+ All Categories
Home > Documents > DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Date post: 02-Feb-2022
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
197
The Pennsylvania State University The Graduate School Department of Geosciences DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS OF OROGENIC DENUDATION IN THE HIMALAYA OF CENTRAL NEPAL A Thesis in Geosciences by Ian D. Brewer 2005 Ian D. Brewer Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2005
Transcript
Page 1: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

The Pennsylvania State University

The Graduate School

Department of Geosciences

DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS

OF OROGENIC DENUDATION IN THE HIMALAYA OF CENTRAL NEPAL

A Thesis in

Geosciences

by

Ian D. Brewer

2005 Ian D. Brewer

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

August 2005

Page 2: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

The thesis of Ian D. Brewer was reviewed and approved* by the following:

Douglas W. Burbank Professor of Geosciences

Thesis Advisor Chair of Committee

Rudy L. Slingerland Professor of Geosciences Head of Department of Geosciences

Peter B. Flemings Professor of Geosciences

Derek Elsworth Professor of Mineral Engineering Associate Dean for Research

Ian D. Brewer
Professor of Geosciences
Ian D. Brewer
*Signatures are on file in the Graduate School
Ian D. Brewer
Ian D. Brewer
Katherine H. Freeman Professor of Geosciences Associate Head for Graduate Programs and Research
Page 3: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

iii

ABSTRACT

This investigation examines the fundamental processes that determine the distribution of cooling

ages observed in detrital minerals eroded from orogenic belts. A detrital cooling-age sample

collected from a riverbed represents an integration of information from the upstream area. Within

orogenic belts that contain glacial cover and high relief, detrital minerals provide an easy method to

sample the range of cooling ages found within a basin. In addition, detrital-mineral

thermochronology can be used to extract information from the foreland stratigraphic record, which

extends the temporal applicability of the technique beyond traditional bedrock thermochronology.

For example, individual mineral grains can be extracted from a stratigraphic horizon and dated.

Following correction for the stratigraphic age of the horizon, the detrital mineral ages provide a

proxy for the erosion rates contained within the catchment area at the time the rock was deposited.

However, before reliable interpretations of the stratigraphic record are made, a modern calibration

of the technique was needed.

We investigated the spatial development of a modern cooling-age signal in the Marsyandi valley

of central Nepal with muscovite grains dated using 40Ar/39Ar thermochronology. Over 500

individual grains were dated from both the trunk stream and tributaries over a ~100-km transect

along the Marsyandi. These provide a database that displays striking contrasts along the length of

the Marsyandi River. The first stage of the investigation focused on the interaction of geological

parameters that control the distribution of detrital cooling ages from an individual basin. The range

of bedrock cooling ages contained within a catchment is determined by the erosion rate and the

depth of the closure isotherm (~350°C for muscovite). With a 2-D thermal model, we investigated

the effects of the vertical erosion rate and topography on the depth of the closure isotherm.

Increasing the erosion rate and/or topographic relief decreased the depth of the closure isotherms

below valley floors, and re-equilibration following sustained changes in the erosion rate took ~10

My. Once the range in cooling ages had been determined for a basin, the distribution of detrital

cooling ages in sediment at the basin mouth was calculated as a function of catchment hypsometry.

Page 4: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

iv

This approach was applied to two sub-catchments of the Marsyandi River. The predicted probability

distribution of cooling-ages matched the observed data better in the more slowly eroding basin, than

in the more rapidly eroding basin.

To understand the more complex distribution of cooling ages from the mouth of the Marsyandi

River, the basin was divided into smaller sub-basins that were modeled individually. To predict the

trunk-stream signal, individual tributaries were mixed as a function of their area, erosion rate, and

the percentage of muscovite in their sediment, which was determined from point counting.

Comparison of the model results with observed data illustrated that the detrital-cooling age signal

evolved downstream in an understandable manner, and suggested that the mechanical comminution

of muscovite was not significant over the length-scale of the basin. The pattern of spatial erosion

seen in the thermochronology – low erosion rates in the Tibetan zone, high erosion rates in the

Greater Himalaya zone, and intermediate erosion rates in the Lesser Himalayan zone – was broadly

similar to calculations of erosion rate based upon the point-counting results. Sample pairs were

dated to assess the temporal and spatial variability of the cooling-age signal within the fluvial

system. Results indicated that the samples were undistinguishable at the 95% confidence level, once

the effects of random selection and the number of grains dated had been accounted for.

A more integrated approach was used to predict the spatial distribution of bedrock cooling ages

within the 3-D landscape, and the distribution of detrital cooling ages resulting from the erosion of

that landscape. A 2-D kinematic-and-thermal model, using the assumption of a single orogen-scale

decollement, was developed to predict the depth of the closure isotherm as a function of the ramp

geometry and the relative partitioning of convergence between the Indian Plate underthrusting and

southern Tibet overthrusting. The thermal result was extrapolated laterally and combined with a

digital elevation model to predict the distribution of bedrock cooling ages. At any site in the

landscape, the cooling age is a function of the distance each rock particle travels after passing

through the closure isotherm and its speed along the trajectory predefined by the underlying

decollement geometry. Once the contribution of each site had been corrected for lithological factors

and the volume of sediment eroded, a theoretical cooling-age probability distribution was calculated

for the Marsyandi by the summation of age-probability for all sites within in the basin. The volume

of sediment was calculated as a function of the regional slope and the angle of the underlying ramp.

Comparison of various model runs with the observed data indicates that the best solution is obtained

Page 5: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

v

by partitioning the total of ~20 mm/yr of convergence between India and southern Tibet into 15

km/my of India underthrusting, and 5 km/yr of Asian overthrusting and subsequent erosion.

However, the exact partitioning is dependent upon the geometry of the decollement. A variant of the

model that assumed the modern Main Central Thrust represented the approximate surface trace of

the orogen-scale decollement produced better results than those runs that assumed the Main

Boundary Thrust represented the surface trace of the orogen-scale decollement. This provides

additional evidence that the MCT has been active recently. The new methodology of integrating

complex kinematic-and-thermal models with digital elevation models can be applied to any

orogenic belt, and it may be used to compare theoretical predictions against easily collected and

analyzed detrital-mineral data.

Page 6: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

TABLE OF CONTENTS

LIST OF FIGURES......................................................................................................ix

LIST OF TABLES .......................................................................................................xiii

ACKNOWLEDGMENTS............................................................................................xiv

INTRODUCTION........................................................................................................1

CHAPTER 1.................................................................................................................5

AN INTEGRATED APPROACH TO MODELING DETRITAL COOLING-AGE POPULATIONS: INSIGHTS FROM TWO HIMALAYAN DRAINAGE BASINS.

Abstract. .........................................................................................................5 1.0 Introduction..............................................................................................6 2.0 Review of Previous work.........................................................................9 3.0 Predicting the distribution of bedrock cooling ages ................................11 3.1 Thermochronology...................................................................................11 3.2 Determination of bedrock cooling ages ...................................................11 3.3 Thermal structure during active erosion ..................................................12 3.4 Steady-state landscapes............................................................................16 3.5 Spatial resolution .....................................................................................17 3.6 Vertical age distribution for a theoretical basin.......................................18 3.7 Distribution in ages at the basin mouth. ..................................................19 3.8 Examining the control of relief and erosion controls on the theoretical PDF of a

single drainage basin ..............................................................................21 4.0 Application to two Himalayan basins..................................................................22

4.1 Geological Background and sample sites ................................................22 4.2 40Ar/39Ar Analytical Protocols.................................................................24 4.3 Detrital cooling-age results and modeling theoretical PDFs ...................25

5.0 The construction of a grab-sample PDF. .............................................................27 5.1 PDF comparison and statistics. ................................................................27 5.2 Resolution of the detrital dating...............................................................28 5.3 Himalayan catchments – synthesis of theoretical PDFs and random sampling.

................................................................................................................30 5.4 Discussion of modeling results ................................................................31

Page 7: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

vii

6.0 Discussion............................................................................................................32 7.0 Conclusions..........................................................................................................35

CHAPTER 2.................................................................................................................49

THE DOWNSTREAM DEVELOPMENT OF A DETRITAL COOLING-AGE SIGNAL, INSIGHTS FROM 40AR/39AR MUSCOVITE THERMOCHRONOLOGY IN THE MARSYANDI VALLEY OF NEPAL.

Abstract......................................................................................................................49 1.0 Introduction..........................................................................................................50 2.0 Previous investigations of detrital thermochronology.........................................52 3.0 Geological Background .......................................................................................54 4.0 Methodology........................................................................................................55

4.1 Sampling strategy ....................................................................................55 4.2 40Ar/39Ar Analytical Protocols.................................................................57 4.3 Point Counting. ........................................................................................58

5.0 40Ar/39Ar results. ..................................................................................................59 6.0 Modeling..............................................................................................................60

6.1 Modeling the detrital cooling age signal .................................................61 6.2 PDF modeling results...............................................................................63

7.0 Discussion............................................................................................................65 7.1 Resilience of the detrital signal................................................................65 7.2 The reliability of the fluvial signal ..........................................................67 7.3 Spatial variations of erosion rate .............................................................69 7.4 Point-counting results ..............................................................................70

8.0 Conclusions..........................................................................................................72

CHAPTER 3.................................................................................................................90

THE APPLICATION OF THERMAL-AND-KINEMATIC MODELING TO CONSTRAINING ROCK-PARTICLE TRAJECTORIES, COOLING AGES OF DETRITAL MINERALS, AND THE TECTONICS OF THE CENTRAL HIMALAYA.

Abstract......................................................................................................................90 1.0 Introduction..........................................................................................................912.0 Geological background........................................................................................933.0 Thermal and Kinematic modeling .......................................................................95

3.1. Constraints on thrust geometry ...............................................................95 3.2. Thermal model ........................................................................................99 3.3. Particle trajectories and detrital cooling-age signals ..............................101

4.0 Modeling Results .................................................................................................106 4.1 Kinematics ...............................................................................................107

Page 8: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

viii

4.1.1 Convergence rates..........................................................................107 4.1.2 Angle of Main Himalayan Thrust ramp.........................................109

4.2 The modeled Marsyandi valley detrital cooling age signal .....................111 4.4 The effects of lithology............................................................................112

5.0 Discussion............................................................................................................113 5.1 Modeling..................................................................................................113 5.2 The single-decollement model.................................................................116 5.3 The stratigraphic record ...........................................................................118

6.0 Tectonic implications for the Himalaya ..............................................................119 7.0 Conclusions..........................................................................................................122

REFERENCES ............................................................................................................143

APPENDIX 1 ...............................................................................................................154

1.0 Thermochronology ..............................................................................................154 1.1 The decay equation ..................................................................................154 1.2 The potassium/argon decay scheme.........................................................156 1.3 The 40Ar/39Ar analytical method..............................................................157 1.4 Closure temperatures ...............................................................................158

APPENDIX 2 ...............................................................................................................161

1.0 40Ar/39Ar results and protocols.............................................................................161

APPENDIX 3 ...............................................................................................................180

1.0 Comparing PDF curves........................................................................................180

Page 9: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

ix

LIST OF FIGURES

CHAPTER 1

Fig. 1. Thermal structure of continental crust with erosion rates of (a) 1.0 km/My (b) 3.0 km/My after 20 My in a landscape with 4 km of relief. In scenario (c) and (d) relief is increased to 6 km with an erosion rate of 1.0 km/My and 3.0 km/My, respectively…………………………………....36

Fig. 2. Average depth of the 350°C isotherm below the valley floors as a function of varying relief and erosion rates after 20 My.……………………………………………………………………....37

Fig. 3. The temporal response of the depth of the 350°C isotherm for a system with 4 km of topographic relief. ……………………….…………….....38

Fig. 4. Construction of a “theoretical” PDF for an individual basin. …………………….………………………………………………….....39

Fig. 5. Relationship between summit ages and valley ages for topographic relief of 2, 4, and 6 km undergoing erosion rates of 0.5 to 3.0 km/My.………………………………………………………………….40

Fig. 6. Effects of uplift rate and relief on theoretical PDFs for a basin with a Gaussian distribution of land area with elevation..……………………..41

Fig. 7. Effects of hypsometry on theoretical PDFs………………………….....42 Fig. 8. Map of the upper Marsyandi drainage basin showing the detrital sample

locations.………..……….………………………………………..…......43 Fig. 9. SPDFs generated from the results of 40Ar/39Ar dating samples from the

upper Marsyandi (Sample 1) and Dordi basin (Sample 2). …………………………………………………………………………..44

Fig. 10. Plot of grain age versus age uncertainty for the 40Ar/39Ar analysis.……………..…………………………………………………..45

Fig. 11. Error calculation for a basin of 4-km relief eroding at 1 km/My.……………….…………………………………………………46

Fig. 12. Number of grains versus the mismatch error from 1000 iterations. …………….…………………………………………………………….47

Fig. 13. A selected range of outcomes from sampling 50 grains from the theoretical PDF (shaded gray) of a basin with 4 km of relief eroding at 1.0 km/My.……………………………………………………………...48

Page 10: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

x

CHAPTER 2

Fig. 1. Simplified geological map of the Marsyadi region.……………………..76 Fig. 2. Map of Marsyandi drainage system based on a 90-m

DEM…………………………………………………………………….77 Fig. 3. Detrital cooling-age PDFs for samples from the Marsyandi drainage.

.…………..……………………………………………………………...78 Fig. 4. Parameters controlling the contribution of an individual tributary to a

trunk-stream cooling-age signal. ……………………………………….79 Fig. 5. Model results compared to 40Ar/39Ar analyses. ………………………...80 Fig. 6. Predicted ages for specified erosion rates.………………...…………….81 Fig. 7. (a) Age versus error (1-σ) for analyses with greater than 40% radiogenic

40Ar. No clear relationship between age and error can be seen. Inset (b) shows a PDF generated from the 1-σ errors..…………………...………82

Fig. 8. Results of repeat sampling to test: a) the spatial variability of the detrital cooling age signal, and; b) the temporal variation of the signal. …………………………….…………………………………………….83

Fig. 9. Spatial variation in erosion rates at the drainage-basin scale. Erosion rates are taken from the results of modeling the detrital cooling age PDFs for individual tributaries. .………………………………………………......84

Fig. 10. The procedure used to convert point-counting results into relative erosion rates..……………………………………………………………85

Fig. 11. Spatial variation in erosion rates at the drainage-basin scale. Erosion rates are calculated from the point-counting data using the methodology illustrated in figure 9.……………….…………………………………...86

CHAPTER 3

Fig. 1. Location of the Marsyandi drainage basin and the study area……….....125 Fig. 2. Conceptual basis for the combined thermal, kinematic, and detrital

model.……………………………………………………………………126 Fig. 3. Constraints used for the kinematic-and-thermal model. ……………......127 Fig. 4. The three components needed to construct cooling ages for the

landscape. ……………………………………………………………….128 Fig. 5. Volume of material eroded in a time increment (dt) depends upon the

aspect of the topography in relation to the particle velocity (V). The gray line mirroring the present topography illustrates the volume of rock eroded in dt with an assumption of complete steady-state conditions.….129

Fig. 6. Calculation of volume of rock eroded in time increment (dt) for one digital-elevation model (DEM) grid cell, assuming a steady-state landscape. ……………………………………………………………….130

Page 11: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

xi

Fig. 7. Relationship between the topographic slope and particle trajectory angle in determining the volume of material eroded from a DEM cell.………….…………………………………………………………...131

Fig. 8. The three linear segments used as a proxy for the regional slope, taken from a transect normal to the strike of the orogen.……………………...132

Fig. 9. Three scenarios for partitioning convergence rate between India and south Tibet. ……………………………...………….………….………..133

Fig. 10. Effects of partitioning the relative convergence rate between India and Asia.………………………………………………………………….......134

Fig. 11. (a) The detrital cooling age signal from the entire mountain front compared against the sample from the mouth of the Marsyandi. (b) Corrected for the age-signal generated specifically from the Marsyandi basin. ……….………….………………………………………………..135

Fig. 12. The steady-state thermal structure with 20 mm/yr of total convergence with (a) 4 mm/yr of total convergence partitioned into Asia, and (b) 8 mm/yr of convergence partitioned into Asia.……….………….………………….………….…………………..136

Fig. 13. The distribution of cooling ages derived from different ramp geometries. ……..……………………………………………………….137

Fig. 14. A comparison of the distribution of detrital cooling ages using (i) a lithological correction and (ii) no lithological-correction factor…...…...138

Fig. 15. Transects…………...……….………………………...……….……….139 Fig. 16. Variation of “apparent” relief as a function of particle trajectory. ……140 Fig. 17. A comparison of the distribution of detrital ages from an orogenic

swath in the study area with: (i) the MHT represented by the MBT being the active fault, and; (ii) the MHT represented by activity on solely the MCT……….…………………………………………………………….141

Fig. 18. A cartoon showing three end-member models for Himalayan evolution. ……………….………………………………………………...………...142

APPENDIX 1

Fig. 1 Diagram illustrating the T-t path of a rock undergoing burial metamorphism and subsequent exhumation.…….…………..………….160

Page 12: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

xii

APPENDIX 2

Fig. 1. Age versus percentage of radiogenic 40Ar for the geochronological analyses presented in this paper.…….………….…….….……………...163

APPENDIX

Ian D. Brewer
3
Page 13: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

xiii

LIST OF TABLES

CHAPTER 1

CHAPTER 2

Table 1. Point-count data……………...………………………………………..87 Table 2. Drainage basin characteristics.………………………………………..89

CHAPTER 3

APPENDIX 1

APPENDIX 2

Table 1. Thermochronology isotopic data.…….…………...…………………..164

APPENDIX

Ian D. Brewer
Ian D. Brewer
3
Page 14: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

ACKNOWLEDGMENTS

This work was funded by National Science Foundation grants EAR-9627865, EAR-9896048, and

EAR-9909647. The Chevron Corporation provided a scholarship and the Krynine Funds from the

Department of Geosciences provided travel support. I would like to thank my advisor, Douglas

Burbank, who has given me much support and advice. Kip Hodges provided his scientific approach

and reviews, which helped improve our approach to the problems herein. My committee, Rudy

Slingerland, Peter Flemings, and Derick Elsworth have provided help full guidance and scientific

insight. Kevin Furlong, Rocco Malserversi and Chris Guzofski provided helpful discussions about

the thermal modeling. I would like to thank Kevin in particular for his excellent insights. Peter

Deines helped further my understanding of isotopic dating. Bill Olszewski, Jose Hurtado, and

Michael Krol provided invaluable help at the MIT laboratory and I would like to thank other

students for making my time in Boston enjoyable. Collaboration with John Garver was rewarding

and enjoyable, and a statistics discussion with Mark Brandon was useful. During my

Undergraduate degree at Oxford University, Mike Searle, Steven Hesselbo, John Dewey and Phillip

Allen passed on much of their enthusiasm. I would like to thank Mike in particular for the great

time we spent mapping in Khumbu. Thanks to friends and faculty at USC who helped me during

my first year in the USA. In Nepal, Dorjee Llama and Chandra Bdr. Niraula provided logistical

support and Pasang Kaji Sherpa and Dawa Tshering Sherpa managed to cope admirably with me in

the field. Dina Bandhu Baral helped prevent my arrest! Thanks to my office mates, Ann Blythe,

Mike Bullen, and Merri Lisa Formento-Trigilio for keeping me sane, and I would lastly like to

acknowledge all my friends here at PSU who have made my life very enjoyable during the course of

this work.

Page 15: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

xv

Little drops of water,

Little grains of sand,

Make the mighty ocean

And the beauteous land.

J.A..Carney (1845)

Page 16: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

1

INTRODUCTION

In 1895 the renowned French physicist Henri Baecquerel discovered radioactive decay, which

revolutionized the science of geology. Not only did radioactive decay provide an additional heat

source for the interior of the earth, supporting theories that the Earth was indeed very ancient, but

when Ernest Rutherford dated the first mineral in Montreal in 1905 geologists could begin to

quantify this antiquity. For the first time, events in Earth history could be placed within a calibrated

temporal framework. From this foundation less than 100 years ago the study of dating rocks,

“thermochronology,” has evolved into a major sub-discipline of Earth Sciences. Many different

radioactive systems are now employed, and a wide range of applications have been developed.

Despite this, the same basic methodology underlies all studies using radioactive isotopes, and

readers unfamiliar with the basics of thermochronology should refer to the overview in Appendix 1.

The initial motivation of thermochronology was to measure the formation age of a rock.

Thermochronometers with high closure temperatures, such as those using the U-Pb decay scheme,

for example, are typically used for such applications because of the similarity between closure and

magma crystallization temperatures. Low-temperature thermochronometry, however, has more

recently developed into an extremely useful tool. The 40K/40Ar series, for example, once primarily

used to date the effectively instantaneous cooling of extrusive igneous rocks, is now used to

constrain erosion and deformation during orogenesis. In active collisional belts, the rate of cooling

is driven by the rate at which a rock particle moves towards the surface. Thus, if the depth of

closure is known, a cooling-age taken from the surface today can be used as a proxy for the erosion

rate.

A limitation of bedrock thermochronology is that the temporal record is limited; only the rocks

found at the surface today can be dated. Detrital-mineral thermochronology is one way to overcome

this problem because sand grains can be preserved in basins. With increasingly accurate mass-

spectrometers, single crystals from the stratigraphic record can now be dated using laser

microprobes. If the cooling age of a grain extracted from the geological record is corrected for the

Page 17: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

2

stratigraphic age of the rock that contained it, it can provide a proxy for erosion rates at the time the

rock was deposited [e.g. Cerveny et al., 1988].

If we can illustrate that detrital-mineral ages produce a reliable indication of the rock uplift and

erosion variability within the contributing area, key questions concerning the tectonic and

physiographic evolution of mountain belts can be answered. “When did structural deformation

initiate in a given range? How rapidly did rock exhumation occur, and was there temporal

variability in the rate or erosion?” Previous studies using detrital thermochronology in the Himalaya

have either focused on constraining the stratigraphic age [Najman et al., 2001; Najman et al., 1997]

or, in an effort to gain insights into the temporal exhumation of the mountain range, have sampled

the stratigraphic record of cooling ages preserved in the Bengal Fan [Copeland and Harrison, 1990]

and the Pakistan Siwaliks [e.g. Cerveny et al., 1988]. The limitation of both the Bengal Fan and the

Pakistan Siwalik investigations is that the sampled sites have such vast upstream areas that only

very general inferences about erosion within the Himalaya can be drawn from these data.

The thrust of this investigation has been to understand modern detrital systems and to provide a

solid foundation for future applications in the stratigraphic record. In doing this we have had to

address some fundamental questions, the most basic being, “What is controlling the detrital cooling-

age signal?” This is a complex question because of the effects of the regional tectonics, distribution

of bedrock cooling ages, lithology, the fluvial system, and the fidelity of our sampling and dating

procedures. We focus on intermediate scale rivers that transect the width of the orogen, with

drainage basins on the order of 10,000 km2. These potentially produce detrital records that are

characterized by high spatial resolution, and are readily preserved in foreland basins within close

proximity to the mountain front. To date, studies of such rivers have neither been widely used to

constrain the modern pattern of erosion, nor to investigate the information preserved within the

stratigraphic record. The latter is because of the uncertainty in interpreting the results, whereas the

former is because bedrock thermochronology has traditionally been used.

In our investigation we examine the modern Himalaya, the preeminent collisional belt of the

Cenozoic. The combination of extreme relief, relative structural simplicity, and high erosion rates

make the Himalaya an excellent natural laboratory for our work. We focus on the Marsyandi valley,

which is located in central Nepal and divides the two 8000-m massifs of Annapurna and Manaslu.

Our thermochronometer of choice is muscovite, dated using 40Ar/39Ar procedures, because it has

Page 18: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

3

been widely used in the Himalaya, is resistant to weathering [Najman et al., 1997], and appears not

to have the same problems of inherited argon as biotite [Clauer, 1981]. In this study, we dated over

500 single crystals of muscovite. These were collected from sites on the main stem of the

Marsyandi River, as well as from many tributaries on either flank of the river. These data show

striking downstream changes in the distribution of ages, as well as clearly contrasting patterns

among tributaries in different parts of the range.

The work here is presented as three separate chapters, which to some extent represents a

progression of thought and complexity, although each chapter addresses different problems. The

first chapter focuses on two individual tributary basins and examines the most basic parameters

controlling the detrital cooling-age signal: the thermal system, erosion rate, and the distribution of

land area with elevation. “How does the closure isotherm respond to changes in erosion rate and

topography? How long does it take the thermal system to respond to changes in these parameters?”

These are key questions we have answered with a thermal model that examines the effects of

erosion rate and topography on the depth of the closure isotherm for muscovite (~350°C). Assuming

that vertical erosion dominates, we can use the results of the thermal model to predict the range of

bedrock cooling ages in a basin. The chance of dating a grain of a particular age from the sediment

at the basin mouth will be dependent upon the amount of land contributing that age. Thus the

probability of land area with elevation (hypsometry) can be used as a proxy for the probability of

dating a grain of a particular age.

One assumption to this basic methodology is that individual basins have uniform erosion rate

across them. Comparing the results to the full range of data collected from the Himalaya during

fieldwork in 1997 and 1998, it became apparent that further elements were needed to extend this

approach from models of individual tributaries, to investigating the complex detrital signal seen in

the trunk-stream entering the foreland basin. Thus, the subsequent chapter addresses key questions

that need to be answered in order to understand the evolution of a detrital signal within a much large

and complex drainage. “What factors control the volumetric contribution of a thermochronometer to

the trunk stream from individual tributaries? How is the spatial variation in denudation rate at the

tributary scale manifest in the trunk-stream signal? Does the downstream comminution of the

thermochronometer play an important role? How does variation in lithology affect the signal?” To

address these questions, a network approach was used, whereby the volumetric contribution of a

Page 19: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

4

predicted detrital cooling-age signal from an individual tributary, was modeled as a function of the

basin erosion rate, area, and percentage of thermochronometer. We found that our models of the

downstream evolution of the trunk-stream signal were broadly consistent with the data. The results

illustrate that erosion rates vary by a factor of at least two across the range and that there is little

evidence for the downstream comminution of muscovite.

Despite the success of these investigations, the differences between observed data and model

predictions focused our thoughts on further questions. “How can the pattern of bedrock-cooling

ages be explained in the context of regional tectonics? Given that the dominant style of Himalayan

deformation is thrusting along low-angle decollements, how does this affect the thermal structure?”

The final chapter represents a large step forward in answering these questions by defining a

methodology for combining complex geodynamic models with digital elevation models to predict

the distribution of detrital cooling ages. We constructed a 2-D kinematic-and-thermal model for the

Himalaya that answers questions such as: “How does the partitioning of underthrusting affect the

thermal conditions in the overthrusting plate, and to what extent does the geometry of the collision

zone manifest itself in the distribution of bedrock cooling ages?” With additional procedures, we

can predict how the distribution of bedrock cooling ages is manifest in the detrital record as a

function of differences in lithology, and how the erosion rate is controlled by the geometry of the

orogen.

Thus this modern calibration has explored and examined many of the assumptions that are needed

for the reliable interpretation of detrital cooling ages. The first two chapters provide a first-order

approach to understanding variations in the detrital-cooling age signals derived from an orogen,

which may be especially useful in regions where the geology is poorly constrained, or to focus

research on a specific geological parameter. The final chapter provides a more complex and

integrated approach that may be applicable to more detailed investigations of regional tectonics.

However, whether investigating the modern detrital signal, or examining the stratigraphic detrital

record, the insights gained from this investigation provide a considerably more solid foundation for

the future geological interpretation of cooling ages.

Page 20: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

5

Chapter 1

An integrated approach to modeling detrital cooling-age populations: insights from two Himalayan drainage basins.

I.D. Brewer and D.W. Burbank

Pennsylvania State University, Department of Geosciences, University Park, Pennsylvania

K.V. Hodges

Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology,

Cambridge, Massachusetts

Abstract.

The distribution of detrital mineral cooling ages in stream sediment can be used to investigate the

erosional history of mountain ranges. We have developed a numerical model that predicts detrital

mineral age distributions for individual basins undergoing vertical erosion. Although this model

requires a restrictive set of assumptions, its judicious application provides an opportunity to explore

the effects of thermal structure, erosion rate, relief, and basin hypsometry on cooling-age

distributions. We illustrate this approach by generating synthetic 40Ar/39Ar muscovite age

distributions for two basins with contrasting erosion rates in central Nepal. We then compare actual

measured cooling-age distributions for stream sediment samples with those predicted by the model.

Monte Carlo sampling is used to assess the mismatch that can be attributed to the number of grains

dated, how well a finite number of grain analyses reproduce the age distribution of the basin. A

Page 21: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

6

good match is produced between observed and predicted detrital cooling ages for a slowly eroding

Himalayan basin. The poorer match for a rapidly eroding basin may result from the wide range in

age uncertainties displayed by the real analyses, or inhomogeneous erosion rates within the basin.

Such mismatches emphasize the need for more accurate thermal and kinematic models to improve

geochronological data interpretations.

1.0 Introduction.

Thermochronology is a technique that allows scientists to place geological events into a quantitative

temporal framework. Although the initial motivation was to date the formation, or crystallization,

age of rocks and minerals, increasing interest has focused on applying low-temperature

geochronometers to investigations of sample cooling history. In active tectonic terrains, cooling

rates obtained through the application of isotope geochronometers are frequently used as proxies for

the unroofing rate of samples (the rate at which the sample moves towards the surface of the Earth).

As a consequence, the goal of many thermochronological investigations is not to strictly examine

the cooling history of a sample, but rather to constrain the rate of erosion that has exhumed it.

It is important to realize, however, that it is necessary to make a number of assumptions to convert a

cooling age into an erosion rate. Calculations typically contain, either explicitly or implicitly,

assumptions of: a) a specified geothermal gradient; b) constant erosion rate through time; and c)

vertical erosion and vertical rock-particle trajectories. This simplistic approach often does not

account for the many complex interactions that occur in nature. The local geothermal gradient, for

example, is a function of topographic relief, erosion rate, and the pattern of deformation in an area

[Henry et al., 1997; Mancktelow and Grasemann, 1997; Stüwe et al., 1994]. In this paper we

examine some of the basic assumptions required to calculate erosion rates, explore the extent to

which they are broadly reasonable, and investigate the interplay of parameters that control the

distribution of bedrock cooling ages.

The most rigorous thermochronological approach to quantifying the erosion history of bedrock is to

employ multiple dating systems to reconstruct detailed temperature-through-time cooling paths [for

Page 22: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

7

review of t-T and P-t-T investigations in the Himalaya see Searle (1996)]. Although excellent

insights are derived from such investigations, there are important limitations imposed on the utility

of such studies. Samples are limited to individual rock samples collected from the surface today,

and in modern orogenic belts that typically experience high vertical erosion rates, the low

temperature thermochronological record may only extend back a few million years. In addition,

detailed temperature-through-time paths from a bedrock sample provide information on the cooling

history of a single point that often has to be extrapolated to a much wider region.

It is a striking fact that a single handful of sand contains approximately half a million grains, each

derived from a slightly different source within the catchment. Detrital mineral geochronology

produces a spatially averaged erosion rate because a sample of sand taken from the mouth of a basin

provides an encapsulation of the spectrum of cooling ages upstream. A modern sand sample can be

rapidly collected from the river system and provides a means to sample the entire range of cooling

ages within the drainage basin. This may be especially useful within basins of extreme relief, or

containing significant glacial cover – areas in which bedrock thermochronology can be impractical.

Detrital geochronology further allows us to exploit the extensive temporal record of detritus shed

from mountain belts and preserved in sedimentary basins. Thus, the temporal applicability of

detrital dating may be extended far beyond traditional bedrock thermochronology to provide critical

information needed for reconstructing orogenic development [e.g. Cerveny et al., 1988; Copeland

and Harrison, 1990]. Precise spatial and temporal variations in tectonics, erosion rates, and the

resulting development of topography are necessary to understand global interactions of mountain

belts with global geochemical cycles [Derry and France-Lanord, 1996; Raymo et al., 1988], and

climate change [Kutzbach et al., 1993; Ruddiman and Kutzbach, 1989].

Detrital thermochronology is a technique, however, that arguably has not been employed to its full

potential because of issues concerning the interpretation, reliability, and significance of results. For

example, the unique, but unknown, specific source of each grain provides both opportunities and

limitations. Whereas it is possible to obtain a full representation of the cooling ages within a basin,

the generation of a detailed cooling history at a single point is precluded. Additionally, before we

can begin to interpret variations in denudation rate, a series of assumptions have to be made to

Page 23: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

8

convert the cooling-age record into erosion rates. Many of these assumptions underlie investigations

of bedrock geochronology, whereas some are specific to the technique of dating detrital minerals.

In this paper we introduce a numerical model that explores the detrital cooling-age signal of a basin.

The model description is divided into two parts to allow a systematic investigation of the controlling

parameters and an assessment of the underlying assumptions. In the first part of the model, we

present a methodology that uses erosion rate and relief to predict thermal structure and the

consequent distribution of ages with elevation. In a new approach, we use the topographic

characteristics of the basin to predict the relative abundance of detrital cooling ages in sediment at

the basin mouth; the theoretical distribution of ages that exactly reproduces the variability of

bedrock cooling ages within the catchment.

The second part of the model evaluates the consequences of the practical fact that we can date only

a finite number of grains and begins to quantify the random nature of detrital grain sampling. It

seems clear that increasing the number of grains produces a more robust analysis, but how can we

quantify this? Apart from initial investigations [e.g. Copeland et al., 1997; Stock and Montgomery,

1996], there have been few tests to determine a) the number of grains required to adequately

represent the entire age signal contained within the river, and b) the nature of the errors associated

with dating different numbers of grains. We use Monte Carlo integration to generate a synthetic

“grab” sample from the known “theoretical” population. This represents the geochronologist

selecting a finite number of grains to date from the millions present at the sample site. Using this

approach, we provide a quantitative measure of how different age distributions, generated from

different individual samples, fit the original cooling-age signal. This is the first methodology that

incorporates the complexity of the detrital signal to determine the uncertainty for a given number of

grains. To compare model results to real data and introduce our Monte Carlo analysis, we present

detrital muscovite ages from two Himalayan catchments. The samples were collected from two sites

within the Marsyandi Valley in central Nepal, and ~35 to 40 grains were dated from each using 40Ar/39Ar thermochronology.

Because our model for generating theoretical age distributions is based on simplifying assumptions,

it is not intended to exactly reproduce the distribution of ages within the data. Instead, if the

assumptions and resulting limitations of the methodology are accepted, we can apply the model to

Page 24: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

9

any basin in order to provide a first-order estimate of the parameters that control the detrital

cooling-age signal. The results of the analysis provide a basis to re-examine the initial model and

decide which assumptions are responsible for the mismatches, which are valid for the particular

region, and which may need further constraint and research.

2.0 Review of Previous work.

Detrital mineral geochronology has been used to investigate: (a) the thermal evolution of basins

[e.g. Green et al., 1996]; (b) source area constraints [e.g. Adams et al., 1998; Garver and Brandon,

1994; Hurford and Carter, 1991; Krogh et al., 1987]; (c) stratigraphic age [e.g. McGoldrick and

Gleadow, 1978; Najman et al., 1997] and; (d) the erosion history of orogens [e.g. Cerveny et al.,

1988; Copeland and Harrison, 1990]. The latter two approaches, in particular, are pertinent to the

problem of Himalayan erosion. Najman et al. [1997] used 40Ar/39Ar dating to constrain the

maximum depositional age of the oldest exposed Gangetic foredeep strata as 28 Ma. This

depositional age was interpreted to define the onset of significant exhumation in the Himalaya.

Although the stratigraphic age provides insight into orogenic exhumation, we focus here on the

detailed interpretation of the distribution of ages within the sediment and how to extract from them

information about the erosional history.

At present, 40Ar/39Ar (muscovite, biotite, k-feldspar, and hornblende), fission-track (zircon and

apatite), and (U-Th)/He (apatite and titanite) dating are the major geochronological methods used to

investigate the cooling history of mountain belts because of their relatively low (< 500°C) closure

or annealing temperatures. One of the first investigations of orogenic erosion using detrital sediment

from the stratigraphic record was undertaken by Cerveny et al. [1988]. In the geological record, the

cooling age of detrital minerals within a sedimentary rock can be corrected for the stratigraphic age

of that rock and then used as a proxy for the erosion rate at the time of deposition. Detrital zircons

from Indus River sediments in the Pakistan foreland indicated that a young 1 to 5 My (at the time of

deposition) cooling-age signal had been persistent since 18 Ma. This suggested that the modern

high-cooling rates experienced by the Nanga Parbat region had been a feature of the Indus River

catchment since early Miocene times. Although an analysis of a large catchment such as the Indus

Page 25: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

10

provides a good overview of the regional deformation, the analysis of the data is limited by the

uncertainty of the detrital source area. Detrital zircon fission-track analysis has also been employed

to investigate the temporal variations in erosion rate of the Olympic range and the British Columbia

Coastal Ranges [Brandon and Vance, 1992; Garver and Brandon, 1994]. In addition, combined

investigations of detrital-zircon fission-track and U-Pb dating using zircons from the same

stratigraphic horizon were used to distinguish the exhumation of igneous and metamorphic sources

in the Khorat basin, Thailand [Carter and Moss, 1999].

Copeland and Harrison [1990] used 40Ar/39Ar analysis of detrital K-feldspar and muscovite grains

from ODP sites on the southern end of the Bengal fan to interpret Himalayan erosion rates through

time. They show that the youngest cooling age at each stratigraphic level is approximately equal to

the depositional age, suggesting that high erosion rates have persisted for the past 18 My. However,

the temporal and spatial pattern of erosion within the catchment area is impossible to determine

from a sample so far from the orogenic front: we know only that rapid cooling was occurring at

some location within the drainage basin.

Despite the value of the aforementioned studies, they reveal little about the combination of physical

processes that control the detrital signal. Stock and Montgomery [1996] studied the theoretical

effects of relief on the range of ages found in the detrital record. They investigated the potential to

constrain paleotopography using the detrital cooling-age signal with the assumptions that: a) there is

good grain-age precision; b) the grains retain their isotopic ages during transport and deposition; c)

the sediment from the basin is well mixed; and d) there are near-horizontal isotherms. They found

that a sample of 40 grains is required to provide a 90% probability of capturing 90% of the relief of

a basin. We use a similar, but more integrated, approach to try to understand the thermal

framework, basin characteristics, and sampling statistics that control the complete cooling-age

signal from a particular basin.

Page 26: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

11

3.0 Predicting the distribution of bedrock cooling ages

We have developed a numerical model that predicts the detrital cooling-age signal from basic basin

characteristics. The model may be divided into two distinct parts and each is addressed separately.

The first part involves constructing a theoretical distribution of cooling ages for a basin, which we

define as the age signal that would be found by dating an infinite number of zero-error detrital

grains that completely sampled every point within the basin. We examine the model results by

comparing them to real data from two Himalayan catchments. In practice, however, only a finite

number of grains can be dated at a given site. Thus, the second part of the numerical model

addresses the uncertainties introduced by the limited number of analyses.

3.1 Thermochronology

This paper is concerned with 40Ar/39Ar dating using muscovite because this has been widely used in

both investigations of both bedrock and detrital-mineral dating. Moreover, there appears to be fewer

problems with excess argon in muscovite than are found with biotite [Roddick et al., 1980]. Our

simplistic treatment assumes that all muscovite samples have the same closure temperature

(~350°C) that is independent of grain size and the rate of cooling through the closure interval [see

Dodson, 1973]. This methodology is typical in geochronology, although when investigating spatial

variation in erosion rates, the latter assumption is clearly violated. In addition, we ignore the

potential effects of inherited radiogenic 40Ar.

3.2 Determination of bedrock cooling ages

In active orogenic belts, cooling ages are determined by the thermal structure and particle velocities

within the crust. The thermal structure of the crust will determine the depth of the mineral closure

temperature. The velocity path that individual rock particles follow to reach the surface will

Page 27: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

12

determine how long rocks take to be transported from the closure temperature to the land surface,

i.e. the cooling age.

The majority of geochronological investigations assume a simplified kinematic geometry such that

rocks pursue a vertical trajectory towards the surface, with erosion occurring perpendicular to that

path. Thus calculations based on vertical erosion typically simplify more complex processes

involving lateral advection that are driven by horizontal tectonics and deformation. Lateral

advection usually lengthens the path that a geochronometer takes to the surface. Hence, the particle

travels further than estimated with solely vertical transportation, and cooling ages could

underestimate erosion rates.

Nevertheless, we follow conventional practice in this study by assuming that particle uplift and

erosion are 1-dimensional processes such that an average unroofing rate (dz/dt) may be calculated

from:

=

ct1.

(dT/dz))T-(T

dtdz sc

(1)

Where Tc is the closure temperature of the geochronometer, Ts is the surface temperature, dT/dz is

an assumed geothermal gradient, and tc is the closure age of the mineral.

3.3 Thermal structure during active erosion

Even with a 1D system, estimating the geothermal gradient is of prime importance in determining

erosion rates. Two end-member approaches may be taken to approximate the depth of the closure

isotherm: a) assume an average geothermal gradient to predict the depth of the isotherm, or; b) use

numerical models to predict the depth of the isotherm. A detailed modeling approach might seem to

be the best solution. However, uncertainties in temporal location and rates of fault movement,

distribution of heat-producing radioactive isotopes, the role of fluids, erosion rate, and subsequent

Page 28: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

13

topographic development, dictate that this approach is often under-constrained and difficult to apply

to many situations. By default, most geochronological investigations use the first approach,

specifying a linear vertical geothermal gradient below the sampling location. The choice of a

geothermal gradient for an active region is, however, problematic. Most continental heat flow data

are taken from boreholes, commonly in basins, and away from tectonically active mountain belts.

Therefore, an “average” orogenic geothermal gradient is difficult to determine, and constraints

applicable to a particular location are even more elusive.

Because of these problems, we use a widely applicable model-based approach to estimate the depth

of the closure isotherm. Two main thermal effects determine the depth of the closure isotherm and

need to be addressed before we can start to analyse geochronological data. In actively eroding

orogenic belts, the near-surface geothermal gradient will be perturbed by erosion and relief. The

erosion rate controls how rapidly the crustal column moves towards the land surface, and thus the

rate of vertical heat advection by the rock mass. More rapid erosion causes more rapid advection of

heat to shallow depths. Relief controls the surface area of the mountain belt and so will affect the

thermal structure. Increasing the relief produces more efficient cooling because there is more

surface area of rock in contact with the atmosphere.

Previous analytical and numerical investigations [Mancktelow and Grasemann, 1997; Stüwe et al.,

1994] have examined the effects of erosion and relief on the thermal structure of a lithospheric

column, with the basal boundary condition fixed at depths of 50 to 100 km. In these models, the

entire lithosphere is heated by the vertical advection of a rock column (with associated heat) that

has mantle temperatures at its base. The models reach steady-state solutions in ~40 My

[Mancktelow and Grasemann, 1997].

We use a slightly different approach because, in reality, areas that experience rapid uplift rates

typically do not undergo rock and heat translation from mantle depths. Instead, in most mountain

belts, movement of rock occurs from the mid to lower crust to the surface along decollements, and

hence the crustal column experiences less overall heating. We use a model that approximates rock

mass that moves in laterally along a deep crustal decollement before eroding vertically towards the

surface. This limits the kinematic portion of the system to mid-to-lower crustal temperatures (depths

of 35 km), rather than upper mantle temperatures (depths of 50 to 100 km in other models).

Page 29: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

14

We investigate the approximate 2-D, steady–state thermal structure using a 2-D combined diffusion

and advection finite-difference scheme [Fletcher, 1991]. Average values for continental crust are

used with surface heat flow of 57x10-3 W/m2 and a uniform radioactive heat production of 1.0x10-6

W/m3 [Fowler, 1990]. The basal boundary condition is fixed at 35-km depth to the temperature at

time t = 0. The sides are zero heat-flow boundaries. Erosion is simulated by the vertical advection of

rock, from the basal boundary, through steady state topography which is generated instantaneously

after t > to. The surface temperature (Ts) is set to 20°C on the valley floor and we assume a lapse rate

(variation of temperature with elevation) of 6.5°C/km [Bloom, 1998]. The topography has straight

hill slopes (in 2D) that are fixed at a 30° angle to model a landslide-dominated landscape, in steady

state, at threshold conditions [Burbank et al., 1996]. The model is run for 20 My because the 350°C

isotherm responds relatively rapidly, and this is a time frame comparable to the longevity of major

structures within many mountain belts. The model output was compared to the basic steady-state

analytical solutions given by Jaeger [1965] and Mancktelow and Grasemann [1997]. We apply the

thermal model to address three classes of problems: 1) the range of erosion rates and relief over

which the assumption of flat isotherms is valid; 2) the depth of the muscovite closure isotherm for

various erosion rates and topographic profiles, and; 3) the temporal response time of the thermal

system to changes in erosion rate.

Stüwe et al. [1994] illustrate that the deflection of isotherms by topography is important in low-

temperature fission-track geochronology. The valleys may have older ages than predicted from

summit ages due to changes in the near-surface geothermal gradient caused by the cooling effect of

valleys and the insulating effect of peaks. In this paper, we argue that the topographically induced

deflection of the 350°C isotherm is negligible for most geologically reasonable erosion rates, given

the errors in the chronological analysis. As an example, we can examine the thermal structure for a

scenario with 4 km of relief undergoing erosion rates of 1.0 and 3.0 km/My (Fig. 1a & 1b). It can be

seen that the depth to the muscovite closure isotherm (considered here to be ~350°C) is very

dependent upon the erosion rate, but the isotherm itself remains effectively horizontal. If the relief is

increased to 6 km and a 3 km/My erosion rate maintained, the isotherms experience only ~400 m of

relief over wavelengths of ~20 km (Fig. 1d). The isotherm deflection from horizontal causes a

maximum discrepancy in the topographic age range of < 0.15 My (calculated using the difference

Page 30: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

15

in depth to the closure isotherm in combination with the specified erosion rate). Because this falls

within the typical analytical error of most single-grain 40Ar/39Ar analyses, we ignore this source of

uncertainty.

To investigate the effects of relief and erosion rate on the deflection of the 350°C isotherm, a larger

sensitivity analysis was conducted with the relief ranging from 0 to 6 km and the erosion rates 0.1 to

3.0 km/My (Fig. 2). Increasing the erosion rate causes an exponential decrease in the depth of the

350°C isotherm. Increasing the relief for a given erosion rate decreases the depth of the isotherm

below the valley floor because cooling is enhanced under larger valleys. These rates and

topographic relief span most geologically reasonable circumstances, but due to the exponential

relationship between isotherm depth and uplift rate, the isotherm deflection will increase markedly

with rates over 3.0 km/My in topographic relief of 6 km or more.

An understanding of the temporal response of the thermal system is important to evaluate the

applicability of the model to real situations. Starting from the initial geothermal gradient and 4 km

of relief, the thermal model can be used to examine the time and depth response of the 350°C

isotherm to various erosion rate scenarios (Fig. 3). Compared to the depth of the isotherm at 20 My,

from the initiation of erosion at 0 My approximately 60 to 80% of the depth response has occurred

after 5 My, and approximately 90 to 95% of the depth response has occurred after 10 My (Fig. 3a).

The system generally takes longer to equilibrate for slower erosion rates, whereas with high erosion

rates of > 2 km/My, 95% of the response has occurred after 10 My.

Additional insight comes from examining the response of the 350°C isotherm to instantaneous

change in the erosion rate at time t = 20 My (Fig. 3b). Increasing the erosion rate from 1.0 to 3.0

km/My has a similar response time as before, with ~90% of the total change occurring within 5 to 6

My. The response to a decrease in erosion rate from 3.0 to 1.0 km/My takes longer, with ~90% of

the total change occurring within 10 My. Therefore, the thermal model results are generally

applicable to systems with 0 to 6 km of relief, which have undergone vertical erosion at uniform

rates from 0.1 to 3.0 km/My, for > 10 My

For easy integration into our detrital cooling-age model, we have adopted a simplified approach to

predict the depth of the isotherm for a specified erosion rate (dz/dt) and relief (zs-zv), and limit our

Page 31: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

16

analysis to applications where the assumption of a horizontal closure isotherm is valid. To calculate

the depth to the closure isotherm (zc), we use an empirical fit to the model results (Fig. 2). The

equation has two exponential functions that represent: a) the exponential increase of the effects of

topography with increasing erosion rate, and b) the exponentially decreasing depth of the closure

isotherm with increasing erosion rate.

We recognize that constraining the thermal structure is of prime importance for exact interpretations

of geochronological data. Our model uses a simplified topography and is only applicable for those

conditions with approximately horizontal closure isotherms and with erosion rates and relief within

the bounds specified. In reality, each mountain belt and potentially individual particles within a

given orogen will experience a different thermal history. Overestimations of the depth of the closure

isotherm will lead to overestimations in the erosion rate, and vice versa. Given this complexity, our

thermal model is a reasonable solution to generalized applications, and it is important to note that

investigations of relative erosion rates within the same orogen may not be hampered by

uncertainties in the closure-isotherm depth to the same degree as comparisons between regions.

3.4 Steady-state landscapes

A key component of the thermal model is the assumption of a steady-state landscape: average

topographic characteristics remain unchanged over timescales of ~107 years. Although this

assumption is found in many other thermal models [Henry et al., 1997; Mancktelow and

Grasemann, 1997; Stüwe et al., 1994] and is embedded in our thermal model, the concept of a

steady-state mountain belt is poorly defined. The basic assumption of geomorphic steady state is

that regional topographic parameters remain invariant at appropriate timescales. This does not

require erosional output flux from each point in the landscape to exactly balance the tectonic input

of rock flux through each point, but merely requires that regional relief, hypsometry, and drainage

density remains steady at time scales equal to, or greater than, major climatic cycles (> 100 Kyr).

Considered within the time frame of a human life, the assumption of steady state may seem absurd.

In the field we see evidence for striking spatial and temporal variations in erosion: landslides,

Page 32: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

17

glacial valleys, and sediment preserved within the mountain belt, all attest to differential denudation

rates. However, at geological timescales, the concept of steady state is necessary, and in the

Himalaya, we can use a simple back-of-the-envelope argument to suggest that a long-term

topographic steady state is likely to exist. If the present convergence rate between southern Tibet

and India (~20 km/My) is representative of the last 10 My, and we assume that half of the total

convergence is due to the Indian plate underthrusting Tibet, then the remaining 100 km of

shortening must be accommodated by the Himalaya. Given that the mean elevation of the Himalaya

is only approximately 2-5 km, a rough balance per unit area is required between rock uplift and

erosion to remove the excess ~9.5 km/My of rock flux. The observations that Himalayan hillslopes

are maintained near the threshold for failure by bedrock landslides [Burbank et al., 1996], and that

major Himalayan drainages appear to predate the main topographic axis [Wager, 1937], also

support the argument that rock uplift is not outpacing the rate of erosion.

Because the response time of the thermal system is longer than short-term fluctuations of elevation

about a mean topography, we can reasonably assume that the closure isotherm does not move

appreciably with respect to the surface during the closure interval of the minerals. The temporal

variations about the average steady-state topography (105 years) are small in comparison to the

cooling ages of most samples. Therefore, it is likely that the minerals will pass though the closure

interval at approximately the same rate and that the resulting cooling ages will provide a good proxy

for the long-term erosion rate, if the other assumptions are valid.

3.5 Spatial resolution

Commonly, geochronologists have interpreted bedrock cooling ages in terms of tectonic zones with

related cooling and erosion histories. The natural division of a landscape into river basins, however,

provides a more readily defined framework within which to consider the detrital cooling-age signal.

For the purposes of this model, we assume that drainage basins are small enough to drain a single

tectonic zone. Therefore, each point in the basin will undergo uniform erosion and share a

genetically common thermal history for the specified relief and erosion rate of the basin. The

Page 33: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

18

geological characteristics of the individual drainage basin will control the distribution of cooling

ages found in sediment at the basin mouth

Whereas assumptions of basin uniformity must be correct at some spatial scale, clearly larger

drainages will be more complex. A large basin can be broken down into many sub-basins to

represent spatial variations. If a sub-basin approach is taken to investigate spatial variations in

erosion rate, then each must be considered separately in terms of erosion rate, relief, and thermal

history. This simplification requires that lateral heat flow due to different erosion rates is negligible.

The validity of choosing a particular size of tributary for modeling will depend upon the geological

constraints available: with a larger number of sub-basins and many unconstrained variables, the

model results must be viewed with increasing skepticism.

3.6 Vertical age distribution for a theoretical basin

If we accept that a) the basin is undergoing vertical erosion in a steady-state landscape and thermal

structure, and b) the depth of the closure isotherm for a given basin can be modeled as a function of

a uniform vertical erosion rate and basin relief, then mineral cooling ages will increase linearly with

elevation. The vertical distance from a horizontal 350°C isotherm to valley bottoms, for example, is

less than the distance to mountain summits, and this is reflected in younger bedrock cooling ages at

the base of a mountain (Fig. 4). Thus the cooling age (tc) of a point in the landscape can be

calculated from:

)/( dtdzzz

t cxc

−=

(2)

Where zx is the elevation of the sample location, zc is the elevation of the closure isotherm, and

dz/dt is the erosion rate. We can examine the consequences of varying the erosion rate on the

predicted age range for basins of 2, 4, and 6 km of relief (Fig. 5). As the erosion rate increases, the

summits and valley floors exhibit younger cooling ages, but in addition, the range of cooling ages

between valley and summits decreases as the closure isotherm becomes shallower.

Page 34: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

19

3.7 Distribution in ages at the basin mouth.

The distribution of detrital ages in a basin can be presented on a probability density plot, which

displays a probability density function (PDF) representing the probability of finding a grain of a

specified age within the overall age distribution of the basin. Equation 2 can be used to predict the

range of ages found in a basin undergoing a specified erosion rate, but not the distribution of

probability within the age range: the number of grains of each age represented at the basin mouth.

If the system of erosion and transportation is totally efficient, the theoretical detrital cooling-age

signal found in sediment at the basin mouth can be considered to be an exact integration of the

bedrock cooling ages within the basin. For a given relief and erosion rate, the probability of finding

a grain of a certain age (a function of the elevation) will be dependent upon the fraction of land at

the corresponding altitude (see Fig. 4). The probability of dating the maximum-age grain in a

detrital sample, for example, will be low because only a small percentage of the drainage area of an

average basin is at the top of a mountain. Therefore, the probability of land occurring at a specific

elevation (Pz), the hypsometry, can be used as a proxy for the probability of dating a grain of a

particular age (Ptc). Thus the elevation PDF can be combined with the age range between the valley

cooling age (tcv) and the summit cooling age (tcs) to produce the theoretical PDF:

( )xPtP ztc =)(

[3]

where, for tcv < tc(t) < tcs,

[ ]vscvcs

cvccvcsv zz

ttttttt

zx −

−−−+= .

)())(()(

[4]

otherwise Ptc(t) = 0. The resulting PDF, with the area normalized to unity, is our “theoretical” PDF

and can be thought of as the hypothetical distribution of ages generated from the zero-error analysis

of grains obtained from all elevations within the basin, in proportion to the frequency of that

elevation.

Page 35: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

20

It is illuminating to assess the three assumptions, in comparison to the analytical errors, that are

imbedded in this approach to modeling the distribution of ages: 1) uniform erosion rates across the

basin; 2) homogeneous distribution of the mineral being dated; and 3) insignificant transport and

storage times within the system. The first assumption requires that every point in the modeled basin

be treated as a source that contributes equally to the basin cooling signal, and the erosion rate is

independent of elevation or position within the basin. Although perhaps not intuitive, this is a

necessary consequence of steady state. If average climatic and tectonic factors remained constant,

the relief in a steady-state landscape would neither change substantially through time due to

preferential peak erosion, nor increase substantially through time due to preferential river incision.

In this context, “steady state” means the topography at present has the same statistically averaged

topographic characteristics as the topography at the time that the sediments in the rivers today

eroded. Although it is unlikely that the topography remains exactly the same, perturbations due to

high-frequency small landslides and short-term shielding of areas from erosion should be

insignificant compared to the overall shape of the hypsometry.

Another prerequisite of this approach is the assumption of a uniform distribution of geochronometer

within the basin. The detrital cooling-age PDF will be weighted towards those areas that are

supplying more geochronometer per unit area. The appropriateness of this assumption has to be

assessed in each setting to which this model is applied. If, from bedrock mapping and petrography,

the spatial distribution of the target mineral for dating is known, then appropriate weightings can be

assigned by combining this spatial distribution with the digital topography to yield a weighted “

target mineral hypsometry”, and thus a corrected basin PDF.

The assumption of no significant storage in the basin must be evaluated because, if a significant

fraction of the sediment is stored while in transit to the sampling site, the basin PDF will be skewed

toward older ages. We use two different arguments to contend that, in most active orogens

experiencing rates of erosion > 0.5 mm/yr, storage cannot be significant (over timescales

comparable to the analytical errors of the geochronological technique). First, at such rates, soil-

mantled hillslopes are uncommon, such that most sediment storage would have to occur along the

valley bottoms. Such sediment storage would mantle the valley floor and prevent river incision from

occurring, thereby preventing sustained high erosion rates. Second, from a volumetric perspective,

Page 36: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

21

given that valley bottoms only occupy a small percent of the total area, they would have to store a

very large thickness of sediments to have any measurable effect: clearly an uncommonly observed

condition in active mountain belts. We can consider the consequences of storing the sediment in the

basin for 0.5 My, which is a timescale of significance in comparison to the analytical errors of the 40Ar/39Ar dating (note that sediment storage since the last glacial maximum, ~20,000 years ago, is

extremely short compared to the analytical errors). A landscape undergoing erosion rates of 1

mm/yr would experience 0.5 km of erosion over the same interval. It would be possible to store

only a small fraction (probably << 1%) of this volume in the valley floors of rapidly eroding

mountains.

3.8 Examining the control of relief and erosion controls on the theoretical PDF of a single drainage basin

We can use the model to generate a theoretical distribution of detrital ages by defining basin

characteristics (relief, hypsometry, and erosion rate). Before examining actual drainage basins, it is

useful to consider the sensitivity of parameters controlling the cooling-age PDF of a single

catchment. For this purpose, we use a hypothetical drainage basin with a Gaussian distribution of

land area with elevation; in this case most of the land area is contained in the middle elevations of

the drainage network, as might be expected with a steep fluvial basin on a mountain flank.

Various uplift and erosion rate scenarios may be examined (Fig. 6). Increasing the relief of a basin

widens the PDF as the vertical separation between the summit and closure isotherm increases,

whereas the vertical separation between the valley floor and closure isotherm decreases. Increasing

uplift rate generates younger ages and narrows the width of the PDF for a given basin relief (Figs. 5

& 6). The summit-to-valley contrast in ages will be subdued, however, when relief is high and

erosion rates exceed 3 km/My because closure isotherms become increasingly deflected by surface

topography.

Once the thermal structure has been constrained, the distribution in ages in a basin is primarily a

function of relief, erosion rate, and hypsometry. Hence, on the basis of an observed distribution of

detrital ages, we have the potential to invert the relief and erosion rate from the geological record

Page 37: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

22

(as suggested for palaeorelief by Stock and Montgomery [1996]). Using the record of detrital

cooling ages from foreland basin sediments, we could in theory test the hypothesis of Molnar and

England [1990] that increased Quaternary incision drove relief production. We would expect the

detrital PDFs from Himalayan basins to move diagonally up and to the right in figure 6, displaying

a younger and narrower range in ages, as relief and erosion rates increased in the Quaternary.

The hypsometry of a basin results from a complex interplay of lithology, erosion rate, tectonics,

climate, surface process, and time [e.g. Keller and Pinter, 1996; Ohmori, 1993]. To explore the

effects of hypsometry, we examine end-member basin morphologies: hypsometries where most of

the basin is situated in the top, base, or middle of the drainage basin. If most of the land area is

concentrated in the lower elevations, the cooling-age signal from the basin will be biased towards

young ages (Fig. 7, case a). If a basin is dissecting the edge of a plateau, land area may be

concentrated in the headwaters, and the age population will be biased towards older ages (Fig. 7,

case e). Basins with land concentrated in the middle reaches (Fig. 7, case d) display normally

distributed cooling-age PDFs like those modeled in figure 6. These might be representative of the

steep fluvial basins experiencing strong tectonic forcing such as those found on the topographic

front of the Himalaya. Alternatively, glacial erosion also appears to concentrate alpine topography

towards the middle elevations of basins with very high relief [Brozovic et al., 1997].

4.0 Application to two Himalayan basins.

4.1 Geological Background and sample sites

In order to illustrate the modeling method outlined above, we consider 40Ar/39Ar detrital muscovite

data for two sediment samples collected in central Nepal from the Marsyandi River and one of its

tributaries, the Dordi Khola (Fig. 8). Along its course, the Marsyandi transects many of the tectono-

stratigraphic zones of the Himalayan-Tibetan orogen [see Hodges, 2000 for review]. Its headwaters

lie within weakly metamorphosed to unmetamorphosed Indian passive margin rocks of the Tibetan

Page 38: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

23

zone. As the river cuts south through the Annapurna and Manaslu massifs, it flows across the

Machhapuchhare Detachment Fault, the Chame Detachment fault (the basal structure of the South

Tibetan fault system in this area [Coleman, 1996]) and over high-grade metasedimentary and meta-

igneous rocks of the Greater Himalayan zone that have been intruded by Oligo-Miocene

leucogranites. Finally, as the river passes through the Himalayan foothills, its bedrock includes

deformed metamorphic rocks of the Main Central Thrust zone and footwall metasedimentary rocks

of the Lesser Himalayan zone.

Sample 1 was collected from the riverbed of the Marsyandi roughly 40 km downstream from the

trace of the Machhapuchhare detachment (Fig. 8). Potential source regions for the muscovite in this

sample include the structurally highest Greater Himalayan gneisses and leucogranites (which yield 40Ar/39Ar muscovite plateau dates of ~17 to 18 Ma in this area [Coleman and Hodges, 1995;

Copeland et al., 1990]) and rare hydrothermal veins in Tibetan zone sedimentary rocks (with 40Ar/39Ar muscovite plateau dates of ~14 Ma [Coleman and Hodges, 1995]). Sample 2 was

collected in the Dordi Khola, roughly 400 m upstream of its confluence with the Marsyandi. Its

drainage basin includes the structurally middle and lower parts of the Greater Himalayan zone, the

Main Central thrust zone, and the uppermost Lesser Himalayan zone rocks. Although bedrock 40Ar/39Ar muscovite data are not available for the middle and lower Greater Himalayan rocks in the

Marsyandi drainage area, samples from equivalent structural levels in the Kali Gandaki drainage

(roughly 80 km to the west) yield ~15 Ma plateau dates [Vannay and Hodges, 1996]. Muscovites in

Sample 2 with a provenance in the Main Central thrust zone or the uppermost Lesser Himalayan

zone should yield considerably younger 40Ar/39Ar dates. Edwards [1995] reported dates of 6.2 ± 0.2

Ma and 2.6 ± 0.1 Ma for muscovites collected from the Main Central Thrust zone in the Marsyandi

valley. Such young cooling ages support other thermochronological evidence for widespread Late

Miocene-Pliocene metamorphism of the Main Central thrust zone and its footwall in the central

Nepalese Himalaya [Copeland et al., 1991; Harrison et al., 1997; MacFarlane et al., 1992;

Macfarlane, 1993].

The catchment areas, topographic parameters, and hypsometric curves for the sampled basins (Fig.

8) were extracted from a 90-m digital elevation model (DEM) of the region using ARCINFO

software. The upper Marsyandi basin (Fig. 8, (i)) contains 6500 m of relief and is 2270 km2 in size,

Page 39: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

24

of which 1230 km2 comprises Tethyan sediments according to the map of Colchen et al. [1986]. For

this investigation, we consider only the remaining 1140 km2 of basin that drains the top of the

Greater Himalaya sequence because the rare hydrothermal muscovites in the Tibetan Zone as well

as the Tethyan carbonate and mudstone lithologies, make a negligible contribution to the detrital

muscovite signal. This assumption is supported by examination of the fluvial detritus from these

areas, which is dominated by rock fragments of fine-grained sediments. Ignoring the Tibetan zone

in the drainage area does not affect the range in ages of the basin as the maximum basin relief is

contained within the Greater Himalayan zone. The hypsometry changes, however, and mean relief

decreases from 4800 m to 4400 m. In addition, limiting the investigation to the area south of the

Machhapuchhare Detachment Fault means that assumptions of uniform erosion rate are more likely

to be valid. The Dordi Khola drains the southern front of the Himalayan topographic axis and is

351 km2 in extent, with 7200 m of relief, and an average elevation of 2900 m (Fig. 8, (ii)).

4.2 40Ar/39Ar Analytical Protocols

The two samples were washed and sieved to a range of grain sizes between 500 and 2000 µm prior

to the commencement of muscovite separation. Individual muscovite grains were isolated by

applying standard gravimetric and magnetic separation techniques, followed by hand-picking under

a binocular microscope. The mineral separates were washed sequentially in distilled water, acetone,

and ethanol before irradiation at the McMaster University research reactor. The irradiation package

included aliquots of the neutron-fluence monitor Fish Canyon sanidine (28.02 Ma, Renne et al.

[1998]), as well as a variety of salts that served as monitors for interfering nuclear reactions.

After irradiation, the muscovites and monitors were analyzed at the 40Ar/39Ar laser microprobe

facility at the Massachusetts Institute of Technology [Hodges, 1998]. Gas was extracted from

individual mica crystals by fusion in the defocused beam of an Argon laser operating at 18 W for a

period of approximately 10 seconds. After purification to remove reactive species, the extracted gas

was analyzed on an MAP 215-50 mass spectrometer using a Johnston electron multiplier. Total

system blanks were measured at the beginning of each analytical session and after every tenth

analysis of an unknown.

Page 40: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

25

Apparent ages (dates) calculated for each muscovite are reported in Appendix2, Table 1, with an

estimated 2-σ uncertainty obtained by propagating all analytical uncertainties. (In order to illustrate

the proportion of this uncertainty that is attributable to uncertainties in the neutron flux during

sample irradiation Appendix2, Table 1, shows uncertainties in apparent ages calculated with and

without the contribution from the irradiation parameter J). Further details on analytical techniques

may be found in Hodges and Bowring [1995].

Given a date (tc), and an analytical uncertainty for that date (σ), a probability density function can

be calculated for each grain assuming that a Gaussian kernel represents the distribution of error [e.g.

Bevington and Robinson, 1992; Deino and Potts, 1992]]. For a sample of N grains collected from a

specific locality, the PDF of the age of each grain (n) can be combined:

∑=

=

−−

=Nn

n

nntt c

en

tP1

))(.2))((

( 2

2

..2).(

1)( σ

πσ

(5)

Once the area of the resulting curve is normalized to unity, a summed probability density function

(SPDF) is generated that represents the distribution of age probability within the sample.

4.3 Detrital cooling-age results and modeling theoretical PDFs

The 40Ar/39Ar results display distinctly different detrital signals originating from each catchment

area (Fig. 9). The upper Marsyandi basin SPDF contains 35 grains that range in age from 11.2 ± 1.4

My to 18.7 ± 1.3 My and is characterized by a sharp peak at ~ 17 My and a “tail” of younger grains

from 10 to 14 My. The peak is comparable to bedrock muscovite 40Ar/39Ar ages from the upper

Greater Himalaya sequence [Coleman and Hodges, 1995; Copeland et al., 1990]. The Dordi Khola

contains 39 grains that range from 2.6 ± 1.2 My to 12.7 ± 0.5 My. The SPDF is characterized by

multiple peaks between 3 and 8 My, and a single peak at ~13 My.

Given the measured hypsometry for the sampled basins, our model can be used to predict the forms

of PDFs that would be expected to be observed for a given vertical erosion rate under steady-state

Page 41: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

26

conditions. By matching the predicted curve to the actual SPDF we determine the solution that

results in the lowest mismatch (as defined below) by varying the erosion rate. With this

methodology, we estimate the approximate erosion rate for the upper Marsyandi basin (Sample 1)

as 0.95 km/My and for the Dordi Khola basin (Sample 2) as 2.15 km/My (Fig. 9). Although the

distribution of ages within the two samples is about what we would expect if the erosion rates were

a factor of two higher in the Dordi Khola basin than the upper Marsyandi basin (cf., Fig. 5), our

ability to reproduce the simple SPDF of Sample 1 is much greater than our ability to match the

more complex function of Sample 2. There are several possible reasons for this. One is that the

riverbed sediments at the Sample 2 locality do not represent the distribution of bedrock cooling ages

in the Dordi Khola basin. This possibility could be examined through additional detrital mineral age

determinations for samples from the Dordi Khola. A second reason is that one or more of the initial

assumptions behind our model is incorrect. For example, the assumptions of uniform uplift rates

and steady-state behavior may be erroneous if the Late Miocene-Pliocene cooling ages in Sample 2

and the lower Marsyandi bedrock reflect episodic reactivation of the Main Central thrust in the

Marsyandi drainage [Edwards, 1995]. Under such circumstances, the pattern of bedrock cooling

ages through the Main Central thrust zone might reflect local complexities in thermal structure and

not the simple, depth-dependent distribution of isotherms required by our modeling scheme. Better

resolution of the bedrock cooling-age distribution would help us test this hypothesis.

For now, we focus on a third reason why the Sample 2 SPDF might be so complex and, moreover,

why even the simpler SPDF for Sample 1 should be viewed with caution: in each case, the

relatively small number of grains we analyzed at each locality might be insufficient to adequately

characterize the true population of detrital muscovite ages. Given the fact that any reasonable

detrital mineral geochronological study involves a random sampling of only a tiny fraction of the

total muscovite grains at a particular site, how confident can we be that such a sample is

representative? Our approach to this problem is to explore the fidelity with which random picks of

grains (which we refer to as a "grab sample") from a synthetic population reproduce the population

PDF as a function of the number of grains in the sample.

Page 42: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

27

5.0 The construction of a grab-sample PDF.

To simulate the random dating of grains from a grab sample of sand containing millions of particles,

Monte Carlo integration [Press et al., 1992] is performed on the theoretical PDF (see Fig. 10). In

this paper we use the random-number generator of MATLAB 5.3. The initial conditions for the

random-number generator are determined by the computer clock at the start of each run. The

integration, with the number of points limited to represent the number of grains dated, represents the

random selection of grains in the grab sample from the river, and the random choice of which

particular grains are used for dating. In practice, micas from rapidly cooling areas require a size

fraction of approximately 500 to 2000 µm (lower-coarse to upper-very-coarse sand) in order to

select grains with enough radiogenic 40Ar for reliable analysis. In this situation, Monte Carlo

integration represents the random picking of muscovite grains within this specified size and mineral

fraction. To construct a synthetic SPDF for each Monte Carlo age pick, a standard deviation

representing the expected analytical error in the apparent age of the pick has to be specified. For the

purposes of this model, we use an estimated standard deviation of 0.64 My (Fig. 10), taken from the

mean uncertainty of our data in Appendix2, Table 1.

5.1 PDF comparison and statistics.

Given a grab-sample SPDF, some quantitative measure is required to compare it to the theoretical

PDF: we want to know how well the grab-sample SPDF reproduces the theoretical PDF. In

particular, if a finite number of grains (20-100) are dated, how well can we expect to reproduce the

characteristics of the theoretical PDF? To quantify the match (or mismatch), the sum of the

difference in the distribution of probability (Pdiff) between the theoretical probability (Ptheoretical) and

grab-sample probability (Pgrab) is computed over each age increment (t), and expressed in terms of a

percentage of total probability:

100*2

)()(0

tPtPP

grabltheoretica

t

tdiff

−=

∑∞=

=

Page 43: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

28

(6)

This provides the percentage mismatch of the entire probability signal in the units of percentage

probability (Fig. 11). For illustration, if the two PDF curves are identical then the error is zero

percent. Alternatively, if we consider an error of 100 % then the two curves would have completely

different age ranges, with no overlap in probability-age space. To investigate the range of SPDFs

that a single theoretical PDF can generate, successive iterations produce many individual synthetic

“grab-sample” curves. Each iteration contains the same number of grains and the misfit between

each iteration and the theoretical curve is recorded. The 95% confidence interval of these errors is

taken as the measure of how well the specified number of grains represents the theoretical PDF for

95% of the time.

5.2 Resolution of the detrital dating

Before examining the observed data, the resolution of the detrital dating methodology can be

investigated. How small a change in the age distribution from a catchment can we detect using a

given number of dates, and how can we be sure that detrital signals are statistically similar or

different? The model can be used to help address the problem of whether SPDF ‘A’ can be

explained by the statistical variability of ‘B’ due to random grain selection. To do this we generate

an theoretical curve that best matches B. A number of iterations are performed (with the number of

grains in each iteration equal to that contained in A) to determine the most likely range of outcomes.

If the misfit between A and B is less than the 2-σ errors on the modeled range of outcomes, we can

be 95% certain that A and B are statistically indistinguishable.

As an example, we use a basin with 4 km of relief eroding at 1.0 km/My to investigate how well

different numbers of grains match the theoretical sample. For each specified number of grains, 1000

iterations produce 1000 grab-sample SPDFs, and the 95% confidence limits are taken from the

range in these (as described in 5.2.). In this example, (Fig. 12, scenario a) increasing the number of

grains decreases the average mismatch of the grab-sample SPDFs from the real PDF. In addition,

increasing the number of grains leads to more certainty in the result as the width of the 95%

Page 44: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

29

confidence window decreases. The reduction in mismatch drops rapidly at first with increasing

numbers of grains, indicating that there is much to be gained from performing additional analyses.

However the rate of improvement declines rapidly after ~50 grains; increasing the number of grains

dated does a better job of constraining the theoretical cooling-age signal, but with diminishing

returns. With expensive dating techniques, there is clearly a trade-off between the cost of dating and

the extra confidence that large numbers of grains provide.

Unfortunately the same result is not applicable to all tributaries; the mismatch of SPDFs will be

related to the shape of the theoretical PDF. Consider, for example, the results of an age PDF

combined from a basin of 4.0-km relief eroding at 1.0 km/My and a basin of 4.5-km relief eroding

at 1.3 km/My (Fig. 12, scenario b). The more complex two-basin PDF (see figure insert) is more

difficult to match than the simple one-basin PDF. For any given number of dated grains, the mean

mismatch of SPDFs is ~3-8% larger for the more complex PDF than the individual basins shown.

The method of representing individual grain errors with Gaussian distributions means that the

precision of the geochronometer, in relation to the shape of the real PDF, will also affect the

mismatch. Low-relief basins with high erosion rates will have a narrow age PDF (Fig. 6) that high-

precision geochronometers will fit well. These geochronometers will have more trouble fitting the

broader peaks characteristic of basins with low erosion rates and high relief. It is therefore easier to

fit the “peaks” of a narrow PDF than the “tails” of broad peaks.

In these scenarios, the statistical analysis illustrates that ~60+ grains (scenario a) and ~90+ grains

(scenario b) are needed to achieve a mismatch of ≤ 15% at the 95% confidence level. Although a

15% mismatch may seem large, in practice it produces a reasonable fit visually (Fig. 13). This is

because error is only calculated where ages are present (if the error is < 100% then two curves have

some ages in common), and because most of the error generally represents misfits with the tails of

the theoretical PDF.

Page 45: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

30

5.3 Himalayan catchments – synthesis of theoretical PDFs and random sampling.

We can also use this approach to explore the differences between the theoretical PDF and observed

data. The Dordi theoretical PDF is generated using an erosion rate of 2.15 km/My that produces a

mismatch error (calculated with equation 6) of 41% of total probability with the observed data . The

upper Marsyandi theoretical PDF has an erosion rate of 0.95 km/My and an error of 10% compared

to the observed data. From the theoretical PDFs of the two basins, 1000 grab samples were

generated that each contained N age selections (where N was the number of grains actually dated in

each basin). The dashed lines in figure 9 indicate the grab-sample SPDF that was the best fit to the

data SPDF from the first 30 iterations of Monte Carlo picks. They are shown to illustrate how the

grab-sample SPDF differs from the theoretical PDF. The best-fit grab-sample SPDF from the Dordi

has an error of 22% when compared to the observed data SPDF, while the best-fit grab-sample

SPDF from the upper Marsyandi has an error of 6%.

The model allows us to examine how well the theoretical PDFs represent the data SPDFs when we

start sampling the sediment. We can find the 95% confidence interval that N grains produce, and

test whether the theoretical curve falls within these limits. A confidence interval of 4% to 19% was

generated from the upper Marsyandi with 1000 iterations of picking 35 grains. The theoretical PDF

of the upper Marsyandi fits the data SPDF well, falling well within confidence limits and indicating

that the two curves are statistically indistinguishable. However, the fit is sensitive to the choice of

the erosion rate because small shifts in the peak probability of the theoretical PDF, away from the

peak probability of the data SPDF, produce larger errors.

The theoretical PDF from the Dordi fits less well, falling outside of the 18% to 26% confidence

interval, and the error is less sensitive to changes in the erosion rate. The very spiked SPDF from

the dated grains is due to ages with low errors (e.g. 5.7 ± 0.2 My), whereas the wide tails are caused

by ages with large associated errors (e.g. 2.8 ± 4.5 My). This makes it difficult for the mean 1-σ

error (0.64 My), used in the production of the theoretical PDF, to fit the curve very closely. The

effects of grain error can be seen as the grab-sample SPDFs fit the observed SPDFs more closely:

the errors associated with individual grains increase the range of ages found in the grab-sample

curve in comparison to the theoretical curve.

Page 46: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

31

5.4 Discussion of modeling results

Although the model does a good job of matching the peak probabilities, both basins show minor age

populations that are problematic to replicate. The upper Marsyandi contains a young age population

(10 to 14 My), and the Dordi shows a minor older age population (9 to 13 My) that the model does

a poor job of representing. These are most likely due to invalid assumptions within the model. The

most likely cause is that erosion rates are not uniform across the basins and hence there are cooling-

age variations.

Differential erosion rates in the Dordi Basin would be geologically reasonable. There is a growing

body of evidence from geomorphology [Lave and Avouac, in review; Seeber and Gornitz, 1983],

structural (ref), and geochronology [Catlos et al., 1999; Catlos et al., 1997; Harrison et al., 1997;

MacFarlane et al., 1992], that the MCT has been active during the past 5-6 million years, or that

subsurface deformation associated with a ramp structure on the Himalayan sole thrust has led to

differential uplift across the Himalayan topographic front. To examine a basic scenario of

differential erosion, the Dordi basin can be divided into two regions, in which the MCT separates

the Greater Himalaya and Lesser Himalaya (see Fig. 8). For example, with erosion rates of ~2.2

km/My in the physiographically high Greater Himalaya and ~1.8 km/My in the Lesser Himalaya

foothills, together with an assumption that the latter is contributing half the amount of muscovite

per unit area, the mismatch may be reduced from 41% to 34%. This still falls outside the confidence

interval and the very young ages are still difficult to represent, perhaps due to hydrothermal

alteration in the MCT zone [i.e. Copeland et al., 1991].

Other potential sources of mismatch include sediment storage in the system, and short-term

imbalances in sediment supplied to the fluvial system. As discussed earlier, the first can be

effectively discounted due to the high erosion rates. In particular, with erosion rates of 1.7 km/My

in the Dordi, long-term sediment storage is unlikely. Short-term imbalances in sediment supply,

however, are more difficult to discount. Large landslides and debris flows have been reported from

the Marsyandi valley [Yamanaka and Iwata, 1982] and may cause an influx of grains of a certain

age (i.e. causing the characteristic 6-My peak in the data from the Dordi that is not seen in the

Page 47: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

32

theoretical curve). However, these landslide fills are mainly concentrated in the lower reaches of the

Marsyandi.

The thermal modeling is an additional source for error as we employed a highly simplified model.

The model is restricted to the vertical erosion of a 35-km-thick crustal block. This matches the

depth of the Himalayan sole thrust that has been seismically imaged at approximately 30 to 40 km

depth beneath South Tibet [the MHT in Nelson et al., 1996]. However, in the Himalaya, erosion

occurs as the column is uplifted along the fault plane, and rocks therefore have a complex thermal

history. Cooling by the underthrusting slab, accretion of material to the hanging wall, shear heating,

and time, will all affect the thermal structure and resulting cooling history. Hence, it is clear that

improving thermal models and the subsequent calculations of the depth to the closure isotherm will

produce better insights into the tectonics.

6.0 Discussion

In this paper we have presented a numerical model that investigates the parameters that control the

detrital cooling-age signal from an individual basin. Numerous assumptions, many of which are

impossible to evaluate in the Himalaya at this time, underpin our treatment, but we believe that it is

an improvement on the simple assumption of a geothermal gradient, as used in many investigations,

because it accounts for the effects of erosion rate and topographic relief when predicting the depth

of the closure isotherm. Another improvement in the calculation of the position of the 350°C

isotherm is that erosion of the crustal column occurs from a depth limited to 35 km. This seems

more reasonable than eroding the entire lithospheric column, given that rock flux into mountain

belts commonly occurs along detachments at such depths. From examination of the temporal

effects, we have seen that for most scenarios the 350°C isotherm has achieved > 90% of its total

response within 10 My.

Despite these insights, the thermal and kinematic structure is typically the most difficult parameter

to constrain when trying to extract erosion rates from geochronology, especially from the geological

record. Although we realize that many the assumptions made are not strictly valid for many areas,

Page 48: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

33

more accurate constraints may be, at best, difficult to determine. Our assumption of uniform and

vertical erosion over large areas is probably the largest simplification for most orogenic systems.

The lateral advection of heat and rock mass into mountain belts [i.e. Beaumont et al., 1994; Willet,

1999] along underlying thrust faults are important processes and will determine the exact pattern of

bedrock cooling ages.

For a predetermined thermal structure, our simplified model predicts the variation of cooling age,

given a uniform post-closure history, as a function of elevation. In a situation with horizontal

isotherms and a uniform distribution of geochronometer, we have shown that the hypsometry may

be used as a proxy for determining the detrital cooling-age signal that results from uniformly

eroding that topography. Assumptions of uniform distribution of geochronometer can be readily

assessed for modern applications in the field, and differences accounted for using GIS applications.

Extrapolation back into the stratigraphic record, however, requires more caution.

If the constraints and assumptions of the model are accepted, the methodology may be used to

predict the detrital cooling-age signal of an individual tributary. If the tributary is now sampled in

the field, and N grains dated, we can use our statistical method to: 1) assess how well N grains can

define the real cooling-age signal; 2) test whether the real data is statistically identical to the model

results (are we able to disprove the assumptions?); 3) decide whether dating 10 more grains, for

example, would produce a more conclusive result, and; 4) discern whether tributary A is eroding

faster than tributary B, given differences in basin characteristics and sampling uncertainty. We use

Monte Carlo techniques to assess this intrinsic variability of grain sampling, and have introduced a

way to measure the difference in two cooling-age distributions.

Our statistical analysis (Fig. 12) shows that, while the exact values of the error are not known when

using forward modeling, the relative increase in accuracy of dating of dating 70 grains, rather than

30, may be considered important, but the increase from 70 to 150 grains may not justify the extra

time and expense. It seems that as a general approximation 50-70 grains provide a good match of

the “grab-sample” SPDF to the theoretical PDF for a simple basin, with mismatch approaching

15%.

Given a series of detrital SPDFs from the stratigraphic record, we can now test if they are

statistically differentiable from one another. For example, if a modelled theoretical curve typically

Page 49: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

34

generates 5-25% error at the 95% confidence level using 40 grains, then SPDFs generated from 40

grains in different stratigraphic horizons cannot be statistically differentiable until the mismatch

errors increase above 25%. As we have seen, (Fig. 12) the 95% confidence interval will be

dependent upon the shape and complexity of the SPDF, the analytical age error, and the number of

grains dated.

One caveat using this technique, however, is that the exact error remains ambiguous because the

true age signal is unknown. Unless the drainage area can be sampled uniformly, at regular sites

throughout the entire altitudinal and spatial range, the exact distribution of bedrock cooling ages

cannot be constrained. Furthermore, the detrital age signal can only be defined once the contribution

of each point within the basin has been corrected for local short-term variations in the erosion rate.

Clearly such an undertaking would be unrealistic for most applications of detrital mineral

thermochronology. Some method of approximating the theoretical signal is, therefore, needed to

assess the uncertainties associated with sampling. As a consequence, however, uncertainty

calculations using forward modeling are not strictly robust for real data.

The theoretical-PDF model is envisioned to be a tool for evaluating the relative influence of

parameters controlling the detrital cooling-age signal. It provides a good first-order approximation

for areas that have limited constraints and focuses the user on assumptions that may not be

applicable. With this approach in mind, we have presented 40Ar/39Ar analysis from two modern

Himalayan catchments. In one instance (the upper Marsyandi sample), the model provides a

reasonable fit to the observed data; in the other, it does not. In the latter case, further analysis

suggests that the mismatch is not the product of sampling bias alone, but instead indicates that one

or more of our assumptions is incorrect. Nevertheless, it is evident that, despite modeling

mismatches, we can confidently state that the average erosion rate varies significantly between the

two basins. Even if the erosion rate is not uniform across each basin, a spatial variation of 10-20%

is insignificant compared to the difference between the two basins considering the uncertainties on

the dates and random sampling. Furthermore, an estimation of the average erosion rate using an

integration of data from the entire basin will probably represent a more accurate value than a single

bedrock cooling age that is then extrapolated to a wider region.

Page 50: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

35

7.0 Conclusions

We have presented an integrated approach to investigate how drainage-basin parameters interact to

produce the distribution of cooling ages found at the basin mouth. Understanding these factors must

underpin a thorough analysis of detrital cooling ages in the modern and stratigraphic record. A

thermal model is used to predict how the interaction of topography and erosion rate control the

depth of the 350°C closure isotherm. Increasing the erosion rate causes an exponential decrease in

the isotherm depth, and increasing the relief compresses the isotherms beneath valley floors. With

respect to the depth of a steady-state isotherm in our model, > 90% of the total depth response to an

increase in erosion has occurred within 10 My.

Monte Carlo analysis provides a useful tool for examining the processes of random selection that

control the detrital age signal. We have provided a new method to constrain the errors of detrital

cooling-age signals and to compare PDFs. The mismatch associated with a grab sample is

dependent upon the precision of the geochronometer, the shape of the SPDF or theoretical PDF, and

the number of grains dated. Therefore, the statistics of each sample need to be assessed separately,

with the caveat that forward modeling is not an exact predictor of the uncertainty.

Application of the model to real data provides a basis for examining the parameters that control the

theoretical cooling-age signal and may emphasize parameters that need further research and

constraint. Our Himalayan data highlights the need for a greater understanding of the kinematic and

thermal structure of an orogen in order to accurately interpret geochronological information. In

addition, further investigation is needed into how individual hinterland drainages interact and

combine to produce the cooling-age signals found in the foreland basin. Only then can we use

detrital mineral geochronological data to investigate the complexities of modern orogenic

deformation, and to extrapolate the results back into the stratigraphic record with confidence.

Page 51: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0

4

8

12

16

20

24

3.0 km/Myr1.0 km/Myrde

pth

belo

w v

alle

y flo

or (

km)

distance (km)

350oC

350oC

a) b)

403020 100

c)

3.0 km/Myr

403020 100

350oC

d)

403020 100

350oC

403020 100

c)

0

4

8

12

16

20

24

dept

h be

low

val

ley

floor

(km

)

1.0 km/Myr

36

Figure 1. Thermal structure of continental crust with erosion rates of (a) 1.0 km/My (b) 3.0 km/My after 20 My in a landscape with 4 km of relief. In scenario (c) and (d) relief is increased to 6 km with an erosion rate of 1.0 km/My and 3.0 km/My, respectively. Relief production in all models is instantaneous, with a steady-state landscape that contains 30o slopes, which simulate threshold conditions for landsliding. Note that the 350°C closure isotherm for 40Ar/39Ar in muscovite is essentially flat for scenarios (a), (b), and (c), and has only ~200 m amplitude of deflection for (d).

Page 52: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Figure 2. Average depth (y-axis) of the 350oC isotherm (zc) below the valley floors as a function of varying relief (x-axis) and erosion rates (labeled on the lines representing equal erosion rate) after 20 My. Each point represents the results of one model run. In the model, relief production is instantaneous, with a steady-state landscape that has slopes of 30o to simulate threshold conditions for landsliding. Range bars on the points indicate the vertical deflection of the 350oC isotherm, about the mean, due to topographic influence. The equation is our empirical fit to the modeled depth of 350oC isotherm as a function of the relief and erosion rate. The relief (R) is the difference between summit and valley elevations[zs - zv]

0

2000

4000

6000

8000

10000

12000

14000

16000

18000

20000

0 1000 2000 3000 4000 5000 6000

Relief, R (m)

0.1 km/My

2.5 km/My

2.0 km/My

1.5 km/My

1.0 km/My

0.5 km/My

3.0 km/My

dept

h, z

(m

)

depth of 350oC closure isotherm (zc). Range bars indicate maximum and minimum depths due to topographic deflection.

zc(dz/dt, R) = (0.18exp(-0.67.dz/dt)-0.34)*R + (19600*exp(-0.28.dz/dt))

37

Page 53: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0

2000

4000

6000

8000

10000

12000

14000

16000

18000

20000

0 5 10 15 20 25 30 35 40

3.0 km/My

2.0 km/My

1.5 km/My

0.5 km/My

1.0 km/My

2.5 km/My

0.1 km/My

1.0 km/My

3.0 km/MyDECREASE

INCREASE

0 to 20 My 20 to 40 My

(b)(a)

(i)

(iii)

(ii)

Time (My)

Dep

th b

elow

val

ley

floor

s (m

)

Figure 3. The temporal response of the depth of the 350oC isotherm for a system with 4 km of topographic relief. (a) Starting from a thermal steady state in the absence of topography and erosion, the lines represent the depth of the 350oC isotherm, at each point in time, for scenarios undergoing uniform erosion rates of 0.1 to 3.0 km/My (labeled). (b) After 20 My, three scenarios show the response to: (i) a decrease in erosion from 3.0 to 1.0 km/My; (ii) an increase in erosion from 1.0 to 3.0 km/My, and (iii) an increase in erosion from 0.5 to 3.0 km/My.

38

Page 54: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Sum

mit

age

Val

ley

age

cooling age (tc)

0 My

Tc

erosion rate (dz/dt)

geothermal gradient (dT/dz)

(zs-zv)

zs

closure

temperature

zx

zv

elev

atio

n (

z)

zc

tcv tcs

z w

ith t

path

tcx(z) = (zx-zc) (dz/dt)

frac

tion

of a

rea

elevation

elev

atio

n

age age

prob

abili

ty

zv zs tcv tcs

zs

zv tcv tcs

HYPSOMETRY x AGE RANGE AGE DISTRIBUTION

(i) (ii) (iii)

Figure 4. Construction of a "theoretical" PDF for an individual basin. A cooling age (tc) is calculated from the depth (zc) of the closure temperature (Tc), which is a result of the thermal modeling, and the erosion rate (dz/dt). The difference between summit elevation (zs) and valley elevation (zv) results in a difference between summit cooling ages (tcs) and valley cooling ages (tcv). The cooling age (tcx) of a sample 'x' collected derived from elevation zx can be calculated using the equation shown. The inset illustrates how the age distribution is governed by the combination of the age range (tcv to tcs), and relationship of land area to elevation, which is shown as a normal distribution here.

39

Page 55: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

6km of relief

4km of relief

2km of relief

erosion rate (km/Myr)

age

(Myr

)

SUMMITS

}summit-to-valley

age ranges

0.5 1 1.5 2 2.5 30

5

10

15

20

25

30

35

40

45

VALLEY FLOORS

6km4km2kmAge ranges for basins of specified relief.

Figure 5. Relationship between summit ages (solid lines) and valley ages (dashed lines) for topographic relief of 2, 4, and 6 km undergoing erosion rates of 0.5 to 3.0 km/Myr. The vertical bars illustrate the age range with elevation for erosion rates of 1.0 and 2.0 km/Myr. The range in ages is a function of erosion rate and the relief. The range narrows as erosion rate increases, and 2 km of relief (black lines) contains a smaller range of ages than 4 km of relief (dark gray lines), which in turn is smaller than 6 km of relief (light gray lines).

40

Page 56: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0.5 1.0 1.5 2.0

2.0

4.0

6.0

increasing erosion rate (Km/My)

incr

easi

ng r

elie

f (K

m)

age

prob

abili

ty

0 50

0 4frac

tion

of la

nd

area

elevation (km)

cumulative fraction

of land area

0 4elevation (km)

Figure 6. Effects of uplift rate and relief on theoretical PDFs for a basin with a Gaussian distribution of land area with elevation (illustrated in the bottom two plots). In the calculations, the depth to the 350oC isotherm is taken from our thermal modeling. The scale on each inset theoretical-PDF plot is the same with the x-axis ranging from 0 to 50 My, and probability on the y-axis.

41

Page 57: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

prob

abili

ty

"time since closure" (tc)

0

1

10 cumulative fractionof elevation

cum

ulat

ive

frac

tion

of a

rea

a

a

b

b

c

c d

d

e

e

a

b

c

d

e

elevation (z)

Fra

ctio

n of

land

are

aat

spe

cifie

d el

evat

ion

0zv

00

tcstcv

zs

cooling age

Figure 7. Effects of hypsometry on theoretical PDFs. The upper-left panel shows the relationship between a specific elevation and the land area at that elevation for 5 basins. The cumulative hypsometric curves are shown in the lower-left panel. The right panel shows the resulting theoretical PDF for the basin. The range in ages (tcv to tcs) is dependant upon the specified erosion rate. Basin (a) contains most land at lower elevations, whereas (e) contains most land at higher elevations. Basin (b) has a uniform distribution of land with elevation, and (c) is biased towards concentration in the middle elevations. Basin (d) has a normal distribution of land with elevation that was used as an approximation in the sensitivity analyses (Fig. 6).

42

Page 58: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

MDF

Manang

TibetN

scaled 45%

% a

rea

0 2000 4000 6000 80000

0.2

0.4

0.6

0.8

1.0

frac

tion

of a

rea

0.5

1

1.5

2

frac

tion

of a

rea

% a

rea

U.MARSYANDI

DORDI

elevation (m)

0

0.2

0.4

0.6

0.8

1.0

0 2000 4000 6000 8000

MCT

Nepal

0.4

0.8

1.2

1.6

Tethyan strata

Greater Himalaya

Lesser Himalaya

0 20km

(i)

(ii)

Figure 8. Map of the upper Marsyandi drainage basin showing the detrital sample locations (black markers). The thick white lines represent the catchment areas upstream of the sample sites within the larger Marsyandi Basin. The Macchupuchare Detachment fault (MDF) is shown with fine dashes, and the Main Central Thrust (MCT) is shown with longer dashes. The inserts depict the hypsometry for (i) the upper Marsyandi (Sample 1) and (ii) the Dordi (Sample 2). The hypsometry is calculated using 50-m elevation bins, and then smoothed over 5 bins. Elevation data for the upper Marsyandi is taken from the area south of the CDF because the area to the north is composed of Tethyan sediments.

43

Page 59: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 5 10 15 20 25 30N=35N=39

u. Marsyandi data SPDF

Dordi data SPDF

theoretical PDF

best fit synthetic "grab sample" SPDF

theoretical PDF

best fit synthetic "grab sample" SPDF

0

0.005

0.01

0.015

0.02

0.025

0.03

age (Myr)

prob

abili

ty

0.035

Sample 1

Sample 2

Figure 9. Diagram showing SPDFs generated from the results of 40Ar/39Ar dating samples from the upper Marsyandi (Sample 1) and Dordi basin (Sample 2). The black lines are the best-fit theoretical PDFs generated by the model for each basin. The dashed gray lines indicate the best-fit grab-sample SPDFs (to the data) from 30 iterations. The number of grains dated (N) varies for each sample, and this determines the number of Monte Carlo picks used for generating each grab-sample SPDF.

44

Page 60: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 2 4 6 8 10 12 14 16 18 20

1-si

gma

erro

r (M

yr)

age (Myr)

mean = 0.64 Myr

0

0.5

1

1.5

2

2.5

Figure 10. Plot of grain age versus age uncertainty for the 40Ar/39Ar analysis of muscovite crystals from both samples. We use mean age error of 0.64 Myr (1-s) for the uncertainties in our model. The generally large age uncertainties are a result of the young grains in our study that contain relatively small amounts of 40Ar*.

45

Page 61: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

10 15 20

20% mismatch

theoretical PDFgrab-sample

SPDF

age (My)

prob

abili

ty

Figure 11. Error calculation for a basin of 4-km relief eroding at 1 km/My, in conjunction with a normal distribution of land area with elevation. The theoretical PDF is outlined in black and shaded, whereas the grab-sample SPDF, containing 50 grains, is outlined by the gray dashed line. The white stars are a cartoon illustration of the Monte Carlo sampling of grains from the theoretical PDF (note that they all fall inside the black curve). The x-axis values of the stars are used as the reported mean age, and with a specified error, are used to produce Gaussian distributed kernels for each grain. The grab-sample SPDF is the summation of the individual grain kernels normalized to unity. The area with diagonal hatching represents the total mismatch between the theoretical PDF and the grab-sample SPDF, which is 40% of the total probability (area) of the theoretical PDF, or an error of 20%.

46

Page 62: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 20 40 60 80 100 120 140 160 180 2000

10

20

30

40

50

number of grains dated (N)

erro

r (%

of p

roba

bilit

y)

mean95 percentilemean95 percentile

2- sigma

a) b)

relief 4.0 km 4.0 & 4.5 kmerosion rate 1.0 km/My 1.0 & 1.3 km/My

5 10 15 20 25 5 10 15 20 25

Theoretical PDF

b)

a)

Figure 12. Number of grains versus the mismatch error from 1000 iterations. The mean is shown with thick lines and the 95% confidence envelope in thinner lines. Two scenarios are illustrated: a) a theoretical PDF from a single basin with 4 km of relief eroding at 1 km/My (solid black line as mean) and; b) a more complex theoretical PDF constructed from a basin with 4.0 km of relief eroding at 1 km/My and a basin with 4.5 km of relief eroding at 1.3 km/My (dashed gray line as mean). Note that it is harder to match the more complex theoretical PDF than the individual basin.

47

Page 63: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

10 15 200

0.002

0.004

0.006

0.008

0.01

0.012

0.014

0.016

0.018

7.5%2.5%

10%

15%

age (My)

prob

abili

ty

20%

5%

theoretical PDF

mismatch withgrab sample

SPDFs

Figure 13. A selected range of outcomes from sampling 50 grains from the theoretical PDF (shaded gray) of a basin with 4 km of relief eroding at 1.0 km/My (as in Fig. 11). Note that while mismatches up to 20% sound large, in reality the grab-sample SPDF still captures the key attributes of the theoretical PDF: the peak-probability age of each grab-sample SPDF varies less than ± 1 My in comparison to the peak-probability age of the theoretical PDF. The mean mismatch for 1000 runs and 50 grains is 9 ± 7% (see Fig. 12) at the 95% confidence limit.

48

Page 64: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

49

Chapter 2

The downstream development of a detrital cooling-age signal, insights from 40Ar/39Ar muscovite thermochronology in the Marsyandi Valley of Nepal.

I.D. Brewer and D.W. Burbank

Pennsylvania State University, Department of Geosciences, University Park, Pennsylvania

K.V. Hodges

Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology,

Cambridge, Massachusetts

Abstract

The nature and variation of the distribution of cooling-ages in modern river sediment may provide

useful constraints on the rates of uplift and erosion within mountainous drainage basins. Such

sediment effectively samples all locations within the catchment area, however remote and

inaccessible, and may be preserved in the foreland basin. We assess the applicability of using

detrital cooling ages to constrain hinterland deformation by examining the modern drainage system

of the Marsyandi valley in central Nepal. Laser fusion 40Ar/39Ar data for detrital muscovite

collected from 12 separate sites illustrates that the downstream development of a detrital signal is

both systematic, and representative of the contributing area. The distribution of bedrock cooling

Page 65: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

50

ages in a sub-catchment and the resulting detrital signal at the basin mouth can be modeled a

function of the erosion rate, relief, hypsometry, drainage area, and the distribution

thermochronometer. Given that independent constraints are available for most of these variables,

the detrital age signal is a robust indication of the spatially averaged erosion rate. In the Marsyandi,

our model predicts ~2 fold differences in erosion rates, with the southern topographic front of the

Himalaya experiencing the most rapid rates of exhumation, exceeding ~2 mm/yr. Over the 100 to

200 km length scale of the Marsyandi basin, there is no significant comminution of the muscovite

grains. Comparison sample pairs from: a) opposite ends of the same sandbank; b) from the modern

river and a fill terrace, are not found to be statistically different at the 95% confidence level,

indicating that at short spatial (10’s of m) and temporal (1000’s of years) scales, the detrital cooling-

age signal appears to be stable.

1.0 Introduction

The growth and evolution of an orogenic belt can have global impact through interactions with

geochemical cycles [Derry and France-Lanord, 1996; Raymo et al., 1988] and climate [Kutzbach et

al., 1993; Ruddiman and Kutzbach, 1989]. Constraining the temporal variation in the development

of topography and the rate of erosion is fundamental to understanding the relationship between

orogenesis and these processes. Despite numerous previous studies of actively deforming mountain

belts, understanding the development of an orogen over geological timescales remains problematic

due to a lack of precise timing constraints. For example, bedrock thermochronology restricted to

rocks currently exposed at the surface today and hence provides a limited temporal record due to the

high erosion rates found in active orogens. Conversely, analysis of basinal sediment can yield a

good temporal record, but is commonly difficult to interpret directly in terms of hinterland erosion

and development.

Detrital mineral thermochronology offers the opportunity to combine the stratigraphic record

preserved in the foreland with the quantitative analysis of thermochronology. Sand grains preserved

in foreland-basin stratigraphy represent an integration of information from the contributing

upstream area at the time of deposition. Such a sample contains millions of sand grains, with each

Page 66: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

51

particle sampling a slightly different point within the basin. Before detrital thermochronological

data can be reliably interpreted, however, the processes controlling the cooling age signal need to be

understood and key questions need to answered: “How are the pattern of erosion, distribution of

lithology, and landscape characteristics manifest within the distribution of detrital cooling ages

observed in the foreland?” In addition, the effects of mechanical breakdown of the

thermochronometer within the river system, and the temporal and spatial reliability of the age signal

are unknown. For example, if we collect two samples from different locations within the river

sediment, will these produce different age distributions or will they represent one homogeneous age

population? If the age signal varies locally, or if a point source or a sub-catchment within the basin

controls the signal, then only a limited amount of information will be extracted from the sediment

preserved. To answer these questions using the stratigraphy, however, is impossible and so the

basic Principle of Uniformitarianism must be applied; we must understand the processes operating

in the modern environment in order to provide a basis for the interpretation of the geological record.

Our previous work [Brewer et al., Chapter 1] investigated factors controlling the modern detrital

signal in two small Himalayan catchments. The examination of individual catchments enabled the

distribution of bedrock cooling ages to be modeled as a function of topographic relief, erosion rate,

and the subsequent geothermal gradient. If isotherms are horizontal (a reasonable assumption at

depths > 8 km below the valley floors), the detrital cooling-age signal resulting from erosion of

topography is determined by the hypsometry (the distribution of area with elevation) of the basin.

These simple models, however, are applicable to drainage basins that are contained within one

tectonic zone and have homogeneous distribution of thermochronometer. In most active orogens

this constraint limits investigations of this type to catchments within larger, orogenic-scale drainage

systems. For example, Brewer et al. [Chapter 1] examined the cooling age signal of two tributary

basins within the larger Marsyandi River system.

In this paper we investigate the more extensive detrital system of the complete Marsyandi River,

and examine how the cooling-age signal of individual catchments combine to produce a modern

foreland signal. We use 40Ar/39Ar analysis of individual muscovite grains to examine how the

lithology, erosion rate, and hypsometry of individual catchment areas vary, and investigate how

these parameters control the evolution of the trunk stream cooling-age signal from the headwaters to

the foreland basin. This allows us to examine how well the cooling-age signal at the basin mouth

Page 67: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

52

represents the contributing area upstream, and hence provides a baseline study for the interpretation

of future detrital investigations.

2.0 Previous investigations of detrital thermochronology

Detrital-mineral thermochronology has been used to investigate: (a) the thermal evolution of

basins; (b) source area constraints; (c) stratigraphic age, and; (d) the erosion history of orogens. The

wide range in closure temperatures for different thermochronometers means that they can be applied

to many types of temperature-dependant geological problems. High-temperature

thermochronometers are typically used to date the crystallization ages of minerals, while low-

temperature thermochronometers are typically used to investigate the late-stage cooling history.

Thermal evolution studies of basins are commonly used to constrain hydrocarbon production

windows. They typically have relied on fission-track dating of detrital apatite [e.g. Green et al.,

1996], although 40Ar/39Ar thermochronology has also been used [Copeland et al., 1996; Mahon et

al., 1998]. The age at which mineral cooling ages are reset by burial metamorphism has been used

to constrain models for the thermal evolution of the basin. Low-temperature thermochronometers

are most commonly used because kerogen breakdown starts to occur at < 400°C [Levorsen, 1967].

Studies of mineral-isotopic systems with high-closure temperatures (e.g. U-Pb zircon or monazite)

are principally aimed at characterizing sediment source areas [i.e. Adams et al., 1998; Krogh et al.,

1993; Krogh et al., 1987; Roddick and van Breemen, 1994]. The low closure or annealing

temperatures of 40Ar/39Ar (hornblende, muscovite, biotite and k-feldspar), (U-Th)/He (titanite and

apatite), and fission-track (zircon and apatite) thermochronometers make them ideal tools to

investigate stratigraphic age and erosion history. When investigating stratigraphic age, the youngest

individual grain age provides a constraint on the maximum depositional age of the sediments. For

example Najman et al. [2001] and Najman et al. [1997] use 40Ar/39Ar analyses from individual

muscovite grains to constrain the age of the oldest exposed foredeep sediments of the Himalaya.

The cooling age of detrital minerals within a sedimentary rock can be corrected for the

stratigraphic age of that rock and used to define cooling rates at the time at which the rock was

deposited. With low-temperature thermochronometers, the corrected cooling rates are typically used

Page 68: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

53

as a proxy for the erosion rate within the catchment at that time. The Himalaya has been a focus for

many such studies. Cerveny et al., [1988] were among the first to use this technique; in Pakistan,

detrital zircons from Indus River sediments provided evidence of a young 1-5 Ma cooling-age

signal that had been persistent for the past 18 Ma. Copeland and Harrison [1990] interpreted 40Ar/39Ar analyses of detrital K-feldspar and muscovite grains from the distal Bengal fan to record

Himalayan erosion rates through time. The youngest cooling age at each stratigraphic level was

approximately equal to the depositional age, suggesting that that rapid cooling was occurring at

some location within the catchment area for the past 18 Ma.

Despite the value of the aforementioned studies, the parameters controlling the detrital signal and

the appropriate inferences are poorly known. Stock and Montgomery [1996] investigated the

potential of the detrital cooling-age signal to place limits on the paleotopography in the source area.

They show that the range in ages produced by a particular basin, in combination with a specified

geothermal gradient, allow the relief to be calculated. Brewer et al. [Chapter 1] extended this study

by investigating how the interactions between geothermal gradient, erosion rate, and relief can be

used in conjunction with the basin hypsometry to predict the distribution of detrital cooling ages.

Brewer et al. [Chapter 1] also address the random nature of detrital sampling and statistically assess

how faithfully specified numbers of grain ages will reproduce a defined detrital signal.

The results of Stock and Montgomery [1996] and Brewer et al. [Chapter 1] indicate that the

detrital cooling-age data can provide useful tectonic-geomorphological insights, although the scale

of such investigations needs to be limited to individual drainages contained within zones of uniform

uplift. Studies of catchments and their detrital signals at the orogenic scale, however, require

consideration of many tectonic zones; a large drainage will produce a more complex signal

representing an integration of detrital grains from multiple tributaries. In this paper, we attempt to

build on previous work to understand the parameters controlling the complete cooling-age signal

from an orogenic-scale basin. Field data from the Marsyandi valley in central Nepal are used to

examine the spatial pattern of erosion in the modern Himalaya.

Page 69: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

54

3.0 Geological Background

The Himalaya are often described as the quintessential continent-to-continent collision zone, and

represent the discrete chain of extreme topography on the southern edge of the Tibetan Plateau and

the more diffuse deformation to the north [Molnar and Tapponier, 1975]. To the north of the

Himalayan topographic axis, the Indus-Tsangpo suture zone marks the surface boundary between

lithologic units of Eurasian plate affinity to the north and Indian plate affinity to the south. After the

initial collision at ~50 to 54 Ma [Rowley, 1996; Searle et al., 1997], the downgoing Indian plate was

imbricated along a series of south-vergent thrust fault systems, and the Himalaya formed as a result

of the subsequent crustal thickening. Deformation continues today and GPS data [Bilham et al.,

1997] indicate that approximately one third of the current convergence between the Indian Plate and

Eurasian Plate, estimated at 58 ± 4 mm/yr from global plate motions [DeMets et al., 1990],

currently occurs across the Himalaya.

The thrust fault systems mark many of the principal boundaries between tectono-stratigraphic

divisions, and have been used to characterize Himalayan geology for decades. The oldest and most

northerly of these, the Main Central Thrust (MCT) system, juxtaposes high-grade metamorphic

rocks and leucogranites of the Greater Himalayan sequence against lower-grade metasedimentary

rocks of the Lesser Himalayan zone. Farther south, the Lesser Himalayan zone is separated from the

foreland basin of the Himalaya by the Main Boundary Thrust (MBT) system. The most foreland-

wards topographic expression of the collision, the Siwalik Hills, corresponds to the Main Frontal

Thrust (MFT) system. The initiation ages of these major thrust systems are progressively younger

from north (20 to 23 Ma for the MCT system) to south (Pliocene to Holocene for the MFT system),

although ample evidence exists for episodic out-of-sequence thrusting along these and other less

significant fault systems in the Himalayan realm over the Miocene-Recent interval [Hodges, 2000].

The surface trace of the MCT system marks the approximate physiographic transition from the

Lower Himalaya to the Higher Himalaya, representing a large increase in mean elevation and relief.

As a consequence, much of the steep southern front of the Higher Himalaya, where erosion rates are

likely to be the highest, is developed on the metamorphic and igneous rocks of the Greater

Himalaya sequence. A fourth important fault system marks the top of the Greater Himalayan

sequence: the South Tibetan Detachment System (STDS). Carrying essentially unmetamorphosed,

Page 70: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

55

Neoproterozoic-Paleaozoic clastic and carbonate sedimentary rocks of the Tibetan zone in its

hanging wall, the STDS incorporates a variety of structures, but chief among them are low-angle,

north-dipping detachments with normal-sense displacement [Burchfiel et al., 1992]. In the study

area, the STDS comprises two splays, the Chame Detachment Fault and Machhapuchhare

Detachment Fault, which are separated by the greenschist-to-amphibolite grade marble of the

Annapurna Yellow Formation [Coleman, 1996; Hodges et al., 1996].

The Marsyandi River system of central Nepal (Fig. 1) has its headwaters north of the trace of the

STDS and drains portions of the Tibetan, Greater Himalayan, and Lesser Himalayan zones over an

area of ~4760 km2. Its major tributaries flow over subsets of these tectonostratigraphic zones and

their streambed sediments thus sample different zones of bedrock in different proportions. As they

flow into the main Marsyandi trunk stream, individual tributary signals are progressively mixed

downstream. The Khansar Khola ("khola" is the Nepali word for river) and Nar Khola

predominately drain Tibetan zone sedimentary bedrock. The Dudh Khola drains Tibetan zone rocks

as well as a major Miocene leucogranite, the Manaslu pluton [Le Fort, 1981]. The Dona Khola

flows over exposures of this pluton as well as a variety of metamorphic rocks of the Greater

Himalayan sequence. The Miyardi, Nyadi, and Khudi rivers have headwaters in the Greater

Himalayan zone and exclusively sample this bedrock before emptying into the Marsyandi. The

Dordi, Chepe, and Darondi rivers flow across the MCT system and thus have sediments with

provenances in both the Greater Himalayan and Lesser Himalayan sequences.

4.0 Methodology

4.1 Sampling strategy

Detritus shed from an evolving mountain belt is primarily transported to the foreland basin by

fluvial systems. The resulting stratigraphy can provide a proximal record of hinterland erosion over

orogenic timescales. As a consequence, detrital cooling-age signals extracted from the foreland

potentially provide one of the highest spatial and temporal resolution records of orogenesis

Page 71: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

56

available. One of our goals in studying the detrital cooling-age signal of the Marsyandi river system

is to evaluate how faithfully samples of sediment from major transverse rivers reflect the pattern of

bedrock cooling ages in the source area. Therefore, it is important to examine how faithfully a

foreland-basin sample portrays the pattern of bedrock cooling ages in the orogenic belt, and

understand the limits of the detrital signal.

In order to constrain the modern foreland cooling-age signal, we need to understand the present

distribution of cooling ages within the hinterland, and how these are eroded and transported

downstream. The sampling strategy in this investigation was designed to: a) maximize the statistical

constraints on contributions by tributary cooling-age signals; b) investigate the downstream

development of the trunk-stream cooling-age signal, whilst; c) using small enough catchment areas

to constrain adequately the spatial variation in cooling ages. With a finite number of age analyses,

there will always be unavoidable trade-offs between obtaining the optimal representation of any

given tributary and reliably reconstructing the evolution of the detrital age signal along the course of

a large river. For example, using very small catchments will increase the resolution of spatial

variation in bedrock cooling ages, but leave fewer analyses to constrain the downstream evolution

of the trunk stream. Conversely, using more analyses to constrain the foreland cooling-age signal

will provide a more reliable characterization the individual sample, but limit the spatial resolution of

the study.

Detrital sand samples were collected within the Marsyandi catchment from sites ranging from the

Tibetan zone to the junction with the Trisuli River in the Lesser Himalayan zone (Fig. 2). At each

sample site, large-grained sand was collected from bars within the modern river channel. Care was

taken to collect samples upstream of sediment-mixing zones at river junctions, and to avoid the

influence of small side tributaries and fill deposits, such as terraces.

While 27 samples were collected for point counting (to characterize the mineralogical constitution

of the sediment), only 14 samples from twelve separate locations were selected for 40Ar/39Ar dating.

Six samples were chosen from the Marsyandi River to investigate the trunk stream, and five

samples were taken at the mouth of major tributaries (Fig. 2). One sample (S-40) represents a sub-

catchment within the overall Darondi Khola (S-37), taken to assess the relative input of ages from

the Greater Himalayan sequence portion of the basin in comparison to the entire basin. To examine

the temporal variability of the detrital signal, two samples were collected from the same location;

Page 72: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

57

one from the modern river bed (S-8), and one from a fill terrace elevated 2 m above it (S-9). To

examine the natural spatial variability of the signal samples were collected 45 m apart on the

downstream (S-53) and upstream (S-52) ends of the same sandbar.

4.2 40Ar/39Ar Analytical Protocols

Sand samples were washed and grain sizes between 500 and 2000 µm separated by sieving. This

investigation focused on detrital muscovite, which has been widely used in detrital mineral

geochronology and appears to have fewer problems with excess argon than biotite [Roddick et al.,

1980]. Individual muscovite grains were extracted from the sieved fraction by applying standard

gravimetric and magnetic separation techniques, followed by hand-picking under a binocular

microscope. Mineral separates were irradiated at the McMaster University research reactor after

being washed sequentially in distilled water, acetone, and ethanol. Aliquots of the neutron-fluence

monitor Fish Canyon sanidine (28.02 Ma, Renne et al. [1998]), as well as a variety of salts that

served as monitors for interfering nuclear reactions, were included in the irradiation package.

Mineral grains and monitors were analyzed at the 40Ar/39Ar laser microprobe facility at the

Massachusetts Institute of Technology [Hodges, 1998]. Individual muscovite crystals underwent

fusion in the defocused beam of an Argon laser operating at 18 W for a period of approximately 10

seconds. After purification to remove reactive species, the extracted gas was analyzed on an MAP

215-50 mass spectrometer using a Johnston electron multiplier. At the beginning of each analytical

session and after every tenth analysis of an unknown, the total system blanks were measured.

Appendix 2, Table 1, shows the apparent ages calculated for each muscovite with an estimated 2-σ

uncertainty obtained by propagating all analytical uncertainties. Hodges and Bowring [1995]

provide additional details on the analytical techniques.

Detrital cooling-age signals are commonly represented as a probability density function, which

represents the probability of finding a grain of a particular age, as a function of the age [Deino and

Potts, 1992]. Assuming that a Gaussian kernel represents the distribution of error [e.g. Bevington

and Robinson, 1992], a probability density function can be calculated for each grain, given the age (

Page 73: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

58

tc) and analytical uncertainty (σ). For a sample of N grains collected from a specific locality, the

PDF of individual grains (n) can be combined:

∑=

=

−−

=Nn

n

nntt c

en

tP1

))(.2))((

( 2

2

..2).(

1)( σ

πσ

(1)

By normalizing the area under the resulting curve to unity, a summed probability density

function (SPDF) is generated. The SPDF represents the distribution of age probability as a function

of all the grains analyzed from the sample.

4.3 Point Counting.

Cooling-age signals can be used to examine how a trunk-stream signal changes downstream, as

successive tributaries contribute varying age populations (as described below). Point counting

provides a complimentary approach to investigating the trunk-stream signal that does not rely on

thermochronology and is less expensive. Detrital minerals can be used as conservative tracers,

whereby the relative abundance of a particular mineral species is used to examine the relative

contribution from an individual tributary. Although this technique has much lower resolution than

the thermochronological approach, it can provide another constraint on the relative erosion rate. In

addition, point counting serves to delineate the relative abundance of the target thermochronometer

within the Marsyandi study area.

The grains used for 40Ar/39Ar analysis are not a complete representation of the fluvial sediment

due to the sample and separation procedure necessary for grain selection. Field sampling restricts

the grain diameter to sand-size particles, and sieving further restricts this to 500 to 2000 µm.

Therefore any analysis of point counting results used in conjunction with cooling-age signals has to

be considered within the context of this range of grain sizes. Hence, in this study, point counting

was used to quantify the distribution of muscovite and other components within the 500-2000 µm

fraction.

Page 74: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

59

Sieved sand samples were stained red for plagioclase feldspars and yellow for alkali feldspars

before being thin sectioned and counted using a mechanized stage. Quartz, feldspars, and micas

were the major minerals counted, with additional minerals grouped together and rock fragments

considered to be an additional species. Crystalline carbonate was considered to be a mineral,

whereas granular carbonate was considered a rock fragment. This method produced an approximate

quantification of the major constituents in each sample (Table 1) while not involving large amounts

of time identifying the broad range of other constituents. In each sample 600 to 900 grains were

identified. All reported errors from the point-counting results are taken from the statistical analysis

of Van der Plas and Tobi [1965]. Repeat counts were performed on three samples (S-1, S-2, S-3:

Table 1) to investigate the consistency of individual counts. Overall, the recounts showed that the

results were indistinguishable at the calculated 2-σ confidence level.

5.0 40Ar/39Ar results.

Before we attempt to interpret the 40Ar/39Ar results using a modeling approach, it is useful to

examine the broad trends shown by the data. The most encouraging observation is that the trunk-

stream age signal illustrates a systematic downstream pattern (Fig. 3). The sample furthest upstream

(S-12) drains the headwaters of the Marsyandi from the edge of the Tibetan Plateau to the crest of

the Annapurna massif. The age signal is dominated by an age population concentrated between 12

and 16 Ma. One source of ages may be rare hydrothermal veins with 40Ar/39Ar muscovite plateau

dates of ~14 Ma that have been reported from Tibetan zone sedimentary rocks that dominate the

bedrock lithology in this area [Coleman and Hodges, 1995]. Alternatively, the catchment contains a

small portion of the Annapurna Yellow Formation and the upper Greater Himalayan sequence,

which also contribute muscovite. The next sample downstream (S-8/ S-9) was collected to the

south of the trace of the STDS. This sample is influenced by two additional major tributaries, the

Dudh and Dona Khola, which drain the top of the Greater Himalaya sequence and the Manaslu

Granite. Whereas the population of ages observed in S-12 is still represented, the cooling-age signal

is dominated by a major age population from 15 to 20 Ma. The weak expression of the < 15 Ma age

population in the downstream sample, and the paucity of muscovite in the upper reaches of the

Page 75: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

60

drainage basin suggest that the area upstream of sample S-12 makes a minor volumetric

contribution when compared to the additional tributaries found upstream of sample S-8/ S-9.

The next trunk-stream sample (S-6) was collected approximately 20 km downstream. This

displays the same 15 to 20 Ma population seen in the upstream sample, but contains additional 5 to

15 Ma ages. After the Marsyandi has crossed the MCT zone (S-3), the 0 to 10 Ma population

becomes more dominant and the 15 to 20 Ma peak represents less than half the total probability.

The Nyadi Khola (S-5), which drains the lower Greater Himalayan sequence, is partly responsible

for the increase in the younger age population as it comprises exclusively the 0 to 10 Ma age

population. Downstream of S-3 the Khudi Khola (S-2) contributes a young population of ages

between 4 and 12 Ma to the trunk stream. Other tributaries, the Dordi Khola (S-44) and Chepe

Khola (S-54) exhibit asymmetric 3 to 10 Ma and 4 to 13 Ma age populations, respectively. The

main Marsyandi River shows a dominance of the 5 to 10 Ma age population after the influx of these

tributaries (sample S-52/S-53).

Two samples were collected from the Darondi Khola: S-40 was collected from above the base of

the MCT zone, whilst S-37 was collected from the basin mouth. Sample S-40 displays a strong 0 to

12 Ma age population. Sample S-37 is similar, but includes a single older age component. The

trunk-stream signal at the basin mouth (S-24) comprises a prominent 5 to 10 Ma signal, a lesser 10

to 15 Ma signal, and a weak 15 to 20 Ma signal.

6.0 Modeling

Given that the results of the 40Ar/39Ar analysis seem to behave in a systematic way within the

Marysandi drainage system, we can use numerical modeling to further our understanding of the

hinterland geology, and interactions within the drainage system. This allows us to examine the

impact of individual parameters and focus on which characteristics of the detrital signal can be

explained by the model, and which cannot due to the limitations of the initial assumptions. Thus, in

combination with the 40Ar/39Ar data, we want to use a numerical model to: 1) assess the spatial

variation of parameters that control the hinterland geology; 2) understand how these parameters are

Page 76: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

61

manifest in the detrital cooling-age signal observed at the basin mouth, and; 3) examine the

reliability and resilience of the cooling-age signal.

To do this we constructed a theoretical tributary PDF for each basin within the Marsyandi

drainage system [Brewer et al., Chapter 1]. This was generated by: 1) putting in the real

topographic characteristics of the basin, and then; 2) finding the optimal match between the

theoretical and observed data PDFs by varying the erosion rate. Once theoretical PDFs had been

generated for each tributary, we modeled the relative contribution from each basin to the trunk

stream. Hence we examined the systematic mixing of age populations in order to understand and

predict the downstream evolution of the cooling-age signal within the Marsyandi Valley.

6.1 Modeling the detrital cooling age signal

Before examining the trunk-stream signal, we need to generate theoretical PDFs for each tributary.

To predict the cooling age of a bedrock sample within a tributary basin the predicted depth of the

closure isotherm, at the time when the mineral passed through the closure temperature, is divided by

the rate of erosion. The depth of the closure temperature, given a crust of predetermined thermal

characteristics, is a function of the topographic relief and the vertical rate of erosion. For virtually

all geologically reasonable erosion rates and topographic relief, this isotherm will experience

negligible deflection due to surface topography [Mancktelow and Grasemann, 1997; Stüwe et al.,

1994]. We use a simplified thermal model [Brewer et al., Chapter 1] to predict the depth of the

350°C closure isotherm as a function of the basin relief and a specified erosion rate. The thermal

model assumes vertical erosion of material containing uniform heat production and conductivity,

from a depth of 35 km, through a steady-state landscape that contains hillslopes at a threshold angle

for landsliding (~30°: Burbank et al. [1996]). These assumptions produce a linear distribution of

cooling-ages with elevation, which forms the basis for calculating the detrital cooling-age signal.

With a uniformly eroding basin, or sub-basin, the PDF is controlled by the distribution of area with

elevation, i.e., the hypsometry. Thus the likelihood of sampling age a particular age at the basin

mouth is a consequence of the fraction of land containing that age.

Page 77: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

62

Using this basic model, the larger Marsyandi Valley was broken into individual tributary basins to

represent the area contributing to each of our tributary 40Ar/39Ar age samples. Treating each sample

separately allowed us to model changes in the cooling-age distribution due to spatial variations in

the erosion rate and topographic relief at the tributary scale. The topographic relief and area of each

basin was extracted from a 90-m digital elevation model (DEM) using Arc/Info software. Thus,

with the other parameters constrained, the basin erosion rate was the only unknown parameter.

Hence, the lowest mismatch [Brewer et al., Chapter 1] between our theoretical PDF and the data

PDF was found by varying the erosion rate within an individual basin. Once the optimal theoretical

PDF had been found for a tributary basin, the erosion rate was fixed for any subsequent analysis.

Given the theoretical cooling-age distributions from the tributaries, we now need an additional

model element to predict how individual basins coalesce to produce the trunk-stream signal.

Ultimately, using all the tributaries, we want to predict the whole-basin PDF that serves as a proxy

for the modern foreland-basin deposit that would be produced by the Marsyandi River. The relative

contribution from an individual tributary is a function of the relative amount of muscovite per unit

time eroded from the basin (Fig. 4).

For our model we assume that a long-term steady-state topography exists whereby the regional

mean relief, hypsometry, and drainage density are essentially invariant over timescales exceeding

0.1 My. As a consequence of steady-state, the flux of material out of a basin will balance the

volume of rock moving into a basin. With vertical denudation, the volume of material derived from

a tributary is a function of the basin area and the erosion rate, and the relative contribution to the

trunk stream will reflect this. For illustration, we can consider basin ‘A’ and basin ‘B’ which are

drained by two individual tributaries that converge downstream to produce a trunk-stream signal. If

the basins have equal area, but B is eroding at twice the rate of A, the relative contribution to the

trunk-stream sample will be A:2B.

With the simple relationship of flux being proportional to area for a given erosion rate, there is an

assumption of uniform lithology. In this study our thermochronometer is muscovite, which may be

heterogeneously distributed through the basin: it is common in high-grade metamorphic rocks for

example, but often absent in carbonates. Therefore, even if basin B is eroding twice as fast, its

Page 78: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

63

contribution to the downstream sample may be negligible if it is dominated by carbonate lithologies.

We use the percentage of muscovite at the basin mouth, taken from the point counting results (Table

1), to calculate a correction factor for each basin. Note that the point-counting results indicate that

the percentage of muscovite varies over two orders of magnitude within the Marsyandi river system,

indicating that a lithological correction factor can profoundly affect the results.

To model the trunk-stream signal at a particular location, we considered only the predetermined

tributary PDFs upstream of the sample. The individual tributary PDFs were combined after they had

been corrected for lithological variation, erosion rate, and area (Table 2). The resulting PDF curve,

with area normalized to unity, represented the theoretical distribution of cooling ages within the

trunk-steam sediment, at that locality. To examine the evolution of the detrital signal, we worked

systematically downstream by combining the calculated trunk-stream signal upstream of a sample

site with the addition from individual tributaries between the two sample sites. In the case where

tributary addition was unconstrained by ages at the mouth of a tributary (for example, the

inaccessible area represented by the Miyardi Khola: Fig. 2), a geologically reasonable erosion rate,

given the surrounding tributaries, was assigned to the area so as to minimize the mismatch of the

trunk-stream signal in the sample directly below the junction. For the purposes of the modeling, the

observed and data PDFs were smoothed using a 2-My-long scrolling window. This reduced the

“peakiness” of the PDFs caused by individual grains, meaning that the calculated mismatch was less

affected by individual grain peaks, but rather reflected the overall pattern of the entire signal.

6.2 PDF modeling results

With our approach of first modeling the distribution of detrital cooling ages from individual

catchments, and then combining them to model the trunk-stream signal, we can see that the overall

pattern of theoretical PDFs is consistent with the observed data (Fig. 5). The main peaks of the

theoretical PDFs generated for each tributary generally align with those of the data PDFs because

we minimize the mismatch by varying the erosion rate. Generally, the tails on either side of the

younger peaks are harder to match with our approach (i.e. see the Khudi Khola theoretical PDF for

illustration). The trunk stream displays a systematic pattern of change downstream as tributary

Page 79: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

64

signals are added. The > 15 My peak that is prominent in the upstream samples becomes diluted

downstream as the 5 to 10 My peak becomes increasingly important, although in comparison to the

data, the predicted basin mouth PDF (represented by S-24) is relatively depleted in the 10 to 15 My

age range.

Some assumptions were made in the modeling in order to find the best solution to the pattern of

observations. Given fixed hypsometries and relief, the erosion rates of the Nar and Khansar were

varied until the best match of the most upstream sample (S-12) was found, yielding an erosion rate

of ~1 mm/yr. However, mixing the contributions from the Dudh and Dona tributaries to the

upstream sample (S-12), according to the initial values, produced a signal that was too dominant in

the 10 to 15 My age range. Therefore, to reproduce the downstream sample S-8/S-9, the relative

contribution from the Nar and Khansar had to be reduced by ~50% of that indicated by the raw data

for the percentage of muscovite. This is geologically reasonable because of the very high proportion

of carbonate rock fragments in the sediment from theses two tributaries. The point counting

indicates that these rock fragments decrease rapidly downstream in the sediment samples below S-

12 and so do not appear to be very resistant. It seems likely, therefore, that significant breakdown of

rock fragments has already occurred within the river before reaching the sample site, S-12. Thus, if

the rock fragments were conservative, we would expect a much higher proportion of them at the

basin mouth, and consequently a much lower percentage of muscovite. In contrast, the percentage

muscovite from the Dudh and Dona tributaries is probably much more representative of the

distribution within the catchment because the large percentage of granitc rock fragments are far

more resistant than carbonate fragments, and thus conservative within the fluvial system.

Furthermore, given that the Tethyan zone contains mostly carbonate rocks and shales, it is likely

that much of the muscovite in sample S-12 is derived from, and representative of, only a small

fraction of the total catchment area situated beneath the Machhapuchhare detachment fault.

Although the model generates a good prediction of the age distribution from tributaries with older

age distributions, it has difficulty reproducing the full range in ages found in many of the tributaries

that yield younger cooling ages. The strong asymmetry observed in the older age tails, in particular,

is difficult to replicate (the 8 to 13 Ma ages observed in the Chepe, for example). Although this

mismatch will be reduced when the model PDF curves are sampled (the analytical errors associate

Page 80: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

65

with a grain age will widen the range of ages at each end), this cannot account for all of the

mismatch between the tails of the observed PDF’s and the modeled PDF’s (Fig. 5). Although we

assume that each basin is uniformly eroding, in reality there are probably variations in erosion rate

in the basins to the south of the range crest because uplift rates are locally controlled by the

geometry of the active subsurface structure [e.g. Lave and Avouac, in review; Pandey et al., 1995;

Seeber and Gornitz, 1983]. However, with no detailed sub-tributary data or bedrock ages from these

regions, we have limited the modeling to the same resolution as our data. The under-represented

tails of the observed data certainly account for some of the mismatch seen at the basin mouth (S-

24). Another source of mismatch would be areas not included in the modeling (stippled in Fig. 8),

yet contributing to the detrital cooling age signal seen in sample 24. There were insufficient

geochronological data from these areas to constrain their erosion rate. Most of these areas lie in the

Lesser Himalaya, and if they have intermediate erosion rates, as their low relief and topography

would suggest, then they may be an additional source of 10 to 15 Ma ages not represented in the

model.

7.0 Discussion

7.1 Resilience of the detrital signal.

One of the main concerns with the application of detrital dating is the survivability of the

thermochronometer in the sediment routing system. If the river network causes a rapid comminution

of grains, then there will be a limited sampling window upstream of the sample site. The size of the

window will be controlled by the rate of attrition within the system. Such comminution will

influence the cooling-age signal preserved in the foreland because information becomes

progressively lost downstream from the headwater area. In the best scenario, an ideal

thermochronometer is neither destroyed nor altered during the weathering and transportation history

Page 81: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

66

of the sediment. Similarly, post-depositional weathering and diagenesis should not affect the

thermochronometer in the resulting sedimentary rock.

Both chemical and physical processes may cause the break down of minerals. The effects of

chemical weathering was investigated by Najman et al. [1997] for muscovite from the Dagshai and

Kasauli Formations in the Himalayan foreland, with depositional ages of approximately < 25 to 28

Ma and < 22 to 28 Ma respectively. They examined thin sections and conducted electron

microprobe analyses on the micas they dated. Most grains were essentially unaltered, with only

slight modification occurring at the edges of a few grains. Because these sediments are: a) much

older than the modern grains we use; b) from the same depositional basin, and; c) the Himalaya

continue to undergo rapid erosion rates today, we assume that chemical alteration of muscovite does

not significantly affect our results. In addition, in a study of modern tropical weathering, Clauer

[1981] found that biotite is more susceptible to breakdown and loss of argon than muscovite, which

displayed little to no detectable weathering effects. Furthermore, some work suggested that the

chemical weathering of micas may produce a congruent loss of K and Ar such that the apparent

ages of altered minerals may be relatively unaffected by the process [Mitchell and Taka, 1984].

The process of physical attrition of muscovite grains during their passage through the fluvial

system is more difficult to assess. With a hardness of 1 to 2 on Mohs scale and a well-developed

basal cleavage, it seems reasonable to suspect that muscovite may be susceptible to physical

breakdown. However, the work of Copeland and Harrison [1990] suggest that muscovites were

capable of surviving transport from the Himalayan front to a distal site in the Bengal Fan, a distance

of up to 2000 km, with little disturbance of their 40Ar/39Ar systematics. Given that the length scale

of the Marsyandi catchment is an order of magnitude smaller, 100 to 200 km, it seems unlikely that

the physical breakdown of muscovite is significant. The persistence of the 15 to 20 Ma age signal

from the upper Greater Himalaya sequence, through all our trunk-stream samples, to the basin

mouth lends support to this assumption. Rather than being a result of breakdown, we interpret the

downstream decrease in the 15 to 20 Ma age signal to be an effect of dilution. Our modeling

assumes a sediment flux proportional to the erosion rate, basin size, and contains no function for the

downstream loss of cooling age signal with distance. Therefore, if comminution were significant

over this length scale, the model would be expected to over-represent the 15 to 20 My age fraction,

Page 82: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

67

particularly in the lower reaches of the river as muscovite grains are progressively destroyed.

Instead, the results approximate the relative proportion of younger to older populations seen in the

data, and indicate that the rate of sediment supply by faster eroding catchments (contributing

younger cooling-age populations) appears to control the downstream signal, as opposed to

significant loss of 15 to 20 My micas.

Given this finding, it is interesting to compare the volume of sediment that our model predicts is

eroded from unit area, for a given aged grain. The results (Fig.6) show that there is significant

differences in the volume of sediment for different cooling ages (representing different erosion

rates). Thus if we consider a basin containing one half eroding at ~2.1 km/My (Fig. 6, i) and the

other half eroding at ~0.8 km/My (Fig. 6, ii), the proportion of sediment contributed by the first half

will be over two-fold greater than that contributed from the second half (indicated by the gray bars).

Thus, if we wish to investigate the relative proportion of a catchment area that is producing a signal

of a specific age, we need to correct for the volumetric contribution of that age. Therefore, although

the 15 to 20 Ma age fraction in sample S-24 is relatively minor, because it comprises older cooling

ages it is areally important in the upstream area. However, note that the relationship between

predicted age and erosion rate (Fig. 6) is not linear in our model because the closure isotherm varies

as a function of denudation rate [Brewer et al., Chapter 1]. Halving the erosion rate from 2 to 1

mm/yr has a large affect, for example, changing the predicted cooling age from 5 to 20 My. The

exponential form of this curve, however, means that when comparing terrains producing cooling

ages of > ~30 Ma, the difference between the volumetric contribution per unit area as a function of

cooling age will be minor. Hence, the probability distribution of age populations in the sediment

will more closely reflect the size of the contributing areas.

7.2 The reliability of the fluvial signal

Two important assumptions in detrital thermochronology studies are: a) that the river is efficiently

mixing the sediment, and; b) that the detrital signal is not prone to point sources causing strong

temporal variations in the age signal. Landslides, rock falls, and localized storms could influence

the latter. If these assumptions are correct, a grab sample should provide a good representation of

Page 83: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

68

the entire signal from the river and the upstream catchment. Two pairs of samples were collected to

investigate the validity of these assumptions.

The first pair of samples was used to study the homogeneity of fluvial mixing, in particular

answering the question, “If a modern sand bar is sampled, are there significant differences in the

distribution of ages from one site to another?” Sample S-53 duplicates trunk stream sample S-52,

but was collected approximately 45 m upstream at the other end of the sand bar. To test if these two

curves are statistically differentiable, we applied the Monte-Carlo methodology described by

Brewer et al. [Chapter 1] to generate random grain ages, and resulting synthetic SPDFs. To

represent the large range in errors seen in our single-grain analyses, a more complex method of

specifying the 1-σ analytical uncertainty was used in our model. Because no consistent relationship

of age uncertainty with grain age was observed, a PDF was constructed to represent the probability

of a specified age uncertainty occurring (Fig. 7). The PDF illustrates that while most the 1-σ

analytical uncertainties were less than 1.5 My, there is a significant portion of grains with larger

associated errors. Therefore, for each modeled grain, the uncertainty PDF (Fig. 7b) was randomly

sampled (using the same Monte-Carlo methodology), and the 1-σ error obtained from that PDF

applied to the model-derived grain age.

Using the modeled theoretical PDF for S-52/S-53 (Fig. 5), 500 “grab-sample” curves were

generated using first 15 grains to represent S-52, and then 22 grains to represent S-53. The

mismatch between the grab-sample SPDF and the theoretical PDF was then calculated [Brewer et

al., Chapter 1] and compared against the mismatch calculated between S-52 and S-53. As with the

modeling, a 2-My smoothing window was used before comparing curves to reduce the influence of

individual grains. The mismatch between S-52 and S-53 was calculated to be 32% (Fig. 8), whereas

the modeling indicated that a mismatch of ~20 ± 14% could be expected given a 95% confidence

limit. Thus the mismatch between S-52 and S-53 is just within the expected range of variability due

to random selection, and the two cannot be proven to be statistically different despite originating

from different positions within the river. Furthermore, the actual range of mismatch may be larger

than the value calculated from this modeling because it is based upon the synthetic PDF illustrated

in figure 5. As mentioned in section 6.2, the modeled PDFs are underrepresented in the 10 to 15 My

Page 84: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

69

age range (when compared to the data), and thus increasing the range of likely probabilities might

be expected to cause a likewise increase in the variability of sampling.

The second pair of samples was designed to evaluate the temporal consistency of the detrital

cooling-age signal at a particular location. Would samples that are separated by 100’s to 1000’s of

years show the same age distribution? If not, the question of sediment storage and production

becomes important. Large bedrock landslides would be one scenario where the style of sediment

production could affect the detrital cooling-age signal. One can imagine that a landslide might

produce a pulse of sediment containing single ages. To test the short-to-medium-term temporal

persistence of the detrital signal in one location, we collected one sample (S-08) from the modern

riverbed, and another (S-09) from an adjacent fill terrace (Fig. 8b). The terrace formed due to

sediment infilling the Marsyandi Valley behind a bedrock-landslide dam. The Marsyandi is now

beginning to incise through the massive landslide blocks, leaving the fill terrace approximately 2m

above the current river level. The same procedure was applied to these samples, resulting in a 5%

mismatch between S-8 and S-9, and an expected mismatch of 21 ± 15% derived from the modeling.

Thus the two PDFs are very similar and cannot be considered statistically different. One caveat to

this analysis is however, that the almost unimodal cooling signal may mask temporal variability in

sediment supply from the catchment area.

7.3 Spatial variations of erosion rate

The cooling ages derived from low-temperature, bedrock thermochronology in orogenic belts are

primarily used as a proxy for the erosion rate. Likewise, the major parameter that we are trying to

extract from the cooling-age signal is the spatial variation in erosion rate. The results of the PDF

modeling indicate that erosion rates vary by over a factor of two within the Marsyandi drainage

system, from 0.9 to 2.3 km/My (Fig. 9). The highest erosion rates of 1.9 to 2.3 km/My are found in

basins draining the topographic front of the Himalaya. Rates decrease to the north, with the TSS

eroding at rates of 0.9 to 1.1 km/My. Areas to the south of the MCT probably have intermediate

rates, but it is difficult to constrain the signal from solely the Lesser Himalaya as most of the rivers

also drain the Greater Himalaya sequence.

Page 85: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

70

Consider the results from the Darondi Khola (Fig. 3). Approximately 40% of the basin area lies

above the MCT zone and is represented by sample S-40. Sample S-37, from the basin mouth, is

very similar to sample S-40. It drains the additional area of the Lesser Himalaya, which is 1.5 times

the size of the Greater Himalaya portion of the basin. The similarity indicates that either: a) the

same erosion rates prevail in both portions of the basin, or; b) the Higher Himalaya sequence is

producing most of the thermochronometer found at the basin mouth. The latter could be explained

either by no effective erosion of the Lesser Himalaya sequence, or low percentages of muscovite in

the Lesser Himalayan lithologies. Point counting indicates that the Greater Himalaya sequence is

probably dominating the signal because Lesser Himalayan catchments have low abundances of

muscovite compared to those draining the Greater Himalayan sequence (Table 1). In addition, lower

rates of erosion in the Lesser Himalaya, compared to higher rates on the topographic front, may be

an explanation for the older PDF “tails” that are difficult to fit with the model, for catchments

draining both regions.

7.4 Point-counting results

The point-counting results (Table 1) indicate that the sediment composition behaves in a

systematic way with the downstream addition of tributary material. Samples from the Khansar and

Nar are very rich in rock fragments, containing up to 80%, mainly limestone, clasts. This is

expected due to the drainage area of predominantly TSS rocks. The percentage of rock fragments

falls rapidly downstream, with some addition of granitic rock fragments, to between 20 and 30%

towards the basin mouth. The Dudh and Dona Khola, draining the Manaslu granite, are quartz- and

feldspar-rich and somewhat under-represented in micas. The percentage of muscovite in the

tributaries increases rapidly southwards and the Nyadi, Khudi and Dordi all contain 5 to 10%

muscovite. The Chepe (S-54) has the largest fraction of muscovite, with over 20% at the basin

mouth. Sample S-41 is taken from the Chepe, at the base of the MCT zone, and shows

approximately 30% muscovite content. This becomes diluted with the addition of quartz and rock

fragments through the lower Himalaya. The two samples in the Darondi show a similar trend of

decreasing muscovite and increasing quartz through the lower Himalaya. However, the percentage

Page 86: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

71

of muscovite at the mouth of the Darondi is approximately 5% compared to the much higher value

of 23% in the Chepe. Samples S-51, S-56 and S-38 are from drainages contained within Lesser

Himalayan rocks. All show quartz contents of over 40%, high fractions of rock fragments, and

muscovite abundances of 0 to 3%. The rock fragments may be a function of the immaturity of the

sediment because the basins containing samples S-38 and S-51 are relatively small in size compared

to the larger tributaries.

As defined by the point counting, the proportion of muscovite in the tributaries is used for the PDF

modeling described above. However, the complete mineral proportions of trunk stream and

tributaries may also be used for an additional calculation of erosion rate. Mixing the sediment of

two rivers together should produce a resulting downstream sample that is representative of the

relative proportions of each of the rivers. In reality, the natural variability of the system and point-

counting errors mean that the data are more complex. Because of this uncertainty, we use a basic

model to examine the general pattern of contributions from each of the inputs and the mixing of the

trunk stream.

The model uses point-counting data from each of the upstream samples and mixes them to

produce a resulting downstream signal. The contribution of each upstream sample is varied from 0

to 100%, and the best solution is picked by finding the minimum residual to the downstream

sample. This procedure assumes ideal mixing and no selective deposition, or attrition, of individual

mineral species. We have already suggested that the carbonate rock fragments are susceptible to

comminution, and hence in some areas they bias the ratios. Therefore, we found that instead of

trying to: a) mix all the species at once using the method described above, or; b) drop out individual

mineral species and recalculate the mixing ratio (with or without renormalizing), the best results

were obtained by calculating the mixing ratios for each mineral species individually and taking the

mean. This eliminated instances where individual mineral species controlled the mixing ratio due to

their large volume (e.g. quartz), and weighted each mineral species equally.

Once mixing ratios were determined, relative erosion rates for individual tributaries within the

Marsyandi were calculated using the catchment areas (Fig. 10). Starting at the top of the basin, the

sediment leaving tributary ‘A’ was calculated assuming an erosion rate of 1unit per unit area per

unit time. The area of tributary A can be measured from a DEM, and the volume of eroded sediment

Page 87: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

72

per unit time may be calculated. Knowing the mixing ratio of A:B, the relative volume of sediment

leaving tributary ‘B’ can therefore be calculated. The resulting volume of sediment leaving tributary

B, divided by the area of B, produces an erosion rate relative to A. For the next junction

downstream, the trunk-stream sample upstream of the junction is taken to have a combined

sediment volume of A plus B. The erosion rate of the tributary ‘C’ can then be calculated in the

same manner as before, using the mixing ratio to find the sediment volume leaving C and dividing

by the area of C. Using this methodology, we can calculate the relative tributary erosion rates for

the whole basin as we work downstream. To produce numbers that are of the same order of

magnitude as erosion rates calculated by the thermochronology, we normalize the relative erosion

rates of a particular basin with the erosion rate derived from thermochronology (in this case using

the Dordi Khola).

The results (Fig. 11) show that, although there are absolute differences between the erosion rates

calculated from thermochronology and those calculated from the point-counting analysis, the

general pattern of low erosion in the north of the region, high erosion rates in basins on the southern

flank of the main topographic axis, and intermediate erosion rates to the south, is the same. Whereas

the PDF modeling suggested variations in erosion rate of up to 2.5 fold, the point-counting model

yields variations in predicted rates of > 10 fold. These discrepancies are not surprising due to the

expected variability of sediment within the river: a) species such as rock fragments will become

comminuted downstream; b) carbonate rock fragments and minerals will experience chemical, as

well as physical, erosion; c) hydraulic sorting will be important for minerals with different densities.

We have shown that the latter process is not too significant for muscovites in the 500 to 2000 µm

size range, however, at least for samples S-52 and S-53.

8.0 Conclusions

In this study we examine the downstream development of a detrital mineral cooling-age signal.

The investigation provides insights into long-standing questions concerning the interpretation of

detrital-mineral ages. In the past a sample collected from the foreland basin might be used to

interpret rapid erosion rates in the upstream catchment area, but where the erosion was occurring

Page 88: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

73

and how much of the basin it represented was unknown. Here, we show that there are systematic

and predictable changes in the detrital cooling-age signal of a large, transverse Himalayan river.

From the analysis of the Marsyandi Valley detrital system, we learn how the inputs from tributary

catchments combine to form the signal at the basin mouth. We can model the cooling-age

distributions from individual tributaries based on observable characteristics of the basin (relief and

hypsometry) to determine a best-fit erosion rate. The numerical modeling indicates that the detrital

cooling-age signal behaves systematically in the trunk stream, whereby the input of an age

population from an individual tributary is a function of the abundance of the thermochronometer per

unit area, the area of the basin, and the rate at which it is eroding. In our analysis, the distribution of

muscovite within the catchment is a critical factor in determining the representation of a particular

area of the basin in the foreland sample. For example, approximately 30% of the Marsyandi

catchment is composed of Tethyan rocks, and yet this represents only a small fraction of the

foreland detrital cooling-age signal because of the paucity of muscovite within the Tibetan zone

lithologies.

Comparison of the detrital-age data versus modeling results indicate that comminution of

muscovite by fluvial processes seems to be insignificant at the 100 to 200 km scale. The 15 to 20

Ma signal derived from the Marsyandi headwaters is persistent downstream, although decreases in

significance. We argue that the calculated volumes of thermochronometer added to the trunk stream

are consistent with the downstream dilution of this signal by younger age populations, rather than

mechanical breakdown of muscovite. This may be attributable to muscovite moving in the

suspended load during the monsoon season, thus encountering less abrasion than during bedload

transportation.

The variability of the cooling-age signal was tested with samples from 1) opposite ends of the

same sandbank, and 2) from the modern river and a fill terrace. Our statistical analysis illustrates

that although there was variability between samples, the sample pairs were not found to be

statistically different at the 95% confidence level: the processes of random selection, from a

common “parent” cooling-age signal, can explain the variation between samples. This indicates

that, for at least these two tests, the detrital cooling-age signal is spatially and temporally reliable in

this area.

Page 89: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

74

The results of the thermochronometry indicate that the topographic front of the Himalaya is

eroding faster than the area around it, generating a ~4 to 10 Ma age signal in those basins draining

the Greater Himalayan sequence Areas to the north of the Himalayan topographic axis, on the edge

of the Tibetan Plateau, experience the lowest erosion rates and produce grain ages between 10 and

15 Ma for those basins sourced in the Tibetan zone, and 15 to 20 Ma for those basins draining the

top of the Greater Himalayan sequence. Intermediate rates are probably found in the lower

Himalaya, although the large tributaries draining this zone have headwaters in, and cooling-age

signals dominated by, the Higher Himalaya. The pattern of erosion from thermochronology is

broadly consistent with the erosion rates calculated from point-counting data. The latter predicts a

greater contrast in the erosion rates that should be interpreted cautiously, however, because of the

intrinsic variability of sediments within the river system.

Two explanations for the overall pattern of the erosion rate are plausible. First, if the MCT and

STDS are active, the high erosion rates of the High Himalaya can be explained by the southwards

extrusion of the Greater Himalaya sequence. Alternatively, if the MCT is relatively inactive, then

the spatial variation in erosion rate is probably a function of the angle of the underlying Main

Himalayan Thrust along which Asia overthrusts India. Investigations of seismicity indicate the

presence of a mid-crustal ramp [Pandey et al., 1995], and increased vertical uplift above a steeper

section, below the main topographic axis, would account for the spatial pattern of erosion rates.

Either of these scenarios introduces the complicating factor of lateral advection. In most

investigations to date, cooling ages are converted into rock exhumation rates using an assumption of

1D thermal and kinematic processes. It is increasingly being recognized that the thermal structure of

an orogen is not a simple model of horizontal isotherms, but is subject to the lateral advection of

rock and heat energy into the system [e.g. Batt and Braun, 1997; Jamieson and Beaumont, 1988;

Jamieson et al., 1998; Willet, 1999]. Rock particles move laterally, not solely vertically, during

erosion and although beyond the scope of this paper, this complicating factor needs to be considered

when assessing the results.

This baseline investigation illustrates that, within the Marsyandi Basin, the detrital cooling-age

signal is a reliable representation of the contributing area and that the downstream evolution of the

trunk-stream signal is understandable, and as a result may be numerically modeled. The

Page 90: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

75

interpretation of the foreland signal, however, is not necessarily simple because the distribution of

cooling ages within the river system is a complex function of the landscape characteristics, erosion

rate, and distribution of thermochronometer within the catchment. For example, the sample at the

basin mouth is dominated by a 4 to 10 Ma population of grain ages and we might argue that those

drainages downstream of sample 6, representing ~40% of the area sampled, control the signal. On

further investigation, however, it seems that they cannot explain all the intricacies of the older

signal. Thus, the area upstream of samples 8 and 9, representing ~55% of the basin area sampled, is

needed to account for the 15 to 20 Ma age population. In this study we would argue that this signal

is so much less dominant in the foreland, despite a larger contributing area, because of the effects of

lithology and erosion rate.

Therefore, while we argue that the distribution of age within the foreland is a good representation

of the upstream area and provides useful and reliable information about the spatial pattern of

erosion rate in the hinterland, some caution is needed when interpreting the source and importance

of age populations from stratigraphic samples. It is apparent that to extract the maximum amount of

information from the stratigraphic record, understanding the parameters that control the modern

cooling-age signal is vital. In this study, we have dated ~500 muscovite grains from 12 locations

that have provided new insights into how lithology and drainage basin characteristics, both easily

observable in other modern locations, influence the detrital signal of the Marsyandi River. Thus,

with some knowledge of the modern signal it should be possible to extract reliable cooling-rate

information from samples within the stratigraphic record, which can subsequently be used to

constrain the temporal and spatial evolution of the orogenic belt.

Page 91: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Pokhara Basin

Tha

kkho

la G

rabe

n

Annapurna II 7939m

Seti Khola

Marsyandi River

Marsyandi/Trisuli junction

MCT

MCT

MCT

Manaslu granite

Dudh

Dona

Nyadi

Dordi

Chepe

Nar

Khansar

Greater Himalayan sequenceLesser Himalayan sequence

Leucogranite (Manaslu Pluton)Tibetan Sedimentary Series (PCam-Ord)Tibetan Sedimentary Series (Sil-Cret)

N

Miyardi

Khudi

Darondi

CDF Manaslu8156m

MDF

Figure 1. Simplified geological map of the Marsyadi region adapted from Hodges et al, [1996], initially from Colchen [1996]. The south-verging Main Central Thrust (MCT) separates the Greater Himalaya sequence from the Lesser Himalaya sequence. Other south-verging thrust faults, the MBT and MFT, are to the south of this diagram. The South Tibetan Detachment system, with normal displacement, forms two splays in this region: the Chame Detachment Fault (CDF) and the Machhapuchhare Detachment Fault (MDF).

76

Page 92: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

POKHARA

N

Tibetan plateau

India

58 mm/yr

17

1612

191021

8/920

6 5

321

4

44

54

41

51

52/5340

56

2437

38

0 20km

MDF

MCT

Sample siteDating site

43a

36

Figure 2. Map of Marsyandi drainage system based on a 90-m DEM. Sample locations are displayed with squares (and gray sample numbers) for point-counting sites and circles (and black sample numbers) for 40Ar/39Ar analysis/point count sites. The Marsyandi drainage (upstream of site 24) is outlined in white. The MCT has triangles indicating south vergence whereas the MDF has half circles indicating down-throw to the north. The MBT/MFT are slightly to the south of this figure. The inset shows the approximate position of the sampling area within Central Asia.

77

Page 93: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

12

0 15105 20 25

0 15105 20 25

0 15105 20 25

0 15105 20 25

0 15105 20 25

0 15105 20 25

0 15105 20 25

0 15105 20 25

0 15105 20 250 15105 20 25

0 15105 20 25

0 15105 20 25Age (My)

Age (My)

Age (My)

Age (My)

n=55

n=35

n=25

n=23

n=33

n=37

n=48

n=23

n=25

n=49

n=45

n=37

Darondi

Dordi

Chepe

Khudi

Nyadi

Nar

Khansar

Dudh Dona

Miyardi

Sample Number

Age (My)

Number of grainsP

roba

bilit

y

n=45

Cooling age PDF

Age (My)

KEY

Trun

k

52+53

12

6

3

2

5

37

54

44

40

24

8+900

N

The Trisuli

Figure 3. Detrital cooling-age PDFs for samples from the Marsyandi drainage. All axes range from 0 to 25 My on the x-axis and have probability on the y-axis. Areas under the PDF curves (shaded black) represent a total probability of one in each case. The individual plots have been arranged topologically to indicate their position within the Marsyandi River system: geographic locations are shown in figure 2. Additional grain ages in sample 24 (229 ± 18 Ma) and sample 54 (27 ± 0.9 Ma) are not illustrated.

78

Page 94: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Yield of thermochronometer per unit area per unit time

Denudation rate

Percentage of thermochronometer

Grain size of thermochronometer

Lithological variation

Trun

k st

ream

age

sig

nal

Tributary age signal contribution

Area

Figure 4. Parameters controlling the contribution of an individual tributary to a trunk-stream cooling-age signal. The foreland signal can be modeled as a specified mix of several such tributaries.

79

Page 95: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 5 10 15 20 25 30

Nar

Khansar

Dudh Dona

Miyardi

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

0 5 10 15 20 25 30

age (my)

Darondi

Dordi

Chepe

Khudi

Nyadi

Trun

k

Sample Number

Age (My)

Pro

babi

lity

Data PDF KEYModeled

PDF

age (my)

age (my)

age (my)

52+53

12

6

3

2

54

44

00

N

The Trisuli

Figure 5. Model results compared to 40Ar/39Ar analyses. Real data PDFs (identical to data used in Fig. 3) are shaded gray and smoothed using a 2-My window. Solid black lines are model PDFs generated using the methodology described in the text, and are also smoothed with a 2-My window. The x-axis of each plot ranges from 0 to 30 My, with probability on the y-axis, and the area under each curve represents a probability of one.

80

8+9

5

37

52+53

24

Page 96: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0

1

2

3

4

5

0 5 10 15 20 25predicted age (My)

eros

ion

rate

(km

/My)

, or u

nit v

olum

eof

sed

imen

t ero

ded

per u

nit t

ime

per u

nit a

rea

Figure 6. Predicted ages from our model for specified erosion rates, Note that changes in the erosion rate between ~0.8 km/My and 2 km/My result in large changes in the predicted ages, which is a function of the isotherm repsonse and the particle speed. Another way to interpret this diagram is to consider the amount of sediment eroded from 1 uinit area of land in one unit time., which is the contribution of each unit land area to a detrital PDF. If we consider a drainage basin containing one half eroding at 2.1 km/My (i), and the other half at ~0.8 km/My (ii), the older grains represent a much smaller signal despite represeting the same amount of land area.

81

(i)

(ii)

Page 97: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 5 10 15 20 250

2

4

6

8

10

12

14

16

18

0

1

2

3

4

5

6

7

8

1-s

erro

r of

gra

in a

ge (

My)

Probability

1-s

erro

r of

gra

in a

ge (

My)

age of grain (Ma)

a)b)

smoothed error PDF

Figure 7. (a) Age versus error (1-s) for analyses with greater than 40% radiogenic 40Ar. No clear relationship between age and error can be seen. Inset (b) shows a PDF generated from the 1-s errors. This PDF is smoothed (solid line) and used in the error determination in the numerical model. It can be seen that most assigned errors will be between 0 and 1.5 My.

82

Page 98: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Age (My)

0 5 10 15 20 25

s-52 smoothed, n=15

s-53 smoothed, n=22

s-52 & s-53 smoothed

prob

abili

ty

0 5 10 15 20 25

s-8 smoothed, n=35

s-9 smoothed, n=10

s-8 & s-9 smoothed

prob

abili

ty

a)

b)

Figure 8. Results of repeat sampling to test: a) the spatial variability of the detrital cooling age signal, and; b) the temporal variation of the signal. For each plot, the shaded PDF is the combined data used in the construction of figure 6, while the PDFs represented by black lines are the contributing data. All curves are smoothed with a 2-My scrolling window.

83

Page 99: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Erosion rate (mm/yr)

Not included in model

N

0.94

1.12

0.94

1.3

1.71.9 2.3

2.1

1.97

2.1

Khansar

Nar

Dudh

Dona

Nyadi

Darondi

ChepeKhudi

MiyardiDordi

MarsyandiRiver

2.0

1.5

1.0

0.0

0.5

84

Figure 9. Spatial variation in erosion rates at the drainage-basin scale. Erosion rates are taken from the results of modeling the detrital cooling age PDFs for individual tributaries. The stippled areas indicate zones not included in the calculations and the dashed gray line indicates the approximate path of the trunk stream. Highest erosion rates occur in the middle areas of the Marsyandi Basin, where rivers drain the steep front of the High Himalayas. Slowest erosion rates occur to the north of the topographic axis, in the rain shadow.

Page 100: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

AB

AB

AB

AB

ABCC

mixing ratio = 1A:2B

mixing ratio = 1AB:1.8C

a=100

a=110

a=140

PROCEDURE:

i) Assumption: basin A is eroding at 1 unit, per unit area, per unit t ime.

ii) In unit t ime, basin A erodes the area multipl ied by the erosion rate, equall ing 100 unit volumes.

i i i ) Given the mixing rat io 1A:2B, basin B must produce 200 uni t volumes in unit t ime.

iv) With an area of 110 units, B must be eroding at 200/110 = 1.8 units, per unit area, per unit t ime.

v) As a result, the total amount of sediment generated by basins A and B combined (at site AB) wil l be 300 unit volumes in unit t ime.

vi) Therefore, given the mixing ratio 1AB:1.8C, basin C must produce 540 unit volumes in unit t ime, and hence is eroding at 540/140 = 3.8 units, per unit area, per unit t ime.

Figure 10. The procedure used to convert point-counting results into relative erosion rates. Point counting sediment samples from the mouth of drainage A and the mouth of drainage B, in conjunction with a downstream sample AB, can be used to calculate a mixing ratio for the two basins. When combined with basin area, extracted from a DEM, this ratio is used to calculate a relative erosion rate.

85

Page 101: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

2.0

1.5

1.0

0.0

Erosion rate (mm/yr)

Not included in model

N

0.04

0.2

0.3

0.82.5

2.1

0.7

1.1

Khansar

Nar

Dudh

Dona

Nyadi

Darondi

ChepeKhudi

MiyardiDordi

MarsyandiRiver

0.5

Figure 11. Spatial variation in erosion rates at the drainage-basin scale. Erosion rates are calculated from the point-counting data using the methodology illustrated in figure 9. Relative erosion rates are normalized by the Dordi Khola to allow direct comparison with figure 8. The stippled areas indicate zones not included in the calculations and the dashed gray line indicates the approximate path of the trunk stream. Note that the overall pattern of erosion is similar to figure 8, with high erosion rates in the middle areas of the Marsyandi Basin, and slowest erosion rates to the north of the topographic axis.

86

Page 102: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

87

Table 1. Point-counting data from the Marsyandi Valley. Abbreviations are as follows: n = number

of grains, qtz = quartz, Pl = plagioclase, Kfs = potassium feldspar, ms = muscovite, Bt = biotite,

frag = rock fragments, opq = opaques, other = any other minerals. All errors are 2-σ and are

calculated with the statistical analysis of Van der Plas and Tobi [1965].

Page 103: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

88

Sample n qtz Pl Kfs ms Bt frag Opq other

S-1 893 43.3 ±3.3 15.4 ±2.4 3.6 ±1.2 5.3 ±1.5 3.7 ±1.3 17.9 ±2.6 0.3 ±0.4 10.4 ±2.0

S-1.2 798 45.6 ±3.5 14.7 ±2.5 4.6 ±1.5 6.5 ±1.7 4.0 ±1.4 18.5 ±2.8 0.1 ±0.2 5.9 ±1.7

S-2 Khudi 790 37.7 ±3.4 16.2 ±2.6 2.3 ±1.1 10.1 ±2.1 9.6 ±2.1 18.5 ±2.8 0.0 ±0.0 5.6 ±1.6

S-2.2 765 42.7 ±3.6 19.6 ±2.9 2.1 ±1.0 11.9 ±2.3 11.2 ±2.3 7.4 ±1.9 0.2 ±0.3 4.3 ±1.5

S-3 802 26.4 ±3.1 19.6 ±2.8 8.4 ±2.0 2.0 ±1.0 1.6 ±0.9 33.9 ±3.3 0.0 ±0.0 7.9 ±1.9

S-3.2 800 30.3 ±3.2 15.8 ±2.6 10.0 ±2.1 2.6 ±1.1 2.4 ±1.1 35.3 ±3.4 0.1 ±0.2 3.4 ±1.3

S-4 691 29.1 ±3.5 13.7 ±2.6 8.7 ±2.1 1.9 ±1.0 3.2 ±1.3 36.2 ±3.7 0.4 ±0.5 6.8 ±1.9

S-5 Nyadi 745 41.2 ±3.6 20.4 ±3.0 8.6 ±2.1 6.3 ±1.8 4.6 ±1.5 12.6 ±2.4 0.3 ±0.4 5.8 ±1.7

S-6 779 27.3 ±3.2 17.1 ±2.7 9.6 ±2.1 3.5 ±1.3 2.6 ±1.1 35.6 ±3.4 0.8 ±0.6 3.6 ±1.3

S-8 747 28.1 ±3.3 21.8 ±3.0 15.4 ±2.6 1.9 ±1.0 0.7 ±0.6 30.7 ±3.4 0.4 ±0.5 1.1 ±0.8

S-10 Dudh 804 40.2 ±3.5 29.9 ±3.2 16.0 ±2.6 1.7 ±0.9 0.0 ±0.0 9.3 ±2.1 0.1 ±0.2 2.7 ±1.1

S-16 Khansar 786 5.6 ±1.6 1.8 ±0.9 1.4 ±0.8 0.1 ±0.2 0.8 ±0.6 79.1 ±2.9 0.9 ±0.7 10.2 ±2.2

S-17 Nar 769 11.8 ±2.3 2.3 ±1.1 1.4 ±0.9 0.5 ±0.5 0.4 ±0.4 79.2 ±2.9 0.7 ±0.6 3.6 ±1.4

S-19 784 15.9 ±2.6 14.0 ±2.5 7.5 ±1.9 1.8 ±0.9 2.0 ±1.0 53.6 ±3.6 0.6 ±0.6 4.5 ±1.5

S-20 808 27.2 ±3.1 21.9 ±2.9 10.3 ±2.1 1.1 ±0.7 1.5 ±0.8 34.9 ±3.4 0.2 ±0.3 2.8 ±1.2

S-21 Dona 777 29.6 ±3.3 21.6 ±3.0 17.4 ±2.7 0.5 ±0.5 2.8 ±1.2 11.1 ±2.3 0.5 ±0.5 16.3 ±2.7

S-24 768 49.1 ±3.6 9.1 ±2.1 3.1 ±1.3 6.8 ±1.8 3.9 ±1.4 23.4 ±3.1 0.5 ±0.5 3.9 ±1.4

S-36 842 43.7 ±3.4 13.2 ±2.3 4.5 ±1.4 5.1 ±1.5 3.7 ±1.3 25.5 ±3.0 0.6 ±0.5 3.7 ±1.3

S-37 Darondi 869 62.7 ±3.3 13.1 ±2.3 1.2 ±0.7 4.6 ±1.4 2.9 ±1.1 12.5 ±2.2 1.0 ±0.7 1.8 ±0.9

S-38 524 44.1 ±4.3 1.1 ±0.9 0.6 ±0.7 0.0 ±0.0 5.2 ±1.9 47.5 ±4.4 0.8 ±0.8 0.8 ±0.8

S-40 624 55.8 ±4.0 13.3 ±2.7 0.6 ±0.6 7.2 ±2.1 3.8 ±1.5 13.8 ±2.8 0.5 ±0.6 4.5 ±1.7

S-41 633 28.6 ±3.6 8.7 ±2.2 0.8 ±0.7 29.7 ±3.6 23.5 ±3.4 2.2 ±1.2 0.2 ±0.3 6.3 ±1.9

S-43a 821 35.7 ±3.3 16.8 ±2.6 10.0 ±2.1 7.9 ±1.9 8.3 ±1.9 16.7 ±2.6 0.6 ±0.5 3.8 ±1.3

S-44 Dordi 777 43.8 ±3.6 14.3 ±2.5 2.7 ±1.2 9.7 ±2.1 9.1 ±2.1 12.6 ±2.4 1.4 ±0.8 6.4 ±1.8

S-51 614 43.3 ±4.0 7.3 ±2.1 4.9 ±1.7 3.1 ±1.4 2.3 ±1.2 34.5 ±3.8 0.2 ±0.3 4.4 ±1.7

S-52 858 31.6 ±3.2 13.6 ±2.3 5.5 ±1.6 10.6 ±2.1 11.0 ±2.1 22.5 ±2.9 0.5 ±0.5 4.8 ±1.5

S-53 897 29.5 ±3.0 10.9 ±2.1 3.7 ±1.3 10.7 ±2.1 16.1 ±2.5 24.2 ±2.9 0.8 ±0.6 4.0 ±1.3

S-54 Chepe 884 40.4 ±3.3 9.3 ±2.0 2.1 ±1.0 23.2 ±2.8 11.9 ±2.2 7.3 ±1.8 0.3 ±0.4 5.4 ±1.5

S-56 598 52.0 ±4.1 9.2 ±2.4 6.9 ±2.1 2.2 ±1.2 3.8 ±1.6 23.2 ±3.5 1.0 ±0.8 1.7 ±1.0

Page 104: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

89

Basin Area (km2) Elevation (m) Lithology

Mean Min Max Relief ~%TSS ~%GH ~%LH

Marsyandi 4760 3332 244 8152 7908

Khansar 713 4794 2634 7824 5190 80 20 -

Nar 884 5209 2634 7097 4463 80 20 -

Dudh 392 4694 1958 7669 5711 10 90 -

Dona 129 4851 1895 8152 6257 100 -

Miyardi 60 4050 1496 5842 4346 - 100 -

Nyadi 215 3440 926 7495 6569 - 80 20

Khudi 136 2565 796 4914 4118 - 90 10

Dordi 351 2885 553 7756 7203 - 70 30

Chepe 309 1808 440 4872 4432 - 50 50

Darondi 609 1470 277 5787 5510 - 40 60

Table 2. Topographic characteristics of the Marsyandi Valley and its associated tributaries. The

approximate litho-tectonic division of the basins are given in percentage area containing: Tethyan

Sedimentary Series (TSS) structurally above the Machhapuchhare detachment fault; Greater

Himalayan sequence (GH) complex, leucogranites, and Annapurna Yellow Formation; Lesser

Himalayan sequence (LH).

Page 105: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Chapter 3

The application of thermal-and-kinematic modeling to constraining rock-particle trajectories, cooling ages of detrital minerals, and the tectonics of the Central Himalaya.

I.D. Brewer and D.W. Burbank

Pennsylvania State University, Department of Geosciences, University Park, Pennsylvania

Abstract

We introduce a new method to use detrital mineral cooling ages in conjunction with a digital

elevation model (DEM) to test numerical models of collisional orogens. We apply this methodology

to the Marsyandi valley, in the central Nepalese Himalaya, where we use a 2-D kinematic-and-

thermal model to predict variations in bedrock cooling ages along an averaged transect within the

Himalaya. The model is based upon a simple ramp-and-flat style decollement, representing the

Main Himalayan Thrust (MHT), and is constrained by the INDEPTH transect, surface geology,

seismicity, and geomorphology. The 2-D kinematic-and-thermal model is extrapolated laterally to

calculate the 3-D distribution of cooling ages predicted for actual Himalayan topography. The

detrital cooling-age signal results from convolving the distribution of cooling ages within a basin,

with the rate at which individual points are eroding, and the distribution of the mineral used as the

thermochronometer. The predicted distributions of cooling ages are compared with detrital 40Ar/39Ar muscovite data to assess varying tectonic scenarios. Model results, assuming that the

Main Boundary Thrust represents the surface expression of the MHT, illustrate that the distribution

of detrital cooling ages is sensitive to how plate convergence is partitioned, with the best-fit model

assigning 5 mm/yr to the motion of the Asian plate and 15 mm/yr to the motion of the Indian plate.

There is, however, a trade off between ramp geometry and convergence rate. A model using the

approximate present position of the Main Central Thrust (MCT) to represent the surface expression

Ian D. Brewer
Ian D. Brewer
90
Page 106: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

91

of the MHT provides a better fit to the observed data and supports evidence for recent activity on

the MCT.

1.0 Introduction

The complexities of the growth and subsequent erosion of orogenic belts have been a focus of

much research in the last decade. The Himalaya, the icon of continental collision, illustrates the

importance of such investigations because the temporal evolution of the orogen has been proposed

as the prominent Cenozoic driving mechanism for strontium and carbon geochemical cycles [Derry

and France-Lanord, 1996; Raymo et al., 1988] and climate change [Kutzbach et al., 1993;

Ruddiman and Kutzbach, 1989]. The development of numerical modeling in geosciences has

greatly increased our insight into how orogenic systems operate [e.g. Beaumont et al., 1992; Koons,

1989; Koons, 1995; Willet, 1999], and while algebraic and numerical descriptions of complex

physical processes are commonly gross simplifications, the results can provide new ideas and

hypotheses to be evaluated with field data.

Such data frequently include bedrock-cooling ages, which are used as a proxy for the erosion rate

in order to measure the strain field. Assumptions are commonly made, however, that limit the

amount of information we can extract from such analyses. Assumptions usually include: 1) vertical

erosion; 2) horizontal isotherms; and 3) an estimated, linear, geothermal gradient. Given the

intricacies of orogenic belts, these assumptions are frequently invalid. Transportation along thrust

faults generally controls deformation, rather than vertical erosion. This lateral rock movement, in

combination with the effects of topography, will deflect isotherms and produce local, non-linear,

geothermal gradients. Thus to extract the maximum amount of geological information from detrital

cooling ages, the complexities of regional tectonics need to be incorporated into numerical

simulations to understand how variations in deformation pathways, convergence rates, topography,

heat production, and lithology influence the spatial distribution of bedrock cooling ages.

Increasingly complex numerical models, however, need to be tested against more comprehensive

sets of field data for corroboration. Thus, in this paper, we focus on the applicability of the modern

detrital cooling-age signal to constraining models of mountain belts. A detrital sample may be more

Page 107: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

92

advantageous than traditional bedrock thermochronology because it is easily collected, separated,

and analyzed, and potentially provides an integration of cooling ages from an entire drainage basin

(which may include a significant portion of a study area when using samples from large transverse

rivers). Opposingly, bedrock cooling ages are restricted to single locations, typically valley floors,

and usually represent a restricted number of samples which are difficult and expensive to collect.

Investigations of individual basins [Brewer et al., Chapter 1; Stock and Montgomery, 1996] and

drainage networks [Brewer et al., Chapter 2] provide evidence that, given some simple

assumptions, the detrital cooling-age signal can be systematically predicted, and detrital ages from

the foreland can yield information about the tectonics of the hinterland. Yet none of these models

account for the increasingly recognized affect of lateral advection of rock mass through the orogen.

Geodynamic models [e.g. Beaumont et al., 1992; Koons, 1989; Koons, 1995; Willet, 1999] have

addressed lateral advection, but do not investigate the detailed distribution of bedrock cooling ages.

Thermal and metamorphic investigations have included lateral advection, but either do not address

the application to cooling ages [Henry et al., 1997], or do not solve for any altitudinal distribution

of ages [Harrison et al., 1998; Jamieson et al., 1998].

In this paper, we have developed a methodology that allows us to combine more complex

geodynamic models with digital elevation models (DEM). We have developed a new methodology

that builds on previous investigations because predicting the spatial distribution of cooling ages

within a landscape, and the resulting detrital signal derived from eroding that bedrock, requires a

more integrated approach. While our thermal model does not have the sophistication of ramp

timing, metamorphism, or melt generation that have been incorporated into previous models

[Harrison et al., 1998; Henry et al., 1997] we have added additional elements in order to predict the

detrital cooling-age signal.

A basic 2-D kinematic-and-thermal geodynamic model, intended to be a simplification of the

Central Nepal Himalaya, predicts the position of the closure isotherm, the distance a rock has to

travel along a path to the surface, and the rate of movement along that path. In conjunction with

digital topography, this 2-D transect is extrapolated into a 3-D model of bedrock cooling ages. The

resulting map of bedrock cooling ages can be manipulated with GIS software to predict the detrital

cooling-age signal from a basin, once corrected for variations in erosion rate and lithology.

Page 108: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

93

Comparison with the modern detrital cooling-age results allows us to constrain the 2-D kinematic-

and-thermal model, thereby providing insights into Himalayan deformation.

Although our area of investigation is the Marsyandi valley in Nepal, the technique of predicting

detrital cooling-age signals from geodynamic models may be applied to other collisional orogens.

Thus using our methodology, it is possible to evaluate tectonic and erosion models using detrital

cooling-age signals derived from orogen-scale drainage basins, which may represent an integration

of perhaps tens to thousands of square kilometers. In addition, when such detrital cooling-age

signals are preserved in the stratigraphic record, they provide quantitative constraints for the

extrapolation of numerical models into the geological record, providing the possibility to investigate

the temporal evolution of mountain belts.

2.0 Geological background

The Himalaya mark the southern boundary of a widespread expression of continental collision

throughout Central Asia. Since collision of India with Asia at 55 ± 5 Ma [Searle, 1996], it has been

postulated that some ~2500 km of subsequent continental convergence has been accommodated by

distributed shortening that is manifested by uplift of the Tibetan plateau [e.g Dewey et al., 1988;

England and McKenzie, 1982], underthrusting of India [e.g Powell and Conaghan, 1973], intra-

continental orogeny [e.g Molnar and Tapponier, 1975], and strike-slip tectonics [e.g. Tapponier et

al., 1986; Tapponier et al., 1982]. Global plate motions calculated from NUVEL-1 [DeMets et al.,

1990] predict a total convergence rate of 58 ± 4 mm/yr between India and Asia, and GPS data

[Bilham et al., 1997] indicate that approximately one third of the total convergence currently occurs

across the main Himalayan chain. This is consistent with spirit-leveling investigations [Jackson and

Bilham, 1994] which predicts that convergence results in vertical-uplift rates of up to 7 ± 3 mm/yr

over the topographic divide of the High Himalaya.

Our study area is located in central Nepal, where the Marsyandi River drains the southern edge of

the Tibetan Plateau before flowing south through the main Himalayan chain (Fig. 1). The Tibetan

zone is characterized by a sequence of lower Paleozoic to lower Tertiary marine sediments [Le Fort,

1975]. These are bound in the south by the South Tibetan Detachment System (STDS), which is a

Page 109: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

94

down-to-the-north normal fault. In the Marsyandi valley, the STDS comprises two detachments that

juxtapose the unmetamorphosed Tethyan carbonates and mudstones of South Tibet against the

Greater Himlaya sequence, with an intervening greenschist-grade marble, the Annapurna Yellow

Formation [Coleman, 1996].

The Greater Himalaya sequence commonly forms the topographic divide and comprises kyanite-

to-sillimanite grade metasedimentary and metaigneous rocks of Neo-Proterozoic [Parrish and

Hodges, 1996] to Cambrian-Ordovician age [Ferra et al., 1983]. Anatectic melting within the

Greater Himalaya, commonly of the lower kyanite-grade schists of Formation I [Barbey et al.,

1996; Harris and Massey, 1994], produces leucogranites that intrude the top of the sequence. In the

study area, the crystallization of the Manaslu leucogranite has been dated at 22.4 ± 0.5 Ma using 232Th/208Pb in monazite [Harrison et al., 1995], and contains an older inherited-Pb monazite

population of ~600 Ma [Copeland et al., 1988]. The Manaslu granite typically yields 40Ar/39Ar

muscovite plateau dates of ~17 to 18 Ma [Coleman and Hodges, 1995; Copeland et al., 1990].

The south-vergent Main Central Thrust zone (MCT) forms the base of the Greater Himalaya

sequence and overthrusts the Lesser Himalayan sequence. The MCT has experienced a poly-phase

history. Earliest motions were synchronous with the regional metamorphism of the Greater

Himalaya sequence, 20 to 23 Ma [Hodges et al., 1996]. Detrital muscovite cooling ages from

tributaries within the Marsyandi valley are typically dominated by populations of 4 to 10 Ma grains

[Brewer et al., Chapter 2] and muscovites collected from the Main Central thrust zone yield dates

of 6.2 ± 0.2 Ma and 2.6 ± 0.1 Ma [Edwards, 1995]. Such cooling ages have been used to argue late-

stage deformation of the MCT [MacFarlane et al., 1992], which this is supported by Th-Pb

microprobe analyses on syn-kinematic monazites [Harrison et al., 1997], although hydrothermal

alteration of the thrust zone has also been proposed [Copeland et al., 1991].

The Lesser Himalayan sequence is predominantly greenschist-grade metasediments [Colchen et

al., 1986] that are Mesoproterozoic to Early Cambrian in age [see Hodges, 2000 for review]. The

southern limit is bound by the south-vergent Main Boundary Thrust (MBT). Movement on the MBT

may have initiated between 9 and 11 Ma [Meigs et al., 1995]. The most recent movement is difficult

to constrain, but must be younger than early Pliocene [DeCelles et al., 1998]. The Main Frontal

Thrust (MFT) represents the distal limit of Himalayan deformation in the foreland, and has an

Page 110: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

95

estimated shortening rate of 21 ± 1.5 mm/yr over the Holocene, indicating that it is currently the

major active fault [Lave and Avouac, 2000].

3.0 Thermal and Kinematic modeling

As opposed to simple vertical motion of rocks, lateral advection during continental collision often

represents the dominant component of the deformation field. Yet within the geochronological

community, cooling ages have typically been interpreted in one dimension, with erosion rates

calculated assuming that the rock column is moving vertically towards the surface. In a recent

study, Harrison et al. [1998] use 2-D kinematic-and-thermal modeling to investigate anatexis and

metamorphism in the Central Himalaya. Thermochronology data can be compared against their

model because bedrock ages are predicted by tracing particle trajectories through the orogen.

In this paper we adopt a similar basic approach, and present a 2-D kinematic-and-thermal model to

determine the depth of the closure temperature and calculate the path of rock particles through an

orogenic transect. To do this we must first define the decollement geometry within the Himalaya

(section 3.1) and secondly specify the thermal characteristics of the orogen (section 3.2). With an

appropriate kinematic-and-thermal model, we can extrapolate the 2-D solution along strike to

predict the 3-D spatial distribution of bedrock cooling ages over the entire landscape (Fig. 2).

Correcting for the volumetric contributions of thermochronometer (section 3.3), GIS software

allows us to use the resulting “age maps” to extract the modern detrital cooling-age signal. We can

compare the predicted cooling-age signal from a number of different scenarios to the observed

detrital cooling ages of Brewer et al. [Chapter 2] to assess which parameters are consistent with the

data and to test the sensitivity of the model to variations in these parameters.

3.1. Constraints on thrust geometry

In an active orogen, a bedrock cooling age represents the time elapsed since a rock particle passed

through the closure temperature and subsequently reached the surface. The distance traveled in this

time, divided by the cooling age, is a proxy for the time-averaged erosion rate. Therefore, in order

Page 111: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

96

to predict a bedrock cooling age we need to know: (1) the position of the closure isotherm; (2) the

particle trajectory, and; (3) the rate of particle transport along this trajectory. Due to the strong

dependence of the thermal structure on the rate of rock advection [Brewer et al., Chapter 1;

Mancktelow and Grasemann, 1997; Stüwe et al., 1994], the kinematic structure of the mountain belt

becomes the primary parameter to constrain. By combining heat production with the velocity (speed

and direction) of particles through the orogen, we can model thermal conditions, and more

specifically, the position of the closure isotherm.

For the purposes of modeling, we use a number of simplifications and assumptions to determine

the kinematic structure. First, we use a 2-D model. As a consequence of a scarcity of accurate

subsurface structural data, we assume that it is possible to extrapolate geometrical constraints along

strike. Given the remarkable lateral continuity in the overall structure of the Himalayan orogen (the

major thrust faults can be traced laterally within Nepal and over much of the 2000-km-long range

front), a 2-D approximation is reasonable for the along-strike scale of 100 to 200 km in our model.

Second, although we recognize that the Himalaya has a complex structural architecture, we use

one major decollement to represent the kinematics of the collisional belt. It has been proposed that a

major plate-scale decollement, the Main Himalayan Thrust (MHT), underlies the structure of the

Himalaya [Seeber et al., 1981]. Surface faults are interpreted to sole out into this decollement, with

individual thrusts representing progressively less net displacement from north to south as structures

become progressively younger. Because our intent is to show the major kinematic characteristics of

the orogen, rather than duplicating local complexities, we adopt this orogenic-scale decollement

model. A similar approach was taken by Henry et al. [1997], who modeled the 2-D thermal

structure of the Himalaya using a single crustal-scale decollement dipping at 10° northwards from

the surface outcrop of the MBT. As outlined below, we use a more complex decollement that has 3

dip domains (Figs. 2 & 3), with kink-band folding in the overlying thrust sheet, to mimic the

assumed large-scale structure of the orogen.

With our simplified structure, the MBT represents the surface expression of the MHT (Fig. 3, a).

The average position of the MBT in the study area is constrained by the geomorphic expression of

the southern edge of the Lesser Himalaya, taken from the 90-m DEM. This appears in a distinctly

different location, on the DEM, from Siwalik deformation to the south associated with the MFT. In

our model the MHT dips shallowly 5° - 6° beneath the Lesser Himalaya (Fig. 3, c). This is

Page 112: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

97

consistent with geological sections [Schelling, 1992], borehole data [Mathur and Kohli, 1964], and

leveling data [Jackson et al., 1992].

From the shallow Lesser Himalayan decollement, we use a 15 to 20° mid-crustal ramp that dips

beneath the High Himalaya (Fig. 3) because a number of sources of evidence suggest that the MHT

steepens beneath the topographic front. Foliation planes in the northern Lesser Himalaya steepen

northwards [Schelling and Arita, 1991] and are interpreted to represent a transition from a flat to

ramp geometry under the topographic front. A cluster of seismicity between 5 and 20 km depth,

centered approximately 80 km north of the MFT (Fig. 3, e) [e.g. Ni and Barazangi, 1984], has

interpreted to represent the stress release on this steeper section of fault [Pandey et al., 1995]. Spirit

leveling indicates short-term rock uplift rates of 4 to 6 mm/yr occur over 40-km wavelengths in the

Higher Himalaya, where 2-D dislocation modeling indicates that this can result from strain

accumulation above a steeper section on a deep decollement [Jackson et al., 1992; Jackson and

Bilham, 1994]. In addition, gravity investigations indicate that the Moho dips at 15-20° under the

topographic front of the Himalayas, suggesting a sharp bend in the Indian Plate [Lyon-Caen and

Molnar, 1983].

The position of the MHT to the north of the mid-crustal ramp is constrained by the INDEPTH

seismic profile. INDEPTH imaged a major reflector dipping northward under southern Tibet that

was interpreted to be the MHT, separating Indian plate from the overthrusting Eurasian plate

[Brown et al., 1996; Nelson et al., 1996]. The ramp is at a depth of 9 seconds (TWT) or

approximately 30-km depth beneath the STDS (figure 3, g). The ramp continues for 65 km to the

north where the reflector disappears, at a depth of approximately 35 km below the surface, under

the southern edge of the Kangmar Dome and ultimately the Yamdrok-Damxung reflector (YDR).

The YDR is thought to represent a partial-melt zone beneath southern Tibet formed by crustal

thickening with resulting anatexis, and the Kangmar Dome has been interpreted as a basement-

cored uplift structure [Nelson et al., 1996].

Because most of the constraints on the deep structure of the Himalaya have to be projected along

strike to the Marsyandi valley area, we use the geomorphology of the Marsyandi valley area to aid

the specific location of ramp-angle domains. Whereas the assumption of a landscape in steady state

is necessary for the thermal model, it is not a prerequisite for this analysis. We simply assume that

the topography reflects a balance between the rock-uplift rate (with respect to the geoid) and the

Page 113: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

98

erosion rate. For illustration, steep slopes and high relief provide maximum potential energy for

erosion processes, but to maintain such a landscape undergoing rapid erosion, the rock influx must

also be rapid.

We use a swath ~130 km long by 0.6 km wide, orientated with the long axis parallel to the strike of

the orogen (Fig. 1), to extract averaged topographic characteristics of the study area from a

smoothed 90-m DEM, and from a slope map with values calculated using a 180 m by 180 m area.

Topographic envelopes in figure 3 illustrate the minimum elevation and maximum elevation within

the swath, and hence represent the along-strike relief within the study area. The minimum elevation

essentially represents the river profiles, and we interpret the relief to be a proxy for erosion rates

[e.g. Pinet and Souriau, 1988]. The mean elevation is the statistical average of all the cell values

contained within the swath at each location along the transect.

We have divided the study area into four regions based on the slope (Fig. 3, top panel) and elevation

characteristics (Fig. 3, bottom panel). In the north, where the rock flux originates, we observe low

hillslope angles (Fig. 3, iv) and low relief (Fig. 3, i) on the edge of the Tibetan Plateau. Despite the

high elevations of the Tibetan Plateau, this region is indicative of low erosion rates. In our model,

material is advected laterally and we use a hypothetical flat decollement (Fig. 3, j) to represent this

section where the MHT disappears under the partial melt zone of South Tibet. To the south we

observe a zone of increasing relief (Fig. 3, f) and increasing slopes (Fig. 3, iii) that form the

headwaters of Himalayan transverse rivers. This is a zone of higher erosion that we represent by a

shallow ~4° ramp section.

As material moves still further south, the effects of the steep MHT section, underlying the High

Himalaya, are observed. As the material is transported over the ramp inflexion point (Fig. 3, g), the

relief increases markedly (Fig. 3, d) and the slopes are uniformly high (Fig. 3, ii). Tethyan

sedimentary rocks become progressively incised as material moves at 4-5° up the MHT before

becoming rapidly stripped off to the south of the inflection point, as the steep ramp section is

encountered. We interpret the high relief (up to 6 km) and high slopes to the north to indicate

hillslopes at the threshold angle for landslide failure [Burbank et al., 1996], representing high

erosion rates due to the increase in the vertical advection component above the steep ramp. The

tectonic forcing is rapid enough above the steep ramp to maintain the hillslopes at the limit of rock

strength, and the landscape erodes rapidly. The minimum elevations show a rapid increase

Page 114: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

99

northwards, above the MHT ramp (figure 3, d), suggesting that the rivers increase grade in response

to an increased rock uplift rate [Lave and Avouac, in review; Seeber and Gornitz, 1983; Whipple

and Tucker, 1999].

Once material passes over the steep section on the MHT, and onto the ~5° flat beneath the Lesser

Himalaya (Fig. 3, c), the landscape has lower relief (Fig. 3, b) and lower hillslopes (Fig. 3, i). We

interpret this change to represent a new balance between erosion processes and rock uplift; due to

the decrease in vertical rock-uplift rates, as lateral advection becomes more dominant, less potential

energy per unit time is added to the landscape by rock uplift. In addition to lower erosion rates, the

average slopes in this zone also decrease southwards due to sediment ponding against uplifting

structures in the south, superficially covering valley floors.

3.2. Thermal model

The geometric structure provides a kinematic framework for the model when coupled with a

convergence rate, but the thermal parameters have to be specified before we can predict the depth of

the closure isotherm. Our thermal model has three main components: (1) the thermal properties

assigned to each thermo-lithological package (radioactive heat production, conductivity, and

diffusivity) and the geometry of that package; (2) the kinematics, controlled by the decollement

geometry; (3) a shear-heating term that represents frictional heating on the main decollement.

Because the results of the model are compared against the 40Ar/39Ar analysis on muscovite from

Brewer et al. [Chapter 2], the closure temperature of interest is considered to be the 350°C isotherm

in this model.

The parameterization of our thermal model (Fig. 3) closely resembles that of Henry et al. [1997],

with a thermally inhomogeneous crust underlain by mantle characterized by negligible heat

production and a conductivity of 3.0 Wm-1K-1 [Schatz and Simmons, 1972]. The Indian crust is

bilayered with a 15-km-thick upper crust with heat production of 2.5 µWm-3, and a 25-km-thick

lower crust with heat production of 0.4 µWm-3 [Pinet, 1992]. With our kinematic model, the Lesser

Himalaya and the Greater Himalaya sequence function as a single tectonic unit, and are assigned a

heat production of 2.5 µWm-3 because of high concentrations of radioactive elements [England et

Page 115: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

100

al., 1992]. The thickness of the Greater Himalaya sequence varies laterally within the Marsyandi

study area, perhaps due to STDS normal faulting at the top of the slab, which is not included in this

model. Therefore we have decided to use a thickness of 22 km for the Greater Himalaya sequence

that is consistent with the INDEPTH geological section, measured from the MHT to the STDS

[Nelson et al., 1996]. In our model, the upper boundary of the Greater Himalaya sequence results in

Tethyan rocks cropping out on the highest peaks, which matches the geology of the range [Colchen

et al., 1986]. Due to the normal faulting and lateral thickness variations, however, the thickness of

22 km is simply considered a thermal parameter for the model, rather than an accurate predictor of

the surface outcrop of Tethyan sediments in the Marsyandi valley. The Tethyan sediments are

assigned heat production of 0.4 µWm-3, because it seems reasonable that they contain a lower

abundance of radioactive isotopes than the Greater Himalaya sequence. Crustal conductivity is set

uniformly to 2.5 Wm-1K-1, and the thermal diffusivity to 0.8 µm2s-1 throughout.

The surface boundary condition is set to 0°C, with the morphology of the interface determined by

the mean elevation calculated by averaging a smoothed 90-m DEM over the previously described

swath (Fig. 1). Due to our 2-D approach, we are ignoring the cooling effects of the relief along the

strike of the orogen. In addition, when extrapolating the thermal model laterally, we are assuming

that there will be no significant deflection of the isotherm by local topography. The basal boundary

is set to a constant mantle heat flow of 15 mWm-2 that is consistent with heat flow values from

Precambrian cratons [Gupta, 1993]. Lateral boundary conditions are set to a constant geothermal

gradient on boundaries with influx of rock mass into model space, whereas they are no heat flow

boundaries if there is a net loss of mass from the system.

A shear-heating term, described by Henry et al. [1997], is used to account for frictional heating

along the basal decollement fault. Two terms are needed to describe the shear heating in both the

brittle and the ductile regime. Heat production is a function of the shear stress and strain rate. Shear

stress is calculated as the minimum of a brittle lithostatic pressure-dependant law (1/10th the

lithostatic pressure) or a ductile temperature-dependant power law. Parameters for the ductile power

law are taken from the moderate friction flow law of Hansen and Carter [1982]. In the ductile

regime the fault zone is 1000-m wide and undergoes uniform strain and heating. This model

Page 116: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

101

predicts that the brittle-to-ductile transition occurs at ~420°C in the undeformed Indian Plate (see

discussion in Henry et al.[1997]).

The initial starting condition is set by calculating a geothermal gradient [Pollack, 1965] for the

thermal structure described above undergoing no advection of heat. The 2-D finite difference

algorithm of Fletcher [1991] is used to calculate the thermal structure after ~20 My of advection of

rock mass through the orogen. This should approximate the steady-state solution, within the area of

interest, given the thermal response times of the 350°C isotherm [Brewer et al., Chapter 1]. They

show that from initial static conditions, 90-95% of the steady-state solution, for a crustal column

undergoing vertical erosion from depths of 35 km, at rates of 0.1 to 3.0 km/My, is obtained in 10

My. The morphology of the surface boundary in our model is not time dependant, and hence there is

an implicit assumption of steady-state conditions, whereby the mean elevation of the study area is

invariant over timescales of 10 to 20 My and across spatial scales of 100-200 km. This implies that,

the rock influx into the orogenic front is necessarily balanced by the denudation rate over these

timescales.

3.3. Particle trajectories and detrital cooling-age signals

Based on the specified geometric architecture, and given a specified convergence rate, we can find a

solution for the steady-state thermal structure of the orogenic belt (Fig. 4a). The thermal steady state

in the overthrusting plate is a balance between three competing processes. The underthrusting plate

comprises relatively cold material and so cools the overthrusting plate from beneath, resulting in a

downwards deflection of the closure isotherm. In contrast, tectonics and erosion advect hot material

into the overthrusting wedge, thus heating the system and moving the closure isotherm towards the

surface. Counteracting this, from the top of the overthrusting plate, conductive heat loss to the

atmosphere cools the orogenic front.

Once the unique steady-state thermal structure has been defined for a particular ramp geometry and

convergence rate, we can use the kinematic framework to predict cooling ages in the overthrusting

plate. From the geometry of the underlying ramp the velocity vector for each point within the

transect can be calculated, and the trajectories of rock particles traced through the orogen. The

Page 117: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

102

distance traveled by a particle between passing through the argon closure temperature for muscovite

(~350°C) and reaching the surface (Fig. 4b) can be converted into a cooling age by dividing by the

velocity of the particle into the orogenic front, which is considered to be in topographic steady state.

Because the cooling age for each location in the landscape is a function of the distance that the

particle has had to travel since passing though the closure temperature, mountain summits will have

older cooling ages than valley floors (Fig. 4c). To predict the cooling ages as a function of

landscape position, we use the kinematic-and-thermal model in conjunction with digital topography

to integrate modern topographic parameters into our analysis: for a particular location in the

transect, we measure the distance along the trajectory from the modeled closure temperature to the

actual elevation of that point. Furthermore, if we assume that the kinematic geometry remains

invariant across the width of the study area, the two-dimensional thermal structure can be

extrapolated laterally. With GIS software, we crop and rotate the DEM so that the Y direction is

normal to the strike of the orogen (Fig. 1). The column of Y cells at each value of X can be treated

as an individual transect and a predicted cooling age calculated at each point along the section (e.g.

Fig. 4b). The areal combination of individual columns creates a “cooling-age map” that predicts the

cooling age for every point in the modern landscape. Figure 2 illustrates such an age map draped

over the modern topography.

Now that the distribution of cooling ages within the landscape can be calculated, we want to

model how this is manifest in the sediments eroding off the orogen. The detrital cooling-age signal

is not a simple function of the areal distribution of ages, because the relative proportion of grains of

a certain cooling-age fraction depends upon both: a) the fraction of land with that cooling age, and;

b) how fast that fraction is eroding. The former is calculated with the kinematic-and-thermal model,

but the latter still needs to be determined. Thus we need to calculate the erosion rate in order to

determine the relative contribution of ages from different parts of the topography.

One approach to calculating denudation rates is to simply calculate the vertical erosion

component, which is a common practice in conventional thermochronometry (where erosion rates

are calculated using the cooling age in conjunction with an assumed geothermal gradient).

However, with the effects of rapid lateral advection, the erosion rate calculated with the vertical

component of the particle velocity commonly underestimates the total amount of denudation.

Page 118: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

103

To compute the total volume of material eroded from a mountain front, both the 2-D velocity field

and the aspect of the topography in relation to it have to be considered. For illustration, we can

consider erosion on the edge of a plateau that is in topographic steady state (Fig. 5). Deformation

occurs as material advects laterally towards the plateau margin, along an underlying decollement,

before passing onto a steeper ramp. With this scenario, the highest erosion rates occur where the

aspect of the topography is normal to the particle trajectory (Fig. 5, i), whereas the lowest erosion

rates occur when the topography is closer to parallel with the velocity field (Fig. 5, iii). The contrast

in erosion rate can be explained by the primarily lateral movement, rather than erosion, of material

until the thrust sheet intercepts the kink bend, at which time it experiences rapid uplift over the

thrust ramp, and subsequent denudation at the topographic front. The geometry clearly illustrates

the difference between the underestimated vertical erosion rate (Fig. 5, (ii)) and the rate of erosion

perpendicular to the transport vector (Fig. 5, (i)). This illustration may be analogous to the

topographic axis of the Himalaya at present: as Tethyan sedimentary rocks of the Tibetan Plateau

move through the kink bend, caused by the ramp in the underlying MHT, they become rapidly

uplifted to cap the highest peaks before being eroded.

Given this relationship, we can calculate the volume of rock eroded for a DEM with cell dimensions

X by Y for direct application to the digital topography (Fig. 6). The expression relates the volume of

rock eroded (V) in time (dt) to the topographic slope and particle velocity in the plane of Y:

XdtvYdtV ).sin(...cos

)( βα

=

(1)

The values β and α are dependant upon the relationship between the topographic surface slope (S),

and the underlying ramp angle, which varies between 0° and 90° and determines (a). If S + a < 90

then β = (S + a) and α = s. If S + a > 90 then β = 180 - (S + a) and α = s. If S is larger than 90, then

a volume can only be calculated if a > (180 - S) in which case β = a – (180 – S) and α = 270 – s,

otherwise the calculated volume becomes negative because the particle trajectory is directed into the

slope, as opposed to out of it. Both angles are measured positively as illustrated in figure 6, and the

Page 119: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

104

equation assumes that the strike of the topography (parallel to the direction of X) is perpendicular to

the particle transport direction.

Using equation 1, the effects of topographic slope and particle trajectory on the volume of material

eroded can be predicted for a grid cell of unit area, undergoing unit erosion for unit time (Fig. 7).

With vertical erosion (particle trajectory angle = 90°), there is unit erosion independent of the

surface slope. Highest erosion rates, per unit horizontal area, for slopes of 30° to 40° (as is typical in

the Himalaya) occur with particle trajectories of 50° to 60°, and the lowest erosion rates occur as

particle trajectories approach 0°.

For our purposes, topographic steady state means that the spatially averaged characteristics of the

landscape (hypsometry, slope distributions, along-strike averaged morphology) do not vary over the

interval of interest, which is several millions of years in this study. For the thermal modeling, we

use the mean elevation as our topographic surface boundary. Thus if the average topography is

invariant, the steady-state thermal structure will be in equilibrium. However, an exact topographic

steady state, whereby the influx of rock into every point in the landscape is exactly balanced by the

erosional flux out of that point, is problematic; it is clearly unrealistic on small-spatial and short-

temporal scale because minor climatic cycles will cause high-frequency variations in the erosion

rate through time, and because erosion processes are intrinsically variable over short distances.

Providing the time-averaged topographic steady state is maintained, small perturbations will not

affect the thermal structure because it has a much slower response time. However, a cell-by-cell

calculation of erosion rates (e.g. equation 1) over wavelengths of 90 m, based on the topography at

the snapshot in time when the DEM image was produced, will be strongly influenced by local

topographic variations.

To minimize the effects of local, short-term deviations from the average topography, we therefore

divide the topography into three zones (Fig. 8) based upon the overall pattern of mean elevation: the

Tibetan Plateau; the Himalayan front, and; the Lesser Himalaya. The aspects of these strike-parallel

zones, with respect to the plate velocity vector, are then used to determine an “erosion-rate map” for

the study area. Each cell within a zone erodes an equal volume of material for the same particle

trajectory angle, which in reality will vary depending upon the predetermined underlying ramp

geometry. The result of this approach is that predicted erosion rates vary as orogen-parallel swaths,

Page 120: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

105

rather than on the cell-to-cell scale of a DEM, which is more consistent with the idea of steady-state

topography as discussed above.

We now have a predicted cooling age and erosion rate for each point in the landscape. In order to

predict the distribution of detrital ages produced by erosion of that landscape, we simply need to

combine the volumetric contribution of a predicted age from each one of these locations. We

represent this as a probability density function (PDF), which represents the probability of a

particular cooling-age being found in the sediment and is equivalent to the theoretical PDF of

[Brewer et al., Chapter 1]. The theoretical PDF for a DEM matrix containing x by y cells is

constructed using equation 2:

∑ ∑=

=

=

=

=

xX

X

yY

Ya YX

dtdvYXaP

0 0),().,()( τ

[2]

where the value of τ has to be computed for each grid cell location (X,Y), for each value of Pa(a): if

the cooling age of a cell (ac) is equal to a, then τ = 1, else τ = 0. Pa is the probability of dating a

grain of a particular age (a) and dv/dt is the volume of material a grid cell contributes per unit time.

Once the area under the resulting curve is normalized to unity, the theoretical PDF describing the

distribution of cooling ages is constructed. In this paper, we apply a 0.5 My smoothing function to

the PDFs to minimize the effects of small perturbations. This approach assumes a steady-state

topography, no sediment storage within the catchment, and that muscovite undergoes no mechanical

comminution within the fluvial system. Within mountain belts undergoing high erosion rates,

significant sediment storage is typically not observed, and over the hinterland-to-foreland spatial

scale, Brewer et al. [Chapter 2] argued that the latter process was negligible in the Marsyandi given

the uncertainty in the analysis of 40Ar/39Ar ages.

Page 121: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

106

4.0 Modeling Results

With the kinematic-and-thermal model, we can now predict the distribution of detrital cooling

ages derived from the erosion of the study area. Two groups of models were run to investigate

relationships between the predicted cooling-age signal and geological parameters. For example, we

can ask: “How are variations in the rate of overthrusting of Asia and underthrusting of India

manifest in the detrital cooling-age signal?” The first set of models examines this question, and the

subsequent comparison of predicted cooling-age signals to actual detrital 40Ar/39Ar age data

[Brewer et al., Chapter 2] allow us to place constraints on these parameters. Likewise, we evaluate

how sensitive the detrital cooling-age signal is to different ramp geometries. We have seen that the

position of the closure isotherm in the overthrusting plate is dependant upon the interaction

between: 1) cooling by the underthrusting plate; 2) heating due to erosion and subsequent lateral

advection of hot material into the system, and; 3) conductive heat loss to the atmosphere (Fig. 4a).

Thus, variations in the ramp geometry underlying the Himalaya should result in different positions

of the closure isotherm, and this should be evident in the predicted distribution of bedrock, and

therefore detrital, cooling ages.

For the second set of models, we examine the effects of drainage basin area and lithological

factors. We want to address the questions: “Is a cooling-age signal from a large, transverse river

representative of the orogen, and to what extent does lithology modify the results?” To investigate

this, we model the detrital signal contributed solely from the modern Marsyandi catchment and

compare the results to the observed data [Brewer et al. Chapter 2]. In order to integrate across the

broadest area possible, we use the most downstream sample collected that we consider to be the best

proxy for the distribution of cooling ages deposited in the foreland. The sand sample was collected

from the modern channel of the Marsyandi River, upstream of the confluence with the Trisuli River

(Fig. 1). Muscovites were separated and 55 grains analyzed at the 40Ar/39Ar laser microprobe

facility at the Massachusetts Institute of Technology.

The 40Ar/39Ar data (Fig. 10) are presented as a summed probability density function (SPDF)

comprising the normalized summation of individual grain PDFs, which in turn represent each age

with a Gaussian distributed analytical error [Deino and Potts, 1992]. Note that the absolute value of

probability displayed on the y-axis of our probability plots is dependent upon the age-bin size

Page 122: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

107

chosen along the x-axis. The age-bin size is constant for all plots and the numbers are retained to

allow the direct comparison of relative probability between plots, and between the data SPDF and

the theoretical PDFs generated from our modeling. The data shows a broad 4 to 8 Ma young age

population that dominates the signal. Secondary 10 to 15 Ma and 15 to 20 Ma populations are also

evident.

4.1 Kinematics

4.1.1 Convergence rates

GPS studies indicate that the convergence rate of India with southern Tibet is 20.5 ± 2 mm/yr

[Bilham et al., 1997]. This is typically considered to be the rate of underthrusting of India beneath

Asia, but is actually a more complex interaction of tectonics and erosion. For this model the

intersection of the MHT decollement plane with surface topography (Decollement/Surface

Singularity (DSS), figure 3) is our reference frame, because this theoretical point is independent of

how total convergence is partitioned between the two plates. To illustrate the effects of partitioning,

we consider three different scenarios: 1) with India fixed, the convergence rate has to be solely

accommodated by overthrusting of Asia at 20 mm/yr with respect to the DSS (Fig. 9a); 2) with Asia

fixed, convergence has to be accommodated by Indian underthrusting (Fig. 9b; and 3) with equal

division of the 20 mm/yr of convergence between the two plates, the Indian Plate moves at 10

mm/yr northwards, and the Eurasian Plate moves at 10 mm/yr southwards towards the DSS (Fig.

9c). In all these cases, the overall slip rate on the MHT remains a constant 20 mm/yr. The

partitioning of convergence will affect the thermal and velocity structure of the system, and so

determine modeled cooling ages. Because other parameters are reasonably well constrained by the

geology, even if they have to be extrapolated along strike, the partitioning of total convergence

between India and southern Tibet becomes a primary unknown variable. Given this, we used the

model constraints illustrated in figure 2 to examine the sensitivity of the detrital signal, to the

relative convergence rates.

Page 123: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

108

The results from a number of different simulations illustrate that the detrital cooling ages are very

sensitive to the relative partitioning of convergence, especially at geologically reasonable rates (Fig.

10). Convergence rates of greater than 10 km/My of Asia, with respect to the DSS, result in very

young age populations with peak probabilities representing ages of < 3 My. Harrison et al. [1998]

use equal convergence rates in their model that requires 10 km/My of erosion perpendicular to the

particle trajectory (5 km/My of vertical erosion above a 30-degree ramp) to maintain a steady-state

topographic condition.

The detrital cooling-age signal is more sensitive to slow convergence rates of Asia with respect to

the DSS (Fig. 10). With 4 km/My of convergence, the peak probability occurs between 5 and 10

My, whereas if the convergence rate decreases to 2 km/My, the peak probability shifts markedly to

20 to 25 My. The general trend also indicates that slowing the convergence rate tends to broaden the

range of ages predicted, both for the peak probability and the overall age signal; the older age ‘tails’

lengthen significantly with slower rates (Fig. 10).

In order to assess the appropriate partitioning, we compare the predicted age distribution for

different convergence rates with the observed age distribution for the Marsyandi River. An

important result from this study is that the most likely range of Asia-to-DSS convergence rates is

between 4 and 6 km/My, because the predicted detrital ages fall within the approximate range of

peak probability in the data (indicated by the shaded band of ages on Fig. 10). This is better

illustrated with a direct comparison to the real data for scenarios with Asia-to-DSS convergence

rates of 4, 5, and 6 km/My (Fig. 11, a). The best fit of model to data is generated using 5 km/My of

Asia-to-DSS convergence. Faster convergence produces a younger peak probability (~3 My) that

matches the youngest observed ages well, but the older tail stops at ~17 My, whereas the actual data

contains a peak at 17 to 22 Ma. Slower Asia-to-DSS convergence rates match the older population

in the peak probability of the data, but under predict the young ages. In addition, the older-age tail

stretches out to 30 My, whereas no grains of this age are observed in the data. The best-fit predicted

PDF, using a 5-km/My Asia-to-DSS convergence rate, matches the average population of the peak

probability observed in the data, and fits the older tail well. However, the width of the observed

peak probability, as well as observed ages < 4 Ma and between 7 and 9 Ma, are underrepresented by

the model.

Page 124: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

109

A general explanation for variations of the detrital cooling ages, caused by different partitioning of

the convergence rate, can be understood within the context of the kinematic-and-thermal system.

Increasing the rate of convergence of Asia (Fig. 12) with respect to the DSS produces younger age

populations by: a) increasing the rate at which rock moves between the blocking temperature and

the surface, and; b) deflecting the closure isotherm towards the surface by advection of hot rock

mass (Fig. 12a). In contrast, increasing the rate of India-to-DSS convergence cools the system by

faster subduction of relatively cold continental plate (Fig. 12b), subduing the effects of warming

due to a thick crust and rapid lateral advection in the overthrusting plate.

4.1.2 Angle of Main Himalayan Thrust ramp

Whereas the kinematic model is a simplification of the complex geology of the Himalaya, it is

useful to examine the sensitivity of the model to other potentially important parameters, such as

possible variations in ramp geometry. The position of the MBT is constrained by surface geology,

and the depth seismicity at the top of the MHT ramp is relatively well known. Therefore, the main

geometric variable is the steepness of the MHT ramp underlying the main topographic front.

Until now, we have used a ramp angle of 18° that represents a reasonable approximation of the

main decollement beneath the Himalaya. However, because of the fact that constraints are

extrapolated along strike, and due to the nature of the data, there is some uncertainty in the exact

structure. Therefore, to investigate whether insight into ramp geometry may be gleaned from detrital

cooling-age data, two additional end-member models were run with ramp angles of 13 degrees (Fig.

13b, i) and 23 degrees (Fig. 13b, ii). The lateral position of the ramp inflection point, on the

northern end, was fixed because we have assumed a link between the TSS outcrop, surface

topography, and the kink bend in the underlying decollement. As a consequence, the depth of the

main decollement under the STDS becomes ~25 km and ~35 km, respectively, as opposed to the

original ~30 km constrained by INDEPTH [Nelson et al., 1996].

The two end-member scenarios were modeled while maintaining the best-fit partitioning of the

convergence rate (5 mm/yr to Asia and 15 mm/yr to India). The steeper ramp angle (Fig. 13a, PDF

i) produces a match to the younger ages, but does a very poor job of representing the 5-10 Ma age

Page 125: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

110

population observed in the data. The shallower ramp angle (Fig. 13a, PDF v) produces a 8 to 10 My

peak, as seen in the data, but contains no population of ages < 6 My, despite their dominance of the

observed ages. Therefore, these results indicate that initial model produces a better overall fit to the

data for a convergence rate partitioned with 5 mm/yr of motion to Asia and 15 mm/yr to India and a

ramp angle of ~18°.

Because the effect of decreasing the ramp angle increases the peak probability age, it seems

reasonable that an improved fit might be obtained by both decreasing the ramp angle and increasing

the convergence rate of Asia with respect to the DSS. In contrast, the opposite scenario is also

reasonable: increasing the ramp angle while slowing the convergence rate of Asia with respect to

the DSS. To test this, two further models were run. Partitioning the total convergence rate into 4

mm/yr Asia and 16 mm/yr India, while maintaining the steep-ramp geometry, produces a PDF with

a peak probability at ~5 My (Fig. 13a, PDF ii). This result is very similar to our initial best fit (Fig.

13a, PDF iii), although is less well represented in the 6 to 13 My age range. Partitioning the total

convergence rate, with a shallow ramp geometry, into 6 mm/yr Asia and 14 mm/yr India produces a

peak probability at ~6 My (Fig. 13a, PDF iv). Again, the younger, < 5 Ma, age population observed

in the data is absent from the resulting PDF.

From this sensitivity analysis, it is clear that predictions of the detrital cooling-age signal from the

orogenic front are dependant upon the relationship between the angle of the thrust ramp underlying

the topographic front of the Himalaya, and the relative partitioning of convergence. The initial

model containing an 18° ramp and a convergence rate of 5 mm/yr Asia and 15 mm/yr India,

produced a reasonable overall fit to the data. However, with the uncertainty in ramp angles,

geometry of the decollement, and the random process of selecting grains for analysis, which

produces a certain amount of uncertainty in the data SPDF [Brewer et al., Chapter 1], it would seem

the convergence rates resulting from our investigation could vary by approximately ± 1 mm/yr,

assuming other parameters are reasonable. Given that our end member ramp angles were ± 5

degrees, more accurate constraints on the ramp geometry would significantly improve the

confidence of the results.

Page 126: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

111

4.2 The modeled Marsyandi valley detrital cooling age signal

Until now, we have been considering the detrital age distribution from a rectangular swath across

the orogen (Fig. 1). A sand sample collected from a riverbed, however, is actually an integration of

the specific points contained within the upstream catchment area. Hence, detrital cooling-age

signals need to be analyzed within the framework of the drainage network: the distribution of

probability within a basin PDF will be modified by the morphology of the basin in relation to the

distribution of bedrock cooling ages. With GIS software, the spatial extent of the area draining any

point in the river network can be calculated from a DEM, and then used as a template to extract the

distribution of cooling ages in the catchment. After correcting for spatial variations in erosion rate

via equation 1, the integrated cooling-age signal can be determined for the specific drainage basin.

Using this methodology, we can reexamine our previous results within the specific area of the

Marsyandi valley (Fig. 11, b). The results show that in comparison to the rectangular swath (Fig. 11,

a) the populations of peak probabilities are enhanced when correcting for the drainage area of the

Marsyandi, while the older age populations become less important. This is because the Marsyandi

contains more areas with younger ages than old, in comparison to the general swath represented by

the model. The effects of changes in catchment morphology need to be considered within the

context of volumetric contributions: a small increase of area within a rapidly eroding zone will have

a much larger effect than the equivalent increase of area within a zone that is producing less

sediment. If we compare the distribution of cooling ages generated using a 5 km/My convergence

rate (of southern Tibet with respect to the DSS) to the observed data, we can see that the relative

importance of the older tail is a better match with the observed data when the basin morphology is

considered, but the 6 to 8 My peak is again underrepresented by the model and the likelihood of

dating the most common age are too high.

The dashed line in figure 12b is a synthetic PDF from Brewer et al. [Chapter 2] that is generated by

a model based upon 40Ar/39Ar data from a further 11 sample sites within the Marsyandi basin.

Individual tributary PDFs are modeled assuming vertical erosion and mixed together, as a function

of basin area and contribution of thermochronometer, to produce a resulting trunk-stream signal (a

theoretical NIB-S24 PDF) [Brewer et al., Chapter 1; Brewer et al., Chapter 2]. While the synthetic

PDF generated by the study underrepresents the 8 to 13 Ma age population found in many

Page 127: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

112

tributaries within the Marsyandi valley and suffers from the limitations of model assumptions, it

provides the only integration of all the data within the basin and is thus shown for comparison. The

theoretical PDF in our model, generated using a 5 km/My convergence rate of southern Tibet with

respect to the DSS, also produces the best fit to this theoretical PDF.

4.4 The effects of lithology

All the results thus far have an implicit assumption of uniform distribution of thermochronometer

across the catchment area. Clearly this is not true in the Himalaya, which contains a wide range of

lithologies ranging from carbonate mudstones to granites. The contribution of thermochronometer

to the fluvial system is dependant upon the percentage of thermochronometer (of the correct size

fraction) per unit volume of material eroded. In this instance, the specification of a correct size

fraction is an analytical constraint: Brewer et al., [Chapter 1; Chapter 2] use a grain fraction of 500

to 2000 µm to ensure that individual muscovite grains contain enough radiogenic 40Ar to detect.

Thus, with some knowledge of the geology within the catchment area, a lithology correction factor

can be applied to produce a refined predicted cooling-age signal at the basin mouth.

Without a detailed investigation of bedrock geology, the distribution of thermochronometer within

the catchment is difficult to constrain. However, in order to examine how lithological contrasts

affect the signal, we use point-counting data [Brewer et al., Chapter 2] to constrain the contribution

of muscovite from individual tributaries within the overall Marsyandi valley. For those tributaries or

areas of the trunk stream that were not sampled and counted, we assign reasonable values based on

averages from the surrounding basins. Individual tributary basins are assigned a lithological

correction factor in ARCINFO, and then combined with the erosion-rate map and cooling-age map

to predict a synthetic cooling-age signal for the basin mouth.

The results (Fig. 14) illustrate that in this case, the lithological correction that we used produced a

worse fit to the observed data than the PDF predicted with uniform distribution of

thermochronometer. The PDF generated with a lithological correction factor tends to concentrate

the probability within younger ages, whereas the older 6 Ma age population observed in the data

becomes even more underrepresented. This might be due to the resolution of our point-counting

Page 128: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

113

data: our approach assumes a uniform distribution of thermochronometer within each tributary

catchment and the strong heterogeneity that probably exists from lithology to lithology, especially

in contrasting litho-tectonic zones, is unconstrained. For example, tributaries were sampled just

upstream of their junction with the Marsyandi River, and those to the south of the Main Central

Thrust typically span both Greater Himalayan and Lesser Himalayan sequences. Point counting

commonly predicted at least 2-fold differences between these two zones. In addition, the effect of

the low contribution by areas in the Tethyan sedimentary series has already been partially accounted

for in each of the models. Due to the ramp geometry, large areas of land to the north of the

topographic divide (in the sedimentary rocks of southern Tibet) have ages that are not reset – they

are advected into the orogen above the closure temperature (Fig. 2). These are not considered when

predicting the age distribution at the basin mouth and therefore a major lithological correction has

been made by default in all of the model runs. However, the general lithological correction

technique we presented above, would almost certainly produce improved results in areas with better

lithological constraints, or areas with dramatic contrasts in the fraction of thermochronometer (that

are not accounted for by un-reset cooling ages).

5.0 Discussion

5.1 Modeling

We have presented a modeling technique for predicting the distribution of cooling ages in sediment

samples from orogenic rivers. The use of a kinematic-and-thermal framework, in conjunction with

the topography, drainage-basin morphology, and lithological characteristics of the bedrock

represents a new approach that can help calibrate and test concepts of orogenic evolution and

thermochronological interpretation. Combining the 90-m DEM with the thermal-and-kinematic

model helps overcome three major problems of geochronological models: 1) the assumption of flat

isotherms; 2) the variation of bedrock cooling ages with elevation, and; 3) the effect of non-vertical

advection. These are solved simultaneously, as an integral part of the model. The assumption of flat

Page 129: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

114

isotherms is replaced by a thermal model that accounts for perturbations caused by 1) the long-

wavelength, strike-normal, topography, and 2) the advection of rock through the system. The

variation of bedrock cooling ages with elevation can be predicted because we know the relationship

between the topography and thermal structure: assumptions of linear and vertical age gradients are

no longer required. The effects of lateral velocity fields are already incorporated into the thermal

model, and the variation of bedrock cooling ages with position in the landscape is a function of the

trajectory of individual rock particles through the orogen.

In this paper, we have illustrated this new approach using a simplified model of Himalayan

tectonics to predict the distribution of detrital-muscovite cooling ages observed at the mouth of the

Marsyandi drainage basin. Despite many uncertainties in the kinematic-and-thermal parameters, and

the simplicity of the single-decollement model, the results mimic the major attributes of the

observed data. Various scenarios allow us to examine the effects of 1) changing the ramp geometry

of the major decollement, and 2) varying the partitioning of Indo-Asian convergence with respect to

the DSS. We take the Decollement/Surface Singularity (DSS) as our reference point because it is

independent of the rate of underthrusting or overthrusting (Fig. 3). The best result was found to be

~5 km/My assigned to Asia-to-DSS convergence, and 15 km/My to India-to-DSS convergence.

However, lessening the ramp angle in conjunction with an increase in the rate of overthrusting, or

increasing the ramp angle and decreasing the rate of overthrusting, also produced reasonable results.

Thus within the likely range of ramp geometry and convergence partitioning, 5 ± 1 mm/yr is the

predicted value for the sustained late Cenozoic overthrusting by Asia. This is in contrast to the 10

mm/yr used in the by Harrison et al. [1998] and so has implications for models of Himalayan

anatexis and metamorphism.

To examine some of the differences between our approach and traditional thermochronology, we

can examine the predictions of our optimal model along a strike-normal transect (Fig. 15). The

predicted bedrock ages (Fig. 15b) increase northwards over the topographic front and southwards

over the lower Himalaya as a function of particle trajectory, depicted in figure 2. We can compare

the distribution of ages to the erosion rate predicted from a) our model of erosion rate (equation 1),

and b) that predicted from the vertical component of the overthrusting vector (Fig. 15c). The former

predicts much higher volumes of sediment eroded from the topographic front region, whereas the

latter predicts uniform volumes of sediment eroded from the width of the orogen. Both, however,

Page 130: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

115

illustrate significant variations in bedrock cooling age across zones of equal erosion. If we consider

the zone of equal erosion shaded in figure 15, we can see that the predicted maximum ages vary

from ~10 to 28 My across the region. This has important implications because traditional

thermochronological approaches, assuming vertical erosion, would yield spurious estimates of the

relative denudation rates. Thus we can see that the affects of lateral advection are very important,

and certainly need to be considered when using cooling rates as a proxy for erosion in active

orogenic belts.

The importance of some of the simplifications incorporated into the kinematic-and-thermal model

need be considered. In the thermal model, the assumption of a 0°C topographic-surface boundary

temperature will produce errors as there is a change in temperature with elevation in the

atmosphere. To investigate this, we ran a model using a 20°C topographic-surface boundary

temperature. This resulting PDF displayed a shift of < 0.5 My for the youngest ages, and had

negligible effect on the older ages. This indicates that for our purposes, this simplification was

reasonable and did not affect our conclusions. However, for more detailed modeling, a lapse-rate

function should be included to account for the changes in temperature with altitude.

The thermal characteristics of the rocks of the Himalayan orogen, and additional thermal processes,

are also simplistically represented by the model. The distribution of radioactive heat-producing

elements within the orogen, the applicability of the shear-heating model, the effects of anatexis, and

the effects of fluids are all largely unknown. In this paper, we have taken the most applicable

estimates for the thermal parameters and have not conducted a full sensitivity analysis of these.

Henry et al. [1997] conducted a more rigorous examination of the effects of various heat production

and heat flow scenarios. Here we are more concerned with introducing the methodology required to

predict the distribution of bedrock cooling ages and the resulting detrital age signal, rather than

producing the best thermal model for the Himalaya.

To extrapolate our 2-D model into 3D, we have to assume that whereas the average strike-normal

topography is accounted for in the model, the strike-parallel topography causes no isotherm

deflection. Brewer et al. [Chapter 1] illustrated that for vertical erosion rates of 3 mm/yr and 6 km

of topographic relief, the isotherm deflection was negligible with respect to the analytical

uncertainty associated with 40Ar/39Ar dating using detrital muscovite. Unfortunately, the

Page 131: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

116

applicability of this result to this investigation is difficult to ascertain. In our thermal model, the

crustal column is not eroding vertically, but erodes normal to the transport direction. Therefore, the

“apparent” relief, that might be applicable, is not measured normal to the geoid, but is measured

parallel to the particle transport direction.

To illustrate this effect in the Himalaya, we can look at the apparent relief for various particle

trajectories (Fig. 16). The bottom panel illustrates that the average relief over the Himalayan

topographic axis is about 8 km for a particle trajectory of 20°. This result, however, taken within the

framework of vertical erosion [Brewer et al., Chapter 1] is not strictly applicable to the situation:

cooling occurs through the top of the thrust sheet, and not just the eroding front edge. As a result,

the whole system will be cooled more, with the closure isotherm deeper and hence less deflected by

topography, than predicted from vertical erosion alone. Thus, although our 2-D extrapolation and

assumption of no topographic-induced deflection of the 350°C isotherm is probably reasonable, a

combined topographic and 3-D, kinematic-and thermal model is necessary to rigorously examine

the thermal interactions between particle trajectory and landscape morphology. This will be

especially important when considering the effects of topography upon lower-temperature isotherms,

such as might be predicted for fission-track or (U-Th)/He investigations.

5.2 The single-decollement model

The evolution of the Himalaya, over the time frame of the thermochronometer, is probably the most

difficult variable to constrain and has large effects on the thermal structure and the trajectory of

particles through the orogen. In this paper we have assumed that deformation field in the Himalaya

can be adequately represented by a single decollement. However, the effects of MCT movement on

the bedrock cooling ages [e.g. Harrison et al., 1998] and the process of tectonic unroofing by the

STDS are difficult to model because the timing, displacement, and duration of faulting is largely

unconstrained except in some rare circumstances [e.g. Hodges et al., 1998; Hodges et al., 1992].

Consequently, we ignore the normal faulting of the STDS, and assume that the Lesser Himalaya and

Greater Himalaya tectono-stratigraphic units are kinematically linked.

Page 132: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

117

The effects of transitory thermal fields on the distribution of bedrock cooling ages have not been

investigated, but we can test the effects of having solely the MCT active and at thermal steady state.

To do this, we take the approximate outcrop of the MCT and extend the ramp north at ~18° from

this (Fig. 13b, (iii)), to create a “paleo-MCT” parallel to our modern decollement. We consider the

cooling-age PDF from the whole range front, because the drainage-basin shape would likely be

different if solely the MCT were active. The resulting theoretical PDF (Fig. 17) is less spiky that

previous predicted PDFs and contains a dominant population from 5 to 10 My, and an older 15 to

20 My population. This result actually fits the broad young population in the data better than

scenarios assuming that the MHT can be approximated by the position of the MBT. To see whether

the predicted cooling-age signal was simply a function of the contributing area, we conducted one

further test. We used the distribution of bedrock cooling ages predicted from the model assuming

that the MBT provides a good approximation for the MHT, and just considered contributions from

the area above the approximate present location of the MCT. This had the effect of reducing the

peak probability of the young age population, but did not resemble the width of the 5 to 10 My peak

seen using the paleo-MCT model.

The predicted age distribution from the model with the paleo-MCT is subject to certain caveats,

however. Firstly, we have to maintain the modern topography in the model, even though if solely

the MCT were active, the Lesser Himalayan landscape would not be so developed. Secondly, the

ramp geometry of the paleo-MCT is poorly constrained. The position of the MCT today may not be

a good proxy for the location of the paleo-MCT. Subsequent movement along the MHT could

translate the paleo-MCT, possibly rotating it as it passes through kink bends caused by the modern

steep ramp underlying the topographic front. Therefore, without isolating the detrital signal from the

Lesser Himalaya, and without a more thorough definition of the true age distribution through dating

more grains, it is difficult to distinguish conclusively between a long-lived, active MCT controlling

erosion and our initial MHT model.

The predicted PDF for the paleo-MCT suggests that our simplification of the MHT may not be the

best solution, and that perhaps both the MCT and MBT are active. Evidence of Late Miocene to

Pliocene metamorphism of the Main Central Thrust zone and its footwall [Copeland et al., 1991;

Harrison et al., 1997; MacFarlane et al., 1992; Macfarlane, 1993] supports theories of recent

activity. Even if the MCT is not actually currently active, we may still be seeing the cooling-age

Page 133: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

118

response to that system, despite the deformation stepping foreland-wards. Certainly, a more

complex kinematic and thermal history, containing multiple fault activity is likely. However, further

geological constraints are needed on the temporal activity of faulting within the study area before

more complex modeling becomes justified.

5.3 The stratigraphic record

One major motivation for trying to understand the generation of cooling-age signals is that in

addition to the possibilities of testing numerical models for modern situations, dating detrital

minerals from the stratigraphic record provide a means to constrain orogenic evolution through

time. Although bedrock geochronological data may be collected to calibrate models of modern

orogenic deformation, constraints through geological time are sparse because bedrock

thermochronology is limited to the rocks exposed at the surface today. Therefore, by dating

minerals from the geological record, we can combine the precise controls of thermochronology with

the record of sedimentation preserved in the foreland basin, thus providing a quantitative temporal

record of orogenic exhumation.

Insights from this investigation indicate that the foreland signal should be representative of the

orogenic signal providing that a major transverse river is sampled. For example, when comparing

the entire swath (Fig. 11a) and the Marsyandi valley (Fig. 11b) it is encouraging to note that

whereas drainage-basin shape does modify the cooling-age signal, the overall pattern remains

consistent. Thus, although one can imagine scenarios in which this will not be true, an average

transverse river should be broadly representative of the overall cooling-age signal derived from a

hinterland that does not vary widely along strike, such as the Himalayan orogen. Additionally, if

one could isolate and examine the stratigraphic record of sediments eroded from the Marsyandi,

small changes in the areal extent of the drainage system through time would result in insignificant

changes to the SPDF when compared to the analytical uncertainty of dating. Whether minor

temporal changes in topographic characteristics, drainage-basin shape, and lithological contribution

could be extracted from the geological record would depend upon the sensitivity of the

thermochronometer and the number of grains dated from the detrital sample [Brewer et al., Chapter

Page 134: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

119

1]. It seems that whereas minor changes in these parameters would not affect the overall distribution

of ages in the sediment, major changes in partitioning the convergence rate (Fig. 10) and ramp

geometry (Fig. 13) should be detectable. To extract maximum information, however, calibrating the

modern detrital cooling-age signal with the modern geodynamics is of prime importance.

6.0 Tectonic implications for the Himalaya

While the main thrust of this paper is developing the methodology of modeling detrital cooling-age

distributions, we discuss the implications of our results in the context of Himalayan tectonics. With

recognition of the importance of the relative partitioning of the total convergence rate, we can

divide models for the kinematics of the Himalaya into three end-member types, categorized on the

relative motion of the DSS towards a marker in South Tibet (i.e. the Indus Suture Zone (ISZ)).

From an initial starting condition (Fig. 18a) the evolution of different models can be examined in

the context of the detrital cooling ages of Brewer et al. [Chapter 2], the kinematic-and-thermal

modeling in this paper, and observable geology.

The first end-member model, and perhaps the most intuitive, is the growth of the mountain belt

through time by thrust-and-fold belt style imbrication of material scraped off the down-going plate

(Fig. 18b). The convergence rate (V) is accommodated by the underthrusting of India beneath Asia.

If the topography is already at threshold conditions, then the DSS will have to move away from the

ISZ at a rate proportional to the material added to the system (Vadd) either by: a) increasing the

width of the orogenic belt with a critical-wedge-type model, or perhaps; b) building elevation to the

limit of rock strength and propagating a self-similar profile as the Tibetan Plateau widens.

The second end-member model requires a steady-state condition, with mass added to the system by

tectonics being removed by erosion (Fig. 18c), while the DSS remains fixed with respect to the ISZ

through time. Finite-element models of orogens predict that deformation can be localized by erosion

[Beaumont et al., 1992; Willet, 1999]. In which case, high erosion rates on the southern front of the

Himalaya might be “sucking” out the metamorphic core of the Himalaya, the Greater Himalayan

Sequence, from beneath (a molten) southern Tibet, with the mountain front maintaining its position

Page 135: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

120

with respect to southern Tibet. This is one proposed explanation of the STDS, the MCT, and the

inverted metamorphic sequence found in the Greater Himalaya.

The third model involves the progressive erosion of southern Tibet, and is the simplified thermal-

and-kinematic model we present in this paper (Fig. 18d). With a fixed India-to-Asia convergence

rate, the relative partitioning of velocity controls the rate of erosion in the Himalaya. With no

normal shear on the STDS, and no internal deformation, the DSS will move towards the ISZ,

resulting in the relative retreat of the southern margin of the Tibetan Plateau.

These three models illustrate end-member scenarios for how the southern margin of the Tibetan

Plateau evolves through time, and although we do not believe that any of them is strictly correct, it

is illuminating to examine the geology within this framework. We can argue that, whereas it is clear

from Himalayan faults and exhumation patterns that periodic mass addition must be occurring (Fig.

18b), this is not the dominant process operating today. At some time in the evolution of the

Himalaya, significant mass-addition is necessary to achieve significant topography by the early

Miocene [France-Lanord et al., 1993; Najman et al., 1997] and Eohimalayan metamorphism [see

Hodges, 2000 for review]. At present, however, the Himalaya are narrow (the distance from the

topographic divide to the deformation front is only ~100 km), with a poorly developed foreland

fold-and-thrust belt considering that collision has been occurring since ~55 Ma. This style of narrow

deformation, deeply-exhumed crust, and an overfilled foreland basin may be typical of windward

orogenic fronts [Hoffman and Grotzinger, 1993], and perhaps indicates that exhumation is localized

to the very edge of the Tibetan Plateau. In addition, if the plateau is widening significantly, the

presence of the TSS capping the high peaks is also problematic; it is difficult to widen the plateau

while maintaining a sedimentary cover across south Tibet and onto the topographic axis.

The presence of the STDS indicates that normal shear had an active role in the evolution of the

Himalaya (Fig. 18c), with evidence that the MCT and STDS moved simultaneously during the

Miocene [Hodges et al., 1992], but there is currently little evidence of significant modern activity

on the STDS. Because normal faulting is a necessary prerequisite of a mass-balance type model,

either it seems unlikely that this mechanism occurs today, or the present role of the STDS has been

significantly underestimated.

If the STDS is inactive, and there is no internal deformation, then mass must be removed from the

edge of southern Tibet (Fig. 18d), resembling the model developed to predict the distribution of

Page 136: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

121

cooling ages. The range front is eroded plateau-wards, with the Greater Himalayan sequence and

Tibetan sediments uplifting passively as the effects of the underlying and northwards-moving MHT

ramp are felt. This model supports the idea that retreat of the Tibetan Plateau has resulted in the

Kathmandu Syncline klippe [Masek et al., 1994] and explains the current association of Tibetan

sedimentary rocks and Greater Himalaya sequence, although it cannot explain the original

juxtaposition of rocks with such contrasting metamorphic grades.

Our model suggests that, in order to match the cooling-age data, ~25% of the current total

convergence across the DSS has to be partitioned into the overthrusting slab since closure (at least 5

to 7 My for the younger ages). This seems to negate a large-scale net mass addition to the system

with deformation moving progressively towards the foreland, as would occur with a traditional fold-

and-thrust belt model. Our model supports evidence for recent movement on the MCT [Harrison et

al., 1997; MacFarlane et al., 1992], rather than Holocene movement on the MFT [Lave and

Avouac, 2000] as being representative of the geodynamics from 0 to 5 Ma. Neither the paleo-MCT

model nor the MHT model required significant normal shear to explain the range in ages observed

in the data. However, if extrusion of the high-grade metamorphic core was occurring due to normal

shear located strictly between the Greater Himalayan sequence and the TSS, this would not be

distinguishable in our model.

Given the complex geological history, the record of detrital cooling ages from the foreland basin

could provide insight into how the Himalayan evolves. If a model of quasi-mass balance were

correct (Fig. 18c), then the detrital cooling age record in the foreland might be expected to show age

populations representing high erosion rates through time. Fluctuations in the older-age populations

might be expected if normal faulting cut through the Greater Himalayan sequence. With a model for

the erosional rollback of the plateau margin, we would predict a fairly uniform signal though time

with relatively constant older age populations controlled simply by the geometry of the detachment.

If an accretion model were representative (Fig. 18b), the age signal would vary with each thrust

sheet added; distinct populations of ages might be expected, perhaps with un-reset ages more

dominant (reflecting the original age of the rock rather than a cooling age). Thus, if we can explain

the modern detrital cooling signal with a specific deformation geometry, the record of ages in the

foreland could provide valuable insights into the growth of the Himalaya.

Page 137: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

122

If the single decollement model is representative of the current kinematics, then Himalayan

topography is a consequence of interaction with the underlying Indian Plate. Thus the rate of

erosion on the topographic front might control the rollback of the Indian Plate inflexion point

towards southern Tibet, and potentially the rate of Indian underthrusting with respect to the DSS. As

a consequence, the erosion rate becomes a key parameter in constraining the kinematics of the

Himalaya. Moreover, changes in kinematics on the southern boundary of the collision zone could

affect the far-field deformation within Central Asia [e.g. England and Houseman, 1988]. As local

bedrock incision rates up to 10 mm/yr have been documented along the Indus River [Burbank et al.,

1996], erosion could potentially accommodate up to one half of the convergence between India and

South Tibet. Our model indicates that one-quarter of the convergence can currently be attributed to

erosion of material off the orogenic front. Therefore it seems likely that changes in the time-

averaged erosion rate, perhaps due to long-term climate change or the initiation of the Indian

Monsoon, could significantly affect the evolution of the Himalaya and consequently the

development of the Tibetan Plateau and far-field deformation.

7.0 Conclusions

Previous kinematics-and-thermal geochronological models [e.g. Harrison et al., 1998] have not

considered either the effects of actual topography upon the distribution of cooling ages, or the

resulting detrital cooling-age signal. In this paper we have introduced a new method of combining

digital elevation models with numerical kinematic-and-thermal modeling to predict cooling-age

distributions. For a given kinematic-and-thermal structure, the spatial variation of bedrock cooling

ages can be predicted. Once corrected for the relative erosion rate, which may be modeled as a

function of the underlying ramp angle in steady-state landscapes, the cooling-age signal can be

determined for any catchment area. Because detrital mineral samples are easily collected, rapidly

dated, and represent an integration of information from a large spatial area, this provides a good

method of testing increasingly complex numerical simulations.

We have applied this new methodology to modeling the detrital cooling-age signal of the Marsyandi

valley in Central Nepal. The results illustrate that the distribution of bedrock cooling ages is

Page 138: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

123

sensitive to the relative partitioning of the total convergence rate between southern Tibet and India.

With a single decollement and the surface outcrop of the MBT representing the modern MHT, the

detrital cooling age data of [Brewer et al., Chapter 2] are most closely matched using a convergence

rate of India 15 km/My, and Asia 5 km/My, with respect to the DSS. There is a trade-off between

ramp angle and convergence rate, but within reasonable geometrical limits, and assuming the

thermal structure is appropriate, the convergence rate can be constrained to ± 1 mm/yr.

Modification of the detrital cooling-age signal by drainage-basin shape and distribution of

thermochronometer is a secondary effect within the Marsyandi drainage basin.

In an attempt to evaluate the relative importance of the Main Central Thrust, two models were

developed: in one, the fundamental Main Himalayan Thrust transfers all motion to the Main

Boundary Thrust; in the second, the motion is transferred to the MCT. Whereas both models fit the

data adequately, given the uncertainties in the observed data SPDF, the active MCT models

produced a better fit. The predicted PDF displays a broader distribution of younger ages that are

more representative of the observed data. Hence our model supports theories suggesting that the

MCT has active recently [Harrison et al., 1997; MacFarlane et al., 1992].

If the simplified two-plate, single-decollement model is considered to be a reasonably good

approximation of Himalayan kinematics, then it has important implications for the development of

the Himalayan orogen. If the STDS is inactive, then erosion rates are controlling the relative

migration of the Himalayan topographic front with respect to southern Tibet. Therefore, the

Himalayan erosion rate is controlling the southern boundary condition of the Tibetan Plateau, and

changes in that boundary condition could affect the deformation field of Central Asia. A change in

climate that resulted in more efficient erosion on the southern boundary would imply that a larger

fraction of the total convergence could be accommodated, thus potentially affecting intra-

continental orogensis many thousands of kilometers away.

This study provides insights into how the detrital cooling-age signal reflects the deformation pattern

within a collisional orogen. Detrital-mineral thermochronology can provide an efficient way to test

ideas of orogenic development, and the methodology introduced in this paper can be combined with

many other numerical models to predict the distribution of detrital cooling-ages. With better

temporal controls on the timing and activity of faults, future kinematic-and-thermal models would

be greatly improved. The effects of topographic deflection of the closure isotherm, particularly

Page 139: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

124

important for lower temperature thermochronometers such as apatite fission-track and (U-Th)/He,

could be investigated with fully 3-D thermal models. In addition, if used in combination with

landscape evolution models, the assumption of topography in steady state could be addressed. The

greatest advantage to detrital, as opposed to bedrock, thermochronology is that the stratigraphic

record provides a window into the past. Therefore, models of orogenic evolution, calibrated by the

modern detrital cooling-age signal, can now be assessed against detailed, quantitative field data

from the foreland.

Page 140: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

N

Marsyandi drainagebasin

South Tibet

Gangetic foredeep

Himalayas

Tibetan plateau

India

45 mm/yr

Topographic axis

65 km

Stu

dy a

rea

TOPOGRPAHIC SWATH

X

Y

Figure 1. Location of the Marsyandi drainage basin and the study area, aligned with the strike of the orogen. The white dot indicates the location of the geochronological sample at the mouth of the Marsyandi catchment (S-24: [Brewer et al., Chapter 2]). The topographic image is derived from a 90-m DEM, and the approximate location of the topographic axis is shown for reference. The hatched area shows a representation of the swath used for the topographic analysis (discussed in text). This swath is scrolled normal to the strike (in the direction illustrated by the arrows) and topographic characteristics calculated over the strike-parallel window.

125

Page 141: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

350oC

0

2

4

6

8

N

S

ab

d

DSS

Ages not reset

MHTage

p

Age (My)

Indian Plate

closure

isotherm

~40 km

Eurasian

Plate

c

Figure 2. Conceptual basis for the combined thermal, kinematic, and detrital model. With a simplified ramp-flat geometry and known convergence rate, the velocity (speed and trajectory angle) of particles through the orogen can be calculated. Within the predetermined kinematic framework, the thermal structure after 20 My is calculated, and the depth of the closure temperature for muscovite (350oC) extracted. Extrapolating the 2D thermal structure into 3D, we use a 90-m DEM to calculate how long it takes each point in the landscape to pass through the closure temperature and reach the surface along the specified particle path (distance of the black arrows divided by the convergence rate of the southern Tibet with respect to the DSS). The youngest ages are created by trajectory (b) because the 350oC isotherm is closest to the surface along this trajectory. Trajectory (a) produces the oldest cooling-ages because it travels along a flat (under the Lesser Himalaya) before reaching the surface. Particles moving along (c), travel the further than (b), and give intermediate ages, whereas trajectory (d) advects rock into the orogen above the closure temperature, and so can be assigned an original-rock age. The insert depicts a hypothetical distribution of detrital ages from the outlined catchment.

126

Page 142: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

10

20

30

0

2

4

6

8

400 12080 160 200 240

40

50-40

distance from MBT (km)

elev

atio

n a.

s.l.

(km

)

MBT

mean topography(min and max envelope)

NS

(a) (b)

(c)

(d)

(e)

(f)

(g)

(h)

(i)

(j)

Indian UC

Indian LC

TS

GH

DSS

(iii)(iv)

10

20

30

mea

n sl

ope

(o)

(i)

(ii)

dept

h b.

s.l.

(km

)

MFT

A = 2.5 m Wm-3

A = 0.4 m Wm-3

A = 2.5 mWm-3

A = 0.4 mWm-3

K = 2.5 Wm-1K-1

K = 2.5 Wm-1K-1

Foreland Lesser Himalayas Higher Himalayas Tibetan Plateau

MantleA = 0 Wm-3 K = 3.0 Wm-1K-1

K = 2.5 Wm-1K-1

Figure 3. Constraints used for the kinematic-and-thermal model. The top panel illustrates the mean slope calculated over a scrolling swath oriented normal to the transect and is divided into four domains. The lower panel shows the thermal conductivity (K) and radioactive heat production (A) assigned to the Tibetan zone sediments (TSS), Greater Himalayan sequence (GH), and the Indian upper crust (UC) and lower crust (LC). The Main Himalayan Thrust (MHT) ramp geometry that we initially use is illustrated, outcropping at the location of the Main Boundary Thrust (MBT). The approximate location of the Main Frontal Thrust (MFT) is indicated. Note the change in scale above sea level to illustrate the minimum, maximum and average elevations. Ramp constraints (section 3.1) and the decollement/surface singularity (DSS) - which is used as a reference marker - are discussed in the text.

127

Seismicity

MHT

Page 143: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

2000

4000

6000

8000

0100 140

elev

atio

n (m

) max.

mean

min.

topographic envelope

5.6 My

6.2 My

4.9 My

20 12040 60 80 100 180160140 200

0

10

20

30

40

dept

h (k

m) 50oC 150oC

250oC

350oC

450oC

Asia

India

a) Thermal model

b) Particle trajectories and cooling ages

c) Topographic influence

(b)

0

10

20

12060 80 100 140

dept

h (k

m)

350oCIndia

Asia

(c)

24 km28 km59 km

Age calculation:(i) 59 km = 11.8 My 5 km/My(ii) 28 km = 5.6 My 5 km/My(iii) 24 km = 4.8 My 5 km/My

(i) (ii) (iii)

5 km/My

15 km/My

120distance (km)

TRAJECTORY

(ii)(iii)

Figure 4. The three components needed to construct cooling ages for the landscape. (a) The depth of the 350oC isotherm is modeled for the specified convergence rate and ramp geometry. In this case the total convergence rate was partitioned into 15 mm/yr of Indian underthrusting and 5 mm/yr of Asian erosion. (b) An enlarged portion of the top panel illustrating how the cooling ages are calculated. The distance a particle travels between passing through the closure temperature and reaching the surface (indicated by the black arrows), is divided by the convergence rate partitioned to southern Tibet with respect to the DSS. (c) Particles following the same flow line in 2D travel different distances along strike, because a 3-D landscape has a range of topography. Hence, when the real topography is used, each incremental change in the transect location along strike will result in a different pattern of predicted bedrock cooling ages. The ages of maximum, minimum, and mean topography are shown for illustration.

128

Page 144: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Thrust sheet

(iii) Advection of topography // to particle direction, hence

minimum erosion.

(i) Advection of topography I to particle direction, hence

maximum erosion.

(iii)

(i)

displacement in time (dt).Volume erodedin dt.

=

=

Kin

k ba

nd

present topography

Particle trajectory

V

V

Vz

(ii)

(ii) Vertical component of erosion

Figure 5. Volume of material eroded in a time increment (dt) depends upon the aspect of the topography in relation to the particle velocity (V). The black topography is the present topographic profile. The gray line mirroring the present topography illustrates the volume of rock eroded in dt with an assumption of complete steady-state conditions. When the particle velocity is normal to the topography, maximum volumes of material are eroded (i). When the particle velocity is parallel to the topography, the topography becomes advected laterally and little erosion occurs (iii). Note that the volume of eroded material is underestimated if solely the vertical component (Vz) of the particle velocity is used in the calculation (ii).

129

Page 145: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

X

Y+a

+S

v.dt

Particle velocity

Average slope Volume of rock

eroded in dt

DEM

v

Figure 6. Calculation of volume of rock eroded in time increment (dt) for one digital-elevation model (DEM) grid cell, assuming a steady-state landscape. The volume of rock eroded for a given particle speed (v) and time (dt) is dependent upon the topographic slope (S) and particle direction, which is controlled by the angle of the underlying thrust ramp (a). The values beta and alpha are dependent upon the relationship between the topographic surface slope, as described in the text. Both angles are measured positively as illustrated above, and the average topographic slope is assumed to be parallel to the Y direction.

130

Page 146: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

10 20 30 40 50 60 70 80 90Particle trajectory angle (ofrom horizontal )

0.2

0.4 0.60.6

0.8

0.8

0.8

1

1

1

1

11

11

1.2

1.21.2

1.21.41.4

1.4

1.6

1.6

1.61.8 1.8

5

10

15

20

25

30

35

40

45

50

55

60

topo

grap

hic

slop

e (o

)

00

Figure 7. Relationship between the topographic slope and particle trajectory angle in determining the volume of material eroded from a DEM cell. The X and Y cell dimensions, dt, and particle speed are set to unity. A key to interpreting this figure is to recall that the surface area of the landscape represented by the 1x1 DEM cell varies as a function of slope. Note that when the particle trajectory is vertical, one unit volume is eroded from the landscape, independently of the topography. As the topographic slope approaches the plane of the particle trajectory, the volume of material eroded approaches zero because material is advected parallel to the slope.

131

Page 147: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

Mea

n el

evat

ion

(m)

distance (km)

-1000

0

1000

2000

3000

4000

5000

6000

0 40 80 120 160 200 240 280

= mean elevation

= best fit line segment

Lesser Himalayas

Tibetan Plateau

Topographic front

Indi

an P

late

Figure 8. The three linear segments (black dashed lines) used as a proxy for the regional slope, taken from a transect normal to the strike of the orogen. The mean elevation, averaged over the ~120 km by 0.6 km swath, is shown by the thick gray line.

132

Page 148: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

AsiaIndia

AsiaIndia

AsiaIndia

a) India fixed

b) Asia fixed

c) Equal partitioning

20 mm/yr

20 mm/yr

10 mm/yr

10 mm/yr

DSS

Figure 9. Three scenarios for partitioning convergence rate between India and south Tibet. (a) India fixed, movement of Asia at 20 mm/yr with respect to the DSS. (b) Asia fixed, convergence accommodated by Indian under thrusting. (c) Equal division between the two plates: the Indian plate moves at 10 mm/yr northwards, and the Eurasian plate moves at 10 mm/yr southwards with respect to the decollement/surface singularity (DSS).

133

Page 149: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 5 15 20 25 301024

68

10

12

140

0.01

0.02

0.03

0.04

0.05

age (My)

prob

abili

ty

Asia convergence rate

(km/My)

shear=36 vertical

modeled PDF

data

22 My

8 My

2 My

1 My

4 My

1.5 My

0 5 10 15 20 25 30age (My)

prob

abili

ty

Observed data

n=55

Figure 10. Effects of partitioning the relative convergence rate between India and Asia, can be observed in the modeled detrital cooling-age signals. The age signals are represented by probability density functions (PDF) and represent the reset-age signal from the width of the study area (Fig. 1). The convergence rate varies between 2 and 14 km/My for Asia, keeping the total convergence rate (20 mm/yr) and all else constant. The gray band (labeled 'data') represents an approximate range of ages for the highest concentration of probability for the observed data PDF (shown in inset: sample S-24 in the analyses of Brewer [in review-b]). The data is a sample collected from 200-m upstream of the confluence of the Marsyandi and Trisuli rivers (Fig. 1).

134

Page 150: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

5 10 15 20 25

Pro

ba

bil

ity

0

0.01

0.02

i) Asia 6, India 14 km/Myii) Asia 5, India 15 km/Myiii) Asia 4, India 16 km/My

i) i i ) i i i )

data

0 5 10 15 20 25 30age (My)

Pro

ba

bil

ity

0

0.01

0.02

0.03

i)

i i )i i i ) data

synthet ic PDF (Brewer et a l , [ in review])

a)

b)

entire swath

Marsyandi basin

3 My

5 My

8 My

3 My

5 My

8 My

Figure 11. (a) The modeled cooling age signal from the entire mountain front compared against the data from the mouth of the Marsyandi (solid black line). Asian overthrusting rate of 4 to 6 km/yr mimic the general pattern of observed detrital data. The PDF generated with Asia = 5 km/My, India = 15 km/My (ii) displays the best fit to the 20 Ma age population observed in the data PDF. (b) The same model scenarios, but corrected for the age-signal generated specifically from the Marsyandi basin. Note that the peak probabilities are enhanced, while the older 'tails' have diminished importance. The dashed line is the predicted PDF at the mouth of the Marsyandi based upon an integration of individual tributary models and assuming vertical erosion [Brewer et al., Chapter 2 ].

135

Page 151: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

20 12040 60 80 100 180160140 200

0

10

20

30

40

dept

h (k

m) 50oC

150oC

250oC

350oC

450oC

Asia

India

20 12040 60 80 100 180160140 200

0

10

20

30

40

dept

h (k

m) 50oC

150oC

250oC

350oC

450oC

Asia

India

distance (km)

4 mm/yr

16 mm/yr

8 mm/yr

12 mm/yr

a)

b)

Figure 12. The steady-state thermal structure with 20 mm/yr of total convergence with (a) 4 mm/yr of total convergence partitioned into Asia, and (b) 8 mm/yr of convergence partitioned into Asia. Note the relative response of the 350oC isotherm.

136

Page 152: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

age (My)

prob

abili

ty

12060 80 100 180160140 200

0

10

20

30

40

dept

h (k

m) Asia

India +/- 5 km depth

MCTLesser Himalaya

distance (km)

15 mm/yr

5 mm/yr

(i)

(ii)

(iii)

a)

b)

data

0 5 10 15 20 25 300

0.005

0.01

0.015

0.02

0.025

Asia India Decollementdepth

(i)(ii)(iii)(iv)(v)

5 mm/yr

5 mm/yr

4 mm/yr

6 mm/yr5 mm/yr

15 mm/yr

15 mm/yr

15 mm/yr

16 mm/yr

14 mm/yr

35km35km

25km25km

30km

(ii) 5 My

(iv) 6 My

(i) 3 My

(iii) 5 My

(v) 8 My

STDS

Figure 13. (a) The distribution of cooling ages derived from different ramp geometries. PDFs (i) and (ii) were run using a steeper 23o ramp (illustrated by panel b, ramp geometry (ii)) with 5 mm/yr and 4 mm/yr assigned to convergence of southern Tibet with respect to the DSS, respectively. PDF (iii), outlined by the dashed line, indicates the original ramp geometry for comparison. PDFs (iv) and (v) use a shallower geometry (illustrated by panel b, ramp geometry (i)) with 6 mm/yr and 5 mm/yr assigned to convergence of southern Tibet with respect to the DSS, respectively. The data curve in panel a is from Brewer et al., Chapter 2 ]. Ramp geometry (iii) illustrated in panel b is used for generating the PDF shown in figure 16.

137

Page 153: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 5 10 15 20 25 300

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

0.045

age (My)

prob

abili

ty

(ii) no lithological correction

(i) PDF withlithological correction HIGH

LOW

Lithological contribution

India 5 mm/yr, Asia, 15 mm/yr

Figure 14. A comparison of the distribution of detrital cooling ages using (i) a lithological correction and (ii) no lithological-correction factor. The insert illustrates the spatial variation in the lithological correction factor over the Marsyandi basin, taken from the point-counting results of Brewer et al. [Chapter 2 ].

138

Page 154: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0

10

20

30

0 50 100 150 200 250 3000

5000

10000

MCT

a) topography

b) ages

c) relative erosion rate

max

minmean

max

min

mean

UNRESET

vertical componentof plate velocitymodeled erosion

rate

Figure 15. Transects illustrating (a) the topography, (b) predicted cooling ages from our model, and (c) a comparison between vertical erosion rates predicted from ramp geometry and our modeled rates (which are a function of the trajectory of rock particles in relation to the averaged topography). The convergence was partitioned into 5 mm/yr of Asianoverthrusting and 10 mm/yr of Indian underthrusting with respect to the DSS. The approximate modern location of the modern MCT is shown in the top panel. The vertical gray bar is a zone of predicted equal erosion (c).

139

distance (km)

elev

atio

n (m

)ag

e (M

y)re

lativ

e ra

te

S N

Page 155: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 40 80 120 160 200 240 2800

5000

10000

15000

distance (km)

appa

rent

rel

ief (

m)

20o40o90o01000200030004000500060007000

elev

atio

n (2

-sig

ma)

(m

)

20

40

90

angle of particle trajectory

apparent relief

10 20 30 40 50 60 70 80 9002468

101214

mea

n re

lief (

m)

particle trajectory (o)

HHWTLH

Figure 16. Variation of "apparent" relief as a function of particle trajectory. Apparent relief (ii) was taken from the 2-s envelope of elevation (i). The actual relief (shown by the shaded gray curve is generated using a particle trajectory angle of 90 degrees. The thin black lines illustrate the apparent relief for the particle trajectories indicated by the arrows. (iii) The average apparent relief as a function of ramp angle for the whole transect (WT), the Lesser Himalaya (LH), and Higher Himalaya (HH).

140

Page 156: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 5 10 15 20 25 300

0.005

0.015

0.02

age (my)

prob

abili

ty

(ii) MCT ACTIVE

(i) MBT ACTIVE

0.01

DATA

Figure 17. A comparison of the distribution of detrital ages from an orogenic swath in the study area with: (i) the MHT represented by the MBT being the active fault, and; (ii) the MHT represented by activity on solely the MCT (with geometry illustrated in Fig. 13b, (iii)). The data from Brewer et al. [Chapter 2 ] is shown for comparison by the black line. Note that with the MCT active, there is a more dominant 6 t 10 My signal.

141

Page 157: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

NSHimalayas

Tibetan Plateau

TS

GHIndian Plate

ISZ

PMZ

a) T0 - Initial condition

b) T1 - Mass addition to the system

c) T1 - Mass balance

d) T1 - Mass removal from the system

MFT

STDS

MHT

V/2

V/2V/2

V/2

V

Vadd

DS

SD

SS

142

Figure 18. A cartoon showing three end-member models for Himalayan evolution. From a starting condition (a) the displacement markers (indicated by black dots) indicate the displacement at T1 after displacement at V mm/yr with models for: (b) mass addition to the system; (c) mass balance; (d) mass removal from the system. The reference point is considered to be the decollement/surface singularity (DSS), which is the intersection between the Main Himalayan Thrust (MHT) and the surface. The Main Frontal Thrust (MFT) is currently the most active fault to the south. The South Tibetan Detachment System (STDS) is found to the north of the topographic axis and separates the Greater Himalayan sequence (GH) from the Tibetan zone sediments (TS) and a partial melt zone (PMZ) has been imaged by INDEPTH. The initial position of the Indus Suture Zone (ISZ) is indicated, with vertical arrows in the north illustrating the position of the ISZ at T1.

Page 158: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

143

References

Adams, C.J., M.E. Barley, I.R. Fletcher, and A.L. Pickard, Evidence from U-Pb zircon and

(super 40) Ar/ (super 39) Ar muscovite detrital mineral ages in metasandstones for

movement of the Torlesse suspect terrane around the eastern margin of Gondwanaland,

Terra Nova. The European Journal of Geosciences, 10 (4), 183-189, 1998.

Barbey, P., M. Brouand, P. Le Fort, and A. Pecher, Granite-migmatite genetic link: the example

of the Manaslu Granite and Tibetan Slab migmatites in central Nepal, Lithos, 38 (1-2), 63-

79, 1996.

Batt, G., and J. Braun, On the thermomechanical evolution of compressional orogens,

Geophysical Journal Int, 128, 364-382, 1997.

Beaumont, C., P. Fullsack, and J. Hamilton, Erosional control of acrive compressional orogens,

Thrust Tectonics - McClay eds, 1-18, 1992.

Beaumont, C., P. Fullsack, and J. Hamilton, Styles of crustal deformation in compressional

orogens caused by subduction of the underlying lithosphere., Tectonophysics., 232 (1-4),

119-132, 1994.

Bevington, P., and K. Robinson, Data reducton and error analysis for the physical sciences,

WCB/McGraw-Hill, 1992.

Bilham, R., K. Larson, J. Freymuller, and P.I. members, GPS measurements of present-day

convergence across the Nepal Himalaya, Nature, 386, 61-64, 1997.

Bloom, A., Geomorphology. A systematic analysis of Late Cenozoic landforms, Prentice Hall,

1998.

Brandon, M.T., and J.A. Vance, Tectonic evolution of the Cenozoic Olympic subduction

complex, Washington State, as deduced from fission track ages for detrital zircons,

American Journal of Science, 292 (8), 565-636, 1992.

Brewer, I., D. Burbank, and K. Hodges, Numerical modelling of detrital cooling ages: insights

from two contrasting Himalayan drainage basins, Chapter 1, this volume.

Page 159: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

144

Brewer, I.D., D.W. Burbank, and K.V. Hodges, The downstream development of a detrital

cooling-age signal, insights from 40Ar/39Ar muscovite thermochronology in the Marsyandi

Valley of Nepal., Chapter 2, this volume.

Brown, Wenjin, Zhao, Nelson, Hauk, Alsdorf, Ross, Cogan, Clark, X. Liu, and J. Che, Bright

spots, structure, and magmatism in Southern Tibet from INDEPTH seismic reflection

profiling., Science, 274 (December), 1688-1690, 1996.

Brozovic, N., D.W. Burbank, and A.J. Meigs, Climatic limits on landscape development in the

northwestern Himalaya, Science, 276, 571-574, 1997.

Burbank, D., J. Leland, E. Fielding, R. Anderson, N. Brozovic, M. Reid, and C. Duncan,

Bedrock incision, rock uplift and threshold hillslopes in the northwestern Himalayas,

Nature, 379, 505-510, 1996.

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

Jiene, The South Tibetan Detachment System, Himalayan Orogen: Extension

Contemporaneous With and Parallel to Shortening in a Collisional Mountain Belt, 41 pp.,

Geological Society of America, Denver, 1992.

Carter, A., and S.J. Moss, Combined detrital-zircon fission-track and U-Pb dating: a new

approach to understanding hinterland evolution, Geology, 27 (3), 235-238, 1999.

Catlos, E., M. Harrison, M. Searle, and M. Hubbard, Evidence for Late Miocene reactivation of

the Main Central Thrust: From Garhwal to the Nepali Himalaya, in Terra Nostra 99/2, 14th

Himalayan-Karakorum-Tibet workshop, pp. 20-22, Cologne, Germany, 1999.

Catlos, E.J., T.M. Harrison, M. Grove, A. Yin, M.J. Kohn, F.J. Ryerson, and P. LeFort, Further

evidence for Late miocene reactivation of the Main Central Thrust (Nepal Himalaya) and he

significance of the MCT-1, EOS (Transactions of the American Geophysical Union), 78,

651, 1997.

Cerveny, P.F., N.D. Naeser, P.K. Zeitler, C.W. Naeser, and N.M. Johnson, History of uplift and

relief of the Himalaya during the past 18 Million years: evidence from fission-track ages of

detrital zircons from the Siwalik Group, in New Perspectives in Basin Analysis, edited by

K.L. Kleinspehn, and C. Paola, pp. 43-61, Springer-Verlag, New York, 1988.

Clauer, N., Strontium and argon isotopes in naturally weathered biotites, muscovites and

feldspars., Chemical Geology, 31, 325-334, 1981.

Page 160: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

145

Colchen, M., P. LeFort, and A. Pêcher, Annapurna - Manaslu - Ganesh Himal, 136 pp., Centre

National de la Recherche Scientifique, Paris, 1986.

Coleman, M., and K. Hodges, Evidence for Tibetian Plateau uplift before 14 Myr ago from a

new minimum age for E-W extension, Nature, 374, 49-52, 1995.

Coleman, M.E., Orogen-parallel and orogen-perpendicular extension in the central Nepalese

Himalayas, Geological Society of America Bulletin, 108, 1594-1607, 1996a.

Coleman, M.E., The tectonic evolution of the central Himalaya, Marsyandi valley, Nepal, Ph. D.

thesis, Massachusetts Institute of Technology, 1996b.

Copeland, P., and M.T. Harrison, Episodic rapid uplift in the Himalaya revealed by 40Ar/39Ar

analysis of detrital K-felspar and muscovite, Bengal fan, Geology, 18, 354-359, 1990.

Copeland, P., T.M. Harrison, K.V. Hodges, P. Maruejol, P. Le Fort, and A. Pecher, An early

Pliocene thermal disturbance of the Main Central Thrust, central Nepal: Implications for

Himalayan tectonics, Journal of Geophysical Research, 96 (B5), 8475-8500, 1991.

Copeland, P., T.M. Harrison, and P. Le Fort, Age and cooling history of the Manaslu granite:

Implications for Himalayan tectonics, Journal of Volcanology and Geothermal Research,

44, 33-50, 1990.

Copeland, P., R. Parrish, and M. Harrison, Identification of inherited radiogenic Pb in monazite

and its implications for U-Pb systematics, Nature, 333, 1988.

Copeland, P., C. Peters, and Anonymous, Dating detrital minerals; how many is enough?,

Geological Society of America, 1997 annual meeting, 29 (6), 420, 1997.

Copeland, P., C. Skirius, B.L. Claxton, and Anonymous, An inorganic thermal maturity indicator

in the gas-condensate to dry gas preservation window; (super 40) Ar/ (super 39) Ar

feasibility study, American Association of Petroleum Geologists 1996 annual convention, 5,

29, 1996.

DeCelles, P.G., G.E. Gehrels, J. Quade, and T.P. Ojha, Eocene-early Miocene foreland basin

development and the history of Himalyan thrusting, western and central Nepal, Tectonics,

17, 741-765, 1998.

Deino, and Potts, Age-probability spectra for examination of single-crystal 40Ar/39Ar dating

results: examples from Olorgesailie, southern Kenya Rift., Quaternary International, 13/14,

47-53, 1992.

Page 161: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

146

DeMets, C., R.G. Gordon, D.F. Argus, and S. Stein, Current plate motions, Geophysical Journal

International, 101, 425-478, 1990.

Derry, L.A., and C. France-Lanord, Changing Riverine 87Sr/86Sr and Neogene Himalyan

Weathering History: Impact on the Marine Sr Record., In press, 1996.

Dewey, J., J, R. Shakleton, C. Chengfa, and S. Yiyin, The tectonic evolution of Tibet, the

geological evolution of Tibet, Royal Scoiety of London, 379-413, 1988.

Dodson, M., Closure temperature in cooling geochronological and petrological systems,

Contributions to Mineralogy and petrology, 40, 259-274, 1973.

Edwards, R.M., 40Ar/39Ar geochronology of the Main Central thrust (MCT) region: Evidence

for Late Miocene to Pliocene disturbances along the MCT, Marsyangdi River valley, west-

central Nepal Himalaya, Journal of the Nepal Geological Society, 10, 41-46, 1995.

England, P., and G. Houseman, The Mechanisms of the Tibetian Plaeau, Phil. Trans. R. Soc.

Lond, 326, 301-320, 1988.

England, P., P. Le Fort, P. Molnar, and A. Pecher, Heat sources for Tertiary metamorphism and

anatexis in the Annapurna-Manaslu region, central Nepal, Journal of Geophysical Research,

B, Solid Earth and Planets, 97 (2), 2107-2128, 1992.

England, P., and D. McKenzie, A thin viscous sheet model for continental defomation, Geophys.

Journal Royal Astron. Soc., 70, 523-295-321532, 1982.

Faure, G., Principles of isotope geology, 1986.

Ferra, G., B. Lombardo, and S. Tonarini, Rb/Sr geochronology of granites and gneisses from the

Mount Everest Region, Nepal Himalaya, Geologische Rundschau, 72, 119-136, 1983.

Fletcher, Computational Techniques for Fluid Dynamics, Springer, 1991.

Fowler, The Solid Earth, an Intoduction to Geophysics, Cambridge University Press, 1990.

France-Lanord, C., L. Derry, and A. Michard, Evolution of the Himalaya since Miocene time:

isotopic and sedimentologic evidence from the Bengal Fan, in Treloar, P.J. and Searle, M.,

eds.,., Himalayan Tectonics: London, Geological Society London., 74, 603-621, 1993.

Garver, J., and M. Brandon, Fission-track ages of detrtial zircons from cretaceous strata,

southern British columbia: Implications for the Baja BC hypothesis, Tectonics, 13, 401-420,

1994.

Page 162: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

147

Green, P.F., K.A. Hegarty, I.R. Duddy, S.A. Foland, and Anonymous, Assessing zircon fission

track analysis as a paleotemperature tool for sedimentary basins, American Association of

Petroleum Geologists 1996 annual convention, 5, 56, 1996.

Gupta, M.L., Is the Indian Shield hotter than other Gondwana shields?, Earth and Planetary

Science Letters, 115 (1-4), 275-285, 1993.

Hansen, F.D., and N.L. Carter, Creep of selected crustal rocks at 1000 MPa, in AGU 1982 spring

meeting, pp. 437, American Geophysical Union, Washington, 1982.

Harris, N., and J. Massey, Decompression and anatexis of Himalayan metapelites., Tectonics, 13,

1537-1546, 1994.

Harrison, T., M. Grove, L. Oscar, and E. Catlos, A model for the origin of Himalayan anatexis

and inverted metamorphism, Journal of Geophysical Research, 103, 27017-27032, 1998.

Harrison, T.M., K.D. McKeegan, and P. LeFort, Detection of inherited monazite in the Manaslu

leucogranite by (super 208) Pb/ (super 232) Th ion microprobe dating; crystallization age

and tectonic implications, Earth and Planetary Science Letters, 133 (3-4), 271-282, 1995.

Harrison, T.M., F.J. Ryerson, P. Le Fort, A. Yin, O.M. Lovera, and E.J. Catlos, A late Miocene-

Pliocene origin for the central Himalayan inverted metamorphism, Earth and Planetary

Science Letters, 146 (1-2), E1-E7, 1997.

Henry, L. Pichon, and Goffe, Kinematic , thermal and petrological model of the Himalayas:

constraints related to metamorphism within the underthrust Indian crust and topographic

elevation., Tectonophysics, 273, 31-56, 1997.

Hodges, K., and S. Bowring, 40Ar/39Ar thermochronology of isotopically zoned micas: insights

from the southwestern USA Proterozoic orogen., Geochemica et Cosmochimica Acta, 59

(15), 3205-3220, 1995.

Hodges, K., S. Bowring, K. Davidek, D. Hawkins, and M. Krol, Evidence for rapid displacement

on Himalayan normal faults and the tectonic denudation in the evolution of mountain

ranges., Geology, 26 (June), 1998.

Hodges, K., R. Parrish, T. Housh, D. Lux, B. Burchfield, L. Royden, and Z. Chen, Simultaneous

Miocene extension and shortening in the Himalayan orogen, Science, 258, 1992.

Hodges, K., R. Parrish, and M. Searle, Tectonic evolution of the Annapurna Range, Nepalese

Himalayas., Tectonics, 1996.

Page 163: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

148

Hodges, K.V., 40Ar/39Ar geochronology using the laser microprobe, in Reviews in Economic

Geology 7: Applications of Microanalytical Techniques to Understanding Mineralizing

Processes, edited by M.A. McKibben, and W.C. Shanks, pp. 53-72., Society of Economic

Geologists, Tuscaloosa, AL, 1998.

Hodges, K.V., Tectonics of the Himalaya and southern Tibet from two perspectives, Geological

Society of America Bulletin, 112 (3), 324-350, 2000.

Hoffman, P., and J. Grotzinger, Orographic Precipitation, erosional unloading, and Tectonic

style, Geology, 21, 195-198, 1993.

Hurford, A.J., and A. Carter, The role of fission track dating in discrimination of provenance,

Geological Society Special Publications, 57, 67-78, 1991.

Jackson, M., S. Barrientos, R. Bilham, D. Kyestha, and B. Shrestha, Uplift in the Nepal Himalya

revealed by spirit leveling, Geophysical Research Letters, 19 (15), 1539-1542, 1992.

Jackson, M., and R. Bilham, Constraints on Himalayan deformation inferred from vertical

velocity fields in Nepal and Tibet, Journal of Geophysical Research, 99 (B7), 13897-13912,

1994.

Jaeger, J., Application of the theory of heat conduction to geothermal measurements., in

Terrestrial heat flow, edited by W. Lee, AGU, 1965.

Jamieson, R., and C. Beaumont, Orogeny an metamorphism: a model for deformation and

pressure-temperature-time paths with applications to the central and southern Appalachians,

Tectonics, 7 (3), 417, 1988.

Jamieson, R., C. Beaumont, P. Fullsack, and B. Lee, Barrovian regional metamorphism: where's

the heat?, in What drives metamorphism and metamorphic reactions?, edited by Treolar, and

O'Brien, pp. 23-51, 1998.

Keller, E., and N. Pinter, Active tectonics; earthquakes, uplift and landscape, Prentice Hall,

1996.

Koons, P., The Topographic evolution of collisional mountain belts: a numerical look at the

southern alps, New Zealand, American Journal of Science, 289 (November), 1041-1069,

1989.

Koons, P.O., Modeling the topographic evolution of collisional belts., Annual Review of Earth

and Planetary Science Letters., 23, 375-408, 1995.

Page 164: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

149

Krogh, T.E., S.L. Kamo, and B.F. Bohor, Fingerprinting the K/T impact site and determining the

time of impact by U-Pb dating of single shocked zircons from distal ejecta, Earth and

Planetary Science Letters, 119 (3), 425-429, 1993.

Krogh, T.E., J.D. Keppie, J.L.e. Bates, and D.R.e. MacDonald, Detrital zircon ages indicating a

North African provenance for the Goldenville Formation of Nova Scotia, in Report -

Department of Mines and Energy ; Mines and Minerals Branch report of activities; 1986,

pp. 208, (Halifax), 1987.

Kutzbach, J.E., W.L. Prell, and W.F. Ruddiman, Sensitivity of Eurasian climate to surface uplift

of the Tibetan Plateau., Journal of Geology., 100, 177-190, 1993.

Lave, J., and J. Avouac, Active folding of fluvial terraces across the Siwaliks Hills, Himalayas

of central Nepal, Journal of Geophysical Research, B, Solid Earth and Planets, 105, 5735-

5770, 2000.

Lave, J., and J. Avouac, Fluvial incision and tectonic uplift across the Himalaya of Central

Nepal, JGR, in review.

Le Fort, P., Himalayas: The collided range. Present knowledge of the continental arc, American

Journal of Science, 275A, 1-44, 1975.

Le Fort, P., Manaslu leucogranite: A collision signature in the Himalaya, a model for its genesis

and emplacement., Journal of Geophysical Research, 86, p. 10545-10568., 1981.

Levorsen, A., Geology of petroleum, 570 pp., 1967.

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

gravity anomalies and a flexural model of the lithosphere, JGR. Journal of Geophysical

Research. B, 88 (10), 8171-8191, 1983.

MacFarlane, A., K. Hodges, and D. Lux, A structural analysis of the Main Central Thrust Zone,

Langtang National Park, central Nepal Himalaya, Geological Society of America Bulletin,

104, 1992.

Macfarlane, A.M., Chronology of tectonic events in the crystalline core of the Himalaya,

Langtang National Park, central Nepal, Tectonics, 12 (4), 1004-1025, 1993.

Mahon, K.I., T.M. Harrison, and M. Grove, The thermal and cementation histories of a

sandstone petroleum reservoir, Elk Hills, California; Part 1, (super 40) Ar/ (super 39) Ar

thermal history results, Chemical Geology, 152 (3-4), 227-256, 1998.

Page 165: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

150

Mancktelow, N.S., and B. Grasemann, Time-dependent effects of heat advection and topography

on cooling histories during erosion, Tectonophysics, 270 (3-4), 167-195, 1997.

Masek, J.G., B.L. Isacks, T.L. Gubbels, and E.J. Fielding, Erosion and tectonics at margins of

continental plateaus., Journal of Geophysical Research., 99, 13,941-13,956, 1994.

Mathur, L.P., and G. Kohli, Exploration and development for oil in India, , 1964.

McDougall, I., and T.M. Harrison, Geochronology and Thermochronology by the 40Ar/39Ar

Method., New York, Oxford University Press., 212, 1988.

McGoldrick, P.J., and A.J.W. Gleadow, Fission-track dating of lower Palaeozoic sandstones at

Tatong, North central Victoria, ournal of the Geological Society of Australia, 24, pp.461-

464, 1978.

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

Main Boundary Thrust in the western Himalaya, Geology, 23, 423-426, 1995.

Mitchell, J.G., and A.S. Taka, Potassium and argon loss patterns in weathered micas;

implications for detrital mineral studies, with particular reference to the Triassic

palaeogeography of the British Isles, Sedimentary Geology, 39 (1-2), 27-52, 1984.

Molnar, P., and P. England, Late Cenozoic uplift of mountain ranges and global climate change:

chicken or egg?, Nature, 346 (July), 29-34, 1990.

Molnar, P., and P. Tapponier, Cenozoic Tectonics of Asia: Effects of a Continental Collision,

Science, 189 (November), 419-425, 1975.

Najman, Pringle, Godin, and Oliver, Dating of the oldest continental sediments from the

Himalayan foreland basin, Nature, 410, 194 - 197, 2001.

Najman, Y.M.R., M.S. Pringle, M.R.W. Johnson, A.H.F. Robertson, and J.R. Wijbrans, Laser

(super 40) Ar/ (super 39) Ar dating of single detrital muscovite grains from early foreland-

basin sedimentary deposits in India; implications for early Himalayan evolution, Geology

(Boulder), 25 (6), 535-538, 1997.

Nelson, W. Zhao, Brown, Kuo, J. Che, X. Liu, Klemperer, Makovsky, Meissner, Mechie, Kin,

Wenzel, Ni, Nabelek, C. Leshou, H. Tan, W. Wei, Jones, Booker, Unsworth, Kidd, Hauck,

Alsdorf, Ross, Cogan, C. Wu, Sandvol, and Edwards, Partially molten middle crust beneath

Southern Tibet: synthesis of project INDEPTH results., Science, 274 (December), 1684-

1687, 1996.

Page 166: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

151

Ni, J., and M. Barazangi, Seismotectonics of the Himalayan collision zone: geometry of the

underthrusting Indian plate beneath the Himalaya., Journal of Geophysical Research, 89

(B2), 1147-1163, 1984.

Pandey, M., R. Tandukar, J. Avouac, J. Lave, and J. Massot, Interseismic Strain Accumulation

on the Himalayan Crustal Ramp (Nepal), Geophysical Research Letters, 22 (7), 751-754,

1995.

Parrish, R., and K. Hodges, Isotopic constraints on the age and provenance of the Lesser and

Greater Himalyan sequences, GSA Bulletin, 108 (7), 904-911, 1996.

Pinet, C., Structure thermique et evolution des continents. Thermal structure and evolution of the

continents, Doctoral thesis, Universite de Paris VII, Paris, 1992.

Pinet, P., and M. Souriau, Continental erosion and large-scale relief, Tectonics, 7, 563-582, 1988.

Pollack, Steady heat conduction in layered mediums: the half space and sphere., Journal of

Geophysical Research, 70 (22), 5645, 1965.

Powell, C.M., and P.J. Conaghan, Plate tectonics and the Himalayas, Earth and Planetary

Science Letters, 20, 1-12, 1973.

Press, W., S. Teukolsky, W. Vetterling, and B. Flannery, Numerical recipies in fortran; the art of

scientific computing, Cambridge University Press, 1992.

Raymo, M.E., W.F. Ruddiman, and P.N. Froelich, Influence of late Cenozoic mountain building

on ocean geochemical cycles, Geology, 16, 649-653, 1988.

Renne, K-Ar and Ar/Ar Dating, in Quaternary Geochronology: Methods and applications, pp.

77-100, AGU, 2000.

Renne, P.R., C.C. Swisher, A.L. Deino, D.B. Karner, T.L. Owens, and D.J. DePaolo,

Intercalibration of standards, absolute ages and uncertainties in (super 40) Ar/ (super 39) Ar

dating, Chemical Geology, 145 (1-2), 117-152, 1998.

Roddick, J., R. Cliff, and D. Rex, The evolution of excess argon n Alpine biotites - a 40Ar/39Ar

analysis, Earth and Planetary Science Letters, 48, 185-208, 1980.

Roddick, J.C., and O. van Breemen, U-Pb zircon dating; a comparison of ion microprobe and

single grain conventional analyses, in Current Research - Geological Survey of Canada;

Radiogenic age and isotopic studies; Report 8, pp. 1-9, 1994.

Rowley, D.B., Age of initiation of collision between India and Asia: A review of stratigraphic

data., Earth and Planetary Science Letters, 145, 1-13, 1996.

Page 167: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

152

Ruddiman, W.F., and J.E. Kutzbach, Forcing of the late Cenozoic northern hemisphere climate

by plateau uplift in southeast Asia and the American southwest., Journal of Geophysical

Research., 94, 18,409-18,427, 1989.

Schatz, J.F., and G. Simmons, Thermal Conductivity of Earth Materials at High Temperatures,

Journal of Geophysical Research, 77 (35), 6966-6983, 1972.

Schelling, D., The tectonostratigraphy and structure of the eastern Nepal Himalaya, Tectonics, 11

(5), 925-943, 1992.

Schelling, D., and K. Arita, Thrust tectonics, crustal shortening, and the structure of the far-

eastern Nepal Himalaya, Tectonics, 10 (5), 851-862., 1991.

Searle, M., Cooling History, erosion, exhumation, and kinematics of the Himalaya- Karakorum-

Tibet orogenic belt, Asian Tectonics, 110-137, 1996.

Searle, M., C. Corfield, B. Stephenson, and J. McCarron, Structure of the North Indian

continental margin in the Landakh-Zanskar Himalayas: implications for the timing of

obduction of the Spontang ophiolite, India-Asia collision and deformation events in the

Himalaya, Geol. Mag., 134 (3), 297-316, 1997.

Seeber, J.C. Armbruster, and J.C. Quittmeyer, Seismicity and continental subduction in the

Himalayan arc, in Zagros-Hindu Kush-Himalaya Geodynamic Evolution, edited by H.K.

Gupta, and F.M. Delaney, 1981.

Seeber, L., and V. Gornitz, River profiles along the Himalayan arc as indicators of active

tectonics, Tectonophys., 92, 335-367, 1983.

Stock, J.D., and D.R. Montgomery, Estimating palaeorelief from detrital mineral age ranges.,

Basin Research, 8 ((3)), 317-328., 1996.

Stüwe, K., L. White, and R. Brown, The influence of eroding topography on steady-state

isotherms. Application to fission track analysis, Earth and Planetary Science Letters, 124,

63-74, 1994.

Tapponier, P., G. Peltzer, and R. Armijo, On the Mechanics of the Collision between India and

Asia., in Collision Tectonics, edited by M.P. Coward, and A.C. Ries, Geological Society,

London, 1986.

Tapponier, P., G. Peltzer, A.Y. Ledain, and R.a.C. Armijo, P. , Propagating extrusion Tectonics

in Asia: New insights from Simple experiments in Plasticene, Geology, 10, 611-616, 1982.

Page 168: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

153

Van der Plas, L., and A.C. Tobi, A chart for judging the reliability of point counting results,

American Journal of Science, 263, 87-90, 1965.

Vannay, J.C., and K.V. Hodges, Tectonometamorphic evolution of the Himalayan metamorphic

core between Annapurna and Dhaulagiri, central Nepal, Journal of Metamorphic Geology,

14 (635-656), 1996.

Wager, Geographic Journal, 89, 239-250, 1937.

Whipple, K.X., and G.E. Tucker, Dynamics of the stream-power river incision model;

implications for height limits of mountain ranges, landscape response timescales, and

research needs, Journal of Geophysical Research, B, Solid Earth and Planets, 104 (8),

17,661-17,674, 1999.

Willet, S.D., Orogeny and orography; the effects of erosion on the structure of mountain belts,

Journal of Geophysical Research, B, Solid Earth and Planets, 104 (12), 28,957-28,982,

1999.

Yamanaka, H., and S. Iwata, River Terraces along the Middle Kali Gandaki and Marsyandi Khola,

Central Nepal, Nepal Geological Society, 2, 95-111, 1982.

Page 169: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

154

Appendix 1

1.0 Thermochronology

The following is intended as a brief overview of thermochronology for readers not familiar with the

basic techniques of 40Ar/39Ar dating. For further explanation, detailed experimental procedures may

be found in McDougall and Harrison [1988]. Renne [2000] provides a useful review, with emphasis

on Quaternary applications.

1.1 The decay equation

Isotopes of some elements are unstable such that they spontaneously disintegrate, undergoing

radioactive decay to form new elements. The statistically averaged rate of disintegration is constant

for a particular isotope, and will not vary with temperature or pressure. If the rate of disintegration

is known, one can use the ratio between the number of parent atoms (the original element present)

and the number of daughter atoms (produced by the decay of the parent atoms) to calculate the time

elapsed since decay initiated, and there were no daughter atoms present.

Page 170: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

155

The rate of decay of parent atoms into daughter atoms (dN/dt) is dependent upon the number of

radioactive atoms present (N) and the decay constant (λ), or half-life, which represents the rate of

disintegration.

λNdtdN

−=

(1)

This can be rearranged to give:

∫∫ =− dtN

dN λ

(2)

Once integrated:

CtN +−=− .ln λ

(3)

The constant of integration (C) can be evaluated from the condition that N = N0, when

t = 0:

0ln NC −=

(4)

Substituting equation 4 in equation 3, we obtain:

0ln.ln NtN −=− λ

(5)

Page 171: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

156

N is the number of parent atoms that remain at time (t) after the disintegration of the original

number of parent atoms (N0) began. N0 is typically calculated by measuring N and the number of

daughter (ND) atoms present.

DNNN +=0

(6)

Thus we can combine equation 5 and 6 and solve for t:

λ−+−

=)ln(ln DNNNt

(7)

1.2 The potassium/argon decay scheme

K4019 decays into both Ca40

20 by β- decay and into 40Ar by electron capture (with release of a neutrino

and γ radiation). The latter scheme represents only 11.16% of the K4019 decay, but is of most interest

to thermochronologists:

γν ++→+ − AreK 4018

4019

(8)

With 40K/40Ar dating, the parent isotope (potassium) and daughter isotope (argon) have to be

determined independently. Potassium may be measured using wet chemistry procedures (amongst

other techniques), whereas 40Ar is measured on a mass spectrometer.

Page 172: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

157

1.3 The 40Ar/39Ar analytical method

The 40Ar/39Ar technique was formalized by Merrihue and Turner, (1966). It is based on the same 40K/40Ar decay scheme, but follows different analytical procedure. The 40Ar/39Ar dating method is

more accurate than conventional 40K/40Ar dating because both isotopes can be measured at once,

and only the ratios rather than the exact amounts of isotopes are measured [see Faure, 1986]. It is

based upon the fact that K can be calculated from the amount of 39Ar produced during neutron

activation induced by irradiation.

1) Samples are irradiated to transform a portion of 39K to 39Ar (denoted as 39ArK):

pArnK k +→+ 3918

3919

(9)

where the amount of 39ArK produced is proportional to the amount of 39K in the sample.

2) As the ratio 39K/40K is known to be constant in nature, the amount of 40K in the sample can be

calculated. Thus, this in theory determines N, and ND (40Ar*; the * indicating a radiogenic source)

can be measured directly on the mass spectrometer. In reality, only the ratio of 40Ar to 39Ar is

needed to calculate the age:

+= 1.*ln.1

39

40

JArArt

(10)

As the exact amount of 39ArK produced during irradiation depends upon the duration of exposure

and neutron flux, a fluence correction factor (J) is calculated from dating monitors of known age.

These are typically placed at several locations within the irradiation package.

Page 173: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

158

1.4 Closure temperatures

Equation 7 is used to calculate a date: the time (t) elapsed since parent isotopes started to decay into

daughter isotopes, since N = N0. At this stage, however, it is useful to consider what our “date”

actually represents. With the 40K/40Ar decay system, our date records the time at which the daughter

isotope (Ar) starts to accumulate within the crystal, and this is a function of temperature. At high

temperatures the argon can escape rapidly out of the crystal by diffusion because of high lattice and

component (Ar) energy. Only when the crystal cools below a certain temperature, the “closure

temperature”, does the argon become trapped and the radioactive “clock” starts to record the

elapsed time. In reality, diffusion processes do not stop (approach zero) abruptly at a specific

temperature, but cease over a range of temperature: the closure zone. The closure temperature

represents the theoretical temperature that yields the same age as if closure occurred at a discrete

temperature [Dodson, 1973]. We take the closure temperature of muscovite to be ~350 ± 25oC, but

this will vary with grain size and the rate of cooling. The “cooling age” is the time elapsed since the

sample cooled through the closure temperature and reach the surface of the Earth.

Thermochrononology is primarily used to measure either the age of formation of a rock (typically

using geochronometers with closure temperatures close to the crystallization temperature of the

rock), or constrain the rate of cooling (typically using geochronometers with lower closure

temperatures to examine the erosion history the sample). To illustrate the effects of closure

temperatures on cooling ages, we can consider the history of a sedimentary rock that undergoes

regional metamorphism in a continent collision zone (Fig. 1). The rock is deposited at the surface of

the earth before being buried as crustal thickening proceeds during the collision (Fig. 1 i). As a

result of crustal thickening, the rock may undergo anatexis at high temperatures (Fig. 1 ii), resulting

in the formation S-type granites such as the High Himalayan leuco-granites observed in the

Himalaya. We can consider two decay schemes in two different minerals: U-Pb in monazite, and

K-Ar in muscovite. The U-Pb system typically has higher closure temperatures because lead is

relatively immobile. Thus the time at which the rock passes through the U-Pb closure isotherm in

monazite (Fig. 1, iii) records the early crystallization phase as the granite begins to cool (providing

it is reset by burial). As the orogen undergoes erosion, rocks approach the surface and experience

progressive cooling. When the rock cools through the K-Ar closure isotherm for muscovite (Fig. 1,

Page 174: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

159

iv), the rock has already undergone significant exhumation and is at ~350°C. It continues to cool

and is sampled at the surface today (Fig. 1, v).

As we know the age of a sample (tc-tp) and can calculate the depth of the closure isotherm (zx-zc)

using an assumed geothermal gradient, we can use the cooling age as a proxy for the erosion rate

(dz/dt). From the arrows over the time axis of figure 1, we can see that the cooling age recorded by

U-Pb decay in monazite and K-Ar decay in muscovite integrate the cooling rate over very different

time periods. Thus, for the applications in this paper, trying to constrain the short-to-medium term

erosion history of an orogen, we use the lower-temperature geochronometer: the K-Ar series in

muscovite.

Page 175: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

300

400

200

100

0

500

600

700

800

time b.p., t K-Ar msc age

Temperature (oC)

appr

oxim

ate

dept

h, z

(K

m)

10

20

FT apatite

FT zirconK-Ar biotite

K-Ar muscovite

K-Ar hornblende

U-Pb monazite

Helium

Geochronometerr

Bur

ial

Exhum

ation

ZONE OF ANATEXIS

U-Pb mzt age

rock trajectory

(i)

(ii)(iii)

(iv)

(v)

Figure 1. Diagram illustrating the T-t path of a rock particle undergoing burial metamorphism and subsequent exhumation. With a geothermal gradient of 35oC the closure temperature of muscovite (msc), ~350oC, occurs at 10 km (zc). Thus, if we have a cooling age (tc) of 10 Ma, for example, an erosion rate can be calculated: 10 km in 10 My = 1.0 km/My. U-Pb in monazite (mzt) has a higher closure temperature, and therefore records a longer cooling histroy than K-Ar in muscovite. Other geochronomters are shown on the left axis at the approximate temperature corresponding to their closure. Fission-track (FT) geochronometers are illustrated for reference, but are not isotope based.

160

Zx

Zc

tp tc 0

Page 176: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

161

Appendix 2

1.0 40Ar/39Ar results and protocols

Muscovite mineral separates were irradiation at the McMaster University research reactor. The

irradiation package included aliquots of the neutron-fluence monitor Fish Canyon sanidine (28.02

Ma, Renne et al. [1998]), as well as a variety of salts that served as monitors for interfering nuclear

reactions. The muscovites and monitors were analyzed at the 40Ar/39Ar laser microprobe facility at

MIT [Hodges and Bowring, 1995]. Gas was extracted from individual mica crystals by fusion in the

defocused beam of an Argon laser. This operated at 18 W for a period of approximately 10 seconds.

The extracted gas was analyzed on an MAP 215-50 mass spectrometer, using a Johnston electron

multiplier, after purification to remove reactive species. At the beginning of each analytical session

and after every tenth analysis of an unknown, the total system blanks were measured. Apparent ages

(dates) calculated for each muscovite are reported in Table 1 with an estimated 2-σ uncertainty

obtained by propagating all analytical uncertainties. In order to illustrate the proportion of this

uncertainty that is attributable to uncertainties in the neutron flux during sample irradiation, age

uncertainties are shown with and without propagated error in the irradiation parameter J (Table 1,

iii). The Number of moles of K-derived 39Ar (39ArK) released during fusion (Table 1, i) and the

percentage of radiogenic 40Ar (40Ar*) in the total 40Ar for each analysis (Table 1, ii) are also

reported.

For all geological analysis, a lower cutoff of 40% 40Ar* was used to eliminate spurious data (Fig.

1). Grains yielding less than 40% 40Ar* have little radiogenic argon, in comparison to 40Ar

contained within the mass spectrometer blank, due to: a) incomplete ablation; b) impure muscovite,

or; c) very young ages. Many of the grains rejected were from the first run of sample S-24 that

Page 177: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

162

experienced some analysis problems. The very young ages observed in the sample are not found in

any of the catchment signals and so were deemed geologically unreasonable. The data have a strong

linear dependence of percentage 40Ar with age below the cutoff (Fig. 3), whilst they are more

scattered above the cutoff. The age of different samples should not be directly proportional to the

amount of 40Ar* because grains of the same age will have 40Ar concentrations in proportion to the

amount of 40K that the mineral initially contained. Based on this interpretation, it might be argued

that this cutoff could occur anywhere between 25% and 60% radiogenic 40Ar. We felt that a 40%

cutoff was the best solution as spurious ages were removed while not cutting too many younger

grains from the results.

Page 178: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

0 5 10 15 20 25 300

10

20

30

40

50

60

70

80

90

100

Age (Ma)

%40

Ar

(rad

ioge

nic)

40% 40Ar cutoff

range in plausible %40Ar cutoffs

Figure 1. Age versus percentage of radiogenic 40Ar for the geochronological analyses presented in this paper. To eliminate spurious ages, a number of plausible possibilities (within the range of the dashed lines) were investigated to find a lower limit of radiogenic 40Ar. A 40% cutoff was judged the most appropriate, and analyses with less radiogenic 40Ar (within the shaded area) were rejected.

163

Page 179: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

164

Sample grain 36Ar/40Ar 39Ar/40Ar 39Ark 40Ar* Age

(x10-4) (x10-1) (x10-14moles)(i) (%)(ii) (Myr)(iii) (with J) (w/o J)

NIB-S2 1 5.64 ± 0.65 6.28 ± 0.32 7.760 83.1 8.64 ± 0.56 0.56

NIB-S2 2 3.07 ± 0.48 5.86 ± 0.44 8.804 90.7 10.09 ± 0.84 0.84

NIB-S2 3 7.13 ± 1.20 7.60 ± 0.45 3.591 78.6 6.77 ± 0.58 0.58

NIB-S2 4 9.00 ± 2.51 8.95 ± 0.17 2.765 73.1 5.34 ± 0.56 0.56

NIB-S2 5 4.21 ± 0.54 6.36 ± 0.36 5.095 87.3 8.96 ± 0.60 0.60

NIB-S2 6 3.57 ± 0.46 5.47 ± 0.33 6.388 89.2 10.64 ± 0.73 0.73

NIB-S2 7 8.47 ± 1.75 6.89 ± 0.12 2.536 74.7 7.08 ± 0.51 0.51

NIB-S2 8 10.04 ± 3.34 6.84 ± 0.32 1.293 70.1 6.70 ± 1.03 1.03

NIB-S2 9 5.58 ± 1.53 6.11 ± 0.21 1.895 83.3 8.89 ± 0.60 0.60

NIB-S2 10 11.57 ± 5.34 10.17 ± 0.76 3.077 65.5 4.22 ± 1.11 1.11

NIB-S2 11 11.74 ± 4.91 6.60 ± 0.41 1.959 65.1 6.44 ± 1.54 1.54

NIB-S2 12 10.12 ± 3.74 7.26 ± 0.25 3.131 69.8 6.29 ± 1.03 1.03

NIB-S2 13 21.04 ± 8.69 9.97 ± 0.66 1.313 37.6 2.47 ± 1.71 1.71

NIB-S2 14 8.00 ± 7.10 6.48 ± 0.13 1.519 76.1 7.66 ± 2.11 2.11

NIB-S2 15 8.36 ± 3.92 7.14 ± 0.16 2.447 75.0 6.87 ± 1.07 1.07

NIB-S2 16 7.03 ± 3.65 5.38 ± 0.13 2.390 79.0 9.58 ± 1.34 1.33

NIB-S2 17 8.87 ± 3.41 7.19 ± 0.23 3.029 73.5 6.68 ± 0.95 0.95

NIB-S2 18 9.09 ± 7.34 7.42 ± 0.22 1.522 72.9 6.42 ± 1.92 1.92

NIB-S2 19 5.82 ± 4.39 6.99 ± 0.18 2.279 82.5 7.71 ± 1.23 1.23

NIB-S2 20 7.14 ± 4.94 7.78 ± 0.22 2.232 78.6 6.60 ± 1.24 1.24

NIB-S2 21 11.49 ± 14.81 6.41 ± 0.20 0.841 65.8 6.71 ± 4.44 4.44

NIB-S2 22 14.94 ± 6.18 8.49 ± 0.27 1.925 55.6 4.29 ± 1.42 1.42

NIB-S2 23 11.83 ± 7.61 7.62 ± 0.17 1.370 64.8 5.56 ± 1.93 1.93

NIB-S2 24 8.10 ± 4.83 11.58 ± 0.08 2.707 75.6 4.28 ± 0.80 0.80

NIB-S3 1 10.82 ± 10.02 8.82 ± 0.36 8.000 67.7 5.02 ± 2.20 2.20

NIB-S3 2 3.60 ± 3.62 3.45 ± 0.07 8.604 89.2 16.81 ± 2.04 2.04

NIB-S3 3 16.94 ± 21.41 7.81 ± 0.53 3.317 49.8 4.17 ± 5.30 5.30

NIB-S3 4 32.45 ± 57.47 8.67 ± 1.47 1.382 4.1 0.31 ± 12.78 12.78

NIB-S3 5 5.43 ± 8.39 4.58 ± 0.18 4.955 83.8 11.91 ± 3.55 3.55

NIB-S3 6 5.29 ± 9.53 4.09 ± 0.18 3.862 84.2 13.40 ± 4.50 4.50

NIB-S3 7 13.71 ± 14.14 7.89 ± 0.46 5.013 59.3 4.91 ± 3.47 3.47

NIB-S3 8 10.60 ± 18.80 4.67 ± 0.32 2.236 68.5 9.57 ± 7.77 7.77

Page 180: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

165

NIB-S3 9 18.35 ± 26.63 8.24 ± 0.76 2.786 45.6 3.62 ± 6.25 6.25

NIB-S3 10 5.87 ± 6.54 4.64 ± 0.20 6.384 82.5 11.58 ± 2.77 2.76

NIB-S3 11 11.22 ± 17.54 3.60 ± 0.26 1.862 66.7 12.08 ± 9.41 9.41

NIB-S3 12 19.11 ± 31.17 10.50 ± 1.08 3.021 43.3 2.70 ± 5.74 5.74

NIB-S3 13 4.17 ± 8.21 3.50 ± 0.16 3.822 87.5 16.26 ± 4.56 4.56

NIB-S3 14 6.76 ± 7.31 4.28 ± 0.20 5.301 79.8 12.15 ± 3.34 3.34

NIB-S3 15 4.67 ± 8.23 3.32 ± 0.16 3.637 86.1 16.84 ± 4.82 4.82

NIB-S3 16 20.85 ± 48.49 10.72 ± 1.54 2.000 38.2 2.33 ± 8.73 8.73

NIB-S3 17 5.51 ± 4.52 10.22 ± 0.28 20.366 83.3 5.34 ± 0.87 0.87

NIB-S3 18 9.47 ± 17.26 11.41 ± 0.87 5.956 71.6 4.11 ± 2.94 2.94

NIB-S3 19 16.28 ± 24.93 9.52 ± 0.88 3.468 51.7 3.55 ± 5.07 5.07

NIB-S3 20 7.56 ± 11.69 4.85 ± 0.34 3.729 77.5 10.42 ± 4.71 4.71

NIB-S3 21 4.81 ± 1.95 8.62 ± 0.07 2.404 85.4 6.48 ± 0.44 0.44

NIB-S3 22 3.13 ± 0.22 4.01 ± 0.03 10.537 90.6 14.70 ± 0.18 0.16

NIB-S3 23 2.28 ± 0.29 3.50 ± 0.04 4.678 93.1 17.28 ± 0.29 0.27

NIB-S3 24 0.95 ± 0.15 3.89 ± 0.05 8.745 97.0 16.21 ± 0.26 0.24

NIB-S3 25 2.86 ± 0.09 4.68 ± 0.04 13.589 91.3 12.72 ± 0.13 0.11

NIB-S3 26 4.61 ± 0.48 9.10 ± 0.12 5.029 86.0 6.18 ± 0.14 0.14

NIB-S3 27 1.70 ± 0.23 3.28 ± 0.03 4.390 94.8 18.77 ± 0.25 0.23

NIB-S3 28 11.78 ± 1.15 9.80 ± 0.05 3.541 64.9 4.34 ± 0.23 0.23

NIB-S3 29 1.27 ± 0.19 3.20 ± 0.05 5.506 96.1 19.48 ± 0.33 0.31

NIB-S3 30 0.95 ± 0.14 3.40 ± 0.05 8.499 97.0 18.57 ± 0.29 0.27

NIB-S3 31 4.63 ± 0.64 12.08 ± 0.15 6.143 85.8 4.66 ± 0.12 0.12

NIB-S3 32 0.82 ± 0.19 3.59 ± 0.02 6.124 97.4 17.64 ± 0.19 0.16

NIB-S3 33 1.84 ± 0.11 3.28 ± 0.04 6.743 94.4 18.70 ± 0.27 0.25

NIB-S3 34 3.50 ± 0.77 13.56 ± 0.09 5.000 89.1 4.31 ± 0.12 0.11

NIB-S3 35 5.42 ± 0.44 11.62 ± 0.17 15.101 83.5 4.71 ± 0.11 0.11

NIB-S3 36 1.36 ± 0.17 3.61 ± 0.06 8.951 95.8 17.25 ± 0.31 0.30

NIB-S3 37 1.52 ± 0.32 3.57 ± 0.05 5.560 95.3 17.35 ± 0.32 0.30

NIB-S3 38 1.80 ± 0.19 3.38 ± 0.06 5.269 94.5 18.16 ± 0.36 0.34

NIB-S3 39 2.64 ± 0.53 5.39 ± 0.03 4.636 92.0 11.12 ± 0.21 0.20

NIB-S3 40 0.91 ± 0.25 3.24 ± 0.07 6.554 97.2 19.49 ± 0.48 0.47

NIB-S3 41 3.04 ± 0.20 3.53 ± 0.15 5.865 90.8 16.76 ± 0.78 0.77

NIB-S3 42 10.36 ± 1.47 6.52 ± 0.75 12.855 69.2 6.93 ± 1.18 1.18

NIB-S3 43 3.89 ± 0.69 3.68 ± 0.43 7.647 88.4 15.63 ± 2.08 2.08

Page 181: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

166

NIB-S3 44 6.97 ± 1.20 3.32 ± 0.50 4.565 79.3 15.52 ± 2.93 2.93

NIB-S3 45 13.38 ± 3.24 10.49 ± 1.99 10.773 60.2 3.76 ± 1.23 1.23

NIB-S3 46 11.30 ± 2.29 6.91 ± 1.14 5.997 66.4 6.27 ± 1.60 1.60

NIB-S3 47 3.83 ± 0.33 3.15 ± 0.16 7.119 88.6 18.30 ± 1.09 1.08

NIB-S3 48 13.88 ± 2.32 9.18 ± 0.86 4.980 58.7 4.19 ± 0.78 0.78

NIB-S3 49 3.67 ± 0.41 3.39 ± 0.20 5.982 89.0 17.05 ± 1.12 1.11

NIB-S3 50 3.67 ± 0.45 3.58 ± 0.09 5.223 89.0 16.17 ± 0.53 0.53

NIB-S5 1 3.14 ± 0.87 9.81 ± 0.18 11.799 90.3 5.98 ± 0.22 0.21

NIB-S5 2 5.97 ± 2.75 13.13 ± 0.44 8.808 81.8 4.06 ± 0.43 0.43

NIB-S5 3 6.45 ± 2.50 8.52 ± 0.25 5.395 80.6 6.15 ± 0.60 0.60

NIB-S5 4 4.85 ± 1.99 8.40 ± 0.24 7.480 85.3 6.60 ± 0.51 0.50

NIB-S5 5 5.87 ± 1.28 12.46 ± 0.25 11.484 82.1 4.30 ± 0.23 0.22

NIB-S5 6 3.07 ± 2.45 10.26 ± 0.24 8.469 90.5 5.74 ± 0.48 0.48

NIB-S5 7 10.79 ± 1.18 9.25 ± 0.18 18.189 67.8 4.77 ± 0.28 0.27

NIB-S5 8 14.90 ± 1.10 4.99 ± 0.07 13.785 55.8 7.25 ± 0.45 0.45

NIB-S5 9 21.36 ± 7.99 9.06 ± 0.88 1.638 36.7 2.64 ± 1.77 1.77

NIB-S5 10 6.03 ± 1.87 8.91 ± 0.42 4.718 81.8 5.97 ± 0.53 0.52

NIB-S5 11 5.64 ± 2.69 13.27 ± 0.57 5.499 82.8 4.07 ± 0.44 0.44

NIB-S5 12 4.67 ± 2.43 8.48 ± 0.30 6.285 85.8 6.58 ± 0.61 0.61

NIB-S5 13 23.73 ± 5.47 7.63 ± 0.40 1.420 29.8 2.54 ± 1.41 1.41

NIB-S5 14 10.35 ± 7.82 12.11 ± 0.70 1.643 69.0 3.71 ± 1.27 1.27

NIB-S5 15 5.13 ± 2.10 10.61 ± 0.73 5.111 84.4 5.18 ± 0.56 0.56

NIB-S5 16 12.19 ± 9.76 11.44 ± 0.68 1.717 63.6 3.62 ± 1.66 1.66

NIB-S5 17 9.79 ± 5.04 8.41 ± 0.30 1.901 70.8 5.47 ± 1.18 1.18

NIB-S5 18 0.30 ± 13.14 15.60 ± 1.33 1.545 98.4 4.11 ± 1.65 1.65

NIB-S5 19 2.81 ± 0.97 6.89 ± 0.30 7.637 91.4 8.61 ± 0.49 0.48

NIB-S5 20 31.46 ± 7.82 13.02 ± 0.19 2.137 7.0 0.35 ± 1.15 1.15

NIB-S5 1 11.59 ± 11.68 12.66 ± 0.29 5.907 65.3 3.36 ± 1.77 1.77

NIB-S5 2 9.87 ± 7.29 12.94 ± 0.71 9.884 70.4 3.54 ± 1.11 1.11

NIB-S5 3 19.86 ± 17.36 9.63 ± 0.23 3.023 41.1 2.78 ± 3.45 3.45

NIB-S5 4 14.11 ± 12.32 9.90 ± 0.57 4.385 58.0 3.81 ± 2.40 2.40

NIB-S5 5 43.41 ± 43.42 14.05 ± 0.77 1.757 0.0 0.00 ± 5.92 5.92

NIB-S5 6 16.47 ± 12.44 8.60 ± 0.36 3.758 51.1 3.86 ± 2.78 2.78

NIB-S5 7 11.50 ± 5.57 13.55 ± 0.70 13.512 65.6 3.16 ± 0.82 0.82

NIB-S5 8 10.81 ± 4.55 9.48 ± 0.16 12.143 67.7 4.65 ± 0.92 0.92

Page 182: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

167

NIB-S5 9 20.34 ± 18.75 10.79 ± 0.64 3.131 39.7 2.40 ± 3.34 3.34

NIB-S5 10 19.77 ± 14.84 9.94 ± 0.64 3.741 41.4 2.71 ± 2.88 2.88

NIB-S5 11 6.10 ± 5.92 8.01 ± 0.12 7.360 81.6 6.62 ± 1.42 1.42

NIB-S5 12 12.22 ± 12.44 14.20 ± 0.45 6.195 63.5 2.92 ± 1.68 1.68

NIB-S5 13 19.28 ± 22.50 8.44 ± 0.28 2.041 42.9 3.30 ± 5.10 5.10

NIB-S5 14 15.56 ± 15.15 7.87 ± 0.18 2.825 53.8 4.44 ± 3.68 3.68

NIB-S5 15 11.95 ± 12.96 10.37 ± 0.23 4.344 64.3 4.04 ± 2.39 2.39

NIB-S5 16 15.71 ± 15.07 10.77 ± 0.26 3.904 53.3 3.22 ± 2.68 2.68

NIB-S5 17 35.61 ± 46.91 11.83 ± 0.70 1.369 0.0 0.00 ± 7.60 7.60

NIB-S5 18 9.92 ± 5.78 7.15 ± 0.22 6.818 70.4 6.40 ± 1.57 1.57

NIB-S5 19 11.99 ± 11.58 8.45 ± 0.27 4.000 64.3 4.95 ± 2.63 2.63

NIB-S5 20 16.52 ± 17.56 9.82 ± 0.27 3.036 50.9 3.38 ± 3.42 3.42

NIB-S6 1 0.42 ± 1.22 3.61 ± 0.17 3.280 98.6 17.66 ± 1.07 1.05

NIB-S6 2 2.47 ± 1.64 6.27 ± 0.50 3.669 92.4 9.56 ± 0.97 0.96

NIB-S6 3 12.32 ± 5.93 5.93 ± 0.19 1.612 63.4 6.94 ± 1.94 1.94

NIB-S6 4 1.10 ± 0.71 3.59 ± 0.19 5.279 96.6 17.37 ± 1.04 1.02

NIB-S6 5 1.91 ± 1.08 3.52 ± 0.17 3.753 94.2 17.30 ± 1.08 1.06

NIB-S6 6 3.74 ± 1.90 3.06 ± 0.15 3.202 88.8 18.72 ± 1.57 1.55

NIB-S6 7 2.01 ± 1.88 3.64 ± 0.17 2.688 93.9 16.68 ± 1.29 1.27

NIB-S6 8 0.85 ± 0.54 3.45 ± 0.17 6.855 97.3 18.21 ± 0.99 0.96

NIB-S6 9 0.47 ± 2.68 3.85 ± 0.16 1.653 98.4 16.52 ± 1.51 1.50

NIB-S6 10 2.18 ± 1.34 3.96 ± 0.21 4.928 93.4 15.26 ± 1.08 1.06

NIB-S6 11 0.37 ± 2.72 4.85 ± 0.18 2.395 98.7 13.18 ± 1.19 1.18

NIB-S6 12 0.14 ± 0.98 3.54 ± 0.18 4.118 99.4 18.13 ± 1.09 1.07

NIB-S6 13 48.18 ± 12.66 13.40 ± 0.66 1.382 0.0 0.00 ± 1.79 1.79

NIB-S6 14 0.11 ± 3.32 3.79 ± 0.17 1.257 99.5 16.96 ± 1.83 1.82

NIB-S6 15 8.29 ± 1.77 3.65 ± 0.08 2.184 75.4 13.37 ± 1.01 1.00

NIB-S6 16 0.43 ± 3.51 4.06 ± 0.10 1.350 98.5 15.71 ± 1.70 1.69

NIB-S6 17 1.96 ± 12.51 4.99 ± 0.16 0.471 94.0 12.19 ± 4.79 4.79

NIB-S6 18 3.59 ± 4.06 10.60 ± 0.70 2.560 88.9 5.46 ± 0.84 0.83

NIB-S6 19 0.27 ± 2.18 3.49 ± 0.09 2.253 99.0 18.36 ± 1.29 1.28

NIB-S6 20 9.95 ± 2.10 3.60 ± 0.09 1.796 70.5 12.68 ± 1.19 1.18

NIB-S6 21 3.71 ± 0.31 3.15 ± 0.03 3.443 88.9 18.24 ± 0.34 0.27

NIB-S6 22 1.78 ± 0.22 3.40 ± 0.08 8.769 94.6 17.96 ± 0.50 0.46

NIB-S6 23 2.07 ± 0.24 3.81 ± 0.02 6.270 93.7 15.91 ± 0.23 0.15

Page 183: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

168

NIB-S6 24 3.85 ± 0.90 13.26 ± 0.18 6.579 88.0 4.33 ± 0.15 0.15

NIB-S6 25 7.81 ± 2.89 10.61 ± 0.22 1.477 76.5 4.69 ± 0.54 0.53

NIB-S6 26 5.75 ± 0.43 3.01 ± 0.02 5.762 82.9 17.79 ± 0.36 0.30

NIB-S6 27 2.66 ± 0.47 3.42 ± 0.05 3.279 92.0 17.36 ± 0.41 0.36

NIB-S6 28 1.42 ± 0.18 3.24 ± 0.06 7.298 95.6 19.09 ± 0.45 0.39

NIB-S6 29 0.77 ± 0.11 3.54 ± 0.03 8.725 97.6 17.83 ± 0.27 0.18

NIB-S6 30 1.21 ± 0.58 3.21 ± 0.03 4.320 96.3 19.39 ± 0.46 0.40

NIB-S6 31 4.19 ± 1.39 6.89 ± 0.06 2.342 87.3 8.23 ± 0.40 0.39

NIB-S6 32 2.04 ± 0.55 4.46 ± 0.03 2.951 93.8 13.60 ± 0.30 0.26

NIB-S6 33 0.86 ± 0.24 3.47 ± 0.04 7.301 97.3 18.14 ± 0.31 0.23

NIB-S6 34 3.35 ± 0.25 7.80 ± 0.14 10.005 89.8 7.48 ± 0.18 0.16

NIB-S6 35 1.69 ± 0.26 3.27 ± 0.02 5.578 94.9 18.74 ± 0.29 0.20

NIB-S6 36 1.07 ± 0.29 3.24 ± 0.04 5.326 96.7 19.29 ± 0.35 0.28

NIB-S6 37 1.33 ± 0.38 3.40 ± 0.07 3.568 95.9 18.23 ± 0.47 0.43

NIB-S6 38 2.27 ± 0.34 4.87 ± 0.03 5.639 93.1 12.38 ± 0.21 0.16

NIB-S6 39 0.94 ± 0.31 3.44 ± 0.07 7.076 97.0 18.21 ± 0.47 0.42

NIB-S6 40 5.22 ± 1.44 11.33 ± 0.08 4.799 84.1 4.83 ± 0.25 0.25

NIB-S6 41 4.24 ± 0.62 9.55 ± 0.08 5.042 87.1 5.93 ± 0.15 0.14

NIB-S6 42 4.14 ± 0.24 6.75 ± 0.12 13.783 87.5 8.41 ± 0.20 0.18

NIB-S6 43 2.25 ± 0.27 3.49 ± 0.05 5.085 93.2 17.24 ± 0.36 0.30

NIB-S6 44 1.34 ± 0.33 3.14 ± 0.03 4.578 95.9 19.69 ± 0.37 0.29

NIB-S6 45 2.15 ± 0.32 4.30 ± 0.09 8.379 93.5 14.08 ± 0.37 0.33

NIB-S6 46 1.40 ± 0.49 3.18 ± 0.04 4.364 95.7 19.44 ± 0.46 0.41

NIB-S6 47 3.72 ± 1.08 5.66 ± 0.02 2.836 88.8 10.18 ± 0.39 0.37

NIB-S6 48 2.51 ± 0.53 3.27 ± 0.03 3.246 92.4 18.26 ± 0.42 0.37

NIB-S6 49 2.16 ± 0.40 4.21 ± 0.08 4.746 93.4 14.35 ± 0.37 0.34

NIB-S6 50 2.75 ± 0.54 4.44 ± 0.04 4.945 91.7 13.35 ± 0.31 0.27

NIB-S8 1 1.17 ± 0.18 3.38 ± 0.19 10.033 96.4 18.41 ± 1.09 1.07

NIB-S8 2 2.65 ± 0.34 3.30 ± 0.12 5.779 92.0 18.03 ± 0.77 0.74

NIB-S8 3 3.63 ± 0.86 3.45 ± 0.42 11.365 89.1 16.71 ± 2.31 2.30

NIB-S8 4 1.45 ± 0.17 3.51 ± 0.19 10.690 95.5 17.59 ± 1.02 1.00

NIB-S8 5 1.49 ± 0.21 3.28 ± 0.21 7.991 95.4 18.77 ± 1.27 1.25

NIB-S8 6 4.02 ± 0.35 3.28 ± 0.21 6.238 88.0 17.34 ± 1.25 1.24

NIB-S8 7 0.92 ± 0.23 3.56 ± 0.14 10.337 97.1 17.62 ± 0.73 0.71

NIB-S8 8 1.02 ± 0.20 3.36 ± 0.19 8.765 96.8 18.61 ± 1.10 1.08

Page 184: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

169

NIB-S8 9 2.46 ± 0.54 3.45 ± 0.13 3.931 92.6 17.33 ± 0.77 0.74

NIB-S8 10 1.15 ± 0.40 3.65 ± 0.13 6.942 96.4 17.09 ± 0.69 0.66

NIB-S8 11 2.63 ± 0.45 3.71 ± 0.30 11.485 92.1 16.05 ± 1.45 1.43

NIB-S8 12 1.50 ± 0.25 3.48 ± 0.19 9.587 95.4 17.70 ± 1.04 1.02

NIB-S8 13 2.45 ± 0.25 3.60 ± 0.19 6.635 92.6 16.61 ± 0.97 0.95

NIB-S8 14 2.01 ± 0.31 3.47 ± 0.16 9.348 93.9 17.49 ± 0.87 0.85

NIB-S8 15 2.43 ± 0.33 3.62 ± 0.16 8.191 92.7 16.55 ± 0.83 0.81

NIB-S8 16 1.14 ± 0.24 3.44 ± 0.19 6.462 96.5 18.10 ± 1.07 1.05

NIB-S8 17 2.17 ± 0.33 3.80 ± 0.10 7.448 93.4 15.89 ± 0.50 0.47

NIB-S8 18 2.00 ± 0.32 3.66 ± 0.24 8.758 93.9 16.60 ± 1.17 1.16

NIB-S8 19 1.47 ± 0.20 3.54 ± 0.10 8.720 95.5 17.43 ± 0.57 0.53

NIB-S8 20 1.21 ± 0.44 3.69 ± 0.18 4.093 96.2 16.85 ± 0.89 0.86

NIB-S8 21 3.26 ± 5.52 3.70 ± 0.19 4.293 90.2 15.75 ± 2.98 2.97

NIB-S8 22 4.74 ± 3.66 3.30 ± 0.09 5.930 85.9 16.83 ± 2.18 2.17

NIB-S8 23 6.38 ± 2.31 3.04 ± 0.06 8.623 81.0 17.22 ± 1.52 1.50

NIB-S8 24 3.64 ± 2.89 3.49 ± 0.12 7.674 89.1 16.50 ± 1.71 1.70

NIB-S8 25 3.92 ± 5.78 3.56 ± 0.18 3.919 88.3 16.02 ± 3.22 3.22

NIB-S8 26 3.39 ± 3.67 5.21 ± 0.22 9.067 89.7 11.17 ± 1.44 1.44

NIB-S8 27 2.79 ± 4.35 3.45 ± 0.16 5.056 91.6 17.18 ± 2.56 2.55

NIB-S8 28 2.64 ± 3.26 3.87 ± 0.17 7.795 92.0 15.38 ± 1.77 1.76

NIB-S8 29 4.78 ± 6.51 3.54 ± 0.23 3.483 85.7 15.68 ± 3.69 3.69

NIB-S8 30 1.73 ± 2.41 3.39 ± 0.06 8.984 94.7 18.07 ± 1.41 1.40

NIB-S8 31 3.59 ± 3.03 3.26 ± 0.12 6.868 89.2 17.69 ± 1.92 1.91

NIB-S8 32 3.87 ± 4.93 3.40 ± 0.14 4.397 88.4 16.81 ± 2.87 2.86

NIB-S8 33 2.64 ± 4.76 3.26 ± 0.15 4.390 92.1 18.26 ± 2.93 2.93

NIB-S8 34 5.72 ± 7.70 3.22 ± 0.17 2.789 83.0 16.65 ± 4.65 4.65

NIB-S8 35 2.40 ± 4.26 3.60 ± 0.12 5.358 92.8 16.64 ± 2.33 2.32

NIB-S9 1 5.85 ± 0.42 3.02 ± 0.05 6.157 82.6 17.56 ± 0.46 0.42

NIB-S9 2 7.14 ± 2.51 3.98 ± 0.32 3.525 78.7 12.71 ± 1.73 1.73

NIB-S9 3 9.09 ± 0.44 2.56 ± 0.03 5.610 73.0 18.33 ± 0.46 0.41

NIB-S9 4 5.75 ± 0.47 3.25 ± 0.01 6.019 82.9 16.38 ± 0.34 0.28

NIB-S9 5 7.49 ± 0.41 3.41 ± 0.03 4.449 77.7 14.65 ± 0.34 0.29

NIB-S9 6 3.98 ± 0.28 3.06 ± 0.02 6.356 88.1 18.50 ± 0.31 0.22

NIB-S9 7 2.19 ± 0.48 3.42 ± 0.03 6.617 93.4 17.56 ± 0.38 0.32

NIB-S9 8 4.29 ± 0.48 3.21 ± 0.02 2.572 87.2 17.42 ± 0.38 0.32

Page 185: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

170

NIB-S9 9 3.53 ± 0.25 3.23 ± 0.05 5.307 89.4 17.77 ± 0.41 0.36

NIB-S9 10 7.64 ± 0.35 2.78 ± 0.04 5.692 77.3 17.83 ± 0.43 0.38

NIB-S12 1 6.92 ± 3.24 3.63 ± 0.08 8.576 79.4 14.06 ± 1.73 1.73

NIB-S12 2 4.71 ± 4.33 3.45 ± 0.05 6.025 85.9 15.99 ± 2.39 2.38

NIB-S12 3 8.74 ± 8.32 4.00 ± 0.12 3.642 74.0 11.90 ± 3.96 3.96

NIB-S12 4 3.83 ± 4.74 4.44 ± 0.14 7.147 88.5 12.83 ± 2.07 2.07

NIB-S12 5 12.08 ± 12.29 3.93 ± 0.24 2.423 64.2 10.50 ± 5.99 5.99

NIB-S12 6 0.51 ± 0.49 4.67 ± 0.18 3.967 98.3 13.53 ± 0.58 0.56

NIB-S12 7 1.66 ± 0.35 4.37 ± 0.18 8.458 94.9 13.97 ± 0.64 0.62

NIB-S12 8 3.82 ± 0.44 3.94 ± 0.10 6.132 88.5 14.44 ± 0.48 0.45

NIB-S12 9 2.12 ± 0.37 4.54 ± 0.09 8.846 93.5 13.26 ± 0.35 0.31

NIB-S12 10 1.82 ± 0.26 3.89 ± 0.08 7.595 94.4 15.61 ± 0.41 0.37

NIB-S12 11 2.66 ± 0.40 4.05 ± 0.07 8.246 92.0 14.58 ± 0.36 0.32

NIB-S12 12 2.22 ± 0.34 3.38 ± 0.10 7.074 93.3 17.70 ± 0.61 0.58

NIB-S12 13 3.62 ± 0.68 4.03 ± 0.11 5.776 89.1 14.23 ± 0.57 0.54

NIB-S12 14 6.04 ± 0.52 3.49 ± 0.03 3.213 82.0 15.10 ± 0.36 0.31

NIB-S12 15 4.60 ± 0.88 3.29 ± 0.03 1.689 86.3 16.84 ± 0.57 0.54

NIB-S12 16 3.44 ± 0.47 4.23 ± 0.08 8.163 89.6 13.64 ± 0.40 0.37

NIB-S12 17 2.73 ± 0.51 4.09 ± 0.03 8.756 91.7 14.41 ± 0.31 0.26

NIB-S12 18 4.93 ± 0.77 4.14 ± 0.11 4.046 85.3 13.25 ± 0.54 0.52

NIB-S12 19 3.01 ± 0.50 3.87 ± 0.03 6.367 90.9 15.09 ± 0.32 0.28

NIB-S12 20 4.70 ± 0.53 3.98 ± 0.13 6.174 85.9 13.90 ± 0.59 0.56

NIB-S12 21 3.51 ± 0.45 3.25 ± 0.03 3.669 89.5 17.70 ± 0.38 0.32

NIB-S12 22 6.97 ± 0.64 3.68 ± 0.02 2.357 79.2 13.86 ± 0.38 0.35

NIB-S12 23 4.00 ± 0.38 3.94 ± 0.01 4.656 88.0 14.35 ± 0.25 0.19

NIB-S12 24 2.26 ± 0.28 4.05 ± 0.11 7.697 93.1 14.78 ± 0.48 0.45

NIB-S12 25 2.58 ± 0.55 3.95 ± 0.07 8.216 92.2 14.99 ± 0.43 0.40

NIB-S24 1 17.56 ± 61.51 8.61 ± 1.60 0.582 47.9 3.57 ± 13.53 13.53

NIB-S24 2 9.94 ± 19.30 10.53 ± 0.74 2.216 70.3 4.29 ± 3.48 3.48

NIB-S24 3 7.60 ± 43.69 10.80 ± 1.49 0.995 77.1 4.59 ± 7.67 7.67

NIB-S24 4 13.61 ± 11.46 6.10 ± 0.28 2.217 59.6 6.26 ± 3.57 3.57

NIB-S24 5 7.66 ± 13.40 7.45 ± 0.40 2.299 77.1 6.63 ± 3.42 3.42

NIB-S24 6 19.83 ± 51.39 8.12 ± 1.32 0.635 41.2 3.26 ± 11.99 11.99

NIB-S24 7 9.25 ± 16.02 8.43 ± 0.61 2.134 72.4 5.51 ± 3.62 3.62

NIB-S24 8 39.45 ± 33.94 6.83 ± 0.86 0.834 0.0 0.00 ± 9.37 9.37

Page 186: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

171

NIB-S24 9 15.14 ± 42.92 12.37 ± 1.85 1.170 54.9 2.86 ± 6.59 6.59

NIB-S24 10 10.38 ± 47.47 11.64 ± 1.74 0.992 68.9 3.81 ± 7.73 7.73

NIB-S24 11 14.18 ± 26.68 9.25 ± 0.79 1.410 57.8 4.02 ± 5.47 5.47

NIB-S24 12 9.42 ± 17.06 8.31 ± 0.49 2.028 71.9 5.55 ± 3.90 3.90

NIB-S24 13 17.59 ± 38.69 7.47 ± 0.93 0.779 47.8 4.11 ± 9.81 9.81

NIB-S24 14 3.83 ± 11.16 4.06 ± 0.25 1.450 88.5 13.94 ± 5.26 5.25

NIB-S24 15 5.11 ± 30.70 13.64 ± 1.34 1.822 84.3 3.98 ± 4.27 4.27

NIB-S24 16 7.67 ± 2.18 6.37 ± 0.31 5.835 77.1 7.76 ± 0.81 0.80

NIB-S24 17 14.03 ± 4.52 10.21 ± 1.14 4.092 58.2 3.67 ± 1.05 1.05

NIB-S24 18 3.40 ± 3.91 6.28 ± 0.27 1.476 89.7 9.14 ± 1.26 1.25

NIB-S24 19 6.71 ± 6.40 12.88 ± 1.19 1.695 79.6 3.98 ± 1.04 1.04

NIB-S24 20 1.93 ± 2.78 4.50 ± 0.29 1.616 94.1 13.37 ± 1.49 1.48

NIB-S24 21 64.48 ± 9.31 16.87 ± 1.85 2.418 0.0 0.00 ± 0.80 0.80

NIB-S24 22 49.89 ± 5.20 8.96 ± 0.52 1.659 0.0 0.00 ± 1.00 1.00

NIB-S24 23 38.25 ± 3.87 9.34 ± 0.60 2.255 0.0 0.00 ± 0.75 0.75

NIB-S24 24 17.49 ± 2.23 9.85 ± 0.86 5.203 48.1 3.14 ± 0.65 0.65

NIB-S24 25 3.61 ± 2.64 7.65 ± 0.62 2.665 89.0 7.46 ± 0.94 0.93

NIB-S24 26 25.33 ± 2.40 7.71 ± 0.62 2.595 25.0 2.09 ± 0.74 0.74

NIB-S24 27 4.78 ± 1.43 8.32 ± 0.63 4.781 85.5 6.59 ± 0.67 0.66

NIB-S24 28 80.25 ± 10.44 10.35 ± 1.27 1.101 0.0 0.00 ± 0.99 0.99

NIB-S24 29 66.94 ± 7.68 3.93 ± 0.39 0.501 0.0 0.00 ± 2.50 2.50

NIB-S24 30 40.10 ± 3.17 12.29 ± 0.61 2.614 0.0 0.00 ± 0.46 0.46

NIB-S24 31 32.50 ± 2.44 6.69 ± 0.42 1.756 4.0 0.38 ± 0.71 0.71

NIB-S24 32 9.07 ± 0.97 6.15 ± 0.40 5.777 73.0 7.61 ± 0.72 0.71

NIB-S24 33 11.35 ± 2.23 3.67 ± 0.25 4.466 66.3 11.57 ± 1.60 1.59

NIB-S24 34 38.16 ± 3.50 8.01 ± 0.44 1.790 0.0 0.00 ± 0.79 0.79

NIB-S24 35 92.19 ± 11.74 14.79 ± 1.75 1.368 0.0 0.00 ± 0.76 0.76

NIB-S24 36 29.76 ± 4.13 11.10 ± 0.59 2.870 12.0 0.70 ± 0.72 0.72

NIB-S24 37 1.29 ± 4.15 12.05 ± 0.75 3.171 95.6 5.10 ± 0.73 0.73

NIB-S24 38 5.28 ± 1.44 2.84 ± 0.20 4.504 84.3 18.90 ± 1.81 1.79

NIB-S24 39 26.52 ± 3.75 9.10 ± 0.83 2.640 21.5 1.52 ± 0.88 0.88

NIB-S24 40 10.12 ± 2.64 6.61 ± 0.46 4.601 69.9 6.78 ± 1.00 0.99

NIB-S24 41 29.54 ± 3.59 10.87 ± 0.89 2.830 12.6 0.75 ± 0.67 0.67

NIB-S24 42 32.68 ± 4.47 9.54 ± 0.66 2.245 3.4 0.23 ± 0.89 0.89

NIB-S24 43 17.99 ± 2.15 6.23 ± 0.40 2.662 46.7 4.81 ± 0.86 0.86

Page 187: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

172

NIB-S24 44 63.33 ± 7.14 8.75 ± 0.73 1.063 0.0 0.00 ± 1.21 1.21

NIB-S24 45 50.75 ± 5.59 8.91 ± 0.60 1.350 0.0 0.00 ± 1.06 1.06

NIB-S24 46 16.57 ± 1.27 5.37 ± 0.31 2.701 50.9 6.07 ± 0.74 0.74

NIB-S24 47 5.99 ± 1.77 5.77 ± 0.42 5.589 82.1 9.11 ± 0.99 0.98

NIB-S24 48 26.25 ± 1.93 10.19 ± 0.39 2.641 22.3 1.41 ± 0.39 0.39

NIB-S24 49 0.81 ± 2.76 10.18 ± 0.50 2.920 97.1 6.13 ± 0.60 0.60

NIB-S24 50 348.34 ± 162.79 0.02 ± 0.01 0.000 0.0 0.07 ± 2689 2689

NIB-S24 51 61.18 ± 5.95 11.85 ± 0.95 1.316 0.0 0.00 ± 0.69 0.69

NIB-S24 52 27.27 ± 3.38 6.06 ± 0.20 1.511 19.4 2.05 ± 1.07 1.07

NIB-S24 53 37.90 ± 3.92 9.76 ± 0.66 1.750 0.0 0.00 ± 0.73 0.73

NIB-S24 54 28.80 ± 2.41 7.93 ± 0.48 1.873 14.8 1.20 ± 0.63 0.63

NIB-S24 55 8.42 ± 0.80 4.85 ± 0.32 3.915 74.9 9.88 ± 0.90 0.89

NIB-S24 56 25.60 ± 1.51 9.54 ± 0.30 2.531 24.2 1.63 ± 0.33 0.33

NIB-S24 57 18.67 ± 3.32 3.25 ± 0.10 2.008 44.7 8.81 ± 1.99 1.99

NIB-S24 58 15.65 ± 1.40 9.11 ± 0.37 3.954 53.5 3.77 ± 0.39 0.38

NIB-S24 59 68.80 ± 8.72 9.61 ± 0.82 0.949 0.0 0.00 ± 1.39 1.39

NIB-S24 60 65.10 ± 7.68 11.22 ± 0.87 1.172 0.0 0.00 ± 1.07 1.07

NIB-S24 61 25.50 ± 2.02 10.23 ± 0.37 2.726 24.5 1.54 ± 0.40 0.40

NIB-S24 62 49.77 ± 5.50 12.26 ± 0.73 1.675 0.0 0.00 ± 0.78 0.78

NIB-S24 63 31.40 ± 3.10 10.09 ± 0.71 2.184 7.2 0.46 ± 0.60 0.60

NIB-S24 64 29.34 ± 2.28 11.33 ± 0.49 2.624 13.2 0.75 ± 0.40 0.40

NIB-S24 65 50.03 ± 6.75 12.40 ± 1.42 1.684 0.0 0.00 ± 0.80 0.80

NIB-S24 66 0.28 ± 0.67 8.91 ± 0.48 7.766 98.7 7.11 ± 0.43 0.41

NIB-S24 67 39.43 ± 5.00 8.36 ± 0.77 1.441 0.0 0.00 ± 1.05 1.05

NIB-S24 68 5.13 ± 1.46 4.56 ± 0.13 3.749 84.6 11.86 ± 0.74 0.72

NIB-S24 69 1.10 ± 3.29 11.29 ± 0.42 3.366 96.2 5.48 ± 0.59 0.59

NIB-S24 70 16.73 ± 2.93 4.07 ± 0.21 1.207 50.5 7.93 ± 1.52 1.52

NIB-S24 71 1.92 ± 2.11 6.87 ± 0.67 1.029 94.0 8.77 ± 1.08 1.07

NIB-S24 72 19.03 ± 1.47 6.10 ± 0.15 2.178 43.6 4.59 ± 0.51 0.51

NIB-S24 73 72.55 ± 7.32 9.80 ± 0.86 0.918 0.0 0.00 ± 0.90 0.90

NIB-S24 74 1.96 ± 0.66 8.24 ± 0.31 5.080 93.8 7.31 ± 0.34 0.32

NIB-S24 75 53.24 ± 7.48 8.97 ± 0.90 1.144 0.0 0.00 ± 1.31 1.31

NIB-S24 76 34.39 ± 3.86 10.26 ± 0.69 2.027 0.0 0.00 ± 0.71 0.71

NIB-S24 77 10.59 ± 1.25 6.66 ± 0.40 4.276 68.5 6.59 ± 0.66 0.65

NIB-S24 78 37.56 ± 3.96 8.51 ± 0.69 1.539 0.0 0.00 ± 0.83 0.83

Page 188: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

173

NIB-S24 79 21.48 ± 1.87 8.86 ± 0.34 2.804 36.4 2.64 ± 0.45 0.45

NIB-S24 80 56.01 ± 8.60 9.76 ± 1.02 1.184 0.0 0.00 ± 1.40 1.40

NIB-S24 81 51.95 ± 5.85 9.87 ± 0.64 1.122 0.0 0.00 ± 1.01 1.01

NIB-S24 82 54.25 ± 5.40 11.12 ± 0.75 1.210 0.0 0.00 ± 0.78 0.78

NIB-S24 83 28.08 ± 2.93 8.53 ± 0.32 1.794 16.9 1.28 ± 0.67 0.67

NIB-S24 84 1.64 ± 0.16 0.25 ± 0.02 0.900 95.2 229.15 ± 17.51 17.26

NIB-S24 85 50.32 ± 5.71 10.53 ± 0.72 1.237 0.0 0.00 ± 0.92 0.92

NIB-S24 86 38.09 ± 3.71 8.80 ± 0.54 1.364 0.0 0.00 ± 0.76 0.76

NIB-S24 87 19.17 ± 1.64 8.71 ± 0.30 2.682 43.2 3.18 ± 0.41 0.41

NIB-S24 88 26.20 ± 2.56 5.04 ± 0.19 1.136 22.5 2.87 ± 1.00 1.00

NIB-S24 89 30.71 ± 2.64 7.54 ± 0.23 1.451 9.2 0.79 ± 0.67 0.67

NIB-S24 90 8.74 ± 1.26 6.38 ± 0.53 4.314 73.9 7.43 ± 0.90 0.89

NIB-S24 91 25.55 ± 2.04 6.33 ± 0.25 1.463 24.4 2.48 ± 0.66 0.66

NIB-S24 92 11.15 ± 3.23 3.44 ± 0.30 3.892 66.9 12.43 ± 2.35 2.34

NIB-S24 93 6.70 ± 1.50 4.46 ± 0.41 6.279 80.0 11.46 ± 1.45 1.44

NIB-S24 94 37.05 ± 3.88 9.96 ± 0.53 1.589 0.0 0.00 ± 0.72 0.72

NIB-S24 95 38.53 ± 3.58 8.56 ± 0.40 1.313 0.0 0.00 ± 0.77 0.77

NIB-S24 96 0.63 ± 0.81 3.00 ± 0.21 3.473 98.0 20.82 ± 1.58 1.55

NIB-S24 97 21.98 ± 1.99 9.02 ± 0.28 2.425 34.9 2.49 ± 0.45 0.45

NIB-S24 98 44.34 ± 4.50 9.51 ± 0.60 1.267 0.0 0.00 ± 0.82 0.82

NIB-S24 99 8.31 ± 3.01 6.14 ± 0.43 3.184 75.2 7.85 ± 1.17 1.17

NIB-S24 100 37.00 ± 4.18 10.17 ± 0.51 1.624 0.0 0.00 ± 0.76 0.76

NIB-S24 101 12.14 ± 2.10 5.18 ± 0.16 3.481 63.9 7.90 ± 0.85 0.84

NIB-S24 102 37.72 ± 4.22 8.74 ± 0.46 1.370 0.0 0.00 ± 0.89 0.89

NIB-S24 103 16.38 ± 1.32 6.30 ± 0.25 2.274 51.4 5.23 ± 0.53 0.53

NIB-S24 104 39.69 ± 4.96 9.35 ± 0.51 1.391 0.0 0.00 ± 0.98 0.98

NIB-S24 105 37.21 ± 3.66 8.37 ± 0.38 1.328 0.0 0.00 ± 0.81 0.81

NIB-S24 106 31.58 ± 2.69 6.76 ± 0.33 1.265 6.7 0.63 ± 0.77 0.77

NIB-S24 107 20.58 ± 1.94 10.02 ± 0.38 2.875 39.0 2.50 ± 0.41 0.41

NIB-S24 108 17.13 ± 1.54 9.97 ± 0.52 3.438 49.1 3.17 ± 0.41 0.41

NIB-S24 109 4.56 ± 1.13 7.33 ± 0.37 3.632 86.2 7.54 ± 0.53 0.52

NIB-S24 110 11.72 ± 1.22 3.33 ± 0.18 1.679 65.3 12.51 ± 1.18 1.17

NIB-S24 111 22.99 ± 2.78 8.90 ± 0.26 2.286 31.9 2.30 ± 0.61 0.61

NIB-S24 112 22.13 ± 2.20 9.52 ± 0.29 2.542 34.4 2.32 ± 0.46 0.46

NIB-S24 113 1.30 ± 2.66 10.77 ± 0.42 2.944 95.6 5.71 ± 0.53 0.52

Page 189: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

174

NIB-S37 1 11.72 ± 8.65 7.25 ± 0.18 7.821 65.1 5.57 ± 2.18 2.18

NIB-S37 2 25.20 ± 39.33 8.90 ± 0.55 2.116 25.4 1.77 ± 8.08 8.08

NIB-S37 3 65.34 ± 124.12 12.16 ± 2.22 0.915 0.0 0.00 ± 18.60 18.60

NIB-S37 4 26.74 ± 25.92 6.26 ± 0.38 2.245 20.9 2.07 ± 7.57 7.57

NIB-S37 5 43.61 ± 82.42 7.24 ± 0.90 0.816 0.0 0.00 ± 20.81 20.81

NIB-S37 6 14.83 ± 31.28 6.33 ± 0.34 1.878 56.0 5.48 ± 9.02 9.02

NIB-S37 7 32.32 ± 82.90 8.22 ± 1.00 0.920 4.5 0.34 ± 18.45 18.45

NIB-S37 8 15.62 ± 30.44 10.53 ± 0.85 3.208 53.5 3.16 ± 5.30 5.30

NIB-S37 9 9.00 ± 13.22 5.83 ± 0.37 4.123 73.2 7.78 ± 4.18 4.18

NIB-S37 10 60.11 ± 135.09 7.70 ± 1.53 0.535 0.0 0.00 ± 32.01 32.01

NIB-S37 11 15.48 ± 29.53 9.43 ± 0.67 2.968 54.0 3.56 ± 5.73 5.73

NIB-S37 12 26.52 ± 57.53 7.59 ± 0.70 1.224 21.5 1.76 ± 13.85 13.85

NIB-S37 13 11.16 ± 19.71 6.73 ± 0.44 3.173 66.8 6.15 ± 5.36 5.36

NIB-S37 14 14.50 ± 14.80 5.03 ± 0.33 3.185 57.0 7.01 ± 5.41 5.41

NIB-S37 15 32.07 ± 58.07 11.96 ± 1.05 1.921 5.2 0.27 ± 8.88 8.88

NIB-S37 16 3.90 ± 0.47 2.33 ± 0.09 5.137 88.4 23.36 ± 1.12 1.09

NIB-S37 17 5.12 ± 0.94 8.51 ± 0.41 5.894 84.5 6.16 ± 0.41 0.40

NIB-S37 18 6.85 ± 0.73 4.29 ± 0.08 3.885 79.6 11.46 ± 0.43 0.41

NIB-S37 19 9.72 ± 5.10 7.63 ± 0.41 1.258 71.0 5.77 ± 1.29 1.29

NIB-S37 20 6.99 ± 2.07 8.67 ± 0.14 2.794 79.0 5.65 ± 0.45 0.45

NIB-S37 21 13.66 ± 1.16 5.23 ± 0.10 3.095 59.5 7.05 ± 0.46 0.45

NIB-S37 22 10.95 ± 6.39 8.16 ± 0.75 0.879 67.4 5.12 ± 1.58 1.57

NIB-S37 23 6.74 ± 2.96 9.54 ± 0.28 1.660 79.7 5.19 ± 0.60 0.60

NIB-S37 24 8.43 ± 1.81 5.83 ± 0.09 2.201 74.9 7.95 ± 0.59 0.59

NIB-S37 25 7.58 ± 2.03 9.01 ± 0.44 10.050 77.2 5.32 ± 0.53 0.53

NIB-S37 26 2433.01 ± 18561.7 205.66 ± 1562.98 1.621 0.0 0.00 ± 14.66 14.66

NIB-S37 27 49.54 ± 6.15 0.16 ± 0.01 0.041 0.0 0.00 ± 65.89 65.89

NIB-S37 28 10.65 ± 3.15 7.81 ± 0.53 2.287 68.3 5.42 ± 0.90 0.89

NIB-S37 29 16.51 ± 7.36 11.66 ± 0.24 1.266 50.9 2.72 ± 1.16 1.16

NIB-S37 30 16.74 ± 4.49 7.90 ± 0.15 1.520 50.3 3.95 ± 1.05 1.05

NIB-S37 31 12.32 ± 4.78 7.58 ± 0.33 1.306 63.4 5.18 ± 1.20 1.20

NIB-S37 32 15.88 ± 1.82 7.26 ± 0.28 3.493 52.9 4.52 ± 0.54 0.54

NIB-S37 33 4.18 ± 8.74 9.10 ± 0.44 0.609 87.2 5.95 ± 1.78 1.78

NIB-S37 34 1.73 ± 0.95 7.02 ± 0.07 4.971 94.6 8.35 ± 0.28 0.26

NIB-S37 35 5.78 ± 2.51 5.15 ± 0.27 1.123 82.7 9.93 ± 1.08 1.08

Page 190: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

175

NIB-S40 1 6.76 ± 13.78 8.90 ± 0.48 4.258 79.7 5.46 ± 2.80 2.80

NIB-S40 2 10.47 ± 22.71 8.93 ± 0.70 2.556 68.8 4.70 ± 4.59 4.59

NIB-S40 3 7.34 ± 3.55 6.16 ± 0.24 11.554 78.1 7.71 ± 1.12 1.10

NIB-S40 4 8.75 ± 12.11 6.51 ± 0.40 3.533 73.9 6.92 ± 3.38 3.38

NIB-S40 5 31.89 ± 57.31 12.89 ± 1.33 1.486 5.7 0.27 ± 7.99 7.99

NIB-S40 6 17.03 ± 30.60 6.82 ± 0.42 1.451 49.5 4.43 ± 8.06 8.06

NIB-S40 7 11.88 ± 20.84 11.70 ± 0.90 3.637 64.5 3.37 ± 3.22 3.22

NIB-S40 8 11.62 ± 25.72 7.79 ± 0.68 1.992 65.4 5.12 ± 5.95 5.95

NIB-S40 9 405.89 ± 1313.25 2.68 ± 4.94 0.016 0.0 0.00 ± 724.56 724

NIB-S40 10 14.95 ± 16.72 7.50 ± 0.55 2.918 55.6 4.52 ± 4.03 4.03

NIB-S40 11 9.27 ± 18.31 8.16 ± 0.45 2.944 72.3 5.40 ± 4.04 4.04

NIB-S40 12 10.04 ± 10.84 9.79 ± 0.77 5.947 70.0 4.36 ± 2.04 2.04

NIB-S40 13 8.04 ± 19.08 6.93 ± 0.59 2.400 76.0 6.68 ± 4.98 4.98

NIB-S40 14 9.22 ± 14.08 8.55 ± 0.62 3.975 72.4 5.17 ± 3.00 2.99

NIB-S40 15 9.25 ± 26.65 9.61 ± 0.78 2.378 72.3 4.59 ± 4.99 4.99

NIB-S40 16 11.09 ± 7.58 6.63 ± 0.18 1.626 67.0 6.15 ± 2.07 2.06

NIB-S40 17 5.56 ± 3.59 6.02 ± 0.06 3.133 83.3 8.42 ± 1.09 1.07

NIB-S40 18 5.10 ± 1.17 6.76 ± 0.10 12.052 84.7 7.62 ± 0.39 0.34

NIB-S40 19 9.32 ± 3.75 7.49 ± 0.13 3.809 72.2 5.87 ± 0.92 0.91

NIB-S40 20 7.32 ± 5.32 6.56 ± 0.11 2.329 78.1 7.25 ± 1.47 1.46

NIB-S40 21 7.48 ± 3.42 8.73 ± 0.09 4.843 77.6 5.42 ± 0.72 0.71

NIB-S40 22 6.67 ± 6.62 8.12 ± 0.19 2.277 80.0 6.00 ± 1.48 1.47

NIB-S40 23 7.49 ± 1.12 5.03 ± 0.10 8.893 77.7 9.38 ± 0.52 0.46

NIB-S40 24 4.23 ± 4.39 5.27 ± 0.06 2.291 87.3 10.06 ± 1.52 1.49

NIB-S40 25 6.62 ± 5.16 7.68 ± 0.08 2.806 80.1 6.36 ± 1.22 1.21

NIB-S44 1 1.74 ± 14.36 8.16 ± 0.74 0.552 94.5 7.05 ± 3.23 3.22

NIB-S44 2 9.02 ± 6.28 8.44 ± 0.17 1.182 73.0 5.27 ± 1.35 1.34

NIB-S44 3 5.70 ± 2.77 8.94 ± 0.20 2.596 82.8 5.65 ± 0.59 0.57

NIB-S44 4 6.35 ± 10.74 12.13 ± 1.03 0.983 80.7 4.07 ± 1.65 1.64

NIB-S44 5 4.55 ± 8.99 7.63 ± 0.09 1.511 86.2 6.88 ± 2.12 2.11

NIB-S44 6 44.65 ± 13.95 9.61 ± 0.75 1.025 0.0 0.00 ± 2.57 2.57

NIB-S44 7 3.08 ± 2.17 7.96 ± 0.66 5.565 90.5 6.93 ± 0.82 0.80

NIB-S44 8 1.43 ± 4.54 8.07 ± 0.22 1.949 95.4 7.20 ± 1.04 1.03

NIB-S44 9 6.32 ± 8.48 9.30 ± 0.16 1.411 81.0 5.31 ± 1.64 1.64

NIB-S44 10 4.18 ± 3.86 7.26 ± 0.55 3.541 87.3 7.32 ± 1.15 1.14

Page 191: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

176

NIB-S44 11 4.76 ± 3.33 8.03 ± 0.67 3.733 85.6 6.49 ± 0.99 0.97

NIB-S44 12 2.46 ± 4.30 10.88 ± 0.88 3.519 92.2 5.17 ± 0.85 0.84

NIB-S44 13 7.77 ± 4.77 8.36 ± 0.40 2.022 76.7 5.59 ± 1.08 1.07

NIB-S44 14 4.06 ± 4.67 8.27 ± 0.19 1.558 87.6 6.46 ± 1.04 1.03

NIB-S44 15 5.56 ± 6.26 5.63 ± 0.16 1.124 83.3 9.00 ± 2.02 2.01

NIB-S44 16 2.01 ± 4.11 9.45 ± 0.74 2.723 93.6 6.04 ± 0.94 0.93

NIB-S44 17 8.21 ± 2.53 10.94 ± 0.43 7.538 75.3 4.21 ± 0.48 0.47

NIB-S44 18 11.48 ± 3.59 9.16 ± 0.75 3.287 65.8 4.38 ± 0.87 0.87

NIB-S44 19 5.73 ± 7.72 7.42 ± 0.41 1.154 82.7 6.79 ± 1.92 1.91

NIB-S44 20 20.07 ± 22.18 8.97 ± 0.71 0.472 40.5 2.76 ± 4.46 4.46

NIB-S44 21 1.83 ± 4.60 11.09 ± 0.31 1.321 94.1 5.18 ± 0.77 0.76

NIB-S44 22 4.86 ± 2.46 10.44 ± 0.14 2.003 85.2 4.98 ± 0.45 0.43

NIB-S44 23 6.61 ± 0.75 8.25 ± 0.23 6.180 80.1 5.92 ± 0.30 0.26

NIB-S44 24 6.87 ± 0.71 7.43 ± 0.07 6.169 79.4 6.51 ± 0.25 0.19

NIB-S44 25 7.21 ± 0.63 8.66 ± 0.13 9.533 78.4 5.52 ± 0.22 0.17

NIB-S44 26 3.59 ± 0.64 4.25 ± 0.07 2.929 89.2 12.75 ± 0.48 0.36

NIB-S44 27 13.28 ± 2.94 9.05 ± 0.33 2.153 60.5 4.08 ± 0.63 0.62

NIB-S44 28 11.42 ± 2.98 12.85 ± 0.64 2.068 65.8 3.13 ± 0.48 0.47

NIB-S44 29 7.19 ± 1.02 8.24 ± 0.12 3.437 78.4 5.80 ± 0.29 0.25

NIB-S44 30 16.81 ± 1.03 7.14 ± 0.08 3.750 50.1 4.28 ± 0.29 0.27

NIB-S44 31 9.55 ± 0.76 7.93 ± 0.10 5.262 71.5 5.50 ± 0.24 0.19

NIB-S44 32 10.26 ± 2.84 10.67 ± 0.55 2.772 69.3 3.97 ± 0.56 0.55

NIB-S44 33 6.75 ± 1.52 8.33 ± 0.15 3.728 79.7 5.83 ± 0.38 0.35

NIB-S44 34 14.96 ± 3.07 8.28 ± 0.16 1.350 55.6 4.09 ± 0.69 0.68

NIB-S44 35 13.45 ± 1.19 6.32 ± 0.09 3.469 60.1 5.79 ± 0.39 0.36

NIB-S44 36 6.40 ± 0.94 6.98 ± 0.02 5.453 80.8 7.05 ± 0.30 0.24

NIB-S44 37 12.78 ± 2.62 8.38 ± 0.21 1.638 62.0 4.51 ± 0.60 0.59

NIB-S44 38 23.05 ± 4.96 7.59 ± 0.29 0.988 31.8 2.55 ± 1.19 1.19

NIB-S44 39 11.74 ± 0.70 6.92 ± 0.03 4.474 65.1 5.73 ± 0.23 0.18

NIB-S44 40 22.83 ± 2.17 4.60 ± 0.18 1.217 32.5 4.29 ± 0.93 0.93

NIB-S52 1 6.62 ± 13.71 3.25 ± 0.15 2.056 80.3 14.84 ± 7.49 7.49

NIB-S52 2 43.06 ± 37.57 8.81 ± 4.66 2.611 0.0 0.00 ± 7.04 7.04

NIB-S52 3 194.53 ± 425.04 0.03 ± 0.02 0.001 0.0 0.04 ± 26808 26808

NIB-S52 4 29.64 ± 63.17 7.15 ± 0.92 0.980 12.4 1.05 ± 15.74 15.74

NIB-S52 5 22.75 ± 57.58 8.37 ± 0.96 1.262 32.6 2.36 ± 12.25 12.25

Page 192: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

177

NIB-S52 6 5.17 ± 1.06 8.17 ± 0.07 3.640 84.4 6.24 ± 0.25 0.24

NIB-S52 7 6.04 ± 2.15 7.18 ± 0.12 1.346 81.9 6.89 ± 0.55 0.55

NIB-S52 8 1.01 ± 0.73 3.14 ± 0.10 2.030 96.9 18.53 ± 0.74 0.71

NIB-S52 9 0.07 ± 1.33 8.19 ± 0.07 2.505 99.4 7.33 ± 0.31 0.29

NIB-S52 10 2.72 ± 0.73 6.24 ± 0.10 3.488 91.7 8.86 ± 0.28 0.26

NIB-S52 11 2.75 ± 0.28 6.86 ± 0.11 11.024 91.5 8.06 ± 0.18 0.16

NIB-S52 12 2.22 ± 0.29 3.29 ± 0.09 6.611 93.3 17.03 ± 0.58 0.54

NIB-S52 13 3.24 ± 0.53 5.20 ± 0.14 6.528 90.2 10.45 ± 0.38 0.36

NIB-S52 14 8.85 ± 1.06 7.74 ± 0.09 3.247 73.6 5.74 ± 0.27 0.26

NIB-S52 15 3.06 ± 0.49 6.71 ± 0.11 6.582 90.7 8.15 ± 0.22 0.19

NIB-S52 16 5.14 ± 1.23 10.79 ± 0.06 3.876 84.4 4.73 ± 0.21 0.21

NIB-S52 17 4.73 ± 1.21 5.30 ± 0.10 2.299 85.8 9.76 ± 0.47 0.45

NIB-S52 18 4.73 ± 0.77 10.15 ± 0.16 6.427 85.6 5.10 ± 0.18 0.16

NIB-S52 19 114.57 ± 13.61 0.29 ± 0.01 0.011 0.0 0.00 ± 81.67 81.67

NIB-S52 20 2.89 ± 0.64 6.67 ± 0.09 5.836 91.2 8.24 ± 0.23 0.21

NIB-S53 1 16.48 ± 41.07 9.43 ± 0.84 2.417 51.1 3.28 ± 7.76 7.76

NIB-S53 2 36.26 ± 89.63 11.93 ± 1.84 1.403 0.0 0.00 ± 13.37 13.37

NIB-S53 3 18.23 ± 34.90 8.16 ± 0.73 2.459 46.0 3.41 ± 7.62 7.62

NIB-S53 4 6.69 ± 9.58 5.50 ± 0.38 6.062 80.0 8.77 ± 3.17 3.17

NIB-S53 5 31.74 ± 56.76 6.91 ± 0.63 1.283 6.2 0.54 ± 14.64 14.64

NIB-S53 6 11.86 ± 30.51 3.03 ± 0.24 1.049 64.8 12.88 ± 17.88 17.87

NIB-S53 7 21.42 ± 35.48 9.86 ± 0.98 2.924 36.5 2.24 ± 6.42 6.42

NIB-S53 8 14.24 ± 39.82 6.71 ± 0.72 1.773 57.7 5.20 ± 10.58 10.58

NIB-S53 9 22.06 ± 22.13 6.17 ± 0.51 2.986 34.7 3.40 ± 6.40 6.40

NIB-S53 10 10.16 ± 26.88 4.39 ± 0.33 1.724 69.8 9.58 ± 10.89 10.89

NIB-S53 11 128.58 ± 307.72 0.30 ± 0.14 0.010 0.0 0.00 ± 1799 1799

NIB-S53 12 40.95 ± 90.38 7.82 ± 1.13 0.913 0.0 0.00 ± 20.57 20.57

NIB-S53 13 13.03 ± 29.41 7.30 ± 0.42 2.615 61.3 5.07 ± 7.17 7.17

NIB-S53 14 273.34 ± 738.60 0.44 ± 0.48 0.007 0.0 0.00 ± 2744 2744

NIB-S53 15 12.65 ± 27.63 9.07 ± 0.49 3.476 62.3 4.16 ± 5.42 5.42

NIB-S53 16 3.05 ± 0.72 8.17 ± 0.07 7.683 90.6 6.70 ± 0.19 0.17

NIB-S53 17 1.56 ± 0.16 3.45 ± 0.03 8.090 95.2 16.57 ± 0.27 0.18

NIB-S53 18 5.21 ± 2.08 7.61 ± 0.09 1.479 84.3 6.69 ± 0.50 0.49

NIB-S53 19 3.89 ± 0.29 7.71 ± 0.18 9.797 88.2 6.90 ± 0.21 0.19

NIB-S53 20 4.21 ± 1.13 3.66 ± 0.02 1.612 87.4 14.37 ± 0.58 0.55

Page 193: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

178

NIB-S53 21 3.41 ± 0.64 3.15 ± 0.02 2.373 89.8 17.10 ± 0.44 0.39

NIB-S53 22 3.56 ± 0.89 6.80 ± 0.05 4.072 89.2 7.92 ± 0.26 0.24

NIB-S53 23 8.61 ± 1.64 3.46 ± 0.03 0.765 74.4 12.92 ± 0.86 0.84

NIB-S53 24 5.67 ± 5.59 7.38 ± 0.08 0.601 83.0 6.79 ± 1.35 1.35

NIB-S53 25 3.49 ± 1.19 2.90 ± 0.05 1.369 89.5 18.51 ± 0.82 0.79

NIB-S53 26 0.08 ± 0.53 3.53 ± 0.06 2.348 99.6 16.97 ± 0.44 0.39

NIB-S53 27 1.57 ± 0.48 7.46 ± 0.07 4.970 95.0 7.69 ± 0.16 0.13

NIB-S53 28 6.75 ± 0.87 7.09 ± 0.04 4.650 79.8 6.79 ± 0.24 0.22

NIB-S53 29 1.42 ± 1.84 7.62 ± 0.07 2.335 95.4 7.56 ± 0.44 0.43

NIB-S53 30 27.81 ± 2.37 10.54 ± 0.11 1.858 17.7 1.02 ± 0.40 0.40

NIB-S54 1 30.02 ± 3.24 0.01 ± 0.00 0.008 11.3 623.84 ± 448.15 448.10

NIB-S54 2 4.78 ± 5.43 7.77 ± 0.08 1.399 85.5 6.58 ± 1.23 1.23

NIB-S54 3 13.58 ± 1.79 5.06 ± 0.34 5.537 59.7 7.05 ± 0.96 0.95

NIB-S54 4 9.78 ± 5.67 8.44 ± 0.09 1.477 70.8 5.02 ± 1.19 1.18

NIB-S54 5 9.98 ± 1.79 5.29 ± 0.31 4.846 70.3 7.94 ± 0.87 0.86

NIB-S54 6 10.17 ± 2.36 3.92 ± 0.26 3.136 69.8 10.63 ± 1.43 1.42

NIB-S54 7 2.43 ± 1.63 5.07 ± 0.40 3.340 92.6 10.90 ± 1.09 1.08

NIB-S54 8 4.15 ± 2.04 10.92 ± 1.10 5.408 87.3 4.79 ± 0.64 0.64

NIB-S54 9 3.30 ± 2.90 5.08 ± 0.05 1.185 90.0 10.58 ± 1.02 1.01

NIB-S54 10 21.98 ± 4.21 7.32 ± 0.16 1.787 34.9 2.85 ± 1.02 1.02

NIB-S54 11 19.36 ± 3.63 3.72 ± 0.18 1.783 42.7 6.85 ± 1.83 1.83

NIB-S54 12 1.26 ± 2.42 6.27 ± 0.49 3.149 96.0 9.15 ± 1.01 1.00

NIB-S54 13 26.06 ± 6.02 6.84 ± 0.06 1.408 22.9 2.01 ± 1.55 1.55

NIB-S54 14 59.39 ± 17.50 11.53 ± 1.22 1.042 0.0 0.00 ± 2.56 2.56

NIB-S54 15 0.22 ± 5.74 5.80 ± 0.21 0.919 99.1 10.20 ± 1.78 1.78

NIB-S54 16 1.65 ± 6.54 6.44 ± 0.07 1.227 94.8 8.79 ± 1.79 1.79

NIB-S54 17 5.60 ± 2.37 5.81 ± 0.36 2.201 83.2 8.55 ± 0.95 0.94

NIB-S54 18 3.29 ± 12.51 6.41 ± 0.60 0.511 90.0 8.39 ± 3.54 3.54

NIB-S54 19 13.87 ± 8.33 6.39 ± 0.06 1.236 58.8 5.50 ± 2.29 2.29

NIB-S54 20 7.00 ± 1.98 7.47 ± 0.73 4.055 79.0 6.33 ± 0.89 0.89

NIB-S54 21 33.83 ± 46.87 0.38 ± 0.07 0.004 -0.3 0.07 ± 217.03 217.03

NIB-S54 22 29.52 ± 24.86 5.42 ± 0.58 0.164 12.7 1.41 ± 8.11 8.11

NIB-S54 23 6.03 ± 2.54 7.54 ± 0.21 2.234 81.9 6.50 ± 0.64 0.63

NIB-S54 24 13.28 ± 2.07 6.47 ± 0.21 1.480 60.6 5.60 ± 0.63 0.63

NIB-S54 25 5.86 ± 0.69 8.18 ± 0.07 10.543 82.3 6.03 ± 0.18 0.16

Page 194: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

179

NIB-S54 26 11.01 ± 1.66 7.47 ± 0.07 3.284 67.2 5.38 ± 0.40 0.40

NIB-S54 27 11.00 ± 2.82 7.03 ± 0.19 2.297 67.3 5.72 ± 0.74 0.74

NIB-S54 28 22.34 ± 1.73 3.90 ± 0.02 2.277 33.9 5.20 ± 0.78 0.78

NIB-S54 29 5.49 ± 1.15 8.01 ± 0.06 6.887 83.4 6.23 ± 0.27 0.26

NIB-S54 30 8.09 ± 2.19 9.55 ± 0.52 2.700 75.7 4.75 ± 0.52 0.52

NIB-S54 31 1.98 ± 1.18 4.45 ± 0.23 1.835 93.9 12.58 ± 0.84 0.83

NIB-S54 32 0.81 ± 0.66 6.92 ± 0.09 5.025 97.3 8.40 ± 0.22 0.20

NIB-S54 33 15.98 ± 3.07 7.04 ± 0.25 1.098 52.6 4.47 ± 0.81 0.81

NIB-S54 34 18.57 ± 3.11 4.85 ± 0.24 0.828 45.0 5.55 ± 1.24 1.23

NIB-S54 35 6.05 ± 1.77 6.47 ± 0.05 1.654 81.9 7.56 ± 0.50 0.49

NIB-S54 36 6.90 ± 1.47 6.31 ± 0.15 3.807 79.4 7.52 ± 0.47 0.47

NIB-S54 37 1.40 ± 0.31 2.08 ± 0.06 3.658 95.8 27.34 ± 0.94 0.88

NIB-S54 38 10.29 ± 4.12 6.97 ± 0.25 1.358 69.4 5.95 ± 1.08 1.08

NIB-S54 39 6.77 ± 1.32 7.79 ± 0.06 3.570 79.7 6.12 ± 0.31 0.30

NIB-S54 40 10.01 ± 2.39 8.85 ± 0.27 1.738 70.1 4.75 ± 0.52 0.51

Table 1. Assigned uncertainty corresponds to 2-σ error. (i) Number of moles of K-derived 39Ar

(39ArK) released during fusion. (ii) Percentage of radiogenic 40Ar (40Ar*) in the total 40Ar for each

analysis. (iii) Age uncertainties are shown with and without propagated error in the irradiation

parameter J. Grains with shaded values are those cut from the analysis because they contained less

than 40% 40Ar*.

Page 195: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

180

Appendix 3

1.0 Comparing PDF curves

To quantify the match (or mismatch) between two PDF curves, we calculated the sum of the

difference in the distribution of probability (Pdiff) between the theoretical probability (Ptheoretical) and

grab-sample probability (Pgrab). This was computed over each age increment (t), and expressed in

terms of a percentage of total probability:

100*2

)()(0

tPtPP

grabltheoretica

t

tdiff

−=

∑∞=

=

(1)

This provides the percentage mismatch of the entire probability signal in the units of percentage

probability. Any two PDF curves may be analyzed this way, although the exact value of mismatch

will be affected by the age increment (t).

Several other statistical comparison techniques were considered for comparing PDFs, but

rejected:

1) Correlation coefficient between idealized PDF and grab sample PDF.

The value of a correlation coefficient is dependant upon the number of discrete points, and so is

very dependant upon the age increment (t). In addition, a simple regression line, in combination

with an r-square test, measures the fit of the two curves to the regression line. We know, however,

Page 196: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

181

that in our analysis the theoretical PDF is “correct” and so only need a measure of how well the

synthetic grab-sample fits this.

2) Forced correlation coefficient.

A correlation coefficient based on the deviation of the grab-sample points from a regression line

forced to intercept at (0,0) and with slope of one (an ideal fit of grab-sample to theoretical PDF

curve would lie exactly on this line). This is a good measure of the deviation from the exact match,

but is hard to interpret directly in terms of probability.

3) Percentage residuals.

Expressing the absolute residuals as a percentage of the real probability at that age :

100.)(

)()()(

−=

tP

tPtPtP

ltheoretica

grabltheoretica

(2)

This is easily understood on the PDF plot, but weights the differences in small probabilities

(primarily on the “tails” of the Gaussian curves) too much.

4) The Bootstrap and Jackknife.

Bootstrapping and Jackknifing are non-parametric tests of a population. When used on a “grab-

sample” population they provide a good test of the stability of the population, but do not measure

how well the grab sample matches the theoretical population.

Therefore, we feel that our comparison of the area under two PDF curves is a reasonable

solution. It is easily visualized on the PDF plot, simply understood in terms of probability, and

weights each age increment equally.

Page 197: DETRITAL-MINERAL THERMOCHRONOLOGY: INVESTIGATIONS …

VITA. IAN D. BREWER

Date Organization Description

1987-1993 Poole Grammar School,

England

GCSEs, A-Levels

1993-1996 Oxford University, England Bachelors Degree in Earth Sciences

1996-1997 University of Southern

California, Los Angeles,

USA

Teaching and Research Assistant.

Started PhD Program (supervisor; Dr

D.W.Burbank)

1997-2001 The Pennsylvania State

University, USA

Teaching and Research Assistant.

PhD Program (supervisor; Dr

D.W.Burbank)

2001-2005 Shell International

Exploration and Production,

The Netherlands and New

Zealand.

Exploration Geoscientist


Recommended