclastics in the Tibetan Tethyan Himalaya:implications for the final breakup of Eastern Gondwana. Sed. Geol. 223, 193–205.

Hu, X., Sinclair, H.D., Wang, J., Jiang, H., Wu, F., 2012. Late Cretaceous–Palaeogenestratigraphic and basin evolution in the Zhepure Mountain of southern Tibet:implications for the timing of India–Asia initial collision. Basin Res. 24, 1–24.

Jaeger, J.-J., Courtillot, V., Tapponnier, P., 1989. Paleontological view of the ages ofthe Deccan Traps, the Cretaceous/Tertiary boundary, and the India–Asiacollision. Geology 17, 316–319.

Jagoutz, O.E., Burg, J.-P., Hussain, S., Dawood, H., Pettke, T., Iizuka, T., Maruyama, S.,2009. Construction of the granitoid crust of an island arc part I:geochronological and geochemical constraints from the plutonic Kohistan(NW Pakistan). Contrib. Miner. Petrol. 158, 739–755.

Jordan, T.A., Watts, A.B., 2005. Gravity anomalies, flexure and the elastic thicknessstructure of the India–Eurasia collisional system. Earth Planet. Sci. Lett. 236,732–750.

Khan, S.D., Walker, D.J., Hall, S.A., Burke, K.C., Shah, M.T., Stockli, L., 2009. Did theKohistan–Ladakh island arc collide first with India? Geol. Soc. Am. Bull. 121,366–384.

Kinny, P.D., Wijbrans, J.R., Froude, D.O., Williams, I.S., Compston, W., 1990. Ageconstraints on the geological evolution of the Narryer Gneiss Complex, WesternAustralia. Aust. J. Earth Sci. 37, 51–69.

Krol, M.A., Zeitler, P.K., Copeland, P., 1996. Episodic unroofing of the KohistanBatholith, Pakistan: implications from K-feldspar thermochronology. J.Geophys. Res.-Solid Earth 101, 28149–28164.

Leech, M.L., Singh, S., Jain, A.K., Klemperer, S.L., Manickavasagam, R.M., 2005. Theonset of India–Asia continental collision: early, steep subduction required bythe timing of UHP metamorphism in the western Himalaya. Earth Planet. Sci.Lett. 234, 83–97.

Liu, Y.-Q., Ji, Q., Jiang, X.-J., Kuang, H.-W., Ji, S.a., Gao, L.-F., Zhang, Z.-G., Peng, N.,Yuan, C.-X., Wang, X.-R., Xu, H., 2013. UiPb Zircon ages of Early CretaceousVolcanic Rocks in the Tethyan Himalaya at Yangzuoyong Co Lake, Nagarze,Southern Tibet, and implications for the Jurassic/Cretaceous Boundary. Cretac.Res. 40, 90–101.

Maas, R., Kinny, P.D., Williams, I.S., Froude, D.O., Compston, W., 1992. The Earth’soldest known crust: a geochronological and geochemical study of 3900–4200 Ma old detrital zircons from Mt. Narryer and Jack Hills, Western Australia.Geochim. Cosmochim. Acta 56, 1281–1300.

McDermid, I.R.C., Aitchison, J.C., Davis, A.M., Harrison, T.M., Grove, M., 2002. TheZedong terrane: a Late Jurassic intra-oceanic magmatic arc within theYarlung-Tsangpo suture zone, southeastern Tibet. Chem. Geol. 187, 267–277.

Molnar, P., Stock, J.M., 2009. Slowing of India’s convergence with Eurasia since20 Ma and its implications for Tibetan mantle dynamics. Tectonics, 28.

Najman, Y., Appel, E., Boudagher-Fadel, M., Bown, P., Carter, A., Garzanti, E., Godin,L., Han, J., Liebke, U., Oliver, G., Parrish, R., Vezzoli, G., 2010. Timing of India–Asiacollision: geological, biostratigraphic, and palaeomagnetic constraints. J.Geophys. Res..

Parrish, R.R., Tirrul, R., 1989. U–Pb age of the Baltoro Granite, Northwest Himalaya,and implications for monazite U–Pb systematics. Geology 17, 1076–1079.

Pearce, N.J.G., Perkins, W.T., Westgate, J.A., Gorton, M.P., Jackson, S.E., Neal, C.R.,Chenery, S.P., 1997. A compilation of new and published major and traceelement data for NIST SRM 610 and NIST SRM 612 glass reference materials.Geostandards Newslett.-J. Geostandards Geoanal. 21, 115–144.

Pedersen, R.B., Searle, M.P., Corfield, R.I., 2001. U–Pb zircon ages from the SpontangOphiolite, Ladakh Himalaya. J. Geol. Soc. 158, 513–520.

Pudsey, C.J., 1986. The Northern Suture, Pakistan: margin of a Cretaceous island arc.Geol. Mag. 123, 405–423.

Ravikant, V., Wu, F.Y., Ji, W.Q., 2009. Zircon U–Pb and Hf isotopic constraints onpetrogenesis of the Cretaceous–Tertiary granites in eastern Karakoram andLadakh, India. Lithos 110, 153–166.

Rehman, H.U., Seno, T., Yamamoto, H., Khan, T., 2011. Timing of collision of theKohistan–Ladakh Arc with India and Asia: Debate. Island Arc 20, 308–328.

Reuber, I., 1986. Geometry and accretion and oceanic thrusting of the SpontangOphiolite. Nature 321, 592–596.

Reuber, I., 1989. The Dras arc: two successive volcanic events on eroded oceaniccrust. Tectonophysics 161, 93–106.

Robertson, A.H.F., Collins, A.S., 2002. Shyok suture zone, N Pakistan; late Mesozoic–Tertiary evolution of a critical suture separating the oceanic Ladakh Arc fromthe Asian continental margin. J. Asian Earth Sci. 20, 309–351.

Robertson, A.H.F., Degnan, P.J., 1993. Sedimentology and tectonic implications ofthe Lamayuru complex; deep-water facies of the Indian passive margin, Indussuture zone, Ladakh Himalaya. In: Treloar, P.J., Searle, M.P. (Eds.), HimalayanTectonics. Geological Society, London, pp. 299–321.

Schärer, U., Copeland, P., Harrison, T.M., Searle, M.P., 1990. Age, cooling history, andorigin of post-collisional leucogranites in the Karakoram Batholith; a multi-system isotope study. J. Geol. 98, 233–251.

Schulz, B., Klemd, R., Braetz, H., 2006. Host rock compositional controls on zircontrace element signatures in metabasites from the Austroalpine basement.Geochim. Cosmochim. Acta 70, 697–710.

Searle, M.P., 1986. Structural evolution and sequence of thrusting in the HighHimalayan, Tibetan-Tethys and Indus Suture Zones of Zanskar and Ladakh,Western Himalaya. J. Struct. Geol. 8, 923–937.

Searle, M., Corfield, R.I., Stephenson, B., McCarron, J., 1997. Structure of the NorthIndian continental margin in the Ladakh-Zanskar Himalayas: implications forthe timing of obduction of the Spontang ophiolite, India–Asia collision anddeformation events in the Himalaya. Geol. Mag. 134, 297–316.

Sláma, J., Košler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M., Horstwood,M.S.A., Morris, G.A., Nasdala, L., Norberg, N., Schaltegger, U., Schoene, B., Tubrett,M.N., Whitehouse, M.J., 2008. Plezovice zircon A new natural reference materialfor U–Pb and Hf isotopic microanalysis. Chem. Geol. 249, 1–35.

Sutre, E., 1991. Les formations de la marge nord-Neotethsienne et les mélangesophiolitique de la zone de suture de l’Indus en Himalaya du Ladakh. L’Universitéde Poitiers, p. 345.

Vermeesch, P., 2004. How many grains are needed for a provenance study? EarthPlanet. Sci. Lett. 224, 351–441.

Wan, X.Q., Scott, R., Chen, W., Gao, L., Zhang, Y., 2011. Early Cretaceous stratigraphyand SHRIMP U–Pb age constrain the Valanginiane Hauterivian boundary insouthern Tibet. Lethaia 44, 231–244.

Wang, J., Hu, X.-M., Jansa, L., Huang, Z., 2011. Provenance of the UpperCretaceous–Eocene deep-water sandstones in Sangdanlin, southern Tibet:constraints on the timing of initial India–Asia collision. Journal of Geology119, 293–309.

White, L.T., Ahmad, T., Ireland, T.R., Lister, G.S., Forster, M.A., 2011. Deconvolvingepisodic age spectra from zircons of the Ladakh Batholith, Northwest IndianHimalaya. Chem. Geol. 289, 179–196.

Wu, F.Y., Clift, P.D., Yang, J.H., 2007. Zircon Hf isotopic constraints on the sources ofthe Indus Molasse, Ladakh Himalaya, India. Tectonics, 26.

Yamamoto, H., Nakamura, E., 2000. Timing of magmatic and metamorphic events inthe Jijal complex of the Kohistan arc deduced from Sm-Nd dating of maficgranulites. In: Khan, M.A., Treloar, P.J., Searle, M.P., Jan, M.Q. (Eds.), Tectonics ofthe Nanga Parbat Syntaxis and the Western Himalaya. Geological Society,London, pp. 313–319.

Zeilinger, G., Burg, J.P., Schaltegger, U., Seward, D., 2001. New U/Pb and fission trackages and their implication for the tectonic history of the lower Kohistan ArcComplex, northern Pakistan. J. Asian Earth Sci. 19, 79–81.

Zhu, D.C., Pan, G.T., Mo, X.X., Wang, L.Q., Liao, Z.L., Jiang, X.S., Geng, Q.R., 2005.SHRIMP U–Pb zircon dating for the dacite of the Sangxiu Formation in thecentral segment of Tethyan Himalaya and its implications. Chin. Sci. Bull. 50,563–568.

Zhu, D.C., Pan, G.T., Mo, X.X., Liao, Z.L., Jiang, X.S., Wang, L.Q., Zhao, Z.D., 2007.Petrogenesis of volcanic rocks in the Sangxiu Formation, central segment ofTethyan Himalayas: a probable example of plume-lithosphere interaction. J.Asian Earth Sci. 29, 320–335.

Ziabrev, S., Aitchison, J.C., Abrajevitch, A.V., Badengzhu, Davis, A.M., Luo, H., 2004.Bainang Terrane, Yarlung – Tsangpo suture, southern Tibet (Xizang, China): arecord of intra-Neotethyan subduction – accretion processes preserved on theroof of the world. J. Geol. Soc., London 161, 523–538.