mountain building, erosion, and the seismic cycle in …avouac/classes/ge277/avouac04.pdf · india...

80
MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN THE NEPAL HIMALAYA JEAN-PHILIPPE AVOUAC Californian Institute of Technology, Pasadena, USA 1. INTRODUCTION The Himalaya is the most impressive example on earth of an active collisional orogen. It combines rapid crustal shortening and thickening, intense denudation driven by the monsoon climate, and frequent very large earthquakes along an incomparably long and high mountain arc. It has therefore been the focus of a variety of investigations that have addressed various aspects of mountain building on various timescales. Geological and geophysical studies give some idea of the structure of the range and physical properties at depth. The long-term geological history of the range, over say several millions to a few tens of millions of years, has been documented by structural, thermobarometric, and thermochronological studies. Morpho- tectonic investigations have revealed its evolution over several thousands or tens of thousands of years; and geodetic measurements and seismological monitoring have revealed the pattern of strain and stress built-up over several years. This chapter is an attempt to show that the results of these investigations can be assembled into a simple and coherent picture of the structure and evolution of the range. The author also intends to illustrate the interplay between these various processes operating at different time- scales. One important example of processes that interact via feedback mechanisms is particularly clear in the Himalaya: the thermal structure of the range, which is a result of the long-term crustal deformation and pattern of exhumation, governs, through its influence on rheology, the pattern of deformation as well as the seismic behavior of the range-bounding thrust fault. In this chapter the key role played by surface processes is emphasized. These processes have carved morphologic features that can be used to deduce vertical displacements. They also have generated the molasse deposits that have filled the subsiding foreland basin, providing a record of mountain building. In addition, they have participated actively in the evolution of the ADVANCES IN GEOPHYSICS, VOL. 46 1 ß 2003 Elsevier Inc. All rights reserved ISSN: 1474-8177 DOI: 10.1016/S0065-2687(03)46001-9

Upload: others

Post on 29-May-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

MOUNTAIN BUILDING, EROSION,AND THE SEISMIC CYCLE IN

THE NEPAL HIMALAYA

JEAN-PHILIPPE AVOUAC

Californian Institute of Technology, Pasadena, USA

1. INTRODUCTION

The Himalaya is the most impressive example on earth of an activecollisional orogen. It combines rapid crustal shortening and thickening,intense denudation driven by the monsoon climate, and frequent very largeearthquakes along an incomparably long and high mountain arc. It hastherefore been the focus of a variety of investigations that have addressedvarious aspects of mountain building on various timescales. Geological andgeophysical studies give some idea of the structure of the range and physicalproperties at depth. The long-term geological history of the range, over sayseveral millions to a few tens of millions of years, has been documented bystructural, thermobarometric, and thermochronological studies. Morpho-tectonic investigations have revealed its evolution over several thousands ortens of thousands of years; and geodetic measurements and seismologicalmonitoring have revealed the pattern of strain and stress built-up overseveral years. This chapter is an attempt to show that the results of theseinvestigations can be assembled into a simple and coherent picture of thestructure and evolution of the range. The author also intends to illustratethe interplay between these various processes operating at different time-scales. One important example of processes that interact via feedbackmechanisms is particularly clear in the Himalaya: the thermal structure ofthe range, which is a result of the long-term crustal deformation and patternof exhumation, governs, through its influence on rheology, the pattern ofdeformation as well as the seismic behavior of the range-bounding thrustfault.

In this chapter the key role played by surface processes is emphasized.These processes have carved morphologic features that can be used to deducevertical displacements. They also have generated the molasse deposits thathave filled the subsiding foreland basin, providing a record of mountainbuilding. In addition, they have participated actively in the evolution of the

ADVANCES IN GEOPHYSICS, VOL. 46

1

� 2003 Elsevier Inc. All rights reservedISSN: 1474-8177

DOI: 10.1016/S0065-2687(03)46001-9

Page 2: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

range by influencing the thermal structure and the stress field throughredistribution of mass at the earth surface. Surface processes must thereforebe taken into account in any analysis of the mechanics of mountain build-ing. They are also probably the major factors that differentiate intra-continental megathrust from subduction zones.

This chapter is not a comprehensive review of the Himalaya. Forpedagogic reasons the author mainly describes studies carried out acrossthe Himalaya of central Nepal because he is most familar with this area thathas attracted particular attention over the last decade.

The author first briefly introduces in Section 1 the geodynamic settingand presents in Section 2 a model of the development of the Himalayanorogen and foreland basin. In Section 3 the structure and kinematicevolution of the Himalaya as constrained from surface geology, geo-chronology, and geophysical investigations is reviewed in more detail.Section 4 describes how the kinematics of thin-skinned deformationalong the Himalayan foothills can be determined from the deformation ofabandoned river terraces. In Section 5 the pattern of river incision acrossthe whole range is described and it is shown that it basically reflects thekinematics of overthrusting. Section 6 discusses geodetic measurements ofcrustal deformation, historical seismicity, and the seismic cycle along theHimalaya. Section 7 discusses some general questions about continentaldeformation and seismicity:

– How is deformation distributed throughout the range, and where are thefaults capable of producing very large recurrent earthquakes?

– What can we learn about future large earthquakes from seismicity anddeformation monitored over a limited period of time?

– What proportion of crustal deformation is expressed in the seismicity?– Starting with recent deformation, measured over a decade with geodetic

techniques, can we extrapolate backwards to explain the long-termhistory of the range as expressed in its structural geology?

– How does the erosion rate compare with tectonic uplift?

2. AN ACTIVE COLLISIONAL OROGEN

2.1. Geodynamical Setting and Key Structural Features

The Himalayan arc is one of the major zones of deformation thathave absorbed the indentation of India into Eurasia (e.g., Powell and

2 AVOUAC

Page 3: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Conaghan, 1973). The collision started about 50 Myr ago and produceda combination of lateral escape and crustal thickening that has givenrise to the highest topographic features on earth (e.g., Molnar andTapponnier, 1975; Harrison et al., 1992; Tapponnier et al., 2001) (Fig. 1).Subsequently, India and stable Eurasia continued converging at a rateof about 5 cm/year (Patriat and Achache, 1984). At present, the 4–5 cm/yearof northward displacement of India relative to stable Eurasia is stillbeing absorbed by a combination of horizontal shear and crustalshortening. This is demonstrated both by the pattern and kinematics ofactive faults in Asia (Molnar and Tapponnier, 1975; Avouac andTapponnier, 1993) and by GPS measurements (Larson et al., 1999)(Fig. 2a), although the respective contribution of these two mechanismsto the overall deformation remains a matter of debate (Tapponnier et al.,2001; Wang et al., 2002). Across the Himalaya of central Nepal a fractionof this convergence (estimated at about 2 cm/year) is absorbed bycrustal shortening, as shown from GPS geodetic campaigns carried out overthe last decade (Fig. 2b). Ongoing crustal shortening across the Himalaya isalso manifested by recurring large earthquakes with magnitude Mw

FIG. 1. Sketch showing how indentation of India into Eurasia, since the onset of the

collision about 50 Myr ago, has been absorbed by a combination of crustal thickening and

lateral escape.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 3

Page 4: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

above 8, such as the Bihar–Nepal earthquake of 1934 or the Kangraearthquake of 1905 (Fig. 3, Table 1).

Relics of the Tethys ocean that used to separate the northern margin ofIndia from the active southern margin of Eurasia can now be tracedalong the Indus-Tsangpo suture zone (ITSZ) (e.g., Burg, 1983; Searle et al.,1987) well north of the Himalayan summits (Figs. 1 and 4). To thesouth, Cambrian to Eocene Tethyan sediments deposited on the northernpassive margin of the Indian continent were sutured to the volcanic andplutonic rocks of the once-active margin of Eurasia (Burg et al., 1987; Searleet al., 1987). Now lying at elevations around 5000 m, they wereintensely deformed (Ratschbacher, 1994), probably in the early ‘‘Alpine’’period of the collision which lead to the development of the North

FIG. 2. (a) Velocities relative to stable Eurasia measured from GPS over about the last

10 years, compiled in 2001 (data from Wang et al., 2001). All velocities were expressed in

ITRF 97. (b) Velocity relative to India determined from GPS measurements by IDYL-

HIM (Jouanne et al., 1999), LDG (Avouac et al., 2001), and CIRES (Larson et al., 1999). All

velocities were expressed in ITRF 97. Insert shows all measurements projected on section AA0

with accounting for the arc curvature as defined from the arcuate shape of the front of the high

range (Bilham et al., 1997).

4 AVOUAC

Page 5: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Himalayan Nappe zone (Burg et al., 1984; Yin et al., 1999). There isevidence that thrust faulting and crustal thickening in southern Tibetpersisted until mid-Miocene times (Yin et al., 1999). Extension perpendi-cular to the range has also been documented along a normal fault thatseparates the Tethyan sedimentary cover from the High Himalayancrystalline units. This fault is called the North Himalayan Normal fault(Burg et al., 1984) or the South Tibetan Detachment (Burchfiel et al., 1992).In this chapter, these extensional processes that are thought to haveresumed between about 15 and 20 Myr ago are not discussed (Searle et al.,1997); rather, the focus is on the subsequent deformation that affectedthe Himalayan orogen. To the south, crustal shortening chiefly resultedfrom deformation on a limited number of major thrust faults: from north tosouth, these are the Main Central Thrust fault (MCT), the MainBoundary Thrust fault (MBT), and the Main Frontal Thrust fault (MFT)(Fig. 4) (e.g., Gansser, 1964; Le Fort, 1975a; Nakata, 1989; Yeats et al.,1992). To the first order, these major thrust faults were activated in aforward propagation sequence. This geometry has prompted the viewthat the Himalaya should be seen as a crustal-scale accretionary prism

FIG. 2. Continued.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 5

Page 6: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

(Davis et al., 1983), with a structure like that produced when a rigidbackstop overthrusts a rigid basement with overlying sand layers (Fig. 5).According to this model, a mountain range is a critical wedge with steady-state geometry reflecting the balance between frictional stresses at base of

FIG. 3. Map of Himalayan arc and Gangetic plain with location of section AA0 and

INDEPTH seismic profiles (e.g., Hauck et al., 1998). The Main Frontal Thrust fault (MFT)

emerges along the front of the Siwalik hills that form the sub-Himalaya. The front of the high

range, defined here as the 3500-m elevation contour line, lies about 100 km north of MFT.

Shaded areas show proposed ruptured areas of the M>8 earthquakes of 1905 and 1934 (see

parameters in Table 1).

TABLE 1. Estimated Parameters of Major Himalayan Earthquakes (Mw>8) since 1897

(Pandey and Molnar, 1988; Chander, 1989; Molnar, 1990; Bilham, 1995; Ambraseys and

Bilham, 2000, Bilham and England, 2001)

Date Epicenter L (km) l (km) M0� 1021 (Nm) Mw

1897 � 26�N � 91�E 200–300 � 100 03 8.0–8.7

1905 33�N 76�E 280 (>120) 80 2 8.2

1934 27.6�N 87.1�E 250–300 100 4 8.4

1950 28.4�N 96.8�E 200 120 8 8.7

6 AVOUAC

Page 7: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

the wedge and shear stress induced by the topographic slope (Chapple,1978; Davis et al., 1983; Dahlen and Suppe, 1988; Dahlen, 1990).Crustal thickening results from frontal accretion by southwardpropagation of the deformation front and from internal thickening to

FIG. 4. Schematic geological section across the Himalaya of central Nepal reflecting early

interpretation of the major thrust faults as crustal-scale parallel faults.

FIG. 5. Example of an analog modeling of growth of a brittle wedge. A rigid backstop

moves to the left over a fixed rigid basement inducing detachment of the sand layers. Note the

forward propagation sequence of thrusting (courtesy of Stephane Dominguez, ISTEEM,

Montpellier).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 7

Page 8: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

maintain a constant critical slope. If erosion is taken into account, then thewhole wedge might reach some steady-state geometry with particletrajectories like those pictured in Fig. 6. Horizontal shortening, andhence active thrust faults, would be distributed throughout the prism. Iferosion is assumed to be distributed uniformly, the pattern of horizontaldisplacement can be expressed analytically based on a simple mass balanceyielding,

VðxÞ ¼ V0hx=ððhþ ðxþWÞtg’Þ, ð1Þ

where the origin of abscissa is taken at the backstop, assumed to be fixed.V0 is the convergence rate, hence the velocity of the leading edge of thewedge. Fig. 6 shows the predicted horizontal surface velocity for h¼ 15 km,W¼ 200 km, tg(�þ �)¼ 4%, and V0¼ 20 mm/year. This set of parameterscorresponds approximately to the geometry and size of the Himalayan

FIG. 6. Particle trajectories in a steady-state critical accretionary prism subjected to erosion

(Barr et al., 1991). Horizontal surface velocities computed from Eq. (1) for h¼ 15 km,

W¼ 200 km, tg(�þ�)¼ 4%, and V0¼ 20 mm/year, and an erosion rate of 1.5 mm/year.

8 AVOUAC

Page 9: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

wedge of central Nepal, assuming that about half of the underthrustingIndian crust is incorporated into the Himalayan wedge. For the wedge to bein steady state, a 2-D erosion flux of 300 m2/year is required (per unit lengthof the Himalayan arc). This is in the range of possible value inferred fromsediment budget over the last 2 Myr at the scale of the whole Himalayanrange (Metivier et al., 1999) or from denudation rates derived from riversediment discharge in central Nepal (Lave and Avouac, 2001). This valuecorresponds to an erosion rate of about 1.5 mm/year on average over thesection.

Over the last few years several models have been proposed, all based onassumption that the crust is detached from a subducting upper mantlewith prescribed kinematics (e.g., Willet et al., 1993). Various rheologies,possibly with some depth dependence, as well as various distributionsof erosion across the surface, have been considered (Williams et al., 1994;Willet, 1999). Although they differ as to the details of the predictedkinematics or steady-state wedge geometry, these various models wouldpredict a more- or less-distributed shortening throughout the wedge, asexpressed by Eq. (1). The geodetic measurements in Fig. 2b show distributedcontraction similar to what is predicted from Eq. (1). However, it shouldbe realized that this comparison does not make any physical sense. Thegeodetic measurements represent the accumulation of interseismic strain,part of which is elastic deformation that will ultimately be released bycoseismic deformation. Equation (1), in contrast, represents unrecover-able strain due to brittle deformation of the sand layers. So, Eq. (1)should rather be compared to the averaged displacement field that results,over the long term, from accumulated interseismic and coseismicdeformation. This analysis therefore requires identification of the majoractive faults across the Himalaya and the determination of the slip rateson these faults.

2.2. Foreland Deposition During Overthrusting

Sedimentation in the foreland basin provides indirect informationon the growth of an orogen. As the Himalayan wedge grows, it over-thrusts and flexes down the Indian basement forming a ‘‘foreland flexuralbasin’’ in which a fraction of the material eroded from the range accumu-lates with a stratigraphic organization that depends on the kinematicsof overthrusting (Lyon-Caen and Molnar, 1985) (Fig. 7). Several kilometers

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 9

Page 10: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

of Cenozoic molasse (Murrees and Siwaliks formations) have thusaccumulated on the Precambrian Indian basement. These molassecrop out along the foothills where they were scraped off the basement andfolded as schematized in Fig. 9. They consist of an upward coarseningsequence (claystone, siltstone, sandstone, and conglomerates) ofupper Miocene to Pleistocene age. As shown in Fig. 7, the developmentof the flexural basin during underthrusting implies some progradationaway from the mountain front of sediment facies and of the contactbetween the basement and the most distal sediments. This modelimplies, as observed, an upward gradation from distal to more proximalfacies.

If the age of the oldest sediments lying over the pre-Tertiary basementis plotted as a function of the distance perpendicular to the range, the rate

FIG. 7. Sketch showing how sediments prograde over the flexed Indian basement during

overthrusting and Himalayan wedge growth. The flexed Indian lithosphere makes a forebulge at

about 250–300 km from the Himalayan front where most recent foreland sediments are starting

to overlap on the Indian basement (X3). The sediment progradation rate, Vpr, can be estimated

by plotting the ages (t1, t2) of the oldest foreland sediments overlying the basement as a function

of distance (x1, x2) perpendicular to the range front. Assuming that the foreland’s geometry is

entirely constrained by loading due to the weight of the Himalayan topography, Vpr depends

on the rate of convergence with the front of the high range, VHR, and on the growth of the

mountain wedge. W is the width of the accretionary prism.

10 AVOUAC

Page 11: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

of sediment progradation can be deduced (Fig. 7). The southwardprogradation over about the last 15 Myr was estimated in this way tobe between 10 and 15 mm/year (Lyon-Caen and Molnar, 1985), based on acompilation of well data in the Gangetic foreland.

In Fig. 8, only the well data are close to the study area shown in Fig. 3.The front of the high range (defined from the 3500-m elevation contourline shown in Fig. 3) is also indicated because the topographic load overthe Indian plate has probably been the main factor controlling the shapeof the flexed basement. This plot is not very well constrained since theage of the youngest sediments on top of the basement was estimated withvery few chronological constraints and with the assumption of constantsedimentation rates (Lyon-Caen and Molnar, 1985). Two additional pointswere added based on the stratigraphic sections along the Surai Khola

FIG. 8. Age of oldest Cenozoic foreland deposits overlying the Indian basement as a function

of their present distance from the front of the high range (defined from the 3500-m elevation

contour line shown in Fig. 3). See Fig. 2 for location of wells (modified from Lyon-Caen, 1985).

Two additional points were added to this plot based on the stratigraphic sections along the

Surai Khola (Appel et al., 1991a) and along the Bakeya (Harrison et al., 1993), as explained in

Fig. 9 (modified from Lave, 1997).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 11

Page 12: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

(Appel et al., 1991b) and along the Bakeya (Harrison et al., 1993) (seelocation in Fig. 3); these were both dated from magnetostratigraphy, andcan be used to infer sedimentation rates and hence progradation rates(Lave, 1997). If the geometry of the foreland basin is assumed to be steadystate, the stratigraphic thickness, h*, of the foreland sediments depositedbefore t1–2 might be used to estimate the distance at time t1–2 of that sectionfrom the front of the high range. As shown in Fig. 9, this estimate requiressome restoration of the fault slip on the MFT. The position of the transitionfrom the Lower to the Middle Siwalik and from the Middle to the UpperSiwalik was both used. Altogether, the plot in Fig. 8 indicates a rate ofprogradation, Vpr, of about 15 mm/year over the last 15 Myr.

Assuming that the geometries of the Himalayan wedge and of theflexed Indian plate have not changed with time, as initially proposedby Lyon-Caen and Molnar (1985), Vpr might be taken as equal tothe overthrusting rate V0 (the convergence rate between India andsouthern Tibet) (Fig. 7). Actually this reasoning ignores how themountain front might have retreated or advanced as a result of erosionor of crustal thickening. Later, Molnar (1987) proposed that internaldeformation of the crustal wedge would account for about 5 mm/yearof additional shortening, and revised the thrusting rate to about15–20 mm/year.

In fact, a more general formulation should account for internaldeformation of the range as well as for possible erosional retreat of the

FIG. 9. Sketch showing how outcropping sections in the sub-Himalaya can be used to infer

horizontal displacement of the foreland due to underthrusting. The present thickness of unit 1,

h*, represents, after decompaction, the depth of the foreland basin, h, at time of the onset of

deposition of unit 2, t1–2. If the geometry of the basin is assumed to have remained stationary,

this section was at that time at a distance (�x1þ�x2) from its present position. We can then

estimate by how much the Indian plate has underthrust the Himalayan topography since

time t1–2.

12 AVOUAC

Page 13: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

front of the high range, taken to represent the edge of the overriding loadthat flexes down the plate (Fig. 7). We should then write,

Vpr ¼ VHR þ dW=dt, ð2Þ

where VHR represents the velocity of the front of the high range withrespect to the foreland, and W is the width of the accretionary prism(Fig. 7). To relate Vpr to the convergence rate V0, it is necessary to takeinto account erosion of the mountain front. The retreat (if positive) oradvance (if negative) of the front of the high range with respect to thehanging wall (southern Tibet being considered as the back stop) is then

Ver ¼ V0 � VHR: ð3Þ

This term might actually be large and could have varied during theHimalayan orogeny. In the absence of any crustal thickening or isostaticresponse, if eroded at a vertical rate e the mountain front would apparentlyretreat by

Ver ¼ e=tg�: ð4Þ

The denudation rate in the High Himalaya is estimated to be of the order of4–8 mm/year (Lave and Avouac, 2001) and the slope of the front of the highrange is about 10%. This would yield a very rapid retreat by as much as40–80 mm/year. In reality, erosion is compensated by some uplift, u, due toisostatic response and crustal thickening, so that the apparent retreat is lessand should be written,

Ver ¼ ðe� uÞ=tg�: ð5Þ

Some estimate of Ver is therefore needed to fully describe the kinematics ofmountain building and its relation to rates of sediment progradation anddeposition in the foreland. Thus, the southward progradation rates ofsediments and subsidence rates in the foreland depend on overthrusting, buta direct estimate of the thrusting rate is not straightforward and someidea of the distribution of erosion and of internal shortening of the wedgeis needed.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 13

Page 14: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

3. STRUCTURE AND LONG-TERM KINEMATICS OF DEFORMATION

AND EXHUMATION OF OROGENIC WEDGE

3.1. Geological and Geomorphologic Setting

The major Himalayan thrust faults separate domains with contrastedgeology and also control the present morphology of the range (Figs. 4and 10). The Indo-Gangetic plain is the 200–300 km wide foredeep atfront of the range, where a fraction of the material eroded from therising topography is trapped, the rest of the material being transferred bythe Ganges drainage system and ultimately delivered to the Bengal fan(e.g., France-Lanord et al., 1993). Several kilometers of Cenozoic molasse,essentially the Miocene to Quaternary Siwalik formation, have accumulatedover the Precambrian Indian basement. The Indian basement is mainlycomposed of Archean to Early Proterozoic metamorphic rocks. Northof the Indo-Gangetic plain, the sub-Himalaya is a zone of thin-skinnedtectonics bounded to the south by the MFT and to the north by theMBT (e.g., Delcaillau, 1986; Mugnier et al., 1999; Lave and Avouac, 2000).It consists of Tertiary siltstones, sandstones, and conglomerates thathave been scraped off the basement, folded, and faulted at the front ofthe advancing range. Although highly erodible, the molasse form steepreliefs which attest to the on-going tectonic activity of this fold andthrust belt. In the study area of central Nepal, the sections along SuraiKhola (Corvinus, 1988; Appel and Rossler, 1994), Tinau Khola (Gautamand Appel, 1994), Arung Khola (Tokuoka et al., 1986), and BakeyaKhola (Harrison et al., 1993) show nearly the same sedimentary sequence,3500–5500 m thick, that was deposited between 14 and 1 Myr ago accor-ding to paleontological and magnetostratigraphic studies. It forms sub-dued reliefs with elevations of the order of a few hundred meters. In places,the fold belt involves slices of pre-Tertiary sediments, reddish-maroonquartzites and gray shales with some doleritic intrusions (Gautam et al.,1995; Lave and Avouac, 2000), very similar to Vindhyan units drilled atRaxaul (Sastri et al., 1971) (see location in Fig. 4). Pre-Tertiary unitsinvolved in the sub-Himalaya fold belt are also reported 50 km east ofthe section along the Bagmati, north of the Kamla Khola (Mascle andHerail, 1982). The decollement underlying the fold belt must therefore lieon top of the Indian basement, at some 5–6 km depth along the sectionacross central Nepal. This hypothesis is fairly consistent with thecomparison of the balanced cross-sections in the sub-Himalaya south ofthe Katmandu basin and the geological well log of Raxaul, which cuts

14 AVOUAC

Page 15: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

through the whole undeformed Tertiary cover over the basement (Lave andAvouac, 2000).

Some authors have proposed that the deformation front might extendfarther south of the MFT as a blind detachment below the Indo-Gangeticplain. This view was initially proposed by Seeber and Armbruster (1983),

FIG. 10. (a) Mean, maximum, and minimum elevation within a 50-km wide swath centered

on section AA0 (see Fig. 1 for location). Several geographical domains are distinguished from

north to south: Tibet, High Himalaya, Lesser Himalaya, sub-Himalaya, and the Gangetic plain.

(b) Geological section across central Nepal Himalaya at the longitude of Katmandu. Thick line

shows the MHT which reaches the surface at the front of the Siwalik hills, where it coincides

with the MFT. The Main Boundary Thrust fault separates Lesser Himalayan metasediments

from the molasse deposits of the sub-Himalaya (the Siwalik hills). The Main Central Thrust

fault (MCT) places the higher-grade metamorphic rocks of the High Himalayan crystalline

units over the Lesser Himalayan metasediments (from Lave and Avouac, 2000).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 15

Page 16: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

based on the intensity distribution of the large historical earthquakes alongthe Himalaya, in particular during the 1924 Bihar–Nepal event. The authorhas never seen any convincing evidence for this possibility. There is nogeomorphic or subsurface geological evidence of any zone of deformationsouth of the MFT along the Himalaya of central Nepal (Lave and Avouac,2001). The zone of high macroseismic effects south of the MFT mostprobably resulted from site effects, i.e., amplification of ground motionand liquefaction within the loose, water-saturated sediment cover of theGangetic plain.

North of the MBT, the Lesser Himalaya units consist of low-grademetasediments: phyllite, quartzite, and limestone of Devonian or older ages(Upreti, 1999). At some places overlying Tertiary units have been preserved,in particular the sandstones and siltstones of the Dumri formation (Sakai,1985). This formation consists of Himalayan foreland sediments that weredeposited between about 16 and 21 Myr ago, as indicated from � 20-Myr-old detritic muscovites and from preliminary results from a magnetostrati-graphic study (DeCelles et al., 1998). The section across the Himalaya ofcentral Nepal near Katmandu is characterized by a crystalline sheet thruston top of the Lesser Himalayan units (Stocklin, 1980). It forms theKatmandu klippe which is the chief reason for the impressive relief of theMahabarat range which reaches elevations up to 2500–3000 m just north ofthe sub-Himalayan fold belt (Fig. 10). The crystalline units, consistingmainly of schist and gneiss intruded by Late Cambrian to Ordoviciangranites (Scharer and Allegre, 1983), are overlain by Cambrian to Eocene‘‘Tethyan’’ sediments (Stocklin, 1980). Thermobarometric studies indicatethat this crystalline sheet overrode the Dumri formation, in particular in theTansen area (Bollinger, 2002). The basal ‘‘Mahabarat’’ thrust can beinterpreted as the southern extension of the MCT, although the possibilitythat it is a distinct thrust fault rooting below the MCT cannot be excluded(e.g., Upreti, 1999). North of the Klippe, the foliation in the LesserHimalaya schist depicts a large antiformal structure, the Pokhara–Gorkhaanticlinorium (Pecher, 1989). This structure has been interpreted as ahinterland-dipping duplex structure as shown in Fig. 10 (Brunel, 1986), thathas been recognized on most sections across the Nepal Himalaya (Schellingand Arita, 1991; Schelling, 1992; DeCelles et al., 2001) and in Kumaon,India (Srivastava and Mitra, 1994). The structural control of the mor-phology in the Lesser Himalaya is generally inverted, with synclinal crestsand anticlinal valleys, suggesting a typically ‘‘mature’’ relief (Valdiya, 1964).

North of a line trending about N108�E, that roughly coincides with thetrace of the MCT, the topography rises abruptly from elevations around

16 AVOUAC

Page 17: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

1000 m to more than 6000 m. The abrupt break-in-slope marks the front ofthe High Himalaya that can be traced all along the Himalayas of Nepal(Fig. 3). The High Himalaya units consist of amphibolite-grade schistintruded by large leucogranitic plutons. Neodymium isotopic provenancesuggests that Lesser Himalaya metasediments were derived from the Indiancraton, while the High Himalayan rocks most probably correspond to anexotic terrane accreted onto India in the early Paleozoic (Robinson et al.,2001).

A singular feature of the range morphology is the position of the frontof the high-range well to the north of the main bounding thrust faultsalong the foothills. This feature has inspired a variety of interpretations.Because this area coincided with a zone of localized ongoing uplift revealedfrom leveling data, it has been attributed to active thrusting at the frontof the high range (Bilham et al., 1997). Some authors interpret it as thepreserved topographic signature of a Late Miocene reactivation of theMCT zone (Harrison et al., 1997). Others have proposed that head-ward regressive erosion along the rivers cutting the edge of the TibetanPlateau would have induced uplift of the Himalayan peaks through iso-static rebound, enhancing orographic precipitation, and hence denudation(Molnar, 1990; Burbank, 1992; Masek et al., 1994; Montgomery, 1994).Such a process would have driven accelerated uplift, independently of thekinematics of active thrust faulting, in response to climate change during theCenozoic.

3.2. Crustal-Scale Structural Models of the Himalaya

Based on surface geology, a variety of crustal-scale sections acrossthe Himalaya have been proposed. One early view is that all the majorthrust faults are crustal-scale faults that developed following a forwardpropagation sequence after the collision along the ITSZ (e.g., Le Fort,1975b; Mattauer, 1975, 1986; Molnar and Lyon-Caen, 1988; Molnar, 1990).According to that view, all faults would be parallel to one another and reachto the Moho as schematized in Fig. 4.

Alternatively, some authors have proposed that all the major faults rootin a common midcrustal decollement (Brunel, 1983; Schelling and Arita,1991). Such a geometry was initially inferred from the Lesser Himalayanantiformal structure, and it appears to be consistent with barometric studies.Indeed, peak metamorphic pressures documented in the Nepal Himalayanever exceed about 8 kbar (see Guillot, 1999 for a review), suggesting that

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 17

Page 18: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

rocks were exhumed from depths of about 30–35 km at most. This is thehypothesis used to construct the cross-section in Fig. 10. This section bearssome similarity with the more elaborated ‘‘balanced’’ sections that havebeen constructed along several transects across the Himalaya of Kumaon(Srivastava and Mitra, 1994), far-western Nepal (DeCelles et al., 2001), andeastern Nepal (Schelling and Arita, 1991). However, it should be realizedthat the geometric rigor used to balance these sections might give amisleading impression of accuracy. The basic assumptions—that deforma-tion occurs by bedding slip with maintenance of constant width and lengthof the various units, and that the footwall remains rigid—certainly do nothold in the case of the Lesser Himalayan units, because these havedeveloped an intense foliation and have experienced a significant amount ofpure shear. Although the details of the balanced sections should therefore beregarded with caution, they have the advantage of integrating a logicalmodel of the structural evolution and do provide a reasonable basis forkinematic interpretation. The main point is that these balanced sectionsall suggest some duplex structure in Lesser Himalaya, with all the majorthrust faults rooting in a midcrustal decollement beneath the high range.The duplex must have developed after the emplacement of the crystallinethrust sheets.

3.3. Moho Geometry

The seismic experiments conducted in southern Tibet all suggest a crustalthickness of the order of 70–80 km (Hirn et al., 1984; Zhao et al., 1993; Kindet al., 1996) about twice the crustal thickness of the Indian shield which hasbeen estimated at about 40 km on the basis of teleseismic receiver functions(Saul et al., 2000) (Fig. 12). The Bouguer anomaly in India and in Tibetprimarily indicates local Airy compensation in keeping with those estimatedcrustal thicknesses. By contrast, important deviations from Airy isostasy areobserved below the Himalayan range and its foreland, as seen in particularfrom the gravity data across the Himalaya of central Nepal (Fig. 11) (Lyon-Caen and Molnar, 1983; Lyon-Caen and Molnar, 1985; Jin et al., 1996;Cattin et al., 2001). The observed values are more negative than expectedfrom local isostasy over the Gangetic plain, indicating some mass deficitthere. By contrast they indicate some mass excess below the range (Fig. 14).These deviations are the typical signature of a flexural support of the range,meaning that the weight of the Himalaya is supported by the strength of theunderthrusting Indian plate. Moreover, the steep gravity gradient of the

18 AVOUAC

Page 19: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

order of 1.3 mGal/km beneath the High Himalaya suggests a locally steeperMoho, coincident with the position of the midcrustal ramp along the MainHimalayan Thrust fault (MHT) (but not necessarily parallel). Flexuralmodeling in terms of a thin elastic plate overlying an inviscid fluid (Lyon-Caen and Molnar, 1983; Lyon-Caen and Molnar, 1985; Jin et al., 1996) cansuccessfully reproduce the observed gravity anomalies, but this requires thatsome forces or momentum be exerted on the flexed plate in addition to theload due to the high topography. This kind of modeling also requires anabrupt decrease in the strength of the Indian plate beneath the high range toaccount for the locally steeper gradient of gravity anomalies. Actually thisnorthward decrease in the apparent flexural rigidity might reflect thethermal structure of the range and its influence on crustal rheology. Due tothe kinematics of underthrusting, the thermal structure implies relativelyhigh temperatures at midcrustal depths favoring ductile flow within the crust(Fig. 12). This induces some decoupling within the upper crust and uppermantle resulting in an abrupt decrease of the flexural rigidity (Burov andDiament, 1995). It turns out that, if the thermal structure and its influenceon crustal rock rheology are taken into account, there is no need foradditional forces other than the weight of the topography. The gravityprofile is indeed relatively well reproduced by a 2-D mechanical model inwhich the Indian lithosphere is flexed down by the advancing topography ofthe range and sedimentation in the foreland (Cattin et al., 2001) (Fig. 11).The computed Moho fits seismological constraints and is consistent with themain trends in the observed Bouguer anomaly. Although the picture inFig. 11 shows a relatively smooth Moho, there is some indication that thestructure of the lower crust might be quite complex with imbrications ofcrustal and upper mantle rocks (Hirn et al., 1984) (Fig. 12).

3.4. Geophysical Constraints on the Geometry of theMain Himalayan Thrust Fault (MHT)

The structure of the crust across the central Himalaya has been invest-igated through a variety of means including gravimetric, magnetotelluric,and seismic techniques. A major feature revealed by the seismic experimentsin southern Tibet is a strong midcrustal reflector at a depth of 35–40 km.The existence of this reflector was first suggested by wide-angle seismicreflexion studies (Hirn and Sapin, 1984); later, it was better imaged from thecommon midpoint (CMP) deep seismic profiles run during the INDEPTHexperiment (Zhao et al., 1993; Brown et al., 1996; Nelson et al., 1996).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 19

Page 20: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

As shown in Fig. 13, this conspicuous reflector was found to coincide withthe midcrustal decollement inferred from the structural sections across therange and was therefore termed the Main Himalayan Thrust fault, MHT(Brown et al., 1996).

FIG. 11. Bouguer anomalies along a N18�E profile across the Himalaya of central Nepal

(modified from Cattin et al., 2001). All the data within a 30-km wide swath were projected onto

the section. Several data sets were merged with accuracies ranging from about 0.5 to 7 mGal

[Bureau Gravimetrique International database (Van de Meulebrouck, 1983; Abtout, 1987; Sun,

1989)]. The Moho is set to 40 km beneath the Indian shield according to teleseismic receiver

functions (Saul et al., 2000), white line shows Moho according to INDEPTH seismic profiles

(Zhao et al., 1993; Brown et al, 1996), and thick vertical bars show Moho picks on receiver

functions (Kind et al., 1996). The thick line shows the expected Bouguer anomaly in the case of

isostatic compensation of a crust with uniform 2.67 density (computed from the topography

smoothed with a 50-km wide Gaussian). The thin line is the result of a mechanical model that

accounts for the low-density sediments in the foreland, the rheological layering of the crust, and

its dependence on the thermal structure (Cattin et al., 2001).

20 AVOUAC

Page 21: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 13. Geophysical constraints on the crustal structure across central Nepal. The

conductivity section was obtained from a magnetotelluric experiment carried out along

the section AA0 across central Nepal (Lemonnier et al., 1999). Also reported are the

INDEPTH seismic sections run about 300 km east of section AA0 (see location in Fig. 3) (Zhao

et al., 1993; Brown et al., 1996; Nelson et al., 1996). All the thrust faults are inferred to root at

depth in a subhorizontal ductile shear zone that would correspond to the prominent midcrustal

reflector.

FIG. 12. Steady-state thermal structure of the range computed from a 2-D finite element

model (Henry et al., 1997) (from Cattin and Avouac, 2000). Erosion is assumed to exactly

balance tectonic uplift, as computed from thrusting of the hanging wall over the flat–ramp–flat

geometry of the MFT–MHT, assuming a rigid footwall. Boundary conditions are a constant

surface temperature of 0�C, a bottom heat flow of 15 mW/m2, and a heat production in the

upper crust of 1.5 mW/m3. Sensitivity tests show that, insofar as the model-fit thermobarometric

constraints [650�C above the MCT (e.g., Hubbard, 1989; Hubbard and Harrison, 1989), the

computed thermal structures show only small differences, so the model shown here is probably a

robust estimate (Cattin and Avouac, 2000; Bollinger, 2002).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 21

Page 22: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

The deep electrical structure of the central Nepal Himalaya was imagedfrom magnetotelluric sounding (Lemonnier et al., 1999) (Fig. 13). Variationsof electrical conductivity in the crust can result from changes in fluidcontent, pore geometry, or lithology. Conductive zones in the crust aregenerally thought to reflect well-connected conductive phases (brines, melts)or conductive minerals such as graphite (Marquis et al., 1995). The sectionshows a high conductivity in the foreland (� 30 �m) consistent with thegeometry of the molassic foreland basin, which contrasts with the resistiveIndian basement (>300 �m) and Lesser Himalayan units (>1000 �m).A continuous shallow-dipping conductor can actually be traced northwardand coincides relatively well with the position of decollement beneath theLesser Himalaya inferred from structural studies. It may well reflect somefluid-rich sediments dragged along the thrust fault, a process expected asthe rugged topography of the Indian basement underthrusts the LesserHimalaya.

Farther north, the magnetotelluric section show a major conductive zone(� 30 �m) at about 15 km depth under the front of the High Himalaya.This zone coincides with the position of the midcrustal ramp beneath thefront of the High Himalaya, as well as with a zone of intense microseismicactivity (Fig. 13) revealed from local seismic monitoring (Pandey et al.,1995, 1999; Cattin and Avouac, 2000).

These geophysical data thus point to the MHT being a major thrust faultthat can be traced nearly continuously from the MFT, along the foothills, tobeneath southern Tibet. Contrary to the early views depicted in Fig. 4, theMCT, MBT, and MFT should not be seen as equivalent major thrust faultscutting through the whole crust. They more probably represent splay faultsrooting in a single midcrustal decollement. The geometry of the MHT ischaracterized by a ramp and flat geometry, probably with two major ramps.One is very shallow and corresponds to where the fault emerges with a dipangle around 30� at the surface along the MFT. The other one, moreconjectural, lies at midcrustal depth beneath the front of the high range andis estimated to dip north by about 15�.

3.5. Chronological Constraints on the Development of theThrust Package

The chronology of the deformation across the Himalaya has beenconstrained from cross-cutting relationships, provenance data from theTertiary molasse (Siwalik and Dumri formations), and thermochronology.

22 AVOUAC

Page 23: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

A complete discussion of this issue is well beyond the purpose of this review,although a few key results are pointed out here:

� compressional deformation north of the Himalaya started in the early‘‘Alpine’’ times of the collision, and there is evidence that crustalthickening in southern Tibet persisted until mid-Miocene times (e.g.,Ratschbacher, 1994; Yin et al., 1999). Farther south, deformation andanatexis in the MCT zone were occurring at about 22 Myr (Copelandet al., 1991; Hodges et al., 1996), an age consistent with slightlyyounger cooling ages in the rocks of the hanging wall, attributed to theemplacement of the MCT (Hubbard, 1989; Copeland et al., 1991;Hubbard et al., 1991; Hodges et al., 1996). Such a timing is alsoconsistent with conventional and isotopic provenance data from theDumri formation that indicate an early Miocene time for the onset oferosion of the High Himalayan crystalline rocks (Robinson et al.,2001). Out-of-sequence reactivation of MCT may have occurred by lateMiocene to Early Pliocene times as indicated from monazite ages ofrocks exhumed from midcrustal depths (Harrison et al., 1997; Catloset al., 2001). This reactivation has been proposed to be responsible forthe inverse thermal gradient and for the steep morphologic front of theHigh Himalaya range (Harrison et al., 1997, 1998).

� there is no direct, reliable estimate of the timing of thrusting on theMBT. Age and provenance of Siwalik sandstones in northern India,west of our study area, show that exhumation of LH have started some10 Myr ago (Meigs et al., 1995). This observation was interpreted toindicate motion on the MBT. However, exhumation of LH rocks couldalternatively have resulted from the development of the LH duplex, soa much younger activation of the MBT, possibly around 5 Myr, couldalso be possible (DeCelles et al., 1998; Robinson et al., 2001).

� deformation in the sub-Himalaya has proceeded from the mid-Mioceneto present.

The chronological constraints discussed above, together with restorationof the various balanced cross-sections across the central Himalaya(Schelling and Arita, 1991; Srivastava and Mitra, 1994; DeCelles et al.,2001), provide some idea of the kinematics over approximately the last20 Myr. Figure 14 presents the kinematic model proposed for far-westernNepal (Robinson et al., 2001) which also holds with some minor variationsfor central Nepal (Bollinger et al., submitted-a).

The development of the crustal wedge, over a geological period oftime thus appears as a combination of frontal accretion at the toe,

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 23

Page 24: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 14. Schematic model of the structural evolution of the Himalayan orogen (modified

from Robinson et al., 2001). This scheme is based mainly on observations along a section

across the Himalaya of western Nepal. MCT, Main Central Thrust; DT, Dadeldhura Thrust;

RT, Ramgarh Thrust; MBT, Main Boundary Thrust; MFT, Main Frontal Thrust; STD,

Southern Tibetan Detachment. (A) 45–16 Ma: Emplacement of the MCT and then of the DT

crystalline thrust sheets, coeval with motion on the STD. Deposition of the Dumri and

Bhainskati foreland formations. (B) 15–6 Ma: Emplacement of the Lesser Himalaya RT thrust

sheet and development of the duplex; deposition of the Lower andMiddle Siwaliks. (C) 5–0 Ma:

Motion on the MBT and MFT; deposition of the Upper Siwaliks; major phase of exhumation

of the LH duplex.

24 AVOUAC

Page 25: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

internal deformation of the wedge, and underplating. Underplating asso-ciated with duplexation of the Lesser Himalayan units at midcrustal depthhas probably been the dominant process in the growth of the Himalayanwedge since about the Middle Miocene. This kinematic scheme is similar tothe pattern of deformation expected for a critical brittle taper subjected toerosion except that maintenance of the critical slope was probably achievedwithout any significant thickening of the upper crustal units.

3.6. Kinematics Model of the Structural Evolution andExhumation of the Orogen

If topography and the thermal structure are assumed to be in a steady-state regime, the pattern of exhumation should provide direct informationon the kinematics of the deforming orogenic wedge.

The most complete picture of the chronology of exhumation in thearea considered here comes from the determination of 39Ar/40Ar ages ofmuscovites (Fig. 15). This technique provides an estimate of the age ofcooling of the rock sample through the muscovite-blocking temperaturewhich is estimated at about 350�. Several authors, in particular, Copelandet al. (2003) have analyzed samples collected along the central Nepal sectionfrom the high range to the southern edge of the Katmandu klippe (Fig. 15).These data indicate that the trailing edge of the Katmandu thrust sheetcooled below 350�C about 20–22 Myr ago, an age consistent with the onsetof deformation and anatexis in the MCT zone as discussed above. Agesdecrease gradually northwards to about 5 Myr at the front of the highrange, following a nearly linear trend corresponding to a slope of about 0.23Myr/km (Fig. 15). Given that the 350�C steady-state isotherm crosses theMHT at the midcrustal ramp (Fig. 16), this linear trend indicates that thehanging wall overthrusts the midcrustal ramp at a rate that corresponds tothe inverse of the age gradient,

Ver ¼ V0 � VHR � 4:3 mm=year: ð6Þ

As a first attempt at interpreting this trend, we may ignore accretion(Fig. 16). In that case, according to equation (2) and the data in Fig. 8, theforeland would underthrust the Himalayan topographic wedge by

Vpr ¼ VHR � 15 mm=year: ð7Þ

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 25

Page 26: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

This is a value consistent with the data in Fig. 9. Moreover, these twovalues add to the total estimated shortening rate across the range ofV0� 20 mm/year. This simple kinematics reconciles the rate of sedimentprogradation and subsidence of the foreland with the retreat of themountain front driven by erosion.

The kinematics may be compared with two end-member cases of a steady-state regime without accretion:

– If the hanging wall is assumed to overthrust an undeformable footwallby 20 mm/year, there would be no sediment progradation and nosubsidence of the foreland, and we should observe a diachronicexhumation corresponding to,

Ver � 20 mm=year, ð8ÞVpr ¼ VHR � 0 mm=year: ð9Þ

FIG. 15. Cooling ages in central Nepal as a function of distance from MBT along a

N018�E section. All samples collected from the LH and the overlying crystalline thrust

sheets are reported. The Ar39/Ar40 muscovite ages appear to approximately follow a

linear trend with ages increasing southward by about 0.2 Myr/km. Data compilation

by Bollinger (2002) (from Macfarlane et al., 1992; Arita and Ganzawa, 1997; Arita

et al., 1997; Copeland, 1997) and Copeland (personal communication). The origin is taken at

the MBT. All samples collected from the LH (triangles pointing downward) and the overlying

crystalline thrust sheets are reported.

26 AVOUAC

Page 27: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

– In the opposite case, if the convergence is assumed to be entirelyaccommodated by underthrusting of the Indian basement below a rigidhanging wall without any erosion at the surface, we get,

Ver ¼ V0 � 0 mm=year, ð10Þ

Vpr ¼ VHR � 20 mm=year: ð11Þ

The kinematic deduced from the geochronological data is clearly closer tothe second end-member. So both the structural and geochronologicaldata require the footwall to be deforming because it must underthrust theMHT (Fig. 16), before it gets accreted to the footwall. The kinematics inFig. 16 might be modified to fit the structural evolution model of Fig. 14 byassuming that some accretion occurs by underplating due to the develop-ment of a duplex at midcrustal depths (Fig. 17). The resulting modelis shown in Fig. 18. To draw this picture, a continuous process of accretionis assumed. The top of the underthrusting plate is accreted while the lowerpart underthrusts the orogenic wedge. In reality the process of accretioncould be discontinuous.

FIG. 16. Kinematic model of overthrusting at rate V0�VHR, and underthrusting at rate VHR,

if the topography and thermal structure are assumed to be in a steady state. These kinematics

predict diachronic exhumation (through the Muscovite-blocking temperature which is

estimated at 350�C) with an age gradient of 1/V0�VHR.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 27

Page 28: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

The assumption of steady-state topography would imply dW/dt¼0 mm/year in Eq. (2), so that Eqs. (7) and (8) are not modified. Thus, themodel still reconciles the rate of sediment progradation and subsidence ofthe foreland with the retreat of the mountain front driven by erosion. It also

FIG. 17. Schematic model of duplexing in the Lesser Himalaya. Thick line shows the

geometry of the active thrust fault at a time just before it migrates to a new position farther

south (dashed line). This migration induces accretion of material from the footwall to the

hanging wall. Thin dashed line shows the geometry of other thrust sheets accreted before time t.

The position of the midcrustal ramp, the ‘‘accretion window,’’ is taken as a reference to define

velocities Vpr and V0�VHR, assuming that it controls the topography of the range and the load

flexing the Indian lithosphere.

FIG. 18. Particle trajectories expected from the model of duplexation shown in Fig. 17,

assuming that ramp migration is a continuous process. Dots show approximate positions

about every 5 Myr. This model also predicts diachronic exhumation with an age gradient of

1/V0�VHR.

28 AVOUAC

Page 29: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

fits the structural evolution model, which requires accretion. For a moredetailed discussion of this model and its comparison to structural,petrological, thermometric, and thermochronologic data, the reader mayrefer to Bollinger (2002), Bollinger et al. (submitted-a) and Bollinger et al.(submitted-c). In the next sections, we compare this kinematic model withdata on deformation and exhumation of the Himalayan orogen over theHolocene.

4. KINEMATICS OF ACTIVE FOLDING FROM HOLOCENE RIVER

TERRACES ACROSS THE SUB-HIMALAYA

The kinematics of thrusting along the front of the Nepal Himalaya canbe documented from the study of deformed river terraces. Here theauthor reviews the methodology and presents results obtained from theterraces along the Bagmati river, which lies approximately south ofKatmandu Basin along section AA0 (see location in Fig. 3). Similar resultshave been obtained along the Bakeya river, which cuts across the samerising anticline about 10 km west of the Bagmati, but these results arenot reviewed here. For more details the reader should refer to Lave andAvouac (2000).

4.1. Structural Section across the sub-Himalaya

Several structural cross-sections were realized along the Bagmati river andnearby rivers (Lave and Avouac, 2000). In this area, the Siwalik hills aredivided in two fold belts (Fig. 19). The southern one is the topographicexpression of the MFT that is inferred to reach the surface along thesouthern limb of the fold belt (Nakata, 1989). The other fold belt, about15 km north of the MFT fold belt, is associated with the Main Dun Thrust.At some places in Nepal the Siwalik hills are more complex with severalfault zones between the MFT and MBT that are assumed to merge with asingle decollement at depth (Mugnier et al., 1992). The MDT fold beltmakes a steep monocline with dips around 65� to the NNW. This beltinvolves about 500 m of pre-Tertiary sediments, mainly reddish-maroonquartzites and gray shales in the lower part. The incompetent shales showmeter-scale folds. The MFT fold belt makes a gently inclined monoclinewith dips varying between 25 and 50� to the NNW (Fig. 19). At the front,

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 29

Page 30: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Lower Siwalik clay- and siltstones are affected by meter-scale faults andfolds. The sections can be balanced by interpreting both folds associatedwith the MFT and MDT as fault-bend folds (Suppe, 1985) associated withcurved ramps that root in the same decollement (Figs. 19 and 10).Palinspastic restorations of the Bagmati and Bakeya sections implyminimum shortening of 11 and 10 km, respectively at the MFT, and 12km at the MDT. The relief of the fold shows that about 90% of the materialuplifted due to slip on the MFT and MDT has been eroded away. So, onaverage over a geological period of a few million years, denudation balancestectonic uplift.

FIG. 19. Structural section with elevation of abandoned terraces along the Bagmati

river across the Siwalik hills south of Katmandu basin, approximately along the section AA0

(see location in Fig. 2) (modified from Lave and Avouac, 2000). Also indicated are ages of

terrace abandonment from C14 dating after calibration to calendar ages. The age of T3

(in italics) was indirectly inferred from the age of the same terrace level along the nearby

Bakeya river.

30 AVOUAC

Page 31: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

4.2. River Incision across the sub-Himalaya

Abandoned fluvial terraces are ubiquitous in the sub-Himalaya ofcentral Nepal, and have been most extensively surveyed along the Bagmati,Bakeya, Narayani, and Ratu rivers (Delcaillau, 1986; Lave and Avouac,2000). They were first recognized and mapped from air photos and Landsatand SPOT images. In the field, they were classified according to theirgeomorphic nature, facies of the fill material, degree of weathering, andthickness of the soil profile. Nearby disconnected treads were also correlatedon the basis of their elevations. The determination of fluvial incisionrequires elevation measurements of datable levels attributable to formerriverbeds. Several different levels may be of interest. If the bedrock showsevidence for fluvial abrasion, in particular if it is still capped by fluvialgravel, the contact between the bedrock and the overlying gravel is calledstrath surface (Bull, 1991). Sometimes, strath surfaces are preserved only inthe form of bedrock benches. Such features are difficult to exploit due to thelack of datable material or criteria for lateral correlation. If the strathsurface is capped by a thin gravel cover (typically <10 m), the terrace iscalled a strath terrace. If the gravel thickness is much larger, it may be calleda fill terrace (Bull, 1991). The difference in elevation between various strathsurfaces and the present riverbed provide some estimate of the incision rate.This estimate is the most useful for tectonic analysis because it is notsusceptible of being influenced significantly by climatic effects. Thedifference in elevation between the top of the terrace and the presentriverbed can also be informative, but it might not relate simply to long-termfluvial erosion because the deposition of the fill and its subsequent incisionprobably a transient fluvial response to a climatic change (e.g., Weldon,1986; Bull, 1991). Such fill terraces can therefore be diachronous, so that theprecise chronology of their deposition and abandonment is difficult toconstrain tightly.

In the Siwalik hills, the strath surfaces are generally covered by a fewmeters of gravel, subrounded boulders, and cobbles (<30 cm) gradingupward into coarse sands. This gravel is very similar to the bedload of themodern rivers. It is overlain by overbank sands and silts, with, at places,mixed colluvium or fanglomerates fed from adjacent valley slopes. TheseHolocene terraces are characterized by a low degree of weathering withbeige to orange color of the silt and sand. Charcoals were found at severalsites within the terrace material and could be dated from 14C. A few sampleswere found within fluvial bedload, and yielded a lower bound on the timewhen the river was flowing at this level. Most of the samples were found in

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 31

Page 32: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

the overlying, less coarse, overbank deposits and providing a lower boundon the age of the underlying strath level.

The chronologic data obtained in this way indicate four major episodes ofstrath-terrace abandonment, dated to about 9.2� 0.15, 6.1� 0.15, 3.7� 0.1,and 2.2� 0.2 ka (14C ages converted to calendar ages). These dates probablycorrespond to wet-to-dry climatic transitions during the Holocene period(Lave, 1997; Lave and Avouac, 2000). The uppermost level, labeled T0 inFig. 19, is generally the most prominent in the landscape because itcorresponds to particularly wide terrace treads. The second most prominentlevel is T3 (dated to about 2.2 ka). Along the Bagmati river the T2 terrace isnot well preserved and could not be mapped and measured easily in the field.Figure 19 shows the elevation above the present riverbed of the three otherstrath surfaces along the Bagmati river, with the locations of the samplingsites where good dating was possible. These data provide an estimate of theaverage incision, I(x, y, t), at any point (x, y) along the river since theabandonment of terrace T at time t incision,

Iðx, y, tÞ ¼ Tðx, y, tÞ � Rðx, yÞ, ð12Þ

where R(x, y) is the present river profile, and T(x, y, t) is the presentelevation profile of the abandoned river terrace.

4.3. Converting Incision Rates to Uplift Rates in the sub-Himalaya

The geometry of the abandoned terraces in Fig. 19 clearly demonstratesthat they were abandoned and warped as a result of fold growth as depictedin Fig. 20. These data provide evidence for active thrusting along the MFT.The terraces are not affected by any clear deformation across the MDT.They also can be traced farther upstream across the MBT, without anyevidence for tectonic activity. The profiles of the various abandoned terracesalong the Bagmati and Bakeya rivers indicate that the MFT has been themajor active thrust fault across the sub-Himalaya of central Nepal over theHolocene period. To quantify uplift rates across the MFT fold, a model ofthe geometry of the terraces before they were deformed is needed. It maysimply be assumed that the rivers have maintained a constant profile duringdeformation, with river incision having counterbalanced tectonic uplift, andthat the fold has been growing steadily during the Holocene period. Thislatter approximation makes sense, given that the mean recurrence interval ofthe earthquakes that could activate the fold is at most a few hundred years.

32 AVOUAC

Page 33: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Accordingly, the tectonic uplift since terrace abandonment would simply beequal to the incision, i.e., the difference of elevation between the abandonedstrath and the present river,

Uðx, y, tÞ ¼ Iðx, y, tÞ: ð13Þ

The term that would arise from the apparent uplift of the riverbed due tothe horizontal component of the displacement field can be neglected due tothe generally low stream gradient of the rivers cutting across the sub-Himalaya.

Given that, on average over many seismic cycles, the fold might beassumed to grow more or less steadily, U(x, y, t) may be assumed to beproportional to the time elapsed since the terrace abandonment. It follows

FIG. 20. Diagram showing the formation, abandonment, and warping of fluvial terraces

during fold growth (modified from Lave and Avouac, 2000). (a) Initially the river has enough

stream power to bevel a wide channel at the rate imposed by the growing fold (tectonic uplift

minus sedimentation rate at the front of the fold). At time t0, due to some decrease of its stream

power (of possibly climate origin), the river starts entrenching into a narrower channel. (b) At

time t1, a paired terrace, corresponding to the river channel at time t0 is preserved in the

landscape (labeled T0). This terrace overhangs the active riverbed by a quantity that depends on

the amount of tectonic uplift induced by the folding between t0 and t1, and on the change of

base level due to sedimentation at front of the fold.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 33

Page 34: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

that in that case the incision deduced from two different terraces abandonedat times t and t0 should match,

Iðx, y, tÞ=t ¼ Iðx, y, t0Þ=t0: ð14Þ

The incision profiles deduced from T0 and T3 along the Bagmati section aresimilar but some mismatch on the northern limb of the fold, due to anincision deficit, was observed (Lave and Avouac, 2000). This suggests thatthe simple assumption of Eq. (13) does not hold.

Actually the cumulative uplift since terrace deposition, at time t, withrespect to a point attached to the Indian lithosphere just south of the MFT,ought to be written at point M(x, y),

Uðx, y, tÞ ¼ Iðx, y, tÞ þDðx, y, tÞ þ Pðx, y, tÞ, ð15Þ

where D(x, y, t) is the local base-level change since time t, and P(x, y, t) is thechange in elevation at point M(x, y) due to possible changes of the rivergradient and sinuosity. The term D(x, y, t) is the sedimentation rate (orpossibly incision rate) at the front of the MFT (Lave and Avouac, 2000).There are no uplifted Holocene terraces south of the MFT, and rivers clearlyaggrade where they come out of the Siwalik hills. The sedimentation ratetherefore must have remained positive during the Holocene. Based on themagnetostratigraphic sections (Appel et al., 1991b; Harrison et al., 1993), itsvalue can be estimated to 0.45� 0.5 mm/year.

In the absence of any way to assess possible changes of the streamgradient during the Holocene, this parameter is assumed to be constant andequal to the present 0.27% value. It should be realized that the Holocenetrend toward a drier climate in this part of Asia, may have forcedhydrological modifications, including gradient changes, but these are noteasily analyzed.

The paleo-channel corresponding to each terrace level can be estimatedfrom the terrace remnants within the steep flanks of the canyon (Lave andAvouac, 2000). The study of these channels reveals that the river sinuosityhas increased during the Holocene. The river was least sinuous (s¼ 1.8) atthe time of T0 strath beveling. At the time of T1 abandonment the sinuosityhad increased to about 2.0. At time of T3 abandonment, it was about 2.1.The present riverbed shows the narrowest and most sinuous channel(s¼ 2.4). This sinuosity increase must have resulted in a gradual increase ofthe apparent slope of the river, and hence an incision deficit, on a projection

34 AVOUAC

Page 35: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

perpendicular to the fold axis. This might explain why no straths but onlyfill terraces are found north of the MFT fold zone, where tectonic uplift hasprobably been too slow to compensate the gradual elevation of the riverbed.

To account for sinuosity changes P(x, y, t) might be written as,

Pðx, y, tÞ ¼ SðdLðx, y, tÞ=dtÞ ð16Þ

where L(x, y, t) is the longitudinal distance along the river at time t and S isthe channel gradient, which is assumed to be constant along that particularriver reach and unchanged with time.

The uplift rate profiles in Fig. 21 were computed from Eq. (15),taking sinuosity and base-level changes into account. The two curves areidentical within the error bars. Together with the uplift rate profiles deducedfrom other terrace levels along the Bagmati and Bakeya rivers (Laveand Avouac, 2000), these data suggest that the anticline has beengrowing more or less steadily during the Holocene. The zone ofactive uplift is 1–14 km wide and the uplift rate reaches a maximum ofabout 1 cm/year (Fig. 21).

Although the terraces we have surveyed were primarily abandonedbecause the rivers entrenched to keep pace with tectonic uplift, it should beemphasized that river incision may not simply relate to tectonic uplift. Riverentrenchment must be analyzed not only in terms of base-level changes andtectonic uplift, but also of sinuosity changes. In addition, under certaincircumstances hydrological changes can drive very rapid and episodic riverincision. If so, this factor will contribute to an apparent ‘base level change’possibly overweighting the tectonic term in equation (15). For example,Holocene climate change has indexed as much as 1 cm/year of incision alongmajor rivers flowing accross the northern piedmont of the Tian Shan rangein Central Asia (Poisson and Avouac, in press). So tectonic uplift can beretrieved from river entrenchment, but some caution is needed to correctlyaccount for nontectonic factors. The effects due to fluvial adjustment,hydrological changes, or variations of sediment supply may be dominant inareas with moderate rate of tectonic deformation.

4.4. Converting Uplift Rates to Horizontal Shortening fromArea Balance

A simple way to deduce horizontal shortening from uplift profiles acrossthe MFT fold is to assume conservation of mass (Molnar et al., 1994).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 35

Page 36: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 21. Uplift rate (top) along the Bagmati section (bottom) deduced from terrace T0 and T3

after correction for 0.45 mm/year base-level change due to sedimentation south of the

MFT and sinuosity changes. Dashed line shows the predicted uplift rate pattern assuming

fault-bend folding. For a cylindrical fold, it would correspond to U(x, y)¼ d(t) sin �(x, y). Ityields a good fit to the observed uplift rates for a horizontal shortening rate across the MFT of

21 mm/year.

36 AVOUAC

Page 37: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Let A be the area between the present profile of an abandoned terrace treadand the profile of the same terrace level at the time t it was abandoned(Fig. 22). Let h be the thickness of the units detached from the basement(Fig. 22). Assuming plane strain and conservation of mass, we derive themean horizontal displacement, d, necessary to account for the deformationof the terrace, as

dðtÞ ¼ AðtÞ=h ¼Z x

0

Uðx; tÞ dx� ��

h: ð17Þ

FIG. 22. Relationship between horizontal shortening and uplift rates, assuming conservation

of mass (top) or fault-bend folding (bottom). A is the area between the abandoned terrace tread

and the profile of the riverbed at the time t it was abandoned. h is the thickness of the units

detached from the basement. Conservation of mass implies that the cumulative shortening, d,

since terrace abandonment is d¼A/h. At point M(x, y), the fault-bend fold model predicts that

local dip angle, �, and local cumulative uplift,U, should relate to the amount of shortening since

terrace abandonment, according to U¼ d sin �.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 37

Page 38: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

This approach requires a continuous terrace tread across the growing fold.The only such terrace is the T0 level along the Bagmati river. Deformation ofthat level corresponds to an area of A(T0)¼ 1.05� 0.25 km2. According tothe structural description above, the depth of the basement is 5.0� 0.3 kmbeneath the MFT and dips northward with a slope of 2.5%. Given that theMFT ramp merges with the decollement about 18 km north of its trace atthe surface, the depth to the decollement beneath the fold is 5.5� 0.3 km.We get a shortening d¼ 192 m (þ 60/�50 m) since 9.2� 0.2 cal. kyr, hencea shortening rate of 21þ 7/�6 mm/year (Fig. 23).

FIG. 23. Plot of horizontal shortening deduced from the various terrace treads along the

Bagmati and Bakeya section as a function of age of terrace abandonment (Lave and Avouac,

2000). These data are consistent with a uniform 21� 1 mm/year shortening rate over the

Holocene period. If the fold is assumed to have been growing by incremental deformation

during large earthquakes, the mean shortening rate is slightly modified since some amount of

elastic straining at the large scale may yet remain to be released, or may not have been released

by the time of emplacement of the different terraces. Assuming that these increments can be as

large as 5 m of horizontal shortening and that all values between 0 and 5 m are equiprobable, we

obtain a long-term averaged shortening rate, offering a best fit to the terrace record, of

21.5� 1.5 mm/year.

38 AVOUAC

Page 39: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

4.5. Converting Uplift Rates to Horizontal Shortening from theFault-Bend Fold Model

Inspection of Fig. 21 suggests that the uplift profiles recorded by theHolocene terraces are closely related to the geological structure of the fold.Indeed, the small variation of the uplift rate within about 1 km in the middleof the fold correlates with variation of bedding dip angles, hence of the dipof the ramp at depth. The correlation between structural geology and recentuplift is in fact expected for a fault-bend fold. According to this model,the hanging wall accommodates deformation imposed by the faultgeometry at depth by bedding plane slip, with no change in the lengthand width of the beds (Fig. 22). It follows that, on the back limb, uplift ratedepends on the dip of the fault at depth, which equals the local bedding dipangle. Assuming cylindrical geometry, the local uplift since time t, U(x, t),and the related horizontal shortening d(t), have a simple geometricrelationship. In the case of a fold with cylindrical geometry (with bedsstriking perpendicular to the section taken to be parallel to the y axis), andassuming that the slip increment since terrace abandonment is small, onemay write

Uðx, yÞ ¼ UðxÞ dðtÞ sin �ðxÞ, ð18Þ

where �(x) is the dip angle at point (x, y). The equation holds insofar asthe dip angle does not vary significantly over a distance of about d,d<<dx/d(sin �(x)).

In the case of a noncylindrical geometry, Eq. (18) should be replaced by(Lave and Avouac, 2000),

Uðx, tÞ ¼ dðtÞ sin �0ðxÞ, ð19Þ

where,

sin �0ðxÞ ¼ sin �0ðxÞ sinð�ðxÞ � �0Þ: ð20Þ

Here � (x, y) is the bedding strike at point M(x, y) and � 0 is the meandirection of transport.

The fault-bend model thus leads to a testable correlation betweenstructural geology and the terrace warping. The uplift profiles deduced from

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 39

Page 40: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

terraces T0 and T3 along the Bagmati river are in close agreement with thebedding dip angles as expressed in Eq. (19) (Fig. 22). The model was alsosuccessfully tested from the terraces along the Bakeya river (Lave andAvouac, 2000).

This model makes it possible to estimate the cumulative shorteningsince the abandonment of each terrace by least-squares adjustment ofthe model to the uplift profiles, d(t) being the only adjustable variable.Each abandoned terrace yields a point on the plot in Fig. 23, whichshows d as a function of elapsed time since the abandonment of eachterrace. The points are seen to follow a linear trend. A least-squares fityields 20.4� 1 mm/year, which may be taken to represent the rate offault slip on the MFT, or alternatively to the horizontal shortening rateacross the fold over the Holocene period. This rate is much betterconstrained than, but consistent with, the estimate deduced fromconservation of mass in the previous paragraph. However, this estimateignores the fact that the slip is not simply a linear function of time. Slip onthe MHT has most probably resulted from recurring coseismic slip events.When this stick-slip behavior is taken into account, assuming slip events ofup to 5 m, the long-term slip rate is revised to 21� 1.5 mm/year (Lave andAvouac, 2001).

5. HOLOCENE KINEMATICS OF OVERTHRUSTING AND RIVER INCISION

ACROSS THE WHOLE RANGE

5.1. Fluvial Incision across the Whole Range

The kinematics of uplift and overthrusting along a whole section acrossthe Himalaya of central Nepal can be inferred from the pattern of riverincision (Lave and Avouac, 2001). As shown above, the strath terracesprovide good constraints on incision rates in the sub-Himalaya where riversare forced to cut down into the rising anticlines and have abandonednumerous strath terraces. In the Lesser Himalaya strath terraces are sparse.There are, however, prominent fill terraces, more than about 100 m thick,probably of Pleistocene ages (Iwata, 1976; Fort et al., 1983). Close to thefront of the high range, the valleys become steep and narrow and theterraces are no longer preserved. Such a terrace pattern suggests a slow rateof fluvial incision in the Lesser Himalaya and an abrupt increase at the frontof the high range. Due to the difficulty of using the terrace record todetermine incision rates, the geometry of modern channels along major

40 AVOUAC

Page 41: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

rivers draining across the range may be used instead (Lave and Avouac,2001). Fluvial incision can be estimated from the shear stress exerted by theflowing water at the bottom of the channel, which can be shown to be agood proxy for river-incision rate. The model was calibrated from a few sitesin the Lesser Himalaya and sub-Himalaya where Holocene strath terracescould be dated. This approach has been shown to yield results consistentwith the terrace record of river incision, based on cases where the twoapproaches could be applied to a same reach. Figure 24 shows averageprofiles obtained from river incision estimated along six major rivers acrossthe Himalaya of central Nepal. It turns out that river incision in the LesserHimalaya is minor, less than a few mm/year, and rises abruptly at the frontof the high range to reach values of about 7–8 mm/year within a 50-km wide

FIG. 24. Rate of river incision across the Himalaya of central Nepal (Lave and Avouac,

2001). Dark shading shows river incision across the Siwalik hills deduced from three Holocene

terraces along the Bagmati river. River incision across the Lesser Himalaya (gray shading) was

derived from the computation of river incision along the major rivers of central Nepal based on

their present hydrological regime.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 41

Page 42: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

zone that coincides with the High Himalayan peaks and with the position ofthe midcrustal ramp.

5.2. Denudation across the Whole Range

In the sub-Himalaya, the analysis of the river terraces makes it clear thatriver incision is keeping pace with tectonic uplift. The present topographyonly accounts for at most 10% of the total volume of rock uplifted byactive folding and thrusting, implying that denudation also must be inequilibrium with tectonic uplift (Hurtrez et al., 1999). Denudation shouldequal fluvial incision if fluvial downcutting is the rate-limiting processthat drives hillslope processes. This should be expected if, as suggested forthe High Himalaya of northern Pakistan, hillslopes are near their criticalslopes for mass movement (Burbank et al., 1996). Mass processes areprobably the dominant factor in the Himalaya of Nepal, especially in thehigh range, where the steep hillslopes often show bare evidence for recentlandsliding. The highly dissected topography seems to be totally controlledby the fluvial network. In that case an estimate of the sediment yield canbe obtained from the fluvial incision rates along the major rivers. Suchestimation was done for 11 catchments in the Himalaya of central Nepal,within the Naryani and Sapt Kosi watersheds, for which measurements ofsuspended load are available (Lave and Avouac, 2001). The values obtainedthis way were found to be too large by about 20%. This misfit is probablyinsignificant in view of the large uncertainties in both the measurements andestimates, and might partly be offset by the contribution of the bedload.

Denudation rates in the Himalaya of Nepal might also be comparedto sediment yield delivered to the foreland basins and to the Bengalfan. The pattern of fluvial incision of Fig. 24 predicts a sediment flux of375 km2/Myr. This may be compared with the 1� 0.5� 106 m3/year erodedfrom the Himalaya over the last 2 Myr and deposited in the Gangeticplain and Bengal fan (Metivier et al., 1999). Given that the Ganges andBrahmaputra drain the Himalaya over an approximately 1800 km longstretch of the Himalaya, this value implies about 280� 110 km2/Myrof denudation on average across a section of the range. Although it isprobable that river incision and denudation of hillslopes have varied due toclimate changes (Burbank et al., 1996; Goodbred and Kuehl, 2000), over thelong term the pattern of denudation must resemble the curve proposed inFig. 24. We conclude that, to the first order, denudation of the wholelandscape matches the pace of fluvial downcutting, and that fluxes have notvaried dramatically over the Quaternary period.

42 AVOUAC

Page 43: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

5.3. Holocene Kinematics of Overthrusting along the MFT–MHT

The average slip rate on the MFT over the Holocene thus appears tobe very close to the total shortening rate across the range as measuredfrom GPS. It also compares well with geological estimates of the shorten-ing rate across the Lesser Himalaya of Nepal, which fall in the rangeof 16–25 mm/year (Hauck et al., 1998; DeCelles et al., 2001) (point 2 inFig. 25, top). This estimate also compares well with the Quaternary rate ofshortening across the Himalaya deduced from E–W extension in southernTibet. If this Quaternary extension is assumed to accommodate the lateralvariation of the direction of thrusting along the arcuate Himalayan front(Baranowski et al., 1984; Armijo et al., 1986; McCaffrey and Nabelek,1998), the estimated extension rate of about 10 mm/year implies a LateQuaternary shortening rate of 20� 10 mm/year (Armijo et al., 1986) (point 3in Fig. 25, top).

The observed pattern of erosion is found to closely mimic uplift aspredicted by slip along the flat–ramp–flat geometry of the MHT. Thetheoretical pattern of uplift might be simply estimated by assuming thatthe hanging wall is thrust along the MHT and only accommodates theMHT geometry by vertical shear (Molnar, 1987). If the topography isapproximately steady state, as argued in the previous section, the flexuralloading of the Indian plate does not vary, so that this model does not requireany computation of the flexural response of the lithosphere. This kinematicapproach predicts an uplift pattern that fits reasonably with the fluvial-incision pattern of Fig. 24 (Lave and Avouac, 2001). In Fig. 25 the observedincision pattern is compared with the patterns of tectonic uplift and erosioncomputed from the 2-D mechanical model described in the next section.It thus seems that over the Holocene period, all the shortening across theHimalaya has been absorbed by slip along one single major thrust fault.It should be noticed at this point that these kinematics, which represent anaverage over the Holocene period, differ from those expected from thecritical brittle wedge model.

5.4. Mechanical Model of Overthrusting: Evidence for Low Friction onthe Decollement and Coupling between Denudation and Uplift

Figure 25 shows the results of a mechanical model of long-termoverthrusting, designed to fit the average kinematics of the deformation overthe Holocene period (Cattin and Avouac, 2000). More details about the

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 43

Page 44: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 25. Velocity, horizontal, and vertical displacement rates (Cattin and Avouac, 2000).

(a) Velocity field relative to India computed from a finite element model which accounts for

the dependency of rheology on local temperature and for erosion–sedimentation at the surface.

The model assumes frictional sliding along the MHT across the brittle upper crust.

(b) Comparison between computed uplift and erosion profiles, and measured river incision

profile of Fig. 24. (c) Horizontal velocity relative to India of (1) sub-Himalaya as derived from

the slip rate on the MFT, (2) the high Himalaya as derived from the rate of sediment

progradation in the foreland and balanced cross-section, and (3) southern Tibet derived from

the E–W extension rate of southern Tibet. These data indicate that deformation of the hanging

wall of the MHT is negligible. The model (continuous line) fits the pattern providing that

friction on the flat portion of the MHT does not exceed 0.2, or 0.3 if a high pore fluid pressure is

assumed.

44 AVOUAC

Page 45: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

model are given in Appendix A. This model assumes a prescribed geometryfor the MFT, with its characteristic ramps and flats geometry, along whichquasistatic frictional sliding is allowed. Elsewhere, the medium is assumed todeform according to a combination of brittle failure and thermally activatedductile flow. Deformation thus depends on the assumed rheology and localtemperature. Various temperature fields and rheological laws have beentested and yielded similar results, showing that the long-term pattern ofdeformation does not provide critical information about these modelparameters. The model considers a section of the lithosphere subjected to20 mm/year of horizontal shortening. According to this model the MFT–MHT roots in a subhorizontal shear zone that fits the prominent mid-crustal reflector imaged from the INDEPTH profiles (Fig. 26). This is not asurprise, however. It is a result of the position of the downdip end of theprescribed fault’s geometry and of the thermal structure that has beenobtained assuming a priori that the fault continues northwards as asubhorizontal shear zone. The modeling should therefore be seen primarilyas a test of the internal consistency of the various data sets and hypothesesused to constrained the model’s geometry and thermal structure.

The mechanical model does provide important additional insights. First,it turns out that the friction along the MFT–MHT must be low. Indeed, theobservation that the overhanging wall does not shorten during overthrustingrequires that the friction at the base of the thrust sheet induces deviatoricstresses that are balanced by the slope of the topography. In other words,the slope of the wedge must be overcritical. If the friction was too high, thewedge would be undercritical and the hanging wall would deform, sothat there would be no slip on the MFT at the surface until the wedgetopography reached its critical value. If the possible effect of pore pressure isneglected, we found that the friction on the flat portion of the MFT must besmaller than about 0.13. A classical interpretation of low basal friction insuch an overthrusting context invokes fluid pressurization in the fault zone(Hubbert and Rubey, 1959). Even if a high-pore pressure is assumed, up to0.9 the lithostatic pressure, we calculate that the friction on the flat portionof the MFT must be lower than about 0.3. This value is significantly smallerthan the standard value for intact rocks of about 0.6 (Byerlee, 1978), but isin the range of empirical values inferred for some intracontinental weakfaults or along subduction zone. By contrast, the friction on the ramp couldbe up to about 0.6, because friction at the base of the thrust sheet is balancedthere by the steep slope of the front of the high range.

Second, the model leads to some natural balance between denudationand uplift rate. This balance is expressed in Fig. 25, in which it can be

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 45

Page 46: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

seen that denudation rate, modeled from a simple linear diffusion equation,approximately equals the uplift rate. It should be recalled that dynamicequilibrium of the topography requires, in fact, that erosion rate, e, balanceslocal uplift rate as well as the apparent uplift due to the advancingtopography, as expressed from Eq. (5). In the case where incision rate, I, isthe observed quantity, the slope of the topography, tg�, should be replacedby the stream gradient, S, so that Eq. (5) becomes

I ¼ U þ V0S: ð21Þ

This equation should be used to assess dynamic equilibrium of thetopography and might be used conversely to relate tectonic uplift todenudation when a steady-state topography is assumed a priori. From thisapproach, the incision rates along the major rivers across the Himalaya of

FIG. 26. Shear–strain rates over the long term computed by assuming a quartz rheology (a)

or a diabase rheology (b) (modified from Cattin and Avouac, 2000). In both cases a

subhorizontal shear zone develops that closely follows the midcrustal reflector imaged by the

INDEPTH experiment (Zhao et al., 1993).

46 AVOUAC

Page 47: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

central Nepal were shown to yield approximately the same pattern oftectonic uplift, and this pattern appears to match the prediction of themechanical model of Fig. 25 (Lave and Avouac, 2001). Sensitivity test showsthat the midcrustal ramp is required for the model to match the pattern ofuplift obtained from Eq. (21) (Cattin and Avouac, 2000; Lave and Avouac,2001). The dynamic equilibrium between denudation and tectonic upliftarises naturally from the coupling between surface processes and ductileflow at depth and contributes to localizing the deformation (Avouac andBurov, 1996).

5.5. Do the Kinematics of Deformation Determined over the HolocenePeriod Explain the Morphology and Structure of the Range?

The morphology of the Himalaya of central Nepal reflects a dynamicequilibrium between present tectonics and surface processes. Activetectonics in the Himalaya are primarily controlled by localized slip alongthe MHT and the topography is close to steady state because denudation,driven by fluvial downcutting, balances tectonic uplift. This scenario impliesthat the present morphology of the front of the high range is also the resultof present tectonics and erosion. The sharp relief together with the highuplift rates in the High Himalaya is interpreted here to reflect thrusting overthe midcrustal ramp rather than an isostatic response to reincision of theTibetan Plateau driven by late Cenozoic climate change (e.g. Burbank,1992), or a Late Miocene reactivation of the MCT (Harrison et al., 1997).Similarly, it is not necessary to invoke deformation near the front of thehigh range (Bilham et al., 1997). Such a scenario is improbable because itwould imply shortening beyond the 21 mm/year absorbed at the MFT, andthe total shortening rate across the range would therefore exceed the valueobtained from GPS measurements.

Now, we may test whether the kinematics of deformation established forthe Holocene period may be extrapolated back in time. This kinematics doesnot involve any accretion and cannot account for the fact that theHimalayan wedge was built through accretion of several thrust sheetsdetached from the Indian plate. In addition, it predicts a diachronousexhumation corresponding to VHR¼V0� 21 mm/year instead of the5 mm/year determined from Fig. 15. Finally, the position of the bulgewould be fixed with respect to the Indian plate and Vpr� 0 mm/year, sothat there should be no sediment progradation on the Indian plate nor

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 47

Page 48: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

subsidence (for a steady-state geometry of the flexed Indian plate). Clearlythis kinematics cannot reflect the long-term process. The simplestexplanation is that accretion occurs due to migration of the midcrustalramp, as described in Section 3.6 and Fig. 17, but that over theHolocene period no episode of accretion has occurred in central Nepal.So although the pattern of exhumation ages in Fig. 15 shows no evidencefor this, the accretion process is most probably discontinuous. Analternative explanation would be that the midcrustal ramp does notexist. The zone of high uplift rates north of the Lesser Himalaya could thenbe interpreted as due to underplating. It would then indicate the position ofthe zone of accretion of Fig. 17. This explanation seems difficult toreconcile with the structural section and, although the existence ofthe midcrustal ramp might appear conjectural, the first interpretation isfavored.

6. THE SEISMIC CYCLE IN THE NEPAL HIMALAYA

6.1. Large Historical Earthquakes in the Nepal Himalaya

The Himalaya has produced four earthquakes with magnitude larger thaneight since the end of the 19th century. Table 1 lists plausible values for theparameters of these events. Considerable uncertainty remains as to the exactlocation and size of the 1934 Bihar–Nepal earthquake, the most recent largehistorical event in the study area (e.g., Chen and Molnar, 1977; Pandey andMolnar, 1988). The epicenter was probably located in the Lesser Himalayaeast of Katmandu (Chen and Molnar, 1977). Assuming a thrust-faultdipping by about 5� to the north, long-period seismograms indicate areleased seismic moment of about 4.1� 1021 Nm, corresponding Mw� 8.4(Molnar and Deng, 1984). Macroseismic intensities and subsidence of theforeland revealed from leveling data suggest that the earthquake ruptureda 250–300 km along-strike segment of the arc (Bilham et al., 1998) (Fig. 3).The rupture area may have extended up to the MFT but probably notfarther to the south (Chander, 1989). The northward extent of the rupture isnot constrained at all.

There are historical indications that Katmandu has been struck byrepeated severe earthquakes in the past (Rana, 1935; Pant, 2002). Majordestructions were reported in 1833, 1681, 1408, 1344, 1255, and 1223. The1833 event might have ruptured about the same arc segment as the 1934

48 AVOUAC

Page 49: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

quake (Bilham, 1995), but the extent of damage in Katmandu suggestsa slightly smaller or more distant event. The 1255 was a major disasterthat caused the death of King Abhaga Molla and killed about one-thirdof the population of Katmandu valley (Pant, 2002). Very little is knownabout the 1344 event, which may have caused the death of King AriMolla (Pant, 2002). When assessing these historical earthquakes it should bepointed that some palaces and temples in Katmandu valley built after themid-15th century remained intact until the 1934 event. These constructionsare characterized by deep foundations (up to 15 m deep) and were reinforcedby wooden frames so that they could resist significant shaking. The templeof Shiva in Bakthapur was such a building, dated to 1458 AD, which wastotally destroyed in 1934 (Pandey and Molnar, 1988). Some other templesbuilt by the end of the 14th century or in the early 15th century werepreserved until they were destroyed in 1934. The 1681 and 1408 events aretherefore probably smaller than the 1934 earthquake. On this basis, thereturn period of earthquakes similar to or larger than the 1934 event mayfall between 100 and 475 years.

Among the large Himalayan earthquakes of the last century, the 1905Kangra event was probably the most similar to the 1934 event (Fig. 3).It ruptured a fault plane whose structural position was probablysimilar (Molnar, 1987; Chander, 1989). The analysis of the seismogramhas lead to a wide range of estimates corresponding to a surface wavemagnitude between 7.5 and 8.2, with the lower estimates being moreprobable (Ambraseys and Bilham, 2000). The area between the locations ofthe 1934 and 1905 events represents a � 800-km long seismic gap. The mostrecent event that ruptured its eastern portion is probably the 1833 event.A large magnitude event along its western portion was reported in 1803(Oldham, 1883). Although interpretation of the historical data is difficult,due to the political situation in the area at that time (Bilham et al., 1995), itmay have been a very large earthquake with magnitude around 8 thatruptured an arc segment between Dehra Dun and far-western Nepal(Bollinger, 2002).

6.2. Microseismic Activity and Geodetic Deformation in theInterseismic Period

Geodetic measurements over the last 10 years (Bilham et al., 1997;Jouanne et al., 1999; Larson et al., 1999) have revealed that, all along theHimalayan front in Nepal, relative displacements between the Gangetic

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 49

Page 50: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

plain and the Lesser Himalaya MFT (MHT) have been small. This isparticularly clear from the 10-point GPS network emplaced along sectionAA0 (Fig. 27). The results indicate horizontal contraction perpendicular tothe strike of the range and essentially confined to a 50-km wide zone at thefront of the High Himalaya, some 100 km north of the MFT. Velocities,relative to northern India, rise from about 3 mm/year around Katmandubasin to about 15-mm/year 50 km to the north. Data collected in southernTibet some 150 km east of the study area (Bilham et al., 1997; Larson et al.,1999) suggest that velocities taper to about 20 mm/year north of the HighHimalaya (Fig. 28).

Local seismic monitoring in the Katmandu area (Pandey et al., 1995,1999) has also revealed intense microseismic activity in this area, with twoMl� 6 events over the last 5 years (Fig. 27). Most events clustered ina narrow zone beneath the front of the high range at depths between

FIG. 27. Seismicity of Nepal recorded by the DMG seismological network

between 04/01/1995 and 10/12/1999 (Pandey and Rawat, 1999). The dashed line follows the

3500-m elevation contour line. Seismic activity north of this zone is mostly related to

active normal faulting in southern Tibet. South of this line seismic activity, due to NS

compression, is probably triggered by interseismic straining (Pandey et al., 1995; Pandey and

Rawat, 1999).

50 AVOUAC

Page 51: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

30 and 5 km (Fig. 28). This zone also coincides with high uplift ratesevidenced from the leveling measurements (Jackson and Bilham, 1994)(Fig. 28). A casual observation is that the seismicity belt abruptly ends to thenorth as soon as the elevation exceeds about 3500 m (Fig. 27). At places theseismicity cut-off can be seen to follow quite closely the sinuous shape of thiselevation contour line, suggesting some cut-off effect due to the topography(Bollinger et al., submitted-b).

These data suggest that over the last few years the MFT–MHT hasremained locked from where it emerges along the foothills to the midcrustalramp, while aseismic deformation has proceeded beneath the HighHimalayas and southern Tibet. This process would have resulted in stressbuildup around the downdip end of the locked fault zone, triggering seismicactivity.

The mechanical consistency of the simple model described above wastested by the calculations described in Appendix A (Cattin and Avouac,2000). If the friction on the fault is increased to a value high enough(about 0.17), frictional sliding along the fault is inhibited while ductileshear extends along the subhorizontal shear zone beneath the high rangeand southern Tibet. The model then predicts horizontal displacement anduplift rates consistent with the geodetic measurements (Fig. 28), as well asstress accumulation near the northern edge of the locked fault. It turnsout that most of the micro-earthquakes fall within this area of enhancedCoulomb stress (Cattin and Avouac, 2000). A simple explanationfor the seismicity cut-off corresponding to a particular elevation(3500 m) is that the vertical stress increase induces a change of thestress tensor such that the deviatoric stresses are no longer affected,or even decrease during interseismic strain buildup. The seismicity releasesan accumulated moment that amounts to less than 1% of the uppercrustal strain revealed from geodetic monitoring (Avouac et al., 2001).These earthquakes thus provide a strain release that is insignificantcompared to crustal deformation over the long term. The measuredgeodetic deformation is therefore either permanent and aseismic, or elastic.The second hypothesis should be preferred because, as discussed in theprevious section, long-term shortening across the range is entirelyaccommodated by localized slip on the MFT–MHT. The deformationmeasured over the last few years across the Katmandu section isnecessarily due to elastic straining of the upper crust that has to beultimately released by slip on the MHT. This stress transfer is probablychiefly the result of recurring major earthquakes such as the Bihar–Nepalearthquake.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 51

Page 52: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 28. (a) Structural section with seismicity located from three three-component seismic

stations installed in 1996 near section AA0. (b) Uplift rates derived from the measurements

along the leveling line in Fig. 2b (Jackson and Bilham, 1994) and computed from the

mechanical model of Cattin and Avouac (2000) (continuous line). (c) Horizontal velocities

52 AVOUAC

Page 53: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

6.3. Lateral Variations of Interseismic Straining

Interseismic stress buildup by elastic straining of the uppercrust is thus probably the main process responsible for the observed beltof microseismicity that can be traced along the front of the highrange all along the Himalayas of Nepal, including the ‘‘seismic gap’’west of Katmandu. Geodetic measurements also provide clear evidencethat the MFT is essentially locked all along the front of the Himalayaof Nepal (Jouanne et al., 1999; Larson et al., 1999; Bilham et al., 2001)and of northwest India (Banerjee and Burgmann, 2002). It seems highlyprobable that this portion of the Himalayan arc also produces largerecurrent earthquakes similar to the 1934 and 1905 events.

There may however be some along-strike variations of elastic strainingand stress buildup. These variations could result from lateral variations inthe geometry of the fault zone, or from seismic coupling with possiblenonstationary loading during the interseismic period. Such variations havebeen observed along subduction zones possibly reflecting stress builduparound major asperities (Dmowska et al., 1996). In the case of anintracontinental megathrust, such as along the Himalaya, heterogeneousinterseismic straining might be more easily detected than along a subductionzone because of the intracontinental setting and shallow dip of the fault, andalso because interseismic straining is thought to trigger the seismicity in thebelt along the front of the high range. The seismic activity over the last fewdecades along the Himalayan arc does show some lateral variations. Atplaces along the Himalayan clusters of arc shallow-dipping thrust eventswith relatively large magnitudes, up to Ms 6.5–7, have been detected(Baranowski et al., 1984). These events falling along the seismicity belt atfront of the high range can also be interpreted to result from interseismicstress buildup. The most recent such events are the 1991 Uttar Kashi(Ms� 7.2) earthquake and the 1999 Chamoli (Ms� 7.3) earthquake. Quite afew slightly smaller events have occurred in far-western Nepal betweenabout 81.5 and 82�E (Fig. 27). These clusters may correspond to somelateral geometric complexities of the Himalayan arc or alternatively reflect

relative to India determined from GPS measurements and computed from the mechanical model

of Cattin and Avouac (2000) (continuous line). Black dots indicate LDG measurements along

section AA0. Other measurements (Jouanne et al., 1999; Larson et al., 1999) were projected

on section AA0 with account for the arc curvature. To account for possible lateral irregularities

the location of the sites not located close to section AA0 are assigned a standard uncertainty

of 20 km.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 53

Page 54: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

different stage in the seismic cycle. In fact, 3-D modeling shows that thepattern of seismicity and the geodetic data can be fitted by a very simplemodel in which the fault is locked everywhere from the MFT at the surfaceto beneath the front of the high range, over a distance of about 100 km,without any significant lateral variation in the rate or direction of shortening(Bollinger, 2002). Variations in the pattern of interseismic straining aretherefore probably rather subtle. In particular, there is no evidence forheterogeneous straining such as might be expected if asperities were buildingup in the interseismic period. Lateral variations in the pattern of seismicitythus most probably reflect variations of the stress field induced by the effectof topography or by the effect of past large earthquakes.

6.4. A Model of the Seismic Cycle in the Central Nepal Himalaya

Motion along the brittle portion of the MHT is thus probably stick-slip asa result of recurring large earthquakes (Fig. 29). The elastic stressaccumulated in the upper crust during the interseismic period must betransferred to slip on the MFT. Some of the largest events along the front ofthe high range probably do participate in this stress transfer and activate aportion of the MFT–MHT. This may be argued, for example, for the UttarKashi event based on the analysis of the accelerometric data (Cotton et al.,1996). The author’s interpretation of such M� 7.5 events is that they did notpropagate all the way to the Himalayan piedmont probably because elasticstresses in the surrounding elastic medium were too low to sustain thepropagation of the seismic rupture. It might then be conjectured thatthe magnitude 7 events contribute to loading of the shallow-dippingdecollement below the Lesser Himalaya, raising elastic stresses there untilone event ruptures all the way to the earth surface. Such M� 8–8.5events would then account for most of the stress transfer to the front of thethrust-fault system (Fig. 28). If so, the average return period of largeearthquakes on the MHT in the study area could also be estimated from acomparison of slip rate on the MHT and coseismic slip. For an averagecoseismic slip between 4 and 6 m, a value that has also commonly beenestimated for other large Himalayan earthquakes (Molnar, 1990), andconsidering the slip rate on the MFT as well as the shortening rate across therange, it should take at least 180–330 years to accumulate the elastic strainreleased during such an event. Given that preseismic, postseismic, orpossibly aseismic creep events may also contribute to some fraction of the

54 AVOUAC

Page 55: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 29. Model of the seismic cycle in the Nepal Himalaya.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 55

Page 56: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

fault slip, the average return period of such an event might actually belarger.

7. DISCUSSION

7.1. Some Implications of the Himalayan Case for the Mechanics ofMountain Building

The pattern of river incision and the deformation of abandoned riverterraces has provided some key constraints on the Holocene kinematics ofdeformation across the Nepal Himalaya. It turns out that crustal shorteninghas essentially resulted from overthrusting of the Himalaya over the MFT–MHT thrust system with little internal deformation. During that processerosion seems to have maintained a balance between denudation andtectonic uplift, probably due to coupling between surface processes andcrustal flow at depth. The kinematics of deformation over a longergeological time period require in addition an accretion of material, probablyas the result of underplating near the brittle–ductile transition, a processdocumented elsewhere from the study of exhumed thrust system (Dunlapet al., 1997). The structure of the range suggests that underplating hasresulted from southward migration of a midcrustal ramp along the thrust-fault system. So, contrary to previous conceptions, the Himalayan wedgehas not grown by frontal accretion due to a forward propagation of thethrust sequence. It does not owe its critical wedge geometry to distributedbrittle thickening of the whole crust but to a combination of brittle–ductileunderplating and erosion at the surface. For a more general discussion ofthe impact of underplating on the dynamics of orogenic wedges, the readermight refer to Platt (1986). The thermal structure is a key factor indetermining these kinematics, because of its influence on the rheology of thecrust. The distribution of erosion is another fundamental factor due to itsinfluence on the thermal structure and its role in mass redistribution at thesurface. Any analysis of the mechanics of mountain building must take thesefactors into account.

7.2. Some Limitations of the Model

The mechanical modeling described here is based on a number of verysimple assumptions: erosion is modeled from linear diffusion; the faultgeometry is prescribed; fault motion is modeled from static stable sliding;

56 AVOUAC

Page 57: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

the thermal structure is assumed a priori; boundary conditions are expressedin terms of prescribed displacements; the mantle below the lithosphere isassumed to simply maintain some hydrostatic support. These assumptionsplace significant limitations on our ability to answer a number of questions:

– Because the fault geometry is imposed, the model shown here does notprovide any particular insight as to the mechanics of accretion over thelong term. In that respect, future investigations should be based on moreelaborated thermomechanical modeling allowing for large strains, aswell as strain localization.

– As a mountain grows, the increasing crustal thickness modifies the stressfield. If far-field tectonic boundary forces are assumed to be constants,the change in thickness should modify the force balance and thusinfluence strain rates. This effect is not included here. So the modelmight be appropriate to simulate a steady-state regime, but cannot beused to assess the long-term evolution of a mountain range.

– The model does a good job in predicting sediment yields that arecomparable to observations. However, it leads to a steady-stateparabolic topography that bears little resemblance to the realtopography of the Himalaya. This is due to the linear diffusion equationused to simulate erosion. More insight on the problem of the evolution,and eventual dynamic equilibrium, of the topography would requiremore sophisticated models of hillslope erosion, river transport, anddowncutting, with proper account of the coupling between theseprocesses. Such models would be required to get some idea of thetransient response of the morphology of a mountain belt, and of thesediment discharge to the lowlands, to possible changes in climateconditions or tectonic boundary forces.

– The present model does not adequately describe a possible timedependence of interseismic deformation, since the rupture dynamics andthe stress transfers associated with the seismic cycle are not taken intoaccount.

7.3. Evidence for Low Friction on the MFT–MHT

The observation that the footwall does not experience any significantshortening during thrusting along the shallow dipping MFT–MHTimplies a low friction of the order of 0.1, or 0.3 at most if a pore pressureis assumed. Another constraint on the friction along that fault zone comesfrom the observation that microseismic activity along the front of the

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 57

Page 58: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

high range shows a cut-off at an elevation of 3500 m. This threshold effectprobably corresponds to the point where Coulomb stresses no longerincrease, or actually decrease, during the interseismic period. This isachieved if the NS deviatoric stress is of the order of magnitude of thevertical stress variation due to the 3500-m difference in elevation withthe lowland. Unfortunately, estimating the effect of the topography onthe stress field at depth is not simple due to isostatic balance and elasticsupport. To the first order, the 3500-m elevation difference shouldcorrespond to a �zz increase of about 90 MPa. Given that the elevation inthe Lesser Himlaya is of the order of 1000 m, we infer deviatoric stresses ofat most 70 MPa, hence shear stresses on the underlying decollement ofabout 35 MPa at most. This is in keeping with a friction of the orderof 0.1, given the depth of the decollement around 10 km. Such a lowfriction could result from highpore fluid pressure. This seems a reasonablescenario if some foreland sediments are dragged along the interface, as theMT data might suggest. Fluid pressurization would then result fromsediment compaction during underthrusting. Dynamic friction effects mightalso play a role. If the thrust sheet only moves during large seismic events,the shear stress on the decollement may, in fact, always remain muchlower than the value required to balance static friction. In that case theapparent low friction would actually represent dynamic friction duringseismic rupture.

7.4. High Midcrustal Conductivity: Evidence for MetamorphismDehydration?

Here the author briefly comes back to the interpretation of theelectrical conductivity of Fig. 13. The comparison with the thermalstructure (Fig. 12) is instructive. Because the midcrustal zone of highconductivity does not follow any particular isotherm, in particular since itdoes not extend to the north, a thermometamorphic control model (e.g.,Marquis and Hyndman, 1992) can be excluded. This conductive body,which becomes prominent where the underthrusting footwall reachtemperatures of the order of 350�C, may instead result from the presenceof fluids released by metamorphic reactions, possibly chlorite breakdown.The released fluids would percolate upward through the zone where intensemicroseismic activity is triggered by interseismic straining (Cattin andAvouac, 2000), resulting in pathways for percolation through the brittleportion of the crust.

58 AVOUAC

Page 59: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

7.5. What Controls the Downdip End of the Locked Portionof the Fault?

The geodetic and seismic data provide some information on the positionof the downdip end of the locked portion of the fault. The mechanicalmodel of interseismic straining shown in Fig. 28 produces a smooth tran-sition from a fully locked fault zone to ductile flow along its northwardcontinuation, where temperatures exceed about 450�C. Close inspection ofthe velocity field reveals that a portion of the fault, near the base of theprescribed fault geometry, undergoes stable sliding. In fact the exactlocation of the transition cannot be constrained tightly from the GPScampaign data, and there is a variety of possible slip-rate distributions—including the one produced from the mechanical modeling—that would fitthe data equally well.

To assess more accurately the position of the downdip end of the lockedfault, and to avoid possible trade-offs among model parameters, we haveadopted a simplified kinematic approach (Durette, 2002). The modelassumes that stable sliding and ductile flow along the downdipcontinuation of the locked fault can be modeled using a slippingdislocation embedded in an elastic medium, an approach that has beenadopted in some previous studies (Jackson and Bilham, 1994; Bilham et al.,1997). For that analysis, we considered only the GPS data, correspondingto the arc segment that probably broke in 1934, i.e., east of the Katmandubasin. We also included velocities determined at four continuous GPSstations along that section and the IGS station in Lhasa (Fig. 30). Themodel parameters are: the dip angle of the dislocation, the position of theupdip end of the dislocation (corresponding to the downdip end ofthe locked fault zone). The best-fitting solution indicates a slip rate of18.5� 1.5 mm/year on a shallow-dipping fault. The position of the updipend of the dislocation is relatively well constrained and, when compared tothe thermal structure of Fig. 12, falls in a range of temperature between300 and 400�C (although large, this range of values ignores the uncertaintyof the thermal structure). Such a temperature range could correspond tothe transition from slip-weakening friction to aseismic stable sliding.Indeed, according to laboratory experiments and/or field observations, thisthermally activated transition occurs at a temperature around 325–350�Cfor quartzo-feldspathic rocks (Blanpied et al., 1991, 1995). This is also inkeeping with the observation that the downdip extent of the seismogeniczone along subduction zones generally coincides with the 350�C isothermif this temperature is reached above the Moho (Oleskevich et al., 1999).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 59

Page 60: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 30. Geodetic measurements of horizontal uplift rates (a) and displacement rates (b)

across the Himalaya of Nepal (Jackson and Bilham, 1994; Jouanne et al., 1999). Also shown are

the velocities determined at four continuous GPS stations along section AA0 that have been

operated since 1977 in collaboration with the Department of Mines and Geology of Nepal, and

at the IGS station in Lhasa (Courtesy of M. Flouzat and collaborators at LDG). These data

were processed using Bernese and Addneq softwares and referenced to ITRF 97 using seven

IGS stations (Durette, 2002). Continuous lines were obtained from least-squares adjustment

60 AVOUAC

Page 61: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

So, the rationale proposed for subduction zone (Hyndman et al., 1997),for cases where the locked fault zone does not extend deeper than theforearc Moho, seems to hold also for intracontinental megathrust faults.The coincidence between the transition to stable sliding and the zone ofhigher conductivity images from the magnetotelluric sounding experimentmay also suggest that the fluids and possible mineral changes caused bymetamorphic reactions may also influence the position of the downdip endof the locked fault portion.

7.6. Is Interseismic Straining a Stationary or Nonstationary Process?

The results described in Section 6 suggest that interseismic strainingmeasured over the last decade is representative of the average pattern andrate of straining during the interseismic period. The consistency, to betterthan about 10%, between long-term slip rate on the MHT and interseismicshortening rate across the range implies that interseismic straining isessentially constant, except possibly during relatively short periods of timecompared to the recurrence interval of large earthquakes. This is alsosuggested by the fact that there is no resolvable difference in the pattern ofinterseismic straining measured east of Katmandu, which presumably brokeduring the 1934 event, and that measured in western Nepal, which has notbroken for at least two centuries. Given the uncertainties in geodeticmeasurements and in the long-term slip rate, we cannot exclude thepossibility that there are lateral variations that could reflect variations in theinterseismic straining pattern and rate during the seismic cycle. If so, thisinformation might provide important constraint on the rheology of thecrust.

Indeed, simple consideration of the mechanics of the lithosphere implythat there should be some time variations in crustal straining in the periodbetween two large recurring earthquakes. To illustrate this point someresults obtained from a simple 1-D model of the seismic cycle is shown

from a 2-D dislocation model (Durette, 2002). The model is based on a 2-D analytical

approximation of surface deformation (Singh and Rani, 1993). Both GPS and leveling data

were weighted according to their standard deviation, assuming a Gaussian distribution of

errors. (c) Geometry of the creeping dislocation (continuous line) with 2� uncertainty in the

position of its end. The best-fitting solution corresponds to a slip rate of 18.5� 1.5 mm/year and

a dip angle of 1� 0.5�. The depth of the updip end of the dislocation is estimated at 15 km, with

a large 2� confidence interval of 8–25 km. The horizontal distance from the MFT is estimated

between 90 and 120 km. The reduced �2 min is 4.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 61

Page 62: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

(Perfettini and Avouac, manuscript in preparation) (Figs. 31 and 32). In thissimple model, the behavior of the thrust-fault system is simplified andmodeled as a simple springs and sliders system (Fig. 31). Three maindomains with different rheologies are considered, and the springs allowstress transfer between them during the seismic cycle. The upper faultportion is assumed to undergo unstable frictional sliding and is modeled byslider 1 with a state-and-rate friction law (Dieterich, 1979; Rice and Tse,1986). Its behavior is obtained by solving the following equations:

�1 ¼ �1f�* þ a1 logðV=V* Þ þ b1 logð�V*=DcÞg ð22Þd�=dt ¼ 1� ð�V=DcÞ ð23Þ

FIG. 31. Simplified spring-and-slider model. This 1-D model is meant to illustrate the effect

of stress transfer between the seismogenic fault portion, which is assumed to show stick-slip

behavior (slider 1), the ductile shear zone at depth, which is modeled from a Newtonian

viscosity (slider 3), and a transition zone where increasing ductility is assumed to promote

stable sliding (slider 2). A constant force is applied to slider 3 so that the long-term velocity

is 21 mm/year. The force corresponds to F¼ 96 MPa for ¼ 10�20 Pa s, and to F¼ 84 MPa for

¼ 10�19 Pa s. The ductile shear zone is assumed to have a thickness of 5 km. Other model

parameters are: a1¼ 0.04, b1¼ 0.005, a2¼ 0.02, �*¼ 0.1, �2¼ 2�1, �1¼ 270 MPa, Dc¼ 10�2 m,

V*¼ 10�10 m/s, h1¼ 10 km, h2¼ 2h1, G¼ 30 GPa. These values were chosen so as to imply

about 5 m of coseismic slip. Brittle creep parameters were assumed to correspond to the values

obtained from the adjustment of postseismic relaxation following the Chichi earthquake in

Taiwan (Perfettini and Avouac, in press).

62 AVOUAC

Page 63: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

FIG. 32. Slip motion of sliders 1, 2, and 3 for a Maxwell relaxation time of 211 (a) and

210 years (b).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 63

Page 64: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

where �1 is the shear stress; �1 is the effective normal stress set to somearbitrary value of 250 MPa (corresponding to a depth of about 10 km); �* isthe steady-state friction coefficient corresponding to some reference-slidingvelocity V*; a1, b1, and Dc are frictional parameters that were chosen inorder to obtain a coseismic slip of the order of 5 m.

The deeper fault portion is assumed to undergo rate-strengthening brittlecreep. This process is modeled from a rate-strengthening friction lawassociated with slider 2,

�1 ¼ �2f�* þ a2 logðV=V*Þg ð24Þ

where a2>0 to ensure stable sliding.The fault roots at depth in a ductile shear zone with viscosity , assumed

to be about w¼ 5 km thick, which is modeled from a viscous slider,

�3 ¼ *V3, ð25Þ

where *¼ /w.Motion of the most frontal slider, slider 1, would represent the slip along

the seismogenic fault portion. Motion of slider 3 represents the convergenceacross the fault system. To solve for the motion of the system with properaccounting for stress transfer during the seismic cycle, we apply a constantapplied force, F, on slider 3. The force was adjusted in order to produce anaverage slip rate of 20 mm/year.

For simplicity we assume that the fault segment with brittle creep is twotimes longer than the seismogenic zone, and lies at a depth about twice asgreat. The elastic modulus is assumed to be constant within the crust. Thesesimple assumptions imply that the stiffness of the second spring should beabout four times smaller than that of the first one so that we consider,

k1 ¼ 4k2 ¼ k ð26Þ

The effective normal stress �1 is set to an arbitrary value of 250 MPa,corresponding to 10 km depth, and �1 is taken to be twice as large.

The behavior of the system is then essentially determined from:

– the coseismic slip: �U;– the duration of the interseismic period: �t¼�U/V0;– the Maxwell relaxation time defined here as: tM¼ k/*;– the brittle relaxation time defined here as: tR¼ a2�2/kV0.

64 AVOUAC

Page 65: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

The model parameters in the simulations of Fig. 32 were adjusted so thatthe model always predicts a coseismic slip, �U, of the order of 5 m, and thefrictional parameters were not varied.

The brittle creep law associated with the second slider was chosenarbitrarily so as to correspond to a 8-year relaxation time, as deduced frompostseismic deformation in Taiwan (Perfettini and Avouac, in press). Fig. 32shows the slip history of all three sliders for two different viscosities thatcorrespond to a Maxwell relaxation time between 200 and 20 years.

Slider 1 has stick-slip behavior with 5 m of slip every 250 years. Slider 2produces some afterslip that decays rapidly over the first few years ofpostseismic relaxation. Later on its motion is controlled by the viscous slipof slider 3. If the viscosity is high, i.e., if the Maxwell time is of the order ofthe duration of the interseismic period, slider 3 has a nearly uniform motion(Fig. 32a). If the Maxwell time is much shorter than the duration of theinterseismic period, then some significant variations are observed.

This simple 1-D periodic model is meant to illustrate the effect of stresstransfers during the seismic cycle. In reality the fault slip and interseismicinterval vary from one event to another, so that the real behavior is notperfectly cyclic. The stress-transfer mechanism assumes a fault with infinitehorizontal extent. In reality the stress transfer occurs in 3-D and coseismicstress drop due to the rupture of a particular arc segment would becompensated by increased stresses on adjacent segment and elastic responseof the whole surrounding medium (not only of the thrust sheet as modeledhere).

The virtue of this model is essentially to illustrate that time variations ofstraining might be expected and, if detected, would provide uniqueconstraints on the rheological properties of the crust.

7.7. Seismic Coupling and Implication for Seismic Hazard

Geodetic measurements and background seismicity can be used jointly todelineate the locked portion of the fault, providing important informationon the potential location and size of future earthquakes. The major thrustfault along the Himalaya of Nepal appears to be locked from the earthsurface, at the MFT, to beneath the front of the high range, a distance ofabout 100 km. Does this means that this locked portion breaks during majorM� 8–8.5 earthquakes, the recurrence of which might be simply estimatedfrom the ratio of the coseismic slip and long-term slip rate (as in the 1-D

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 65

Page 66: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

model above)? This would be true if coseismic slip was the only mode oftransient slip on the MFT.

This question might be addressed by considering the seismic momentreleased by the historical earthquakes along the Himalayan arc as a whole.Since 1897, the large Himalayan earthquakes in Table 1 have released acumulative moment of about 17� 1021 Nm. For comparison, the momentreleased from all the seismic events recorded along the Himalaya over thelast 30 years amounts only to about 4� 1019 Nm. So the large seismic eventsaccount for nearly all, to within a few %, of the seismic moment releasedover the long term. If we assume that the width of the locked fault zone iseverywhere about 100 km and that the 21 mm/year slip rate in central Nepalis a reasonable average for the whole arc, we find that slip along the MFT–MHT should release about 1.5� 1018 Nm/year for a 100% seismiccoupling. So, the total seismic moment released since 1897 is close to thetotal amount to be released over a century for a 100% seismic coupling.However, if we now extend the same reasoning to a longer time period, saysince 1803, we come up with a very different estimate. If the 1803 and 1833earthquakes are assumed to have released seismic moments of the order of2–4� 1021 Nm, the large earthquakes could in fact amount to only about70% of the value obtained assuming a 100% seismic coupling. Suchan estimate would imply that the return period of large events like theBihar–Nepal earthquake should be revised to about 340 years, a value moreconsistent with the historical record of earthquake damage in the Katmandubasin.

So, by considering the historical earthquake records, seismic slip on theMHT accounts for 70–100% of the geologically determined slip rate on theMHT. It is therefore possible that part of the slip on the MHT could beaccommodated by some aseismic slip that might take place during thepostseismic phase or by any kind of transient events.

Some insight might be gained from the comparison of the Himalayancontext of central Nepal with the western flank of the central range inTaiwan, where the tectonic setting is very similar and where the Mw 7.6Chichi earthquake occurred in 1999 (e.g., Kao and Chen, 2000). Thisearthquake broke a shallow-dipping fault along the piedmont of the range.GPS measurements show that this fault was previously fully locked over adistance of about 40 km from the surface to a depth where it also roots in asubhorizontal aseismic shear zone (Dominguez et al., 2003). The earthquakewas followed by some afterslip along the downdip extension of the rupturedzone (Hsu et al., 2002), a process that can be modeled from the response ofthe stable-sliding portion of the fault to the coseismic stress change as

66 AVOUAC

Page 67: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

described using the springs-and-sliders system of the previous section(Perfettini and Avouac, in press). In fact, the earthquake ruptured only afraction of the previously locked fault (Dominguez et al., 2003) with a ratherheterogeneous slip distribution (Kao and Chen, 2000; Huang, 2001;Johnson et al., 2001). Although significant, afterslip was not sufficient tosmooth out coseismic slip heterogeneities. Also it did produce only a smallamount of slip along the portion of the fault that was locked before theearthquake but did not break during the main shock. This fault portionprobably slips at a rate around 40 mm/year to transfer deformationfrom ductile fault portion to the shallow faults that splay upwardincluding the Chelungpu and Changhua faults. It is unclear whetherslip along that fault portion is ultimately taken up by seismic events byother modes of transient slip. A dominantly seismic mode of slip ispossible, although no historical events would suggest that this faultportion might have produced any large earthquake in the past. Thepossibility for aseismic slow events in such a context, as has been observedon some occasion along subduction zones (Sacks et al., 1981; Dragert et al.,2001; Lowry et al., 2001) or on the San Andreas fault (Linde et al., 1996),should not be discarded. The point is that since asperities do not build up inthe interseismic period, recurring transient slip events (slow events andearthquakes) must sum up so that heterogeneities are smoothed out inthe long term.

We infer that on an intracontinental megathrust such as along theHimalaya of Nepal or along the central range of Taiwan, (1) a futureearthquake could break only a fraction of the previously locked fault zone,and (2) seismic coupling might be small, possibly of the order of 70%. Forseismic hazard assessment, a conservative hypothesis is to assume a 100%seismic coupling. Conversely, determination of crustal deformation basedon the seismic moment release is highly uncertain, even in case where thegeometry of the locked fault zone is known, due to the uncertainty in seismiccoupling.

8. CONCLUSION

The Himalaya is probably a unique place to address a variety ofgeological problems associated with mountain building processes. Someprogress has been made over the last decades due to various multi-disciplinary efforts that have provided important information on thesubsurface structure of the range and on the kinematics of deformation over

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 67

Page 68: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

different timescales. At this point, the available data can be assembled in arelatively consistent way in the light of our current understanding of themechanics of the lithosphere and of the seismic cycle. However, mostcurrent models in seismotectonics or morphotectonics are based on rather‘‘static’’ concepts: interseismic straining is considered uniform; thetopography is assumed to be steady state due to a balance between erosionand tectonic uplift; denudation of hillslopes is assumed to equal riverincision; fault zones geometry do not vary with time; and so on. We shouldbe able to refine our understanding of the dynamics underlying theseprocesses. It is believed that major advances will come from betterobservational constraints as well as improved physical models of theevolution over time of these various processes.

ACKNOWLEDGMENTS

My understanding of the Himalayan orogen has greatly benefited from interactions and

discussions with many former students, postdoc and colleagues in my group (Jerome Lave,

Laurent Bollinger, Rodolphe Cattin, Hugo Perfettini, S. Dominguez, Mireille Fouzat, Thierry

Heritier, Frederic Perrier) as well as with my Nepali colleagues, M. R. Pandey and his

collaborators from the Department of Mines and Geology.

APPENDIX A. MECHANICAL MODELING

The mechanical simulation of deformation during the interseismic periodor due to long-term overthrusting along the MFT–MHT, shown in Figs. 25,26, and 28, was computed from a 2-D finite element code (Hassani et al.,1997) initially designed to account for the mechanical layering of the crust.The code was modified to incorporate surface processes and the dependencyof rheology on local temperature (Cattin and Avouac, 2000).

A.1. Boundary Conditions

The model considers a 700-km long section (Fig. A1) that approximatesthe N18�E section through the Himalaya of central Nepal of Fig. 10. Themodel is initially loaded with the present average topography along theswath considered here. Only the brittle portion of the MHT is prescribed.It includes the flat beneath the Lesser Himalaya and the midcrustal rampat the front of the High Himalaya. Boundary conditions mimic an Indian

68 AVOUAC

Page 69: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

mantle lid that underthrusts southern Tibet (Fig. A1). Vertical andhorizontal displacements are excluded at the southern end of the model.At the northern end, vertical displacements are free, and for depths above40 km a southward horizontal velocity of 20 mm/year is imposed. Theupper crust is thus subjected to 20 mm/year of horizontal shortening.At the northern end and for depths below 40 km we exclude horizontaldisplacements so that the Indian mantle lid does not shorten. Thisdiscontinuity was arbitrarily placed at a depth of 40 km in the lowercrust. The model is long enough that the results in the Himalayan portionof the model are insensitive to the exact position of this discontinuity. Themodel is loaded with gravitational body forces (g¼ 9.81 m/s2) and issupported at its base by hydrostatic pressure to allow for isostatic restoringforces.

A.2. Surface Processes

Following Avouac and Burov (1996) denudation is modeled using a 1-Dlinear diffusion equation in the range, i.e., north of the MFT,

@h

@t¼ k

@2h

@x2, ðA1Þ

where k is the mass diffusivity coefficient, expressed in units of area per time(m2/year), h is the elevation, and x is the distance from the MFT. This lineardiffusion law assumes that the southward flux of sediments at the surface isproportional to the local slope. The value of k was chosen to insure adenudation rate consistent with the 250–675 m2/year flow of sediments

FIG. A1. Boundary conditions (modified from Cattin and Avouac, 2000).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 69

Page 70: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

eroded from the range (Lave and Avouac, 2001). In order to test thesensitivity of the model to erosion k was varied between 103 and 105 m2/year.

In the foreland (x� 0), as argued by Hurtrez et al. (1999), we assume thaterosion balances tectonic uplift or subsidence. This implies that denudationexactly compensates uplift at the MFT and sedimentation maintains a flatforeland at a constant elevation south of the MFT. This model is only anapproximation that was designed to account as simply as possible for thedependency of denudation on topography and sedimentation in the foreland(see Avouac and Burov, 1996 for discussion).

A.3. Rheology of the Continental Lithosphere

A depth-varying rheology has been incorporated with elasto-brittledeformation in the upper part of the crust and ductile deformation in thelower part. The empirical rheological equations and laboratory-derivedmaterial properties are used under the assumption that they can beextrapolated to geological conditions.

Under low stresses, rocks deform elastically. For isotropic materialsthe relationship between each component of strain ("ij) and stress (�ij) iswritten as

"ij ¼ 1þ

E�ij �

E�kk�ij ðA2Þ

where E and are the Young’s modulus and the Poisson’s ratio, respectively.Beyond the elastic domain, rocks deform brittlely or ductilely. We use an

elasto-plastic pressure-dependent law with the failure criterion of Drucker–Prager (Hassani et al., 1997). Failure occurs if

1

2ð�1 � �3Þ ¼ cðcot�Þ þ 1

2ð�1 þ �3Þ

� �sin� ðA3Þ

where c is the cohesion, � is the internal friction angle.Ductile flow in the lithosphere is empirically described by a law which

relates the critical principal stress difference necessary to maintain a steady-state strain rate to a power of the strain rate (Carter and Tsenn, 1987; Kirbyand Kronenberg, 1987; Tsenn and Carter, 1987). This power-law creep iswritten as

_"" ¼ APð�1 � �3Þn expð�EP=RTÞ, ðA4Þ

70 AVOUAC

Page 71: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

where R is the universal gas constant, T is the temperature, EP is theactivation energy, and AP and n are empirically determined ‘‘constants,’’assumed not to vary with stress and ( p,T) conditions. The ductile-flow lawis thus strongly dependent on rock-type and temperature.

We only distinguish the crust and upper mantle. For the upper mantle, anolivine-controlled rheology is assumed (see parameters in Table A1). For thecrust we have considered two end members, by considering either a quartz-controlled or a diabase-controlled rheology (Table A1). The quartz-likerheology implies a relatively thin elastic core in the middle crust and a thicklower crust that favors decoupling between mantle and crust, while adiabase rheology results in a stronger crust with a higher-viscosity lowercrust (Burov and Diament, 1995).

A.4. Frictional Sliding along the Fault

The MHT is assumed to follow a simple static friction law,

j�Tj � �0ð�N � pÞ � 0

Q j�Tj � ��N � 0, ðA5Þ

where �T and �N are the normal and the shear stress on the fault, and � isthe friction coefficient, �0 is the effective friction coefficient, and p is the porepressure.

TABLE A1. Physical Parameters and Material Properties used in the Mechanical Models

Parameter Quartz Diabase Olivine

� (kg/m3) 2900 2900 3300

E (GPa) 20 20 70

0.25 0.25 0.25

c (MPa) 10 10 10

� 30� 30� 30�

AP (Pa�n/s) 6.03� 10�24 6.31� 10�20 7� 10�14

n 2.72 3.05 3

EP (kJ/mol) 134 276 510

�, density; E, Young’s modulus; , Poisson’s ratio; c, cohesion; � , internal friction angle; AP, power-law

strain rate; n, power-law exponent; EP, power-law activation energy (Carter and Tsenn, 1987; Kirby and

Kronenberg, 1987; Tsenn and Carter, 1987).

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 71

Page 72: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

REFERENCES

Abtout, A. (1987). La determination des anomalies des champs de potentiel (magnetisme,

gravimetrie) et la structure de la lithosphere continentale en Asie.

Ambraseys, N., and Bilham, R. (2000). A note on the Kangra earthquake of 4 April 1905. Curr.

Sci. 79, 46–50.

Appel, E., and Rossler, W. (1994). Magnetic polarity stratigraphy of the Neogene Surai Khola

section (Siwaliks, SW Nepal). Himalayan Geol. 15, 63–68.

Appel, E., Muller, R., and Widder, R. W. (1991a). Palaeomagnetic results from the Tibetan

sedimentary series of the Manang area (north central Nepal). Geophys. J. Int. 104, 255–266.

Appel, E., Rossler, W., and Corvinus, G. (1991b). Magnetostratigraphy of the Miocene-

Pleistocene Surai Khola Siwaliks in West Nepal. Geophys. J. Int. 105, 191–198.

Arita, K., and Ganzawa, Y. (1997). Thrust tectonics and uplift process of the Nepal Himalaya

revealed from fission-track ages (in Japanese with English abstract). J. Geogr. (Tokyo

Geographical Society) 106, 156–167.

Arita, K., Dallmeyer, R. D., and Takasu, A. (1997). Tectonothermal evolution of the Lesser

Himalaya, Nepal: Constraints from 40 Ar/39 Ar ages from the Kathmandu Nappe. The

Island Arc 6, 372–385.

Armijo, R., Tapponnier, P., Mercier, J. L., and Tonglin, H. (1986). Quaternary extension in

Southern Tibet. J. Geophys. Res. 91, 13803–13872.

Avouac, J.-P., and Burov, E. B. (1996). Erosion as a driving mechanism of intracontinental

mountain growth. J. Geophys. Res. 101, 17747–17769.

Avouac, J.-P., and Tapponnier, P. (1993). Kinematic model of active deformation in central

Asia. Geophys. Res. Lett. 20, 895–898.

Avouac, J.-P., Bollinger, L., Lave, J., Cattin, R., and Flouzat, M. (2001). Le cycle sismique en

Himalaya. C. R. Acad. Sci. 333, 513–529.

Banerjee, P., and Burgmann, R. (2001). Convergence across the northwest Himalaya from GPS

measurements. Geoph. Res. Lett. 29 (13), 10.1029/2002 GL015184.

Baranowski, J., Armbruster, J., Seeber, L., and Molnar, P. (1984). Focal depths and fault plane

solutions of earthquakes and active tectonics of the Himalaya. J. Geophys. Res. 89,

6918–6928.

Barr, T. D., Dalhen, F. A., and McPhail, D. C. (1991). Brittle frictional mountain building 3.

Low-grade metamorphism. J. Geophys. Res. 96, 10319–10338.

Bilham, R., and England, P. (2001). Platean ‘pop-up’ in the great 1897 Assam earthquake.

Nature 410 (6830), 806–809.

Bilham, R. (1995). Location and magnitude of the 1833 Nepal earthquake and its relation to

the rupture zones of contiguous great Himalayan earthquakes. Curr. Sci. 69, 101–128.

Bilham, R., Blume, F., Bendick, R., and Gaur, V. K. (1998). Geodetic constraints on the

translation and deformation of India: Implications for future great Himalayan earth-

quakes. Curr. Sci. 77, 213–229.

Bilham, R., Bodin, P., and Jackson, M. (1995). Entertaining a great earthquake in western

Nepal; historic inactivity and geodetic test for the development of strain. J. Nepal Geol.

Soc. 11 (Special issue 1), 73–88.

Bilham, R., Gaur, V. K., and Molnar, P. (2001). Himalayan seismic hazard. Science 293,

1442–1444.

Bilham, R., Larson, K., Freymueller, J., and members, P. I. (1997). GPS measurements of

present-day convergence across the Nepal Himalaya. Nature 386, 61–64.

72 AVOUAC

Page 73: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Blanpied, M. L., Lockner, D. A., and Byerlee, J. D. (1991). Fault stability inferred from granite

sliding experiments at hydrothermal conditions. Geophys. Res. Lett. 18, 609–612.

Blanpied, M. L., Lockner, D. A., and Byerlee, J. D. (1995). Frictional slip of granite at

hydrothermal conditions. J. Geophys. Res. 100, 13045–13064.

Bollinger, L. (2002). ‘‘Deformation de l’Himalaya du Nepal’’ (PhD thesis), p. XI.

Bollinger, L., Avouac, J. P., Beyssac, O., Catlos, E. J., Harrison, T. M., Grove, M., Goffe, B.,

Vidal, O., and Sapkota, S. Thermal structure and exhumation history of the lesser

Himalaya, submitted-a.

Bollinger, L., Avouac, J. P., Cattin, R., and Pandey, M. R. Stress build up along the Himalayan

Megathrust, submitted-b.

Bollinger, L., Avouac, J. P., and Henry, P. Mountain building in the Himalaya: Reconciling

recent kinematics and long term geological deformation, submitted-c.

Brown, L. D. et al. (1996). INDEPTH deep seismic reflection observation of a regionally

extensive high-amplitude basement reflector, and associated ‘‘bright spots’’ beneath the

northern Yadong-Gulu rift, Tibet. Science 274, 1688–1690.

Brunel, M. (1983). ‘‘Etude petro-structurale des chevauchements ductiles en Himalaya

(Nepal oriental et Himalaya du Nord-Ouest)’’ (These doctorat d’Etat thesis), Paris VII,

Paris.

Brunel, M. (1986). Ductile thrusting in the Himalayas: Shear sense criteria and stretching

lineations. Tectonics 5, 247–265.

Bull, W. (1991). ‘‘Geomorphic Response to Climatic Change.’’ Oxford University Press,

326 pp.

Burbank, D. W. (1992). Causes of recent Himalayan uplift deduced from deposited patterns in

the Ganges basin. Nature 357, 680–683.

Burbank, D. W., Leland, J., Fielding, E., Andesron, R. S., Brozovic, N., Reid, M. R., and

Duncan, C. (1996). Bedrock incision, rock uplift and threshold hillslopes in the

northwestern Himalayas. Nature 379, 505–510.

Burchfiel, B. C., Zhiliang, C., Hodges, K. V., Yuping, L., Royden, L. H., Changrong, D., and

Jiene, X. (1992). The southern Tibetan detachment system, Himalayan Orogen: Extension

contemporaneous with and parallel to shortening in a collisional mountain belt. Geol. Soc.

Am. Special Paper 269, 1–41.

Burg, J. P. (1983). Himalayan orogen and global tectonics seen from the Tsangpo suture zone of

Tibet (China). In ‘‘Himalayan Orogen and Global Tectonics’’ (A. K. Sinha, ed.). Oxford &

IBH Publ. Co., New Delhi, pp. 35–44.

Burg, J. P., Brunel, M., Gapais, D., Chen, G. M., and Liu, G. H. (1984). Deformation of

leucogranites of the crystalline Main Central Sheet in southern Tibet (China). J. Struct.

Geol. 6, 532–542.

Burg, J. P., Leyreloup, A., Girardeau, J., and Chen, J. M. (1987). Structure and metamorphism

of a tectonically thickened continental crust: The Yalu-Tsangpo suture zone (Tibet). Phil.

Trans. R. Soc. Lond. A321, 67–86.

Burov, E. B., and Diament, M. (1995). The effective elastic thickness (Te) of continental

lithosphere: What does it really mean? J. Geophys. Res. 100, 3905–3927.

Byerlee, J. (1978). Friction of rocks. Pure Appl. Geophys. 116, 615–626.

Carter, N. L., and Tsenn, M. C. (1987). Flow properties of continental lithosphere.

Tectonophysics 136, 27–63.

Catlos, E. J., Harrison, T. M., Kohn, M. J., Grove, M., Ryerson, F. J., Manning, C. E., and

Upreti, B. N. (2001). Geochronologic and thermobarometric constraints on the

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 73

Page 74: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

evolution of the Main Central Thrust, central Nepal Himalaya. J. Geophys. Res. 106,

16177–16204.

Cattin, R., and Avouac, J.-P. (2000). Modeling of mountain building and the seismic cycle in

the Himalaya of Nepal. J. Geophys. Res. 105, 13389–13407.

Cattin, R., Martelet, G., Henry, P., Avouac, J. P., Diament, M., and Shakya, T. R. (2001).

Gravity anomalies, crustal structure and thermo-mechanical support of the Himalaya of

Central Nepal. Geophys. J. Int. 147, 381–392.

Chander, R. (1989). Southern limits of major earthquake ruptures along the Himalaya between

longitudes 75� and 90�E. Tectonophysics 170, 115–123.Chapple, W. M. (1978). Mechanics of thin-skinned fold-and-thrust belts. Geol. Soc. Am. Bull.

89, 1189–1198.

Chen, W. P., and Molnar, P. (1977). Seismic moments of major earthquakes and the average

rate of slip in Central Asia. J. Geophys. Res. 82, 2945–2969.

Copeland, P. (1997). The when and where of the growth of the Himalaya and the Tibetan

plateau. In ‘‘Tectonic Uplift and Climatic Change’’ (W. F. Ruddiman, ed.). Plenum Press,

New York, pp. 19–40.

Copeland, P., Harrison, T. M., Hodges, K. V., Marujeol, P., Le Fort, P., and Pecher, A. (1991).

An early Pliocene thermal disturbance of the Main Central Thrust, central Nepal:

Implications for Himalayan tectonics. J. Geophys. Res. 96, 8475–8500.

Corvinus, G. (1988). The Mio-Plio-Pleistocene litho- and bio-stratigraphy of the Surai khola

Siwaliks in west Nepal: First results. Comptes rendus des seances de l’Academie des Sciences

Paris, Serie D 306, 1471–1477.

Cotton, F., Campillo, M., Deschamps, A., and Rastogi, B. K. (1996). Rupture history

and seismotectonics of the 1991 Uttarkashi, Himalaya earthquake. Tectonophysics 258,

35–51.

Dahlen, F. A. (1990). Critical taper model of fold-and-thrust belts and accretionary wedges.

Annu. Rev. Earth Planet. Sci. 18, 55–99.

Dahlen, F. A., and Suppe, J. (1988). Mechanics, growth and erosion of mountain belts. Special

Paper Geol. Soc. Am. 218, 161–178.

Davis, D., Suppe, J., and Dahlen, F. A. (1983). Mechanics of fold and thrust belts and

accretionary wedges. J. Geophys. Res. 88, 1154–1172.

DeCelles, P. G., Gehrels, G. E., Quade, J., Ojha, T. P., Kapp, P. A., and Upreti, B. N. (1998).

Neogene foreland basin deposits, erosional unroofing, and the kinematic history of the

Himalayan fold-thrust belt, Western Nepal. G.S.A. Bull. 110, 2–21.

DeCelles, P. G., Robinson, D. M., Quade, J., Ojha, T. P., Garzione, C. N., Copeland, P., and

Upreti, B. N. (2001). Stratigraphy, structure, and tectonic evolution of the Himalayan

fold-thrust belt in western Nepal. Tectonics 20, 487–509.

Delcaillau, B. (1986). Dynamique et evolution structurale du piemont frontal de l’Himalaya: les

Siwaliks du Nepal oriental. Rev. geol. dyn. geogr. phys. 27, 319–337.

Dieterich, J. H. (1979). Modeling of rock friction, 1, experimental results and constitutive

equations. J. Geophys. Res. 84, 2161–2168.

Dmowska, R., Z. G., and R. J. R. (1996). Seismicity and deformation at convergent margins

due to heterogeneous coupling. J. Geophys. Res. 101, 3015–3029.

Dominguez, S., Avouac, J. P., and Michel, R. (2003). Horizontal co-seismic deformation of the

1999 Chi-Chi earthquake measured from SPOT satellite images: Implications for the

seismic cycle along the western foothills of Central Taiwan. J. Geophys. Res. 108, DOI:

10.1029/2001JB000951.

74 AVOUAC

Page 75: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Dragert, H., Wang, K., and James, T. S. (2001). A silent slip event on the deeper Cascadia

subduction interface. Science 292, 1525–1528.

Durette, B. (2002). ‘‘Traitement et analyse des series GPS continues dans l’Himalaya du

Nepal’’. Mem. D.E. A., Ec. Norm. Sup.-Universite Paris Sud., 59 pp.

Fort, M., Freytet, P., and Colchen, M. (1983). The Quaternary sedimentary evolution of the

intra-montane basin of Pokhara in relation to the Himalaya Midlands and their hinterland

(west central Nepal). In ‘‘Contemporary Geoscientific Researches in Himalaya’’ (A. K.

Sinha, ed.). Wadia Institute of Himalayan Geology, Dehra Dun, India, pp. 91–96.

France-Lanord, C., Derry, L. A., and Michard A. (1993). Evolution of the Himalaya

since Miocene time: Isotopic and sedimentologic evidence from the Bengal Fan. In

‘‘Himalayan Tectonics’’ (P. J. Treloar and M. Searle, eds.), Geological Society, London,

pp. 603–621.

Gansser, A. (1964). ‘‘Geology of the Himalayas.’’ Interscience Publishers, London, 289 pp.

Gautam, P., and Appel, E. (1994). Magnetic polarity stratigraphy of Siwalik Group

sediments of Tinau khola section in west central Nepal, revisited. Geophys. J. Int. 17,

223–234.

Gautam, P., Upreti, B. N., and Arita, K. (1995). Paleomagnetism and petrochemistry of the

Dowar khola volcanics, central Nepal sub Himalaya. In ‘‘9th Himalaya-Karakorum-Tibet

workshop’’ (B. N. Upreti, and M. R. Dhital, eds.). Journal of Nepal Geological Society,

Katmandu, pp. 179–195.

Goodbred, S. L., and Kuehl, S. A. (2000). The significance of large sediment supply, active

tectonism, and eustasy and evolution of the Ganges-Brahmaputra delta. Sedimentary Geol.

133, 227–248.

Guillot, S. (1999). An overview of the metamorphic evolution in central Nepal. J. Asian Earth

Sci. 17, 713–725.

Harrison, T. M., Copeland, P., Hall, S. A., Quade, J., Burner, S., Ojha, T. P., and Kidd, W. S. F.

(1993). Isotopic preservation of Himalayan/Tibetan uplift, denudation, and climatic

histories of two molasse deposits. J. Geol. 101, 157–175.

Harrison, T. M., Copeland, P., Kidd, W. S. F., and Yin, A. (1992). Raising Tibet. Science 255,

1663–1670.

Harrison, T. M., Grove, M., Lovera, O. M., and Catlos, E. J. (1998). A model for the

origin of Himalayan anatexis and inverted metamorphism. J. Geophys. Res. 103,

27017–27032.

Harrison, T. M., Ryerson, F. J., Le Fort, P., Yin, A., Lovera, O. M., and Catlos, E. J. (1997). A

late Miocene-Pliocene origin for the central Himalayan inverted metamorphism. E. P. S. L.

146, E1–E8.

Hassani, R., Jongmans, D., and Chery, J. (1997). Study of plate deformation and stress in

subduction processes using two-dimensional numerical models. J. Geophys. Res. 102,

17951–17965.

Hauck, M. L., Nelson, K. D., Brown, L. D., Zhao, W., and Ross, A. R. (1998). Crustal

structure of the himalayan orogen at 90deg east longitude from project INDEPTH deep

reflection profiles. Tectonics 17, 481–500.

Henry, P., Le Pichon, X., and Goffe, B. (1997). Kinematic, thermal and petrological model of

the Himalayas: Constraints related to metamorphism within the underthrust Indian crust

and topographic elevation. Tectonophysics 273, 31–56.

Hirn, A., and Sapin, M. (1984). The Himalayan zone of crustal interaction. Ann. Geophys. 39,

205–249.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 75

Page 76: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Hirn, A., Jobert, G., Xu, Z. X., and Gao, E. Y. (1984). Main features of the upper lithosphere in

the unit between the High Himalayas and the Yarlung Zangpo Jiang suture. Ann. Geophys.

2, 113–118.

Hodges, K. V., Parrish, R., and Searle, M. P. (1996). Tectonic evolution of the central

Annapurna range, Nepalese Himalaya. Tectonics 15, 1264–1291.

Hsu, Y. J., Bechor, N., Segall, P., Yu, S.-B., Kuo, L. C., and Ma, K. F. (2002). Rapid afterslip

following the 1999 Chi-Chi, Taiwan earthquake. Geophys. Res. Lett. 29, 10.1029/

2002GL014967.

Huang, B. S. (2001). Evidence for azimuthal and temporal variations of the rupture

propagation of the 1999 Chi-Chi, Taiwan earthquake from dense seismic array

observations. Geophys. Res. Lett. 28(17), 3377–3380.

Hubbard, M. S. (1989). Thermobarometric constraints on the thermal history of the Main

Central Thrust zone and Tibetan Slab, Eastern Nepal. J. Metamorphic Geol. 7, 19–30.

Hubbard, M. S., and Harrison, T. M. (1989). 40Ar/39Ar age constraints on deformation and

metamorphism in the Main Central Thrust zone and Tibetan Slab, Eastern Nepal

Himalaya. Tectonics 8, 865–880.

Hubbard, M. S., Royden, L., and Hodges, K. (1991). Constraints on unroofing rates in the

High Himalaya, Eastern Nepal. Tectonics 10, 287–298.

Hubbert, M. K., and Rubey, W. W. (1959). Role of fluid pressure in mechanics of overthrust

faulting. I. Mechanics of fluid-filled porous solids and its application to overthrust faulting.

Geol. Soc. Am. Bull. 79, 115–166.

Hurtrez, J. E., Lucazeau, F., Lave, J., and Avouac, J.-P. (1999). Investigation of the

relationships between basin morphology, tectonic uplift, and denudation from the study of

an active fold belt in the Siwalik Hills, central Nepal. J. Geophys. Res. 104, 12779–12796.

Hyndman, R. D., Yamano, M., and Oleskevich, D. A. (1997). The seismogenic zone of

subduction thrust faults. The Island Arc 85, 863–889.

Iwata, S. (1976). Late Pleistocene and Holocene moraines in the Sagarmatha region, Khumbu

Himal. J. Jpn. Soc. Snow Ice 38, 115–119.

Jackson, M., and Bilham, R. (1994). Constraints on Himalaya deformation inferred from

vertical velocity fields in Nepal and Tibet. J. Geophys. Res. 99, 13897–13912.

Jin, Y., McNutt, M. K., and Zhu, Y. (1996). Mapping the descent of Indian and Eurasian

plates beneath the Tibetan Plateau from gravity anomalies. J. Geophys. Res. 101,

11275–11290.

Johnson, K. M., Hsu, Y. J., Segall, P., and Yu, S. B. (2001). Fault geometry and slip

distribution of the 1999 Chi-Chi, Taiwan earthquake imaged from inversion of GPS data.

Geophys. Res. Lett. 28(11), 2285.

Jouanne, F., Mugnier, J. L., Pandey, M., Gamond, J. F., Le Fort, P., Serrurier, P., Vigny, C.,

Avouac, J. P., and members, I. (1999). Oblique convergence in Himalaya of western Nepal

deduced from preliminary results of GPS measurements. Geophys. Res. Lett. 26,

1933–1936.

Kao, H., and Chen, W. P. (2000). The Chi-Chi earthquake sequence: Active, out-of-sequence

Thrust faulting in Taiwan. Science 288, 2346–2349.

Kind, R., Ni, J., Zhao, W., Wu, J., Yuan, X., Sandvol, E., Reese, C., Nabelek, J., and

Hearn, T. M. (1996). Mid-crustal Low-velocity zone beneath the southern Lhasa block:

Results from the INDEPTH-II earthquake recording program. Science 274, 1692–1694.

Kirby, S. H., and Kronenberg, A. K. (1987). Rheology of the lithosphere: Selected topics. Rev.

Geophys. 25, 1219–1244.

76 AVOUAC

Page 77: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Larson, K., Burgmann, R., Bilham, R., and Freymueller, J. T. (1999). Kinematics of the India-

Eurasia collision zone from GPS measurements. J. Geophys. Res. 104, 1077–1093.

Lave, J. (1997). ‘‘Tectonique et erosion: L’apport de la dynamique fluviale a l’etude

sismotectonique de l’Himalaya duNepal central’’ (These de doctorat thesis), Paris 7, 226 pp.

Lave, J., and Avouac, J.-P. (2000). Active folding of fluvial terraces across the Siwaliks Hills,

Himalayas of central Nepal. J. Geophys. Res. 105, 5735–5770.

Lave, J., and Avouac, J. P. (2001). Fluvial incision and tectonic uplift across the Himalayas of

Central Nepal. J. Geophys. Res. 106, 26561–26592.

Le Fort, P. (1975a). Himalaya: the collided range: Present knowledge of the continental arc.

Am. J. Sci. 275A, 1–44.

Le Fort, P. (1975b). Himalayas: the collided range: Present knowledge of the continental arc.

Am. J. Sci. 275A, 1–44.

Lemonnier, C., Marquis, G., Perrier, F., Avouac, J.-P., Chitrakar, G., Kafle, B., Sapkota, S.,

Gautam, U., Tiwari, D., and Bano, M. (1999). Electrical structure of the Himalaya of

central Nepal: High conductivity around the mid-crustal ramp along the MHT. Geophys.

Res. Lett. 26(21), 3261–3264.

Linde, A., Gladwin, M., Johnson, M., Gwyther, R., and Bilham, R. (1996). A slow earthquake

sequence on the San Andreas fault. Research Review; CSIRO Exploration and Mining 383,

65–68.

Lowry, A. R., Larson, K. M., Kostoglodov, V., and Bilham, R. (2001). Transient fault slip in

Guerrero, southern Mexico. Geophys. Res. Lett. 28, 3753–3756.

Lyon-Caen, H., and Molnar, P. (1983). Constraints on the structure of the Himalaya from an

analysis of gravity anomalies and a flexural model of the lithosphere. J. Geophys. Res. 88,

8171–8191.

Lyon-Caen, H., and Molnar, P. (1985). Gravity anomalies, flexure of the Indian plate and the

structure, support and evolution of the Himalaya and Ganga Basin. Tectonics 4, 513–538.

Macfarlane, A., Hodges, K. V., and Lux, D. (1992). A structural analysis of the Main Central

Thrust zone, Langtang National Park, Central Nepal Himalaya. Geol. Soc. Am. Bull. 104,

1389–1402.

Marquis, G., and Hyndman, R. D. (1992). Geophysical support for aqueous fluids in the deep

crust—seismic and electrical relationships. Geophys. J. Int. 110(1), 91–105.

Marquis, G., Jones, A. G., and Hyndman, R. D. (1995). Coincident conductive and reflective

middle and lower crust in southern British Columbia. Geophys. J. Int. 120, 111–131.

Mascle, G. H., and Herail, G. (1982). Les Siwaliks: le prisme d’accretion tectonique associe a la

subduction intracontinentale himalayenne. Geol. alpine, Grenoble 58, 95–103.

Masek, J. G., Isacks, B. L., Gubbels, T. L., and Fielding, E. J. (1994). Erosion and tectonics at

the margins of continental plateaus. J. Geophys. Res. 99, 13941–13956.

Mattauer, M. (1975). Sur le mecanisme de formation de la schistosite dans l’Himalaya. Earth

Planetary Sci. Lett. 28, 144–154.

Mattauer, M. (1986). Intracontinental subduction, crust-mantle decollement and crust-stacking

wedge in the Himalayas and other continental belts. In ‘‘Collision Tectonics’’ (M. P.

Coward, and A. C. Ries, eds.). Geological Society of London, Special Publications,

London, pp. 37–50.

McCaffrey, R., and Nabelek, J. (1998). Role of oblique convergence in the active deformation

of the Himalayas and southern Tibet plateau. Geology 26, 691–694.

Meigs, A. J., Burbank, D. W., and Beck, R. A. (1995). Middle-late Miocene (>10Ma)

formation of the main boundary thrust in the western Himalaya. Geology 23(5), 423–426.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 77

Page 78: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Metivier, F., Gaudemer, Y., Tapponnier, P., and Klein, M. (1999). Mass accumulation rates in

Asia during the Cenozoic. Geophys. J. Int. 137, 280–318.

Molnar, P. (1987). Inversion profiles of uplift rates for the geometry of dip-slip fault at depth,

with examples from the Alps and the Himalaya. Ann. Geophys. 5, 663–670.

Molnar, P. (1990). A review of the seismicity and the rates of active underthrusting and

deformation at the Himalaya. J. Him. Geol. 1, 131–154.

Molnar, P., and Deng, Q. (1984). Faulting associated with large earthquakes and

the average rate of deformation in Central and Eastern Asia. J. Geophys. Res. 89,

6203–6227.

Molnar, P., and Lyon-Caen, H. (1988). Some simple physical aspects of the support, structure,

and evolution of mountain belts, in: Processes in continental lithospheric deformation.

Geol. Soc. Am. Spec., Rap. 218, 179–207.

Molnar, P., and Tapponnier, P. (1975). Cenozoic tectonics of Asia: Effects of a continental

collision. Science 189, 419–426.

Molnar, P., Thorson, E. T., Burchfield, B. C., Qidong, Deng, Xianyue, Feng, Jun, Li,

Raisbeck, G. M., Jianbang, Shi, Zhangming, Wu, Yiou, F., Huichan, You (1994).

Quaternary climate change and the formation of river terraces across growing anticlines on

the north flank of the Tien Shan, China. J. Geol. 102, 583–602.

Montgomery, D. R. (1994). Valley incision and uplift of mountain peaks. J. Geophys. Res. 99,

13913–13921.

Mugnier, J.-L., Leturmy, P., Mascle, G., Huyghe, P., Chalaron, E., Vidal, G., Husson, L., and

Delcaillau, B. (1999). The Siwaliks of western Nepal 1: Geometry and kinematics. J. Asian

Earth Sci. 17, 629–642.

Mugnier, J. L., Mascle, G., and Faucher, T. (1992). La structure des Siwaliks de l’Ouest Nepal:

un prisme d’accretion intracontinental. Bulletin de la Societe Geologique de France 163,

585–595.

Nakata, T. (1989). Active faults of the Himalaya of India and Nepal. In ‘‘Tectonics of the

Western Himalayas’’ (L. L. Malinconico, Jr., and R. Lillie, eds.). Geol. Soc. of America,

pp. 243–264.

Nelson, K. D., Zhao, W. J., Brown, L. D., Kuo, J., Che, J. K., Liu, X. W., Klemperer, S. L.,

Makovsky, Y., Meissner, R., Mechie, J., Kind, R., Wenzel, F., Ni, J., Nabelek, J.,

Chen, L. S., Tan, H. D., Wei,W. B., Jones, A. G., Booker, J., Unsworth,M., Kidd,W. S. F.,

Hauck, M., Alsdorf, D., Ross, A., Cogan, M., Wu, C. D., Sandvol, E., and Edwards, M.

(1996). Partially molten middle crust beneath southern Tibet: Synthesis of Project

INDEPTH Results. Science 174, 1684–1688.

Oldham, T. (1883). Catalog of Indian Earthquakes. Memoir Geol. Survey India 19, 163–215.

Oleskevich, D. A., Hyndman, R. D., and Wang, K. (1999). The updip and downdip limit to

great subduction earthquakes: Thermal and structural models of Cascadia, South Alaska,

SW Japan and Chile. J. Geophys. Res. 104(B7), 14965–14991.

Pandey, M. R., and Molnar, P. (1988). The distribution of intensity of the Bihar-Nepal

earthquake of 15 January 1934 and bounds on the extent of the rupture zone. J. Nepal

Geol. Soc. 5, 23–45.

Pandey, P., and Rawat, R. S. (1999). Some new observations on the Amritpur granite series,

Kumaun lesser Himalaya, India. Curr. Sci.

Pandey, M. R., Tandukar, R. P., Avouac, J.-P., Lave, J., and Massot, J. P. (1995). Interseismic

strain accumulation on the Himalaya crustal ramp (Nepal). Geophys. Res. Lett. 22,

741–754.

78 AVOUAC

Page 79: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Pandey, M. R., Tandukar, R. P., Avouac, J.-P., Vergne, J., and Heritier, T. (1999).

Seismotectonics of Nepal Himalayas from a local seismic network. J. Asian Earth Sci.

17, 703–712.

Pant, M. R. (2002). A step toward a historical seismicity of Nepal. Adarsa 2, 29–60.

Patriat, P., and Achache, J. (1984). India-Eurasia collision chronology has implications for

crustal shortening and driving mechanism of plates. Nature 311, 615–621.

Pecher, A. (1989). The metamorphism in Central Himalaya. J. Metamorphic Geol. 7, 31–41.

Perfettini, H., and Avouac, J.-P. Postseismic relaxation driven by brittle creep: A possible

mechanism to reconcile geodetic measurement and the decay rate of aftershocks,

application to the Chichi earthquake, Taiwan. J. Geophys. Res., in press.

Powell, C. M. A., and Conaghan, P. J. (1973). Plate tectonics and the Himalayas. Earth

Planetary Sci. Lett. 20, 1–12.

Rana, B. S. J. B. (1935). ‘‘Nepalako mahabhukampa (1990 sala) Nepal’s Great Earthquake.’’

Ratschbacher, L. e. a. (1994). Distributed deformation in southern and western Tibet during

and after the India-Asia collision. J. Geophys. Res. 99, 19917–19945.

Rice, J. R., and Tse, S. T. (1986). Dynamic motion of a single degree of freedom system

following a rate and state dependent friction law. J. Geophys. Res. 91, 521–530.

Robinson, D. M., DeCelles, P. G., Patchett, P. J., and Garzione, C. N. (2001). The

kinematic evolution of the Nepalese Himalaya interpreted from Nd isotope. E.P.S.L. 192,

507–521.

Sacks, I. S., Linde, A. T., Snoke, J. A., and Suyehiro, S. (1981). A slow earthquake sequence

following the Izu-Oshima earthquake of 1978. In ‘‘Earthquake Prediction’’ (D. W. S., and

P. G. Richards, eds.), AGU IV Series, Washington, DC, pp. 680.

Sakai, H. (1985). Geology of the Kali Gandaki Supergroup of the Lesser Himalayas in Nepal.

Memoirs of the Faculty of Science, Kyushu University, Series D, Geology 25, 337–397.

Sastri, V. V., Bhandari, L. L., Raju, A. T. R., and Datta, A. K. (1971). Tectonic framework and

subsurface stratigraphy of the Ganga basin. J. Geol. Soc. India 12, 222–233.

Saul, J., Kumar, M. R., and Sarkar, D. (2000). Lithospheric and upper mantle structure of the

Indian shield, from teleseismic receiver functions. Geophys. Res. Lett. 27, 2357–2360.

Scharer, U., and Allegre, C. J. (1983). The Palung granite (Himalaya); high resolution U-Pb

systematics in zircon and monazite. E. P. S. L. 63, 423–432.

Schelling, D. (1992). The tectonostratigraphy and structure of the eastern Nepal Himalaya.

Tectonics 11, 925–943.

Schelling, D., and Arita, K. (1991). Thrusts tectonics, crustal shortening and the structure of the

Far Eastern Nepal Himalaya. Tectonics 10, 851–862.

Searle, M. P., Parrish, R. R., Hodges, K. V., Hurford, M. W., Ayres, M. W., and

Whitehouse, M. J. (1997). Shisha Pangma leucogranite, South tibetan Himalaya: Field

relations, geochemistry, age, origin, and emplacement. J. Geol. 105, 295–317.

Searle, M. P., Windley, B., Coward, M., Cooper, D., Rex, A., Rex, D., Tingdong, L.,

Xuchang, X., Jan, M., Thakur, V., Kumar, S. (1987). The closing of Tethys and the

tectonics of the Himalaya. Geol. Soc. Am. Bull. 98, 678–701.

Seeber, L., and Armbruster, J. G. (1981). Great detachment earthquakes along the Himalayan

arc and long-term forecasting. In ‘‘Earthquake Prediction: An International Review’’

(D. E. Simpson and P. G. Richards, eds.), The American Geophysical Union, Maurice

Ewing Series 4.

Singh, S. J., and Rani, S. (1993). Crustal deformation associated with two-dimensional thrust

faulting. J. Phys. Earth 41, 87–101.

MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE 79

Page 80: MOUNTAIN BUILDING, EROSION, AND THE SEISMIC CYCLE IN …avouac/classes/GE277/Avouac04.pdf · India from the active southern margin of Eurasia can now be traced along the Indus-Tsangpo

Srivastava, P., and Mitra, G. (1994). Thrust geometries and deep structure of the outer and

lesser Himalaya, Kumaon and Garhwal (India): Implications for evolution of the

Himalayan fold-and-thrust belt. Tectonics 13, 89–109.

Stocklin, J. (1980). Geology of Nepal and its regional frame. J. Geol. Soc. London 137, 1–34.

Sun, W. (1989). ‘‘Bouguer Gravity Anomaly Map of the People’s Republic of China.’’.

Suppe, J. (1985). ‘‘Principles of Structural Geology.’’ Prentice Hall, Inc., Englewood Cliffs, NJ,

537 pp.

Tapponnier, P., Xu, Z. Q., Roger, F., Meyer, B., Arnaud, N., Wittlinger, G., and Yang, J. S.

(2001). Oblique stepwise rise and growth of the Tibet plateau. Science 294(5547),

1671–1677.

Tokuoka, T., Takayasu, K., Yoshida, M., and Hisatomi, K. (1986). The Churia (Siwalik) group

of the Arung khola area, west central Nepal. Memoirs of the Faculty of Science, Shimane

University 20, 135–210.

Tsenn, M. C., and Carter, N. L. (1987). Upper limits of power law creep of rocks.

Tectonophysics 136, 1–26.

Upreti, B. N. (1999). An overview of the stratigraphy and tectonics of the Nepal Himalaya.

J. Asian Earth Sci. 17, 577–606.

Valdiya, K. S. (1964). The unfossiliferous formations of the Lesser Himalaya and their

correlation. In ‘‘22nd International Geological Congress.’’ Delhi, pp. 15–36.

Van de Meulebrouck, J. (1983). Reconnaissance geophysique des structures crustales de deux

segments de chaıne de collision: le Haut Allier (Massif Central francais) et le Sud Tibet

(Himalaya).

Wang, C., Li, X., Hu, X., and Jansa, L. F. (2002). Latest marine horizon north of Qomolangma

(Mt Everest): Implications for closure of Tethys seaway and collision tectonics. Terra Nova

14(2), 114–120.

Wang, Q., Zhang, P., Freymueller, J. T., Bilham, R., Larson, K. M., Lai, X., You, X. Z.,

Niu, Z. J., Wu, J. C., Li, Y. X., Liu, J. N., Yang, Z. Q., Chen, Q. Z. (2001). Present-day

crustal deformation in China constrained by global positioning system measurements.

Science 294(5542), 574–577.

Weldon, R. J. (1986). Late Cenozoic geology of Cajon Pass; implications for tectonics and

sedimentation along the San Andreas fault. California Institute of Technology.

Willet, S. (1999). Orogeny and topography. J. Geophys. Res. 104, 28957–28981.

Willet, S., Beaumont, C., and Fullsack, P. (1993). A mechanical model for the tectonics of

doubly-vergent compressional orogens. Geology 21, 371–374.

Williams, C. A., Connors, C., Dalhen, F. A., Price, E. J., and Suppe, J. (1994). The effect of

brittle-ductile transition on the topography of compressive mountain belts on Earth and

Venus. J. Geophys. Res. 99, 19947–19974.

Yeats, R. S., Nakata, T., Farah, A., Mizra, M. A., Pandey, M. R., and Stein, R. S. (1992). The

Himalayan frontal fault system. Annales Tectonicae, Special Issue (Supp. to Vol. 6), 85–98.

Yin, A., Harrison, T. M., Murphy, M. A., Grove, M., Nie, S., Ryerson, F. J., Feng, W. X., and

Le, C. Z. (1999). Tertiary deformation history of southeastern and southwestern Tibet

during the Indo-Asian collision. G. S. A. Bull. 111(11), 1644–1664.

Zhao, W., Nelson, K. D., and project INDEPTH Team (1993). Deep seismic-reflection evidence

continental underthrusting beneath southern Tibet. Nature 366, 557–559.

80 AVOUAC