Sunteți pe pagina 1din 409

Tropical Geomorphology

Although similar geomorphic processes take place in other regions, in the tropics these
processes operate at different rates and with varying intensities. Tropical geomorphology
therefore provides many new insights regarding geomorphic processes. This textbook
describes both the humid and the arid tropics. It provides thoroughly up-to-date concepts
and relevant case studies, and emphasises the importance of geomorphology in the man-
agement and sustainable development of the tropical environment, including climate
change scenarios. The text is supported by a large number of illustrations, including satel-
lite images. Student exercises accompany each chapter.
The book highlights three areas:
• Geology, landforms and geomorphic processes in the humid and arid tropics
• Source-to-sink passage of water and sediment from the mountains to the sea
• Anthropogenic alteration of natural geomorphic rates and processes, including climate
change.
Tropical Geomorphology is an ideal textbook for any course on tropical geomorphology
or the tropical environment, and is also invaluable as a reference text for researchers and
environmental managers in the tropics.

Avijit Gupta is a Honorary Principal Fellow at the University of Wollongong, Australia


and a Visiting Scientist at the Centre for Remote Imaging, Sensing and Processing,
National University of Singapore. He was educated at Presidency College, Kolkata, and
Johns Hopkins University. He has held university positions in India, the USA, Singapore
and the UK. His research interests focus on fluvial geomorphology in the tropics, riv-
ers with high-magnitude floods, large rivers, urban geomorphology and the application of
remote sensing in geomorphology. Dr Gupta has served as a Committee Member of the
International Geographical Union Commission on Measurement, Theory and Applications
in Geomorphology (COMTAG) and the International Association of Geomorphologists
(IAG). He is currently the Chair of the IAG Working Group on the Effect of Climate
Change on Large Rivers and Deltas. He is a member of the American Geophysical Union,
the Association of American Geographers, the Society of Sedimentary Geologists and the
International Association of Hydrological Sciences. He is also a corresponding member
of the Académie Royale des Sciences d’Outre-Mer, Belgium. He is on the editorial board
of two journals on geomorphology. Dr Gupta has written eight books and over seventy
research papers. He recently edited The Physical Geography of Southeast Asia (2005) and
Large Rivers:€Geomorphology and Management (2007).
Tropical Geomorphology

Avijit Gupta
Centre for Remote Imaging, Sensing and Processing
National University of Singapore
and
School of Earth and Environmental Sciences
University of Wollongong, Australia
cambridge university press
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Tokyo, Mexico City

Cambridge University Press


The Edinburgh Building, Cambridge CB2 8RU, UK

Published in the United States of America by Cambridge University Press, New York

www.cambridge.org
Information on this title:€www.cambridge.org/9780521879903

© Avijit Gupta 2011

This publication is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.

First published 2011

Printed in the United Kingdom at the University Press, Cambridge

A catalogue record for this publication is available from the British Library

Library of Congress Cataloging in Publication data


Gupta, Avijit.
Tropical geomorphology / Avijit Gupta.
p.â•… cm.
Includes bibliographical references and index.
ISBN 978-0-521-87990-3 (hardback)
1.╇ Geomorphology–Tropics.â•… I.╇ Title.
GB446.G87 2011
551.410913–dc22
2011011810

ISBN 978-0-521-87990-3 Hardback

Cambridge University Press has no responsibility for the persistence or


accuracy of URLs for external or third-party internet websites referred to in
this publication, and does not guarantee that any content on such websites is,
or will remain, accurate or appropriate.
In Memory of Reds
Contents

Preface page xiii

Part I╇ The tropical environment

1╇ Introduction 3
1.1 Geomorphology in the tropics 3
1.2 Traditional tropical geomorphology 7
1.3 Modernisation of tropical geomorphology 8
1.4 Structure of tropical geomorphology 9
1.5 Structure of the book 11

2 Geological framework of the tropical lands 13


2.1 Introduction 13
2.2 A brief introduction to plate tectonics 13
2.3 Major landforms across the tropics 20
2.4 Interrelationships:€plate tectonics, landforms, erosion and sediment
production 27
Questions 29

3 Tropical hydrology 31
3.1 The tropical climate:€a brief review 31
3.2 Temperature 31
3.3 Wind circulation 32
3.4 Precipitation 34
3.5 Tropical disturbances 37
3.6 Miscellaneous factors 43
3.7 Water balance 44
3.8 Climate and geomorphology in the tropics 47
Questions 48

4 Erosion and land cover in the tropics 49


4.1 Erosion from tropical rainfall 49
4.2 Distribution of natural vegetation in the tropics 52
4.3 Tropical rain forests 52
4.4 Tropical deciduous forests, grasslands and deserts 55

vii
viii Contents

4.5 Anthropogenic alteration of the tropical vegetation 57


Questions 58

Part II╇ Process geomorphology in the tropics

5 Weathering in the tropics 61


5.1 Introduction 61
5.2 Sub-processes of weathering:€a brief review 63
5.3 Products of weathering 69
5.4 Weathering and vertical zonation 75
5.5 Pans and crusts 77
5.6 Effects of weathering 78
5.7 Tropics and weathering 80
Questions 81

6 Slopes:€forms and processes 82


6.1 Properties of a slope 82
6.2 Mass movement on hillslopes 84
6.3 Running water on hillslopes 93
6.4 Storage and transfer of surficial material on tropical slopes 94
6.5 A general description of tropical slopes 99
Questions 99

7 Rivers in the tropics 101


7.1 Components of a river system 101
7.2 Water in river channels 101
7.3 Sediment in river channels 104
7.4 Channel geometry 108
7.5 Channel network and nodes 119
7.6 River systems of the humid tropics 120
Questions 127

8 Alluvial valleys 129


8.1 Fluvial depositional environment 129
8.2 The alluvial valley 130
8.3 The channel alluvium 132
8.4 Bars 135
8.5 Floodplain 136
8.6 Terrace 139
8.7 Valley margins 140
8.8 Sediment transfer along the valley axis 141
Questions 141
ix Contents

9 Large rivers in the tropics 143


9.1 Introduction 143
9.2 Characteristics of a large river 143
9.3 The Amazon 147
9.4 The Zambezi 152
9.5 The Ganga–Brahmaputra system 155
9.6 The Mekong 164
9.7 The importance of major tropical rivers 169
Questions 169

10 The tropical coasts 170


10.1 Introduction 170
10.2 Types of coast 171
10.3 Moving water:€tides, waves and currents 172
10.4 Rocky coasts 179
10.5 Non-rocky coast 181
10.6 Coastal sand dunes 186
10.7 Coastal tropics 186
10.8 Coral reefs 189
10.9 Tropical coasts and time 191
Questions 194

11 Deltas in the tropics 195


11.1 Introduction 195
11.2 Distribution of deltas in the tropics 196
11.3 Age and evolution of deltas 197
11.4 Delta morphology 201
11.5 Delta sediments and sedimentary structures 202
11.6 The Ganga–Brahmaputra Delta:€a case study 203
11.7 Deltas in the tropics: a summary 207
Questions 208

12 The arid tropics 209


12.1 Arid areas 209
12.2 Geological characteristics of arid lands 210
12.3 Arid hydrology 211
12.4 Arid landforms 215
12.5 The rock desert 215
12.6 Running water in arid lands 216
12.7 Aeolian geomorphology of sandy areas 223
12.8 Conclusion 230
Questions 230
x Contents

13 Tropical highlands 232


13.1 Importance of highlands 232
13.2 Glaciation in tropical mountains 234
13.3 Mechanics of mountain glaciation 236
13.4 Glacial forms and processes 240
13.5 Slopes and valley floors in high mountains 246
13.6 Rivers in the tropical mountains 247
13.7 Sediment from tropical mountains 249
13.8 Conclusion 251
Questions 252

14 Volcanic landforms 254


14.1 Introduction 254
14.2 Types of volcano and the related landscape 254
14.3 Lava and pyroclastic deposits 257
14.4 Volcaniclastic flows: debris avalanches and flows 262
14.5 Landscape on flood basalts 266
14.6 Conclusion 268
Questions 268

15 Tropical karst 270


15.1 Introduction to karst 270
15.2 Karst in the tropics:€the geographical distribution 271
15.3 Karst hydrology 272
15.4 Dissolution of karst rocks 273
15.5 Karst landforms 274
15.6 Karst in the tropics 281
15.7 Tropical karst as an environment 284
Questions 286

16 Quaternary in the tropics 287


16.1 Introduction 287
16.2 History and structure of the Quaternary 288
16.3 Quaternary glaciation in the tropics 290
16.4 Climate change 292
16.5 Sea-level change 295
16.6 The Ganga River system:€Quaternary adjustments 298
16.7 Quaternary changes around the Sunda Shelf 299
16.8 Conclusion 302
Questions 303
xi Contents

Part III╇ Anthropogenic changes

17 Anthropogenic alteration of geomorphic processes in the tropics 307


17.1 The beginning 307
17.2 Deforestation, land use changes and rural migration 311
17.3 Temporal and seasonal patterns of sediment transport 314
17.4 Spatial transfer of sediment 315
17.5 Impoundments along rivers and their effects 319
17.6 Application of geomorphology towards a better environment 323
Questions 324

18 Urban geomorphology in the tropics 325


18.1 Introduction to urban geomorphology 325
18.2 Urbanisation in developing countries 328
18.3 Three examples of geomorphic hazards and their amelioration 330
18.4 The general nature of urban geomorphological problems 337
18.5 Geomorphology and urban management 338
Questions 341

19 The future with climate change 342


19.1 Climate change and the future 342
19.2 A robust prediction of the effects of climate change in the tropics 343
19.3 Geomorphological adjustments in the tropics from climate and sea-level
changes 345
19.4 The noise effect of anthropogenic changes 346
19.5 Tropical geomorphology in the near future 347
Questions 348

References 349
Index 374
Colour plates appear between pages 212 and 213.
Preface

This is an introduction to a very large part of the world’s surface with rich and varied land-
forms. The tropics include high mountain ranges, major rivers, ancient surfaces, large allu-
vial plains and deltas, arid landscapes, and wonderful examples of volcanic landforms. The
tropical oceanic coverage is huge and it influences the world’s climate. It is surprising that,
in spite of a recent spurt in case studies, our knowledge regarding the geomorphology of
the tropics remains limited and that case studies from the tropics have hardly been used for
generalisation and theory construction. This lacuna is fascinating, especially as all world
maps on sedimentation rates indicate that huge amounts of sediment are pouring into the
oceans from certain parts of the tropics, as a result of events happening inland.
No single template can exist for tropical geomorphology given the wide variations in
regional geology, climate, and land cover. A major part of the tropics carries old subdued
landscapes that have evolved since the Gondwana era, whereas other parts, including active
plate boundaries and large alluvial plains, are much younger and may record a rapid rate
of erosion and sedimentation. The original land cover is changing drastically and the cur-
rent rates of geomorphic processes are no longer natural everywhere. The old images of a
chemical-weathering driven, deep regolith-covered landscape of large plains and isolated
hills are only partly correct.
This book is an attempt to present the tropics in their rich and varied reality. This object-
ive has determined the selection and arrangement of topics included for discussion and the
level at which they have been covered. The book updates the concept of tropical geomorph-
ology in stressing the increasingly important anthropogenic alterations of the landscape.
The book ends with an attempt to look into the future of the tropics, given current modifica-
tions such as climate change. The approach of the book also defines the expectations from
its users. It is written primarily as an advanced undergraduate textbook and assumes that
the readers already have a basic background in physical geography or geology. It makes no
further demands apart from an interest in the tropics.
Two aspects of the book should be mentioned. It is well illustrated, as a book in geo-
morphology should be, and the illustrations are of equal importance to the text. The illus-
trations include several high-resolution satellite images from IKONOS. Given the 1 m
resolution of these images, they are a wonderful tool for future geomorphological studies,
a tool which is probably not as well used as it should be. We now even have commercial
satellites capable of providing images at 50 cm resolution. The second aspect of the book
is the set of questions at the end of the chapters. These are a mixed batch; some for prob-
lem solving, others encouraging the reader to think in more detail beyond the text. Like the
illustrations, these questions complement the text.

xiii
xiv Preface

The formulation of this book started decades ago, when I was a student in Presidency
College, Kolkata and had trouble matching what I read in books with the landscape
around me or even with the landscape displayed on the topographical sheets of India. I am
immensely grateful to the late M. G. Wolman for allowing me to do my PhD fieldwork in
Jamaica which was a liberating educational experience. A long stay in Singapore permitted
field access to Southeast Asia. My various friends and colleagues completed my education
by allowing me to work with them in the field in various parts of the tropics. I am indebted
to all of them.
The introductory chapter has benefited tremendously from the comments of Professor
Wolman. Parts of the book were also read and commented on by Liew Soo Chin, Jean-
Claude Thouret, Richard Corlett, and Anthea Fraser Gupta.The Centre for Remote Imaging,
Sensing and Processing (CRISP), National University of Singapore very kindly allowed
me to use satellite images from their archives. I have also been permitted to use illustra-
tions from a number of publications, which are acknowledged in specific places. I should
acknowledge the kindness of Lee Li Kheng in drafting the figures and of David Appleyard
in transforming my photographs into publishable material. I am grateful to Jean Rollinson
for her careful copy-editing and to Laura Clark of Cambridge University Press for guiding
me through the production stage.
Avijit Gupta
Part I

The Tropical Environment


1 Introduction

The tropics are my element, and I have never been so constantly healthy as in the last
two years.
Alexander von Humboldt

1.1╇ Geomorphology in the tropics

In tropical geomorphology we are constantly surprised by new discoveries. Currently, we


have a limited understanding of the geomorphic processes, landforms and sediment in the
tropics. Furthermore, the rapid, ongoing anthropogenic development in the tropics con-
tinues to modify landforms and operating processes and change the rates of erosion and
sedimentation from the expected natural norms. Geomorphology in the tropics provides
twin opportunities to discover new facts and to apply such information to managing the
environment for a sustainable future. Tropical geomorphology thus has a tendency to look
forward rather than look back exclusively at past landforms.
The tropics are in essence a climatic region, although the only shared meteorological
component across the belt of low latitudes is high temperature. Considerable climatic vari-
ations exist across the tropical zone:€the most impressive of which is the variation in rain-
fall. The annual total, the seasonal pattern and occasional synoptic disturbances all vary
across the tropics. The Amazon lowlands, the Rift Valley of Africa, Raub al Khali of the
Arabian Peninsula, the Ganga–Brahmaputra Delta, the wetlands of eastern Sumatra and a
considerable part of the Red Heart of Australia are all areas of low elevation in the tropics,
but they exhibit huge differences in rainfall.
The tropics can be divided into two primary units based on annual rainfall:€ the
humid tropics and the arid topics. The transition between the two can be sharp (for
example, where an orographic barrier prevails), or gradational (with a subhumid zone
in between). About half of the tropical land surface is humid, with the annual rainfall
exceeding annual evapotranspiration. The rest is subhumid or arid. Certain climatic
characteristics, such as high temperature, high intensity of rainfall and high potential
evapotranspiration are generally associated with the tropics but do not occur with the
same intensity everywhere.
Formally, the tropics can be defined as an area of radiative surplus at the Earth–Â�
atmosphere interface, bounded by anticyclonic circulations near the 30° north and south
latitudes (Reynolds, 1985). The margin of the tropics is best perceived in a pragmatic
fashion as a fluctuating boundary, between 30 and 35 degrees of latitude. It is an area
3
4 Introduction

Fig. 1.1 Major pressure belts and wind systems of the Earth. Shaded area approximately indicates tropics and subtropics. The
vertical circulation of air is shown diagrammatically towards the right of the figure

over which the Hadley Cell operates, and the three-dimensional pattern of air movement
shows a distinct separation between the Hadley and the Ferrel Cells at these latitudes (Fig.
1.1). The two traditional latitude markers, identified as Tropic of Cancer and Tropic of
Capricorn, are not effective boundaries. Thus, the tropics, as defined in this book, also
include areas conventionally identified as the subtropics.
Nearly 60 per cent of the total surface area of our planet lies between the 35° N and S
�latitudes. The tropical oceanic expanse is huge and plays an important role in �influencing
the climate of the world (Graham et al., 1994). In spite of such extensive coverage,
case studies from the tropics have contributed very little to mainstream theories in
geomorphology.
Like climate, landforms and operating geomorphic processes are not the same across
the tropics. The tropics are an assemblage of active tectonic belts, ancient cratons, allu-
vial valleys and subsiding deltas (Fig. 1.2). The early tropical geomorphologists did not
always recognise such wide-ranging geologic variations, putting too much emphasis on
the hot and humid climate as the prime controlling factor. As a result, the tropics used
to be perceived as a set of climo-morphogenetic landforms, where physical features are
primarily controlled by the ambient climate. The characteristics of the landscape were
generally explained by assuming that they evolved in a hot and humid location over a
very long period of time. The arrival of the theory of plate tectonics finally destroyed
such concepts.
5 Geomorphology in the tropics

(a) (b)

(c) (d)

Fig. 1.2 Varieties of tropical landforms:€(a) Western Himalaya Mountain (high tectonic mountains); (b) arid Central Australia
(arid landscape on ancient craton); (c) Auranga River, India (seasonal river on Gondwana rocks). From Gupta and Dutt,
1989. (www.borntraeger-cramer.de); (d) Mekong Delta (river-dominated part of a Holocene delta). Photographs:
A. Gupta

Landmasses that are now in the tropics were once part of a single large continent on
Earth, known as Pangaea, before its break-up about 200 million years ago. The present trop-
ical landmasses (Australia, part of Southeast Asia, the Indian subcontinent, Africa, South
America) and a number of islands of various dimensions together with the cold Antarctica
constituted the southern part of Pangaea, known as Gondwana or Gondwanaland (Box
1.1). The present physical landforms reflect the entire geological history from Pangaea.
This includes the break-up, the nature of drifting of the separated landmasses towards
the equator, the geomorphic processes that began to operate once they reached the trop-
ical zone, and the current highly active anthropogenic modification of both landforms and
processes. Traditionally, the Indian Peninsula was viewed as an excellent example of an
assemblage of tropical landforms that had developed over a long period of time in a warm
humid climate. The peninsula, however, had a near-polar location before the disintegra-
tion of Pangaea. It then drifted through a wide range of latitudes to reach its present pos-
ition and collided with the Eurasian Plate to form the Himalaya Mountains and fuse with
the Asian landmass (Fig. 1.3). The present monsoon system and operating geomorphic
processes developed mostly after the collision. Landforms and subsurface sediment in the
tropics were also modified by climate and sea-level changes during the Pleistocene. Like
elsewhere on Earth, landforms in the tropics are multifactorial in origin. The history of the
landmasses goes back in time beyond Pangaea, but we will start our narrative from this
particular supercontinent.
6 Introduction

Box 1.1 Gondwana/Gondwanaland


From the Upper Palaeozoic to the Cretaceous, the entire continental landmass of the Earth formed a single
continent called Pangaea. The southern half of Pangaea is collectively known as Gondwana or Gondwanaland,
both names being used interchangeably. Gondwana was located south of the equator and a very large part of
it stretched to high southern latitudes near the South Pole, which, as expected, was a very cold location.
As Pangaea broke up, Gondwana disintegrated into several units which moved away from each other.
These fragments of Gondwana are now known as Antarctica, Australia, India, part of western Asia, Africa,
Madagascar and South America. Except for Antarctica, all land masses drifted towards the equator and to
warmer locations. The evidence of their former connections lies in the similarity of the rocks of the Permo-
Carboniferous age found on all these continents.
At the base of such rocks is a glacial sediment, clay with boulders, the lithified form of which is called
a tillite. This occurs on all continents that were part of Gondwana, indicating the former cold location and
presence of glacial ice. The rocks also contain a characteristic cold-weather flora called Glossopteris (from the
Greek for tongue-like, as the leaves of the flora resembled long tongues) and other associated flora such as
Gangamopteris. Glossopteris is found in rocks of this age on the continents that constituted Gondwana, thereby
indicating their past union and the shared glacial climate. From the Upper Carboniferous to the Jurassic, simi-
lar climatic conditions and rock types occurred on the continents that formed part of Gondwana. Such rocks
were studied in central India by H. B. Medlicott, who first used Gondwanaland as a term in 1872 in an unpub-
lished report (Krishnan, 1982). The name is derived from a part of central India which carries the diagnostic
geology and used to be the kingdom of the Gonds. The tillites were followed by thick deposits of fluviatile or
lacustrine origin, in places deposited in structural basins and bearing coal seams. These in turn were followed
by other sedimentary rocks but without coal. Volcanic lavas are found on top of the Gondwana rocks in a
number of places. The similarity of geology in locations now widely separated from each other indicates the
existence of a single landmass which later disintegrated.
The present physical features reflect geological history. For example, extensive areas in Australia, the
Indian Peninsula, Africa and South America are inherited from Gondwana. As a result these areas are under-
lain by old, hard, generally metamorphosed rocks on which weathering, soil formation and erosional proc-
esses operate very slowly. These are the cratons. Elsewhere the edges of the continents may reflect the history
of their passage and collision with other plates. The Himalayan range was formed following the collision of
India with Eurasia. The Andes was formed by subduction of the Cocos, Nazca and Antarctic Plates underneath
the South American Plate. These active margins now form tectonic mountains which are the source areas of
big rivers that drain the continents. The rivers flow from high active margins across the continent towards the
continental margin that is tectonically passive. These are quiet places where the big rivers deposit their sedi-
ment and build deltas. The Amazon, the biggest of all, is an excellent example. The present river came into
existence after the formation of the Andes to flow eastward along a structural low and to build a huge delta
on the passive side of the continent into the South Atlantic Ocean. In contrast, Africa has not drifted much and
its margins are passive, with tectonic mountains occurring only at the extreme northern and southern ends.
Africa, therefore, is a continent with old surfaces at various levels. The present landscape on any continent
thus reflects its past history inherited from Gondwana, marginal modifications that reflect the nature of its
movement since the break-up of Pangaea, and other happenings over time such as large eruptions of basalt
(western India) or the formation of rift valleys (East Africa).
7 Traditional tropical geomorphology

Fig. 1.3 Drifting of the Indian plate. Adapted from US Geological Survey figure

1.2╇ Traditional tropical geomorphology

We know very little about the geomorphological knowledge that existed before the arrival
of the Western maritime powers in the tropics. There must have been some, given the
widespread successful utilisation of water as a resource, often with designed structures.
After the establishment of global maritime sailing routes, the tropics were visited from
the last decade of the eighteenth century with astonishment, enthusiasm and insight by
Western scientific explorers including Alexander von Humboldt, Alfred Russel Wallace
and Charles Darwin. This revolutionised primarily the biological sciences, but remarkable
observations on regional geology and geomorphology were also registered. We may recall
Humboldt’s work in tropical South America or Wallace’s description of the landforms of
Southeast Asia. In 1859, the year On the Origin of Species was published, Darwin was hon-
oured by the Royal Society for his research, and the citation listed his contribution to the
geology of the Andes. We should also recognise the work of German geologists in various
parts of the tropics: Sapper, Passarge, Bornhardt and others, and that of Dana (1850) from
the United States in the Pacific. Dana provided a very early discussion on fluvial processes
in the tropics. Later, as part of colonial governance, land- and soil-related information
was collected, processed and published in a number of countries. Some of these are now
classics, such as reports by Buchanan on Indian laterite, and Mohr and Van Baren on trop-
ical soils (Buchanan, 1807; Mohr and Van Baren, 1954). Topographical maps and official
8 Introduction

geological reports also started to appear. All these later became very useful for studying
local geomorphology.
Up to the second half of the twentieth century, or even later, the teaching of geomorph-
ology in the tropical countries was carried out using standard textbooks written for European
or North American students (Gupta, 1993). These did not reflect the ambient landscape and
local or regional examples were rarely furnished. Large Andean landslides, coral reefs
or tropical karsts were mentioned, but the common reference to tropical geomorphology
was the climo-morphogenetic regions of Peltier (1950) or De Martonne (1951) or Büdel
(1982). Krynine’s (1936) approach was different; he studied the relationship between geo-
morphology and sedimentology in the humid tropics. The general emphasis was on land-
forms, and climatic characteristics were used to explain their nature and distribution. For
years, geomorphological studies in the tropics were driven by two concepts:€the climo-
�morphogenetic region and the Davisian cycle of erosion. The typical tropical landscape
was perceived as a stable erosion surface, studded with low hills (called inselbergs or born-
hardts) and underlain by soil and deeply weathered rock. Geomorphology was a descrip-
tion of landforms.

1.3╇ Modernisation of tropical geomorphology

Tropical geomorphology started to lose its essentially climate-based approach in the second
half of the twentieth century. The writings of L. C. King (1951, 1962) on the South African
scenery led the practitioners to think in terms of rock types and geological history as
important explanations for the ambient landscape. About this time, a number of case studies
started to appear that were more in line with process-based approaches prevalent in other
parts of the world. Early examples include Ruxton and Berry (1957) on granite weathering
in Hong Kong, Fournier (1960) on erosion rates, Simonett (1967) on earthquake-generated
landslides that eroded the mountains of New Guinea, and Coleman’s study (1969) on the
morphology and sedimentation of the Brahmaputra River. These papers highlighted ration-
ality, methodology and excitement in tropical geomorphology. Case studies became easily
available from this time, due to the publication of several regional collections (Jennings
and Mabbutt, 1967; Davies and Williams, 1978; Dardis and Moon, 1988; Warner, 1988).
Prior to these, regional studies were difficult to find as they were published in local jour-
nals, commonly with a limited circulation. All this overlapped with the diffusion of mod-
ern mainstream concepts, techniques and textbooks (Horton, 1945; Strahler, 1952a, 1964;
Leopold et al., 1964; Young, 1972) to the practitioners in the tropics. From the 1970s, such
diffusion, along with the presence of process-oriented geomorphologists in the tropics, led
to a tremendous increase in the quantity and quality of papers in tropical geomorphology.
Textbooks in process geomorphology with a tropical slant started to appear (Douglas,
1977; Faniran and Jeje, 1983; Thomas, 1994), and tropical case studies also were men-
tioned in mainstream textbooks (Schumm, 1977). The new research indicated that the same
geomorphic processes operate in the tropics as elsewhere, but they operate at different rates
and with varying intensities (Selby, 1993). Research publications were generally in English
9 Structure of tropical geomorphology

but, given the large number of tropical countries, research and institutional reports were
also published in a number of other languages. This at times constituted linguistic demands
on the tropical geomorphologist. For example, a researcher on the lower Amazon Basin is
handicapped without knowledge of Portuguese. The diffusion of ongoing research across
the tropics was limited before electronic communication became common. The free distri-
bution of the Tropical Geomorphology Newsletter, published from 1986 to 1996, used to
help. As computer-based searches became common, the problem of inadequate diffusion
of knowledge largely disappeared.

1.4╇ Structure of tropical geomorphology

Tropical geomorphology highlights three areas:


1. geology, landforms and geomorphic processes across the tropics
2. the passage of water and sediment from the mountains to the coast, mainly via river
systems:€a large volume of moisture is in circulation over the humid tropics
3. anthropogenic alteration of the natural rates and processes, associated environmental
degradation, and related geomorphic principles for better environmental management.
We can break these down to a list of specific topics that should be studied (Table 1.1). As
this table indicates, a combination of common and exotic factors characterises tropical
geomorphology.
It is necessary to emphasise that although climate acts as an important control on land-
forms and processes in both humid and arid tropics, the operating processes and the charac-
teristic landforms at a particular location reflect a diversity of causes. Gardner et al. (1987),
writing on locations in Central America and the Caribbean, focused on three selected topics
among a diversity of regional geomorphic processes and landforms:€ karst; alluvial fans;
and tectonism along convergent plate margins. This was justified, as these three are the
most characteristic of the region, representing the dominance of lithology, a rapid fluvial
depositional process and neotectonics. Elsewhere in the Caribbean, other factors such as
hurricanes or volcanoes dominate local geomorphology. The fascination and challenge of
present-day geomorphology in the tropics lies in the recognition of such regional diversity.
Geomorphological rates are high in the humid tropics, especially in high-relief areas
(Fig. 1.4). A number of world maps of suspended sediment yield have been published since
the 1980s (Milliman and Meade, 1983; Walling and Webb, 1987; Milliman and Syvitski,
1992). All such maps show elevated rates for most of the humid tropics. Working with a
data set of 280 rivers, Milliman and Syvitski found very high sediment rates for South Asia
and high oceanic islands, and identified basin relief as a very important factor behind the
high sediment yield. They interpreted relief as a surrogate for tectonism and proposed that
‘the entire tectonic milieu of fractured and brecciated rocks, oversteepened slopes, seismic
and volcanic activity … promotes the large sediment yields from active orogenic belts’
(Milliman and Syvitski, 1992:€539–540). The high sediment yield is also due to intense
tropical rain falling on such mountains, for at least part of the year.
10 Introduction

Table 1.1╇A summary description of tropical geomorphology


Topic Description

Major controls in geomorphology Location of tectonic belts, volcanoes, cratons,


alluvial valleys, deltas, etc. as determined by
plate tectonics
Wind pattern and rainfall systems (especially
tropical storms)
Distribution of vegetation cover
Deforestation, agricultural expansion, urbanisa-
tion and channel controls
Major operating processes; same as in other Tropical weathering, and its effect on slope
parts of the world, but different in rates material and river load
and relative importance Mass movements on tropical slopes
Rivers, a number of which are seasonal and
prone to flooding
Glacial, glacio-fluvial and fluvial processes
operating on high mountain slopes
Fluvial and aeolian processes in the arid tropics
Coastal processes, presence of mangroves, salt
marshes and coral reefs
Tectonic movements and volcanism
Quaternary inheritance Pleistocene glaciations of the tropical
mountains
Climate change
Sea-level changes affecting coasts and lower
river reaches
Present and future changes Common anthropogenic changes
Global warming and climate change

The natural rates of erosion, however, are not necessarily in operation everywhere. Parts
of the tropics have a long history of intense human occupation leading to modification
of the natural landscape. This process has accelerated since about 1950 and the present
landforms and processes may be natural only in certain areas. For example, destruction of
the tropical rain forest has increased the annual sediment yield in Southeast Asia from less
than 102 tkm−2 to more than 103 tkm−2 (Gupta and Krishnan, 1994; Gupta, 2005a). Transfer
and storage of this excessive sediment sequentially affect hillslopes, gullies, rivers and
coastal forms. Satellite images, as on the cover of this book, show sediment plumes at the
mouth of rivers whose basins are undergoing significant anthropogenic changes (Gupta
and Krishnan, 1994). The high rate of erosion and sedimentation therefore is a function of
relief, climate and land use − all three.
Destruction of vegetation cover and accelerated instability of slopes lead to excessive
sediment transfer and storage in the river systems. In contrast, thousands of large dams
globally block the passage of sediment downstream, the total world volume rising to about
50 km3 of sediment each year (Mahmood, 1987). We need to keep all such modifications
11 Structure of the book

Fig. 1.4 Change of state of landforms. Note according to his estimate, a combination of high relief and climate leads to
approximately the same amount of change in landforms in ten years in the Himalaya Mountains as has occurred since
the end of the last glacial advance in Western Europe. The graphs for New Zealand compare the role of the third factor
in accelerating geomorphic processes:€land use. From Selby (1993). By permission of Oxford University Press

in mind. Environmental problems in the developing countries that constitute most of the
tropics involve land and water degradation to a large extent. The recognition and solution
of such problems depend on geomorphological knowledge of the area. Current tropical
geomorphology therefore combines scientific research with the application of its findings
for the betterment of people. Such a linkage will become crucial with climate change that
is expected to alter the rate and intensity of operating geomorphic processes.
In one of the crime fictions of Ellery Queen (1934), a detective was described as a
prophet looking backwards. A tropical geomorphologist frequently needs to be that in order
to explain a particular landform. At the same time, the geomorphologist also needs to oper-
ate in the conventional prophetic fashion, predicting the future determined by anthropo-
genic alterations including climate change. It is a challenging but wonderful occupation.

1.5╇ Structure of the book

This book is divided into three sections. The first section introduces the physical back-
ground of the tropics against which the operating geomorphological processes are sub-
sequently presented. An introduction to the geological framework of the tropics and
landforms (Chapter 2) is followed by a discussion of tropical hydrology (Chapter 3), and
an account of vegetation cover and land use including a brief introduction to anthropogenic
modification of the land and water (Chapter 4).
Part 2, dealing with geomorphic processes, starts with weathering that prepares slope
material and river sediment for subsequent soil formation and removal (Chapter 5). This is
followed by chapters on slopes and slope processes (Chapter 6) and tropical rivers (Chapter
7–9). The coastal environment is discussed next:€first as a general case (Chapter 10) and
12 Introduction

then regarding the special case of tropical deltas (Chapter 11). Hydrology, river systems and
aeolian processes of the arid tropics are presented and compared with the forms and proc-
esses of the humid tropics in Chapter 12. The next chapter describes the tropical mountains,
both glaciated and non-glaciated, as sources of water and sediment to rivers and the storage
and transfer of both along river systems. Chapter 14 is an exposition on volcanic landforms
in the tropics as a special case and Chapter 15 is on tropical limestone geomorphology. The
section on geomorphic processes ends with a brief discussion on Quaternary inheritance
and its inertia effects (Chapter 16).
The third part of the book deals with anthropogenic influences of various kinds:€appli-
cation of geomorphology in the management of the tropical environment, the urban land-
scape and the possible future scenarios due to climate change (Chapters 17–19). Although
the chapters are structured separately for easy comprehension, the processes discussed
individually usually come together to shape the land and to determine the passage of water
and sediment from the higher parts of a drainage basin to the sea. In large parts of the humid
tropics, this transfer is naturally efficient as mapped by Milliman and Syvitski (1992) and
other researchers. Anthropogenic modifications and climate change accelerate such rates,
and comprehension of such changes rises out of our understanding of tropical landforms
and processes.
2 Geological framework of the tropical lands

Civilization exists by geological consent, subject to change without notice.


Will Durant

2.1╇ Introduction

A little more than a century ago, William Morris Davis described the form of the land as
a function of three variables:€structure, process and time (Davis, 1899). The statement is
equally true for tropical landforms. Geology is usually the primary determinant of land:€its
elevation, steepness and stability. Understanding of tropical landforms should start with
an introduction to the geological framework. The distribution of tectonic mountains that
produce large quantities of sediment; stable areas underlain by old rocks (cratons) that
produce very little; and large river valleys where the sediment is temporarily stored while
being transported to the sea are all determined by plate tectonics. A familiarity with the
distribution of plates and their boundaries in the tropics is necessary to understand regional
landforms and operating processes.

2.2╇ A brief introduction to plate tectonics

The theory of plate tectonics started to develop in the 1960s. The surface of the Earth is
divided into several three-dimensional solid bodies called plates that drift slowly, usually
at the rate of several centimetres a year. It has been pointed out that this is the rate at which
our fingernails grow. The top surface of a plate may form the ocean floor or a continent or a
combination of both. Everything on a plate moves with it. The plates come in various sizes.
The largest plates are seven in number:€the Australian–Indian Plate, the Antarctic Plate, the
Eurasian Plate, the African Plate, the Pacific Plate, the North American Plate and the South
American Plate. Plates of smaller size include examples such as the Philippine Plate, the
Caribbean Plate and the Nazca Plate (Fig. 2.1).
The plates are able to move because temperature and pressure increase with depth inside
the Earth. At a depth of about 120 km the temperature is approximately 1350°C. This sof-
tens the ambient material, allowing rigid plates to move slowly through it. The rigid layer
of rocks at the surface is known as the lithosphere; and the softer part of the upper mantle
through which the plates move is the asthenosphere.
13
Fig. 2.1 Map of the tropical latitudes, showing the major tectonic plates
15 A brief introduction to plate tectonics

(a)

(b)

(c)

(d)

Fig. 2.2 Types of plate boundary. (a) Divergent boundary; (b) Convergent boundary between an oceanic and a continental
plate; (c) Convergent boundary between two continental plates; (d) Transform fault boundary

There are three types of inter-plate boundary (also called margins) (Fig. 2.2):€divergent
boundary (where two plates move away from each other), convergent boundary (where
two plates move towards each other and ultimately collide) and transform fault boundary
(where two plates move past each other). In certain places, three plates meet at a point; this
special case is known as the triple junction.
16 Geological framework of the tropical lands

2.2.1╇ Divergent boundary

Here two plates move away from each other, allowing an opening to develop on the litho-
spheric surface between them. The African Rift Valley is an example where plates begin
to split. With time, plates on either side of a new rift move away from the break to form a
linear sea. The Red Sea is a good example of a linear sea. Eventually, after tens of millions
of years, the plates move away far enough to transform a linear sea to an ocean that fills
the wide depressed space between the plates, e.g. the Atlantic. Divergent boundaries below
oceans are marked by a mid-ocean ridge, actually two parallel ridges separated by a linear
depression where lava emerges from below the crust.
New molten rock material emerges in the opening from underneath, solidifies, and forms
new crust at the edge of the plates as they diverge. Thus new ocean floors are formed and
the emerging new material usually solidifies as the volcanic rock basalt. The floor of a rift
valley, a linear sea or an ocean is therefore made of basalt of a comparatively young age.
The emergence of new material from below is known as lava eruption. Lavas are usually
very fluid and erupt easily without any explosive activity. The standard examples of diver-
gent boundaries are submerged volcanic ridges below the oceans. Part of the ridge may
emerge above water to form volcanic islands. For example, the mid-Atlantic Ridge runs
north–south through the middle of the Atlantic Ocean and locally emerges to form islands
such as Iceland and Ascension.

2.2.2╇ Convergent boundary

In this case, two plates move towards each other leading to a collision. The convergence can
happen between plates of comparable or different densities. In the first instance, as between two
continental plates, mostly made of granitic rocks, the former ocean between them is squeezed
out before the plates collide. The ocean bottom sediments crumple but do not disappear under-
neath the crust. Instead they and the ocean floor rocks are folded into mountains, along with
emplacement of granitic rocks below and through them. A mountain is thus formed between
the two plates, which are now sutured together. The Himalaya is an excellent example. In
this process, volcanic rocks of the old sea floor, marine sediments and fossils, and other types
of sedimentary rock are folded, squeezed, and thrust over each other. As a result, limestones
formed under shallow marine conditions have ended up high on Mount Everest. The thrusts
may give rise to large regional faults. Seismic activities may be common. Lines of old colli-
sions and joining (sutures) can be identified even when geologically very old mountains, once
very high, have been eroded down to a landscape of small hills and plateaus.
In contrast, when convergence takes place between two plates of unequal density, e.g.
one oceanic and one continental, the heavy one (oceanic) slides and sinks below the light
one (continental) in the form of a wedge. This process is known as subduction. Where sub-
duction takes place, the floor of the ocean is pulled down, forming an ocean-floor trench.
Such trenches form the deepest parts of oceans. As the subducted plate sinks, it is stressed
because of the friction between the plates, and the episodic release from such friction gen-
erates earthquakes. The heat generated from the friction also melts rocks. This molten
material, originating from the sinking plate and rising because of the lower density of the
17 A brief introduction to plate tectonics

melt, passes through the top plate. As the hot melt rises, it reacts with the silica-rich con-
tinental plate, and the chemical composition of the melt is altered before it emerges on the
surface or is stored below. The addition of silica to the molten material makes these melts
viscous (sticky) and less fluid than basalt. As they rise, cool and solidify into rigid rocks,
the viscosity causes the rocks to stick together to form steep-sided volcanoes. The more
silica and alkali there is in their composition, the steeper the sides. The mouth of the vol-
cano (crater) may even solidify for a time, leading to a build-up of pressure underneath, as
the melt rises with steam and various kinds of gas:€CO2, SO2, H2S, etc. Eventually, parts
of the volcano burst open, breaking the rocks into hot little fragments. This is known as a
pyroclastic (fiery fragments) eruption and can be very destructive.
Many varieties of volcanic rock can be formed due to different scales of mixing between
the materials of the two plates, but the common ‘mixed’ rock is andesite (named after the
Andes Mountains, which were formed by subduction). At times, the melt does not emerge
on the surface but cools and solidifies inside the crust to form other varieties of rock. If it is
silica-rich and coarse-grained from slow cooling deep inside the crust, it is usually granite
or a variation of it. Granite cools to form three-dimensional bodies, arching up the surface
to form huge, steep-sided, round-topped hills and mountains. The continental plates are
granitic in composition.
Plate convergence with subduction thus leads to deep offshore trenches, frequent earth-
quakes and a line of mountains with explosive volcanoes. For example, the Pacific Ocean
is ringed by subduction trenches and a line of volcanoes, popularly described as the ‘ring
of fire’.
Subduction controls the topography. A deep trench, the Java Trench, lies offshore south-
west and west of Sumatra, Indonesia where the Indian–Australian Plate subducts below the
Sunda Plate. In the Java Trench, the Indian Ocean floor drops to 6000 m below sea level.
A line of islands occurs towards Sumatra, parallel to the trench, formed by an accumu-
lation of sediment scraped off the colliding plates and rising magma. This is the fore-arc
ridge associated with subduction. These islands are separated from Sumatra by a narrow
sea, the fore-arc basin, where the sea floor is pulled down by the effect of subduction.
Inland, within tens of kilometres of the west coast of Sumatra rises the long range of the
Barisan Mountains with a number of volcanoes, forming the spine of Sumatra and classi-
fied as a volcanic arc. East and northeast of this slightly curved mountain range is the low
eastern plain of Sumatra, filling a back-arc basin by sediments derived from the Barisan
Mountains. The back-arc basin reaches the stable Sunda Plate to the east (Fig. 2.3).
Plate convergence and subduction not only explain the topography of the region, but also
determine that the region will be affected by earthquakes, volcanic eruptions and tsunamis.
The large tsunami of 26 December 2004 eroded back the Aceh coast of northwest Sumatra
for about 500 m, removing almost all the beaches, low sand dunes and wetlands behind the
beaches. Within a couple of years, new beaches, dune and wetlands were built resembling
the pre-tsunami coast (Liew et al., 2008). Large and episodic alterations on the landscape
may thus be modified over time by common geomorphic processes. The appearance of
the landscape at a point in time depends on how long and how successfully the common
geomorphic processes have been operating after the last large episodic alteration like a vol-
canic eruption or an earthquake.
18 Geological framework of the tropical lands

Fig. 2.3 Subduction-related topography of Sumatra and the islands and ocean to its southwest. See text for explanation

New oceanic crust is created at divergent boundaries, but the old oceanic crust is
destroyed by subduction at convergent margins. Thus the total crust of the Earth maintains
its finite area and, as T. J. Wilson remarked, no part of the floor of ocean is older than 160
million years, whereas the age of the planet is 4600 million.

2.2.3╇ Transform fault boundary

Two plates slipping past each other without prominent divergence or convergence give rise
to this kind of boundary. Ragged plate edges tend to lock, leading to a build-up of stress
19 A brief introduction to plate tectonics

that is relieved by episodic tectonic movements. Thus strike-slip faults occur near the sur-
face and commonly with a broad shear zone underneath. Most examples of transform fault
boundaries occur between oceanic plates, but two land examples with destructive earth-
quakes have acquired notoriety. The first is the San Andreas Fault in California between
the Pacific Plate to the west and the North American Plate to the east. The second example
is the North Anatolian Fault in Turkey between the Eurasian and Anatolian Plates. A good
example within the tropics is the movement of the North American and Caribbean Plates
past each other in opposite directions giving rise to earthquakes, faults and other structural
lineaments and landslides. Landforms on the islands of the Greater Antilles display such
effects. The magnitude 7 earthquake of 12 January 2010 in Haiti occurred at this transform
plate boundary at a shallow depth of 13 km, and along the Enriquillo–Plantain Garden
Fault System that runs through southern Haiti, passing near Port-au-Prince (Showstack,
2010).

2.2.4╇ Plate tectonics and landforms

Plate tectonics determines the basic template over which geomorphic processes like riv-
ers, winds and slope failures operate. In general, active earth movements do not tend to
occur in the interior of plates where the old rocks are. Such stable areas inside continents
are called cratons. Parts of cratons that are underlain by metamorphic rocks are shields. If
floored by sedimentary rocks, they are called platforms. These are areas where old rocks
are close to the surface and the relief is usually not high; the surface is commonly in the
form of low, eroded plateaus or hills. Such areas are difficult to weather or erode because
of the underlying old, hard, often metamorphosed rocks, and not much sediment is derived
from these areas. In contrast, convergent plate boundaries are active areas where tectonism,
often manifested in earthquakes, volcanism and mountain building, maintains a high relief
and a steep gradient.
Simonett, working in Papua New Guinea, estimated that the erosion rate of the
�earthquake-prone Toricelli Mountains (where frequent landslides are linked with tectonics)
is about 100 cm per 1000 years. In comparison, the denudation rate is 22 cm per 1000 years
in the neighbouring Betwani Mountains which are less affected by earthquakes (Simonett,
1967). Steep tectonic mountains are prone to slope failures, contributing a large propor-
tion of sediment that travels down major rivers whose headwaters originate in mountains
associated with active plate margins. Slope failures in the tropical mountains tend to con-
tribute most of the river sediment, irrespective of the size of the river. Larsen and Torres
Sánchez (1992) calculated that 81 per cent of the total sediment transported out of the small
Mameyes Basin in eastern Puerto Rico comes via mass movements.
The stressed rocks in these mountains adjust by faulting, folding and flowage. Thus
steep and high mountains are subjected to frequent slope failures along plains of structural
weakness and, furthermore, if such mountains form orographic barriers against moisture-
bearing winds, floods tend to be common. Even a slight tilting of a river slope increases its
energy, causing it to deepen its channel and flow through a gorge. Explosive volcanic flows
that emerge from the craters of convergent zone volcanoes end as sediment flows along
20 Geological framework of the tropical lands

river channels draining the lower slopes. As a result, convergent zones are erosion-prone
and produce large volumes of sediment.
Divergent plate boundaries are not common on the continents, but where they occur, rift
valleys are formed with possible volcanic eruptions, as in East Africa. Transform boundar-
ies give rise to strike-slip faults which disrupt pre-existing rivers and ridges giving rise to
various types of offset. A linear trough occurs along the fault. As a result, small depressions
such as sag ponds and stream diversions are formed along the linear feature. The vertical
barriers created by the fault tend to disrupt former drainage lines creating scarps, springs
and interrupted river systems. Small ridges create obstacles across drainage lines. These
are called shutter ridges.

2.3╇ Major landforms across the tropics

A variety of landforms occur across the tropics, depending on their past history. Parts
of the tropics consist of older rocks, inherited from the Gondwana. These areas usually
form uplands or basins at various levels, generally made of old metamorphosed rocks that
are difficult to erode. The Brazilian Shield is a good example. Parts of the old uplands
are formed by flood basalts (the Deccan Plateau of India) and granitic bodies (the Main
Range of Malay Peninsula) which having undergone considerable weathering and erosion
have given rise to flat-top plateaus and steep rounded hills respectively. The movement
of plates after the fragmentation of Pangaea gave rise to convergent plate boundaries,
which on the active side of moving continental plates created high mountains such as
the Andes and the Himalaya. Large structural downwarps have been filled with sediment
to form wide plains such as the Amazon and the Ganga valleys. Sediment derived from
the higher parts of the continents has built coastal plains over the shallower parts of the
continental shelf. A higher variation of landforms occurs at a smaller scale. The structure
and rock types therefore induce a range of landforms across the tropics which, given the
ambient climate, have undergone and are undergoing appropriate geomorphic processes
to produce the tropical landforms we are familiar with. We will go through a very brief
account of the geographical distribution of large-scale landforms in the tropics as an aide-
mémoire for the subsequent material in the book. This is a quick summary and not a
detailed description.
The active side of South America (the west) is dominated by the spectacular Andes
Mountains, formed at the convergent plate boundary between the South American Plate to
the east and the Nazca, Cocos and Antarctica plates to the west. The northern and central
parts of this long mountain chain fall within the tropical latitudes. The building of the Andes,
however, happened over a long period of time starting from the Mesozoic and includes
folding, thrusting, crustal thickening, uplift, block faulting, granitic intrusions and volcanic
eruptions. The history of mountain building differs along the length of the Andes.
The Andes is a long and high mountainous region which includes several lines of ranges,
not all of which are continuous, and with the higher peaks in the central region. From west
to east, the first line of ranges is the Coastal Cordillera, which rises from a narrow coastal
21 Major landforms across the tropics

Fig. 2.4 Major morphological divisions of tropical South America. From Gansser 1973. By permission of the Geological Society
and A. Gansser

plain bordering the Pacific. Eastward, this range is succeeded by the Western Cordillera.
The Central Cordillera lies towards the north, and the line of the Eastern Cordillera, which
comprises elevated but narrow ranges, marks the eastern Andes. A large structural basin
between the Western and Eastern Cordilleras has formed a high plateau, the Altiplano.
Granitic intrusions and volcanic activities form high peaks. Hundreds of volcanic peaks
rise to 5000–7000 m in the high Andes. This is a landscape of high elevation and relief,
steep slopes and deep valleys, and tectonically shattered rocks and volcanic material pro-
ducing a huge amount of sediment that is drained eastward by the drainage systems of
the Amazon and the Orinoco. The derived sediment filled a foreland basin created by the
topographic load developed by crustal thickening and the eastward extension of the central
Andes (Horton and Decelles, 2001). The large alluvial plains slope gently away eastward
from the mountains (Fig. 2.4).
Beyond the foreland, the Amazon and the Orinoco flow through huge lowlands; the
Amazon between the ancient rocks of the Guyana and Brazilian Shields (Fig. 2.5) and
the Orinoco between the ancient rocks of the Guyana Shield and the younger ones
of the eastern Andes and the mountains of northern Venezuela. It is the large-scale
22 Geological framework of the tropical lands

Fig. 2.5 The old landforms on the low-relief Brazilian Shield with inselbergs. Photograph: A. Gupta

structural features and the past geological history that created the landforms and also
determined the nature and volume of sediment drained by the two large rivers to the
Atlantic Ocean.
Northwards, the narrow landmass of Central America is marked by a central mountain
system, alluvial fans, small coastal plains and karst. Similar features continue northwards
into Mexico, dominated by steep mountains and intermontane basins. Coastal plains are
generally limited in area except towards the Yucatan Peninsula, where an excellent devel-
opment of a karst landscape has happened. Episodic high-magnitude events are com-
mon and important for regional geomorphology. For example, a large mudslide following
heavy rains devastated hillslopes in southern Mexico on 28 September 2010.
The Caribbean islands of the Greater Antilles (Cuba, Hispaniola, Jamaica and Puerto
Rico) have central highlands with steep slopes, large areas of karst landscape and coastal
plains of varying size. The islands of the Lesser Antilles are smaller and either domi-
nated by a volcano (Martinique) or karst landscape (Barbados). Both features are com-
mon. The geomorphology of the Caribbean Islands also needs to respond to episodic
events of landslides and floods following the arrival of tropical storms and Atlantic
hurricanes.
The landforms and tectonics of Africa are different. Africa is distinguished by a mean
elevation which is higher than that of other continents. A very large part of the contin-
ent consists of high extensive plateaus and large basins. Since the Palaeozoic, Africa
has been controlled by crustal extension and rifting. Significant compressional move-
ments have occurred only at the extreme north (the Atlas Mountains) and the extreme
south (the Cape Fold Belt). Southern Africa forms the higher part of the continent.
A number of very large escarpments mark the edge of the African highlands, e.g. the
Great Escarpment of Southern Africa. Highlands above such escarpments are generally
characterised by low relief. As a result, internal or centripetal drainage operates over
parts of Africa and a low volume of sediment reaches the seas. The aridity of the North
(Sahara) and the Southwest (Kalahari, Namib) is also responsible for a low sediment
transfer. A considerable volume of sediment originating from the volcanic and rifted
23 Major landforms across the tropics

highlands of East Africa used to travel down the Nile, but anthropogenic impoundments
have ended that.
Repeated rifting episodes attributed high-relief landforms to parts of Africa
(Lambiase, 1989). Two major periods of uplift and rifting should be recognised: a
Jurassic–Cretaceous rifting associated with the fragmentation of Gondwana (180–130
million years) and a rifting phase that started 35–25 million years ago and is still con-
tinuing (Summerfield, 1991). The East African Rift System has given rise to steep slopes
and deep lakes such as Lake Tanganyika. Not all the rifts are associated with volcanism
(Lambiase, 1989) but some are. The eastern branch of the Late-Cenozoic East African
Rift System and parts of several older rifts are associated with volcanism. Volcanism
is not always associated with rifting, as the Neogene volcanic rocks of the Tibesti and
Ahaggr Plateaus demonstrate.
Parts of West Asia that fall within the tropical zone display a variety of landforms due to
the presence of active plate boundaries and variations in annual precipitation. Three active
plate margins affect this area. The Arabian Plate is moving away from the African Plate
along the Red Sea and Gulf of Aden. The Arabian Plate is also moving northwards and col-
liding with the Eurasian and Turkish Plates, giving rise to various mountains and basins of
West Asia. River systems drain relatively humid regions, but elsewhere huge areas are arid,
forming rock and sand deserts and coastal salt flats.
South Asia consists of three large physiographic zones:€the Himalaya Mountains, the
Indian Peninsula and the alluvial Indo-Gangetic Plain between the two. The Himalaya
Mountains consist of a series of asymmetrical ranges with steeper southern slopes, sep-
arated by discontinuous elongated valleys. The mountains bend at both the eastern and
western ends, where the rivers Brahmaputra and Indus flow through in deep gorges.
Two high peaks occur at these bends, the Namche Barwa near to the Brahmaputra and
the Nanga Parbat next to the Indus. Between the two peaks the 250–400 km wide moun-
tains extend for 2500 km. The parallel ranges of the Himalaya increase in elevation
from south to north; the main range, known as the Great Himalaya or the Himadri,
exceeds 6500 m in elevation and includes the highest peak in the world, Mount Everest
(8847 m).
The Himalaya can be divided into several units according to lithology, tectonics and
physiography. The units are divided by regional faults and huge thrusts with varying tec-
tonic activity (Valdiya, 2002). The southernmost range, the 900–1500 m Siwalik Hills are
separated from the Indo-Gangetic Plain to the south by the Himalayan Frontal Fault (HFF).
The Siwalik Hills are made of Late Tertiary to Early Quaternary sedimentary rocks known
as molasses. The next range to the north, the 500–2500 m high Lesser Himalaya, is sepa-
rated from the Siwalik Hills by the Main Boundary Thrust (MBT). The Lesser Himalaya
is made of Middle and Late Proterozoic to early Cambrian sedimentary rocks, huge over-
thrusts of Early Proterozoic metamorphic rocks from the north and granitic intrusions. The
6500–7000 m high Great Himalaya lies to the north, separated from the Lesser Himalaya
by the Main Central Thrust (MCT). It is a complex of high-grade metamorphic rocks and
mid-Tertiary granites.
India broke away from the part of Pangaea now known as Madagascar about 86–87€Ma
and rapidly moved towards the Eurasian Plate. The Himalaya started to rise when the
24 Geological framework of the tropical lands

Indian plate collided with the Eurasian plate, starting in 65 Ma. The suture zone is marked
by Cretaceous basic and ultrabasic rocks from the ocean floor. The contact slowed down
the northward movement of India and the welding of the two landmasses was completed
by 55–50 Ma (Patzelt et al., 1996). It is generally accepted that after the collision, the
lower part of the crust under India slid below Asia, probably raising the elevation of the
current Tibetan Plateau, and the upper part detached itself ending in an upwarp.
Valdiya (2002) described two results of this collision. First, the land immediately to the
north of the collision zone was elevated to form a highland (the Kailas–Mansarovar region
of southwestern Tibet), from which large rivers drained in different directions. These riv-
ers, the Indus, Satluj, Karnali and Tsangpo (the Brahmaputra in Tibet), persisted in deep
gorges through subsequent stages of Himalayan orogenesis in the next 35–45 Ma that cre-
ated high mountain chains across their flow. Second, crustal sagging created an elongated
depression south of the rising mountain chain, which temporarily became a shallow marine
basin in which rivers from the north and south deposited sediment. The rise of the moun-
tains also probably intensified the monsoon system (Valdiya, 2002). As Valdiya (2002)
described it, the new river systems established themselves on the rising mountain, and the
eroded sediment started to travel generally southwards.
By 20 Ma ago, the Himalaya formed an orographic barrier and a water divide. Tectonic
disturbances, especially on the Lesser Himalaya, continued at intervals. Particularly strong
movements at the beginning of the Quaternary generated huge slope failures and a num-
ber of debris flows that carried mud and gravel to form a thick boulder conglomerate
southwards. The Siwalik Hills were created towards the Late Pleistocene by intense com-
pression of such sediment. A narrow northern part of the pre-existing basin south of the
Himalaya was compressed into the rising Siwalik Hills, and its wider southern part sub-
sided to form an elongated basin that was subsequently filled up with sediment to form the
Indo-Gangetic Plains (Valdiya, 2002).
The Himalaya continues to be a very mobile mountain range with a number of active
faults. Levelling across the mountain has indicated that the ranges are rising at an annual
rate of several millimetres per year, with the highest value for the Great Himalaya. The
denudation rate is also high. Burbank et al. (1996) measured an annual erosion rate of
2–12 mm for the Indus River near Nanga Parbat, a rate that has speeded up over time as
determined by cosmogenic isotope dating. Four huge earthquakes with magnitudes over
8.0 occurred in 1897, 1904, 1934 and 1950 in different parts of the mountains. Smaller
earthquakes are frequent.
The Himalayan Ranges are very large mountains with high relief, steeper southern slopes,
varying lithology, shattered rocks and clearly demarcated zones of structural weakness.
Furthermore, the mountains are frequently disturbed tectonically. The Himalaya, because
of its elevation, was extensively glaciated during the Pleistocene. All the ranges were gla-
ciated in the past and ice still covers the higher elevations. Mass movements are common,
especially during rainfall in the wet monsoon. A huge volume of sediment is released that is
temporarily stored in valley bottoms and ultimately transported out of the mountains to the
Indo-Gangetic Plains (Fig. 2.6). Almost the entire discharge of water and sediment occur
during the southwestern monsoon in the summer.
25 Major landforms across the tropics

Fig. 2.6 The Himalaya Mountains and Ganga Plains:€a tectonically active high-relief environment. From NASA/MODIS.
See also colour plate section

The Indo-Gangetic Plain lies between the Himalaya and the ancient craton of the Indian
Peninsula. As mentioned, it is a structural foredeep filled by the alluvium of three major
rivers, the Brahmaputra, the Ganga and the Indus, and their tributaries. The relief is sub-
dued except for ridges of ancient rocks emerging in a few places. A number of alluvial fans
soften the contact between the Himalaya and the plains.
The Indian Peninsula, which consists of ancient rocks, lies south of the Indo-Gangetic
Plain, and is separated from it by a series of sandstone and lava plateaus. A line of hills
and plateaus runs parallel to the west coast of the peninsula, marking the drainage div-
ide. This is the Western Ghats, the northern part of which consists of flood basalt (the
Deccan Traps) and the southern part of which is made of Archaean gneisses, schists
and charnockites (a type of metamorphosed igneous rock). The 1600 km long Western
Ghats acts as an orographic barrier to the southwestern monsoon and short steep streams
flow down its western slopes to the Arabian Sea. In contrast, longer rivers such as the
Godavari, Krishna and Kaveri flow eastward to the Bay of Bengal through a range of
disconnected hills of ancient rocks, known as the Eastern Ghats. The only major west-
flowing streams, the Narmada and the Tapi, flow through rifted valleys towards the north
of the peninsula.
Further to the east, Southeast Asia is a corner of Asia which starts from a mountain-
ous region and ends in an assemblage of peninsulas, archipelagos and partially enclosed
seas. From the mountainous region of the northwest that includes the eastern Tibetan
Plateau and the eastern Himalaya Mountains, a number of large elongated river basins
26 Geological framework of the tropical lands

Fig. 2.7 Northern Australia, an old arid landscape. From NASA/MODIS. See also colour plate section

extend north–south or northwest–southeast. These are the basins of the Irrawaddy (or
Ayeyarwady), Salween, Chao Phraya, Mekong and Sông Hóng (also known as the Red
River). An east–west traverse of the mainland of Southeast Asia therefore crosses a ser-
ies of alluvial valleys of large rivers separated by mountain chains or plateaus. Tapponier
et al. (1986) have explained this pattern as related to an eastward lateral extrusion of
Southeast Asia due to the Himalayan orogenesis. Large-scale strike-slip faulting opened
up the basins in Southeast Asia. This also created the present drainage systems of the
region. It has been suggested that a large pre-orogeny drainage net of the Sông Hóng (Red
River) disintegrated (Brookfield, 1998; Clark et al., 2004) due to progressive river capture
following the uplift of the Tibetan Plateau and the present separate large river systems
(Chang Jiang, Mekong, Salween, Irrawaddy) coming into existence. The Chang Jiang is
also known as the Yangtze.
The hills and the valleys end in coastal plains, rocky peninsulas or deltas. Beyond are the
arcuate islands of Indonesia and the Philippines with steep volcanic slopes, intermontane
basins and flat coastal plains, primarily controlled by subduction-related convergent plate
margin activities.
Australia is a low, arid continent (Fig. 2.7) with a mean elevation of 330 m and about
40 per cent of the land area below 200 m (Goudie, 2002). The relief is low over most of
the continent with flat plains dominating the landscape (Twidale and Campbell, 1993).
Australian rocks, and even the surface, tend to be aged, a large fraction being derived from
the Gondwana. Even certain sand dunes have been reported to be Tertiary in age (Benbow,
1990). In general, a low surface with weathered crusts has developed on the ancient rocks
with plateaus and mesas rising from it and inland basins occurring at lower elevations. The
27 Interrelationships

uplands are in bedrock with an association between their forms and rock type. Domes,
tors and pinnacles have formed on granite; escarpments, ridges and cuestas on quartzites
and sandstones; and dissected uplands on metamorphic rocks. The heart of the continent
is mostly stony or sandy deserts, but narrow coastal plains provide a contrast. Aridity, low
relief and hard ancient rocks combine to restrict the volume of sediment carried by the riv-
ers of Australia into the sea.
This is a very brief summary of the major landforms of the tropics, primarily controlled
by plate tectonics. In Part II we examine the various types of geomorphic process that oper-
ate in the tropics on this template.

2.4╇ Interrelationships:€plate tectonics, landforms, erosion and


sediment production

The worldwide delivery of sediment from the land to the ocean via rivers has been
reviewed several times (Milliman and Meade, 1983; Walling and Webb, 1983; Milliman
and Syvitski, 1992; Meade, 1996; Hovius, 1998). In general, basins with high relief cor-
relate with high sediment production, and section 2.3 helps us to understand the sources
and sinks of sediment in the tropics. Milliman and Syvitski (1992) indicated that elevation
and relief is possibly a surrogate variable for tectonism. Relief, as measured from elevation
data, probably reflects, as they describe it, ‘the entire tectonic milieu of fractured and brec-
ciated rocks, oversteepened slopes, seismic and volcanic activity’ (Milliman and Syvitski,
1992:€540) that occurs in high mountains. Active orogenic belts, positioned by plate tecton-
ics, are eroded heavily, becoming primary sediment sources. This is true for the sediment
travelling along large rivers such as the Amazon, Ganga, Brahmaputra and Mekong. The
Amazon is over 6000 km long, but more than 90 per cent of its sediment is derived from
the Andes Mountains and its foothills and carried across this long distance to the South
Atlantic (Meade, 2007).
In tropical South America, four major rivers (Amazon, Magdalena, Orinoco and Paraná)
collectively drain more than half the entire continent, each with their headwaters in the
Andes. The sediment they deliver to the oceans has travelled for thousands of kilometres
taking hundreds, perhaps thousands of years. In contrast, the rivers draining west from the
Andes are short and steep. Sediment transfer is also affected by the generally dry condi-
tions on the western slopes of the Andes. Significant sediment transfer in these rivers there-
fore takes place only when large flows occur sporadically, as during the El Niño Southern
Oscillation (ENSO). Little sediment is sourced from the lowlands or uplands of ancient
rocks such as the Guyana Shield or the Brazilian Shield. This is the common pattern unless
the regional lithology is highly soluble (e.g. consists of carbonate rocks or evaporites) or
can be eroded very easily (Stallard, 1985; Meade, 2007).
The number is high for the east-draining Amazon (about 1500 tkm−2yr−1) but not for
the rivers draining west from the Andes. According to Milliman and Meade, the annual
28 Geological framework of the tropical lands

sediment yield from Central America is a little higher than these west-draining rivers,
being about 230 tkm−2.
The pattern of mountain-derived sediment travelling along short steep rivers in epi-
sodic events such as tropical cyclones is found in many rivers of Central America and the
Caribbean islands. There the mountains are located near the coast due to plate tectonics.
The sediment deposited on a high-gradient coast forms a steep delta resembling a fan,
called a fan-delta.
A comparatively small amount of sediment seems to be deposited in the surrounding
seas by the rivers of Africa (Milliman and Syvitski, 1992). Although the southeastern and
eastern part of tropical Africa is high in elevation and bounded by steep escarpments, the
rocks are ancient and metamorphosed and not as prone to erosion as the fractured rocks of
young folded high mountains. In tropical West Africa, most of the basins of major rivers,
such as those of the Senegal and Niger, are not mountainous. The same pattern frequently
holds true for small river basins. The Congo drains a huge central basin and then goes
over a steep escarpment to flow into the Atlantic Ocean. Although its headwaters drain
the steep slopes of the Western Rift, little sediment reaches the sea. The current average
annual suspended sediment discharge for the Congo is estimated as 43 million tonnes.
Sediment discharged by the Orinoco, which drains a basin less than one-third of the area
of the Congo, is about 150 million tonnes (Meade, 1996). In tropical Africa, rivers that
carry high sediment discharges are those that drain the high relief associated with rifting
in East Africa:€Nile, Zambezi, Limpopo and Rufiji. Even so, many rivers of the world sur-
pass these in sediment discharge. Anthropogenic changes in the channel and river basins
may have a greater influence on the amount of sediment load than the natural environment
in Africa. The impounded Nile is a good example.
In tropical Asia, rivers with high sediment discharge are large rivers that originate in
the Himalaya Mountains:€Ganga, Brahmaputra, Irrawaddy, Salween, Mekong, Zhujiang
and upper Changjiang, (also known as Yangtze). Short streams that drain the moun-
tainous islands of Indonesia, Philippines and Papua New Guinea also contribute high
volumes of sediment given their drainage area. High mountains and volcanoes on these
islands are eroded heavily, giving rise to large sediment loads. Small streams draining
volcanic islands in the Indian or Pacific Oceans may carry a very high sediment load
for their size; the annual contribution from their basins being as high as 3000 tonnes
for each square kilometre (Milliman and Syvitski, 1992). The world map of suspended
sediment discharge prepared by Milliman and Meade (1983) indicates that the highest
annual sediment yield of approximately 3000 tkm−2 is derived from the hilly islands of
Southeast Asia and the Himalayan drainage to the Bay of Bengal. Rivers draining the
Southeast Asian mainland annually contribute more than 900 tkm−2. In contrast, rivers
draining the old landmass of South India were shown to release less than 300 tkm−2yr−1
to the Arabian Sea, a figure comparable to that from East Africa. The rest of Africa
erodes at an even slower rate, the estimated volume of annual sediment yield barely
reaching 100 tkm−2.
Milliman and Syvitski (1992) recognised topography, i.e. tectonics and volcanism, as
the prime explanation of high sediment production, being more crucial than precipitation
29 Questions

and runoff, although the latter are also important factors. Thousands of years of human
occupation of land has, however, significantly changed the natural rate of sediment pro-
duction in two ways. Erosion and sediment production have increased dramatically with
deforestation and farming. Saunders and Young (1983) estimated that such increases
range from 2–3 times (in what they have described as moderately intense land use) to
about 10 times or more with intensive land use and associated soil erosion. On the other
hand, the multiplicity of dams and reservoirs across the world results in the storing of
a huge amount of sediment in the reservoirs, depriving downstream rivers of their sedi-
ment load.
It is a safe generalisation that rivers with high sediment load originate in the mountains.
Most of the large rivers originate in mountains associated with active plate margins (front
edge of a moving plate) but discharge to the sea at passive margins (trailing back edge of
a moving plate), building big deltas with their high sediment load. If the mountains (vol-
canic or otherwise) are located close to a coast, a number of small rivers drain their steeper
shorter side. Similar streams drain mountainous islands. As many of these short rivers flow
into an active plate margin, their sediment deposits may be subducted and deposition forms
at their mouth could be limited in size, perhaps forming a small, steep fan delta. This pat-
tern, however, may change if these slopes are affected by large storms, volcanic eruptions
or earthquakes. Such large events may bring down a huge amount of sediment that needs a
number of years to be reorganised.
Erosion and sedimentation depend on a range of environmental aspects:€ tectonics,
relief, climate, land cover, anthropogenic modification of the land and geomorphic proc-
esses. We look at some of these aspects briefly in the next two chapters before moving on
to a detailed discussion of the operation of geomorphic processes in the tropics.

Questions

1. Which parts of the tropical world originate from Gondwana? Which parts do not?
2. What surface features should you expect on a craton? What would be a reasonable
Â�volume of sediment eroded from a craton? Sediment is commonly measured in tkm−2yr−1.
It is known as the sediment yield.
3. A large volume of sediment is usually derived from high steep mountains forming oro-
graphic barriers to moisture-laden winds. Name any five locations in the tropics where
it happens.
4. List (A) the principal source areas of sediment in the tropics, (B) the large rivers (at least
1000 km in length) that transfer high volumes of sediment to the sea, and (C) where the
sediment is finally deposited by these rivers.
5. Examine Figure 2.6 which is a MODIS (Moderate-Resolution Imaging Spectroradiometer)
image taken from space. (A) Identify the expected areas of high erosion. (B). Which are
the locations where large amount of sediment is expected to be stored?
30 Geological framework of the tropical lands

6. Compare the sediment-producing capacity of the land shown in Figure 2.7 with Figure
2.6.
7. Milliman and Syvitski (1992) identified tectonics and volcanism as the prime factors
behind high sediment production. Do you agree?
8. Is it reasonable to expect the current rates of erosion and sedimentation to be natural
rates?
3 Tropical hydrology

A set of ‘rainfall rules’ to apply in Africa has been suggested by some cynic, perhaps a
tropical meteorologist but certainly a person with experience in the tropics. These rules
are:€(a) the rainfall is seasonal in nature; (b) the amount increases as one approaches
the equator; (c) do not put too much faith in (a) and (b).
J. F. Griffiths

3.1╇ The tropical climate:€a brief review

Climatic characteristics do affect local geomorphology in the tropics, although not as


deterministically as was presented by many geomorphologists in the past. The warm and
moisture-laden condition of the humid tropics certainly accelerates the process of weather-
ing, and the high-intensity rainfall of the tropics gives rise to frequent slope failures and
flooded rivers. In contrast, physical weathering and shorter rare floods from cloudbursts
are common in the arid tropics. This chapter is not on the principles of climatology but
reviews the basic nature of the tropical climate with special reference to the factors that
influence tropical hydrology and geomorphology (for a detailed review of tropical climate
please see McGregor and Nieuwolt, 1998).

3.2╇ Temperature

Climate in the tropics varies according to location. In 1918, the Austrian meteorologist
Köppen characterised the tropics as the region where the average temperature of the cold-
est month is above 18°C. This would be the A-type and part of the B-type in the Köppen–
Geiger classification of world climates (Trewartha, 1954). But the tropics also include
highlands and high mountains where the temperature certainly drops below this designated
threshold, and seasonal differences in temperature with a clearly defined winter become
perceptible at latitudes higher than about 10°. Near the equator a thermal uniformity pre-
vails with year-round high temperature at low altitudes. Here the diurnal range in tem-
perature is usually larger than the annual range. Cloudiness, wherever it occurs, tends to
reduce the diurnal range. If rainfall is strongly seasonal, as it usually is at latitudes above
10°, temperature may drop during the wet season due to both cloudiness and rain. Köppen
described this pattern as the Ganges-type temperature distribution, where the temperature
31
32 Tropical hydrology

Fig. 3.1 Distribution of average annual precipitation in the tropics

rises to the highest level in early summer, then drops at the summer solstice which falls
within the rainy season, and rises again to a post-rain secondary peak before the arrival of
a cool mild winter. Both the annual and diurnal ranges of temperature are high in the arid
tropics. The annual range of temperature is as much as 20–30°C, higher than elsewhere
in the tropics. The annual range may even equate the diurnal range, although the latter is
usually large.

3.3╇ Wind circulation

From the lower latitudes, where there is surplus solar radiation, heat is dissipated to the
higher parallels by wind. This is seen as a three-dimensional circulation known as the
Hadley Cell, a generalised version of which is shown in Fig. 1.1. The surplus heat energy
of the solar radiation is transferred to kinetic energy near the equator, and also evapo-
rates moisture from the ocean. The heated air mass rises with high moisture content. As it
rises, condensation forms clouds, and in some instances rain, releasing part of the latent
heat, which takes part in further warming and lifting of the air mass. This rising mass of
air is then carried to higher latitudes by upper air currents. The upper air currents slow
down around 20–30° of latitude. This is followed by subsidence, building of regional high-
�pressure zones (the subtropical highs) and air mass warming. As a result, drier atmospheric
conditions prevail around 30° latitudes where the arid tropics are located.
At low atmospheric levels, a large and steady movement of air takes place from the
subtropical highs towards the equator thus closing the circulation (Fig. 1.1). As moving air
tends to turn to the right north of the equator and to the left south of it, this low-level wind
steadily blows over the surface of the tropics from the northeast or southeast (depending
on the hemisphere), and is collectively known as the trade winds. The trade winds are
especially steady over the ocean, bringing moisture westward to land surfaces. The north-
east and southeast trade winds converge at a belt of low pressure round the globe near
the equator, known as the Inter Tropical Convergence Zone (ITCZ). This completes the
circulation of the Hadley Cell and contributes to the lifting of moisture-bearing air at the
ITCZ to higher altitudes, leading to equatorial rainfall. The ITCZ shifts its position slightly
between the seasons. The Hadley Cell does not operate quite as simplistically as described
here but this account provides a basic understanding of the general movement of moisture
33 Wind circulation

in the tropics. Oceanic islands with high volcanoes are therefore wetter on their eastern
side, which is the direction from which the moisture-bearing trade winds arrive, and much
drier on the west. The Hawaiian Islands are good examples.
Over the ocean, the trade winds usually operate throughout the year with minor shifts
in latitude and small variations in intensity. In contrast, there is a seasonal rhythm over
the continents and coastal waters, which, being warmer in summer and colder in winter,
influence a seasonal variation in wind direction, particularly in the higher tropics. This
is a gigantic manifestation of land and sea breezes. In winter, subtropical high-pressure
cells develop over the cooler continents, which in summer are replaced by centres of low
pressure over heated landmasses. This gives rise to a seasonal reversal of wind systems
between land and sea, leading to alternate dry and wet seasons, and winds characterised by
this direction reversal are known as monsoons. There are always two monsoons coming
from different directions at different times of the year. Monsoon systems of reversal are
seen over northern Australia, tropical Asia (South, Southeast and parts of East Asia), East
and West Africa, and parts of southwestern North America.
The Asian monsoon is influenced by two factors:€(1) the seasonal thermal contrast over
land and (2) the east–west alignment of the high Himalaya Mountains. In the winter sea-
son, winds near the surface flow round the Himalaya and the Tibetan Plateau, moving east-
ward towards the coast of East Asia where they meet the moisture-carrying trade wind. The
combination turns southeast to bring rain over the eastern coasts of Indochina and Malay
peninsulas. This is the northeastern monsoon. When the wind systems reach Indonesia,
further south, and the equator, they are well saturated, warm and unstable, and continue to
produce rain. On crossing the equator, these winds change direction and blowing from the
northwest bring seasonal rainfall to Northern Australia. In contrast, winds over Northern
Australia come from the southeasterly direction during the southern hemispheric winter.
Rainfall over Northern Australia is seasonal and individual disturbances with embedded
cells of high rainfall arrive in the wet season, causing bursts of heavy rain. This is a com-
mon characteristic of the wet monsoon, the monsoon that brings rain.
Over Asia, the summer monsoon arrives at different times in different places. It is a deep
circulation, reaching 6000 m over India and even higher over Myanmar (Nieuwolt, 1977).
Most of the rain that falls over South Asia, western coasts of Southeast Asia and East Asia
arrives during the summer monsoon. The Indian Ocean and the Western Pacific are the main
sources of moisture. For example, moisture-bearing winds from the East China Sea move
inland up the Chang Jiang (Yangtze) Basin, bringing monsoon rain inland progressively
later. Rainfall from the moisture-bearing air is triggered by a number of factors:€orographic
lifting, convergence and atmospheric disturbances. The disturbances travel embedded in
the monsoon wind system. The passage of a depression is followed by a slackening of rain
for several days until the next depression arrives. This type of pulsating nature is a charac-
teristic of the monsoon system of rainfall. It is also reflected in river flows.
Rain-bearing depressions arrive embedded in the monsoon system of winds during the
wet season. For example, the east coast of the Indian Peninsula is subjected to the passage
of a series of depressions between June and September. Das (1968) mentioned that on
average, eight cyclonic depressions pass from the Bay of Bengal to the eastern coast of
India between June and September. Widespread rain occurs in the southwestern quadrant
34 Tropical hydrology

of these depressions (Rao, 1981). The rain is periodically concentrated by higher-intensity


precipitation pulses. Sikka estimates that 8–10 storms arrive in an average season, with a
lifetime of 2–5 days. Monsoon storms of up to 9 days' duration have been recorded. Such
storms tend to be 1000–1300 km across, with a 7–9 km deep cyclonic circulation with an
anticyclonic outflow on top (Sikka, 1977). Arriving in the middle of the wet monsoon,
when the soil is already saturated and rivers high, these high-rainfall events often create
floods and failures on hillslopes. Tropical disturbances arrive in a similar fashion over the
Indochina Peninsula, usually late in the summer monsoon when the ground is wet and the
rivers are high.
The African monsoon is not as deep as the Asian version. In the northern hemispheric
winter the ITCZ moves close to the equator and dry stable northeasterly winds arrive over
West Africa, frequently carrying dust particles from the Sahara Desert. In contrast, during
the northern hemispheric summer, high temperature and a thermal low build up over West
Africa, and the ITCZ moves northwards to about 15°N. South of it, warm moist south-
westerly winds from the Atlantic Ocean bring rain. In East Africa both monsoons arrive
from the east, and whether they are locally northeasterly or southeasterly depends mostly
on the seasonal position of the ITCZ. The winds, however, are relatively dry and either
convergence or orographic uplift is required for precipitation. Thus, the rainfall over Africa
is not uniformly distributed. For example, the eastern mountain slopes of Madagascar, first
reached by southeasterlies from the Indian Ocean in the southern hemispheric summer,
receive higher rainfall than the continent of Africa (Nieuwolt, 1977). Orographic condi-
tions cause a general increase in rainfall. Seasonal wind reversals and wet and dry differ-
entiations also occur near the coast of the southwestern United States and western Mexico,
known locally as the monsoon.
The arrival of the wet monsoon and the amount of precipitation both vary each year.
Various explanations, such as control by upper-air jet streams or pressure oscillations in
El Niño or La Niña years, etc, have been offered for such variations. Geomorphologically
such variations lead to a high number of slope failures and floods in the wetter years and
less effective erosion and sediment transfer in drier ones.

3.4╇ Precipitation

Average annual total precipitation over the tropics is shown in Fig. 3.1. The amount of
rainfall decreases with distance from the equator unless local factors, such as windward
location on orographic barriers, override this general rule. The ITCZ remains close to the
equator throughout the year producing high annual rainfall with two maxima separated by
two relatively less wet periods (Fig. 3.2). From 10–15º of latitude, the pattern changes pro-
gressively to one rainy season and a long dry period, although the annual rainfall remains
high. At higher latitudes in the tropics, the rainy season becomes shorter, the amount of
rainfall drops, and at around 25–30° the influence of the subtropical high-pressure cells
becomes dominant. Semi-arid and arid tropics appear at these locations. Rainfall inten-
sities are commonly high in the arid tropics, although the duration of rainfall is usually
35 Precipitation

Fig. 3.2 Examples of variations in annual temperature and rainfall in the tropics. Entebbe (Equatorial highland), Jodhpur
(arid subtropics), Mumbai (monsoon) and Townsville (seasonal climate, southern hemisphere). From McGregor and
Nieuwolt, 1998. By permission of Wiley

brief, lasting only for tens of minutes. Over the oceans, the latitudinal pattern of rainfall is
clearer than over the continents, where convection and orographic lifting add complexity
to the general pattern. Even over the oceans, heavy rainfall from disturbances such as trop-
ical cyclones and easterly waves (discussed below) produces deviations from the general
model.
Rain falls over the tropics with high intensity, causing slopes to fail, floods to occur and
land to erode unless it is covered with vegetation. Even on a daily basis, the mean rainfall
intensity may reach figures in tens of millimetres. The short-term intensity is much higher
(Table 3.1). A high percentage of this rain comes in short-duration tropical disturbances
which cause continuous and intense rain while they last. Tropical disturbances that may
last for several days bring in even higher amounts, always with strong intensity. To illus-
trate, rainfall over southern Mexico in late September 2010 exceeded 700 mm in two days,
36 Tropical hydrology

Table 3.1╇ Examples of recorded very heavy one-day rainfall in the humid tropics
Amount Cause
Date (if known) Location (mm) (if known) Source

15–16 Mar. Cilaos, Reunion 1870 Landsberg, foot-


1952 note in Flores &
Balagot (1969)
10–11 Sept. Pai Shi, Taiwan 1248 Typhoon Landsberg, foot-
1963 note in Flores &
Balagot (1969)
14–15 Jul. 1911 Baguio, 1168 Typhoon Flores & Balagot
Philippines (1969)
23 Jan. 1960 Bowden Pen, 1109 Frontal Vickers (1967)
Jamaica
14 Jun. 1876 Cherrapunji, 1036 Monsoon Das (1968)
India depression
31 Aug. 1911 Funkiko, 1034 Jennings (1950)
Taiwan
22 Jan. 1960 Bowden Pen, 977 Frontal Vickers (1967)
Jamaica
6 Nov. 1909 Silver Hill, 775 Incipient Vickers (1967)
Jamaica hurricane
25 May 1889 Cinchona, 718 Mixed Vickers (1967)
Jamaica
Rochambeau, 596 Snow (1976)
Fr. Guinea
Nkhotakota, 570 Torrance (1972)
Malawi
Mumbai, India 548 Rao (1981)
Jaffna, Sri 520 Rao (1981)
Lanka
Zomba, Malawi 509 Torrance (1972)
Diego-Suarez, 508 Griffiths &
Madagascar Ranaivoson
(1972)
Santo Domingo, 508 Portig (1976)
Dominican
Rep.
Darjeeling, 493 Monsoon Rao (1981)
India depression
Dehradun, India 490 Dhar et al. (1984)
Bhuj, India 468 Rao (1981)
Akyab, Burma 465 Nieuwolt (1981)
Tana Tave, 442 Griffiths &
Madagascar Ranaivoson
(1972)
37 Tropical disturbances

Table 3.1 (cont.)


Amount Cause
Date (if known) Location (mm) (if known) Source

10 Apr. 1955 Rungwe, 425 Temple and Rapp


Tanzania (1972)
Oct. 1966 Vango, Kenya 404 Griffiths (1972)

Note:€This table includes only rainfall over 400 mm in a day. Examples of lower figures are too
common to be listed. Figures are rounded to the nearest millimetre. Some figures are one-day
rainfall, while others are rainfall recorded in any 24-hour period. In most cases rain probably
continued for more than a day. Generalised from Gupta, 1988

with about 400 mm falling in a single day. Rainfall from the super-typhoon Megi (October
2010) rose to almost 1200 mm in 48 hours in Taiwan. Such rain is an effective erosive
agent, especially without ground protection. Deforestation removes the protection given by
dense vegetation to the slopes of the humid tropics from such heavy and intense rainfall.

3.5╇ Tropical disturbances

Temporary low-pressure systems of varying size, duration and intensity are superimposed
on the general atmospheric circulation system in the tropics. These disturbances cause
intense and high rainfall which could be associated with surface runoff, floods and slope
failures. Locally, cells of convection may give rise to thunderstorms, which are the smallest
of the tropical disturbances. Thunderstorms also develop when two air masses converge,
forcing the convergent air to rise and condense. A number of separate thunderstorms may
occur along a strong zone of convergence, forming a linear system of squalls often recog-
nised as a line of discrete cumulonimbus clouds rising into the air. Such linear systems may
form due to the convergence of sea breezes over a narrow neck of land or orographic lifting
along mountain crests or coastlines. These squall lines could cover hundreds of kilometres
in length and 10–30 km in width, including zones of both strong and weak thundershowers.
In West Africa, linear thunderstorms may produce a significant amount of rain at the two
extremities of the wet monsoon season. They often arrive in the pre-monsoon summer in
northeastern India and Bangladesh. The seasonal sumatras of southwestern Malaysia and
Singapore are a regionally important linear system of storms. They originate over the east-
ern lowlands of Sumatra but tend to reach the southwestern coast of the Malay Peninsula
in early morning bringing copious rainfall (Nieuwolt, 1977). They comprise a 200–300 km
long band of cumulus and cumulonimbus clouds (McGregor and Nieuwolt, 1998).
Over parts of the tropics, thunderstorms may occur frequently at any time of the year,
producing a large proportion of local rainfall. Thundershowers that instead of travelling
become stationary produce even heavier rainfall. For example, 390 mm of rain fell in
three days in the Virgin Islands in May 1960 resulting in floods (Portig, 1976). Henry
38 Tropical hydrology

Table 3.2╇ Some very heavy multi-day rainfall in the humid tropics
Amount
Date Location (mm) Cause Source

11–19 Mar. 1952 Cilaos, Reunion 4130 Landsberg, in


Flores &
Balagot (1969)
9–16 Jun. 1876 Cherrapunji, 3388 Monsoon Jennings (1950)
India depression
24–30 Jun. 1931 Cherrapunji, 3213 Jennings (1950)
India
22–25 Jan. 1960 Bowden Pen, 2789 Frontal Vickers (1967)
Jamaica
4–11 Nov. 1909 Cinchona, 2287 Hurricanes Jamaica Weather
Jamaica Reports (1909)
14–20 Jul. 1911 Baguio, 2210 Tropical storm Jennings (1950)
Philippines
18–20 Jul. 1913 Funkiko, Taiwan 2071 Tropical storm Jennings (1950)
3–8 Oct. 1963 Tacajo, Cuba 2025 Hurricane Flora Vickers (1967)
5–7 Oct. 1963 Silver Hill, 1524 Hurricane Flora Vickers (1967)
Jamaica

Note:€Generalised from Gupta, 1988

(1974) examined tropical rainstorms (defined as precipitating at least 10 mm of rain) over


parts of Southeast Asia and tropical America. He expected a ‘typical tropical rainstorm’
to cover at least 30 km in diameter. But within it, rain falls with variable intensity in time
and space. Thus rainfall distribution within the area covered by these storms can be patchy,
with inter-storm rain gauges receiving little or no rain. The rainfall at the core of a typical
storm exceeds 25 mm, and more than 60 per cent of the daily rainfall is derived from one
storm. Other studies confirm this picture of a short heavy downpour, in which most of the
rain falls within the first 30 minutes or so, although the storm may tail off slowly. More
than 200 thundershowers a year can be expected over much of Africa (Griffiths, 1972).
Such storms are important in geomorphology because they can cause intense soil erosion.
According to Hudson (1971), the erosive power of tropical rain is about 126 times more
than from rainfall in the temperate areas.
The largest tropical disturbances are tropical cyclones (Box 3.1). Very high-intensity
rain (Table 3.2) has continued for a day or more in the tropical cyclones described below.
These are impressive high-magnitude events that have caused some of the world’s heaviest
rainfall, especially when tropical cyclones have stalled or confronted orographic barriers.
Such heavy rainfall usually causes numerous slope failures and high-magnitude flooding.
Typhoon Morakot in early August 2009 produced rainfall in excess of 600 mm over nearly
the entire southern half of Taiwan. Certain areas received more than 1000 mm of rain on
the western slopes of the central mountain range, giving rise to massive flooding, devas-
tating mudslides, and huge losses of life and property. The geomorphic effects of tropical
storms including the cyclones are discussed in Chapter 7 and also their common locations.
39 Tropical disturbances

Box 3.1 Tropical cyclones


A tropical cyclone is a very large low-pressure system that occurs in the tropics and subtropics and has a
revolving wind velocity of at least 119 kph (33ms−1). Clouds that develop with this system resemble a funnel
with a clear, circular and relatively calm centre, known as the eye. The diameter of the eye ranges between
5 and 60 km. Very low pressure is found in the eye, the lowest value recorded falls in the range 860–890 mb
(Simpson and Riehl, 1981), which contrasts remarkably with normal surface atmospheric pressure of 1013 mb.
Generally, the pressure in the eye of a tropical cyclone is around the 950 mb level. Rapidly circulating winds
occur round the eye with the highest velocity about 1–2 km from its boundary. In most tropical cyclones, a
vertical wall of cloud develops round the eye, which rises into the top of the funnel about 5–10 km above the
surface and then curves outwards, identified as a vortex in satellite images (Fig. 3.3). This huge vortex revolves
counterclockwise in the northern hemisphere and clockwise in the southern. The extent of a tropical cyclone
could be about 1000 km. Generally, two arms of clouds extend out further as long spiral bands. The outflow at
the top is balanced by a strong inflow of air at the lowest levels.
Tropical cyclones develop from persistent but travelling rain disturbances over warm oceans which have
temperatures of 26.5–27°C. Hence they develop only in summer, particularly in late summer when the sea sur-
face temperature (SST) is especially high. Two factors are essential for their formation:€Coriolis force and latent
heat. As the Earth’s Coriolis force is zero at the equator, tropical cyclones are not formed within 5° north and
south of the equator. Once the system is seeded and pulling in evaporated moisture from the ocean surface, it is
driven by the energy released by the latent heat of condensation. The growth of tropical cyclones is also helped
by (a) the presence of an anticyclonic circulation in the upper troposphere that facilitates outflow at the top of
the rotating system and (b) absence of a vertical wind shear allowing development of the tall cumulonimbus
convection (Fig. 3.4).
The small disturbances or lows that may develop into a tropical cyclone usually come from small vorti-
ces near the ITCZ or easterly waves (McGregor and Nieuwolt, 1998). These summer storms form over tropical
oceans except the South Atlantic where the cooler temperature, northerly position of the ITCZ and vertical
wind shear restrict their growth. Table 3.3 summarises their estimated annual frequency over various oceans.
Tropical cyclones are known under different names in different regions:€hurricanes in the tropical Atlantic,

Fig. 3.3â•… Tropical cyclone Katrina over the Gulf of Mexico, 2005. From NASA/MODIS
40 Tropical hydrology

Box 3.1 Continued

Fig. 3.4â•… Diagram of a tropical cyclone. From Simpson and Riehl, 1981. By Permission of Wiley/Blackwell

Table 3.3╇ Estimated annual frequency of tropical cyclones


From Simpson and From Reynolds
Ocean of origin Riehl (1981) (1985)

Northwest Pacific 22 26
Northeast Pacific 15 13
Australian seas 5 10
Northwest Atlantic 8 9
Southern Indian Ocean 5 8
Northern Indian Ocean 8 6
South Pacific 5 6

Note: Generalised from Gupta, 1988

Caribbean, and Northeast and South Pacific; typhoons in the Northwest Pacific; and tropical cyclones in the
Indian Ocean and Australian seas. These huge revolving storms tend to travel from their area of origin, embed-
ded in the trade winds. They may either reach a landmass where they tend to dissipate after destructive activ-
ities on the coastal areas or follow a near-parabolic path circumventing the subtropical anticyclones over the
oceans to reach the westerly winds of temperate latitude and curve back eastward as extra-tropical storms.
The wind strength of a tropical cyclone varies across the system, possibly as a result of the vortex and the
steering air current such as the trades. As a result the front right quadrant of the tropical cyclone, where the
directions of both coincide, has the strongest wind. The clouds are seen as long streets spiralling inward from
the outer edges. A tropical cyclone is capable of producing huge amounts of rainfall (Tables 3.1 and 3.2). This
rain mostly comes from the part of the cyclone where wind velocity is high and thus is difficult to measure with
conventional rain gauges. It is better determined by numerical computations or from satellite radar images.
In general, precipitation from a tropical cyclone shows a logarithmic distribution when the amount is plotted
against the distance from the eye, with the maximum concentration of the rain just outside the core. However,
the absolute volume of rainfall is very high even on the periphery because of the huge size of the storm. Many
41 Tropical disturbances

Table 3.4╇ The Saffir–Simpson damage potential scale


Category of Central Maximum Storm surge
tropical pressure sustained (m above
cyclone (mb) winds (kph) normal) Damage

1 ≥ 980 119–153 ~ 1.5 Damage to shrubbery, trees


and unanchored mobile
homes. Low-lying coastal
roads inundated
2 965–979 154–178 ~ 2–2.5 Considerable damage to trees,
some damage to build-
ing roofs, coastal roads
flooded, possible damage
to moored shipping in
unprotected sites
3 945–964 179–209 ~ 2.6–3.9 Large trees down, small
buildings damaged, serious
flooding at coasts and low
areas, structural damage
near coast
4 920–944 211–249 ~ 4–5.5 Trees down, extensive dam-
age to buildings, structures
near coasts damaged,
coasts and low areas
flooded, beaches eroded
5 (called super- >920 >249 >5.5 Trees down, extensive dam-
typhoons in age to buildings, roof
W. Pacific) failures, small buildings
overturned, extensive
flooding

Generalised from Simpson and Riehl, 1981

tropical cyclones therefore generate rainfall over a substantive part of a river basin and create huge floods. If
the tropical cyclone slows down over a mountainous area, a huge amount of cumulative rainfall occurs over
several days, as happened from Hurricane Flora in 1963 in Cuba and Typhoon Morakot in 2009 in Taiwan.
In geomorphology, tropical cyclones are considered high-magnitude low-frequency events that are cap-
able of causing large landform alterations. Such changes occur due to the strong wind, high rainfall and
storm surges associated with them. They are classified by their properties and ability to damage natural
and anthropogenic features according to the Saffir–Simpson Damage Potential Scale (Table 3.4). As tropical
cyclones make landfall, wind drives huge waves inland and the super-elevation of such waves during high
tide levels are known as storm surges. These are extremely destructive on beaches, barrier islands and dunes.
The level of the storm surge rises in bays and estuaries as these become narrower inland causing even greater
damage. The wind round the eye, and especially in the front right quadrant of the storm, is very high and
42 Tropical hydrology

Box 3.1 Continued


capable of destroying forests, individual trees, buildings and other structures. Apart from the sustained wind
speed of the storm (average wind speed over time, usually measured in minutes), gusts (high instantaneous
wind speed) also cause damage. When hurricanes or typhoons confront slopes, the destruction of vegetation
by wind is followed by huge rainfall, a combination that leads to extensive slope failures, gullies and river
floods (Chapter€7). Tropical cyclones lose their power after crossing the coastline, as they are no longer able
to evaporate water from the ocean surface, a principal source of their energy. They also decay from friction as
they cross mountain barriers across their track. These, however, are times when huge amounts of precipitation
may occur, leading to large-scale flooding. Tropical cyclones thus are effective geomorphologic agents in parts
of the humid tropics.
The Atlantic tropical cyclones were given women’s names from 1953. The system has been modified. At pre-
sent men’s and women’s names alternate, in alphabetical order. There are six standing lists of names which are
used in rotation. For example, the list for 2011 hurricanes runs Arlene, Bret, Cindy, Don, Emily, etc. This list will
be used again in 2017. If a hurricane is particularly destructive, its name is retired and replaced by another one
starting with the same letter. Tropical cyclones for every ocean have their own special set of names. A complete
list is available from the National Hurricane Center (USA) website:€http://www.nhc.noaa.gov.

They are not only powerful but also expected. Over a period of 70 years between 1884 and
1953, 22 storms arrived on average each year in an area bounded by 5°–30° N latitudes and
105°–150°E longitudes, an area off the coasts of South China, the Philippines and Vietnam
(Nieuwolt, 1981). In the 70 years between 1891 and 1960, a total of 360 tropical cyclones
arrived over the Indian subcontinent, averaging 6 per year. About 70 per cent of these
storms were between June and November, when the ground was already saturated and
the rivers high. These tropical cyclones repeatedly cause slopes to fail and rivers to flood,
thereby strongly affecting local landforms and geomorphic processes.
Perturbations in the trade winds may show a wave in the pattern of isobars forming a
low-pressure trough at right angles to the direction of the trade wind. These are known as
easterly waves as these waveforms move with the easterly trades (Fig. 3.5). The waveform
leads to convergence followed by rain, which may reach a large amount if the convergence
takes place near an orographic feature. Heavy rainfall arrives from time to time from east-
erly waves in the Caribbean and eastern Philippines along orographic barriers. Tropical
cyclones may originate from easterly waves as they move across the Atlantic, from the
African shores towards the Caribbean.
Expectations of high-rainfall events have been computed for certain regions. For
example, the 5-year 24-hour rainfall over an area of 280 km2 in eastern Jamaica was
estimated to be between 350 and 600 mm (Lirios, 1969). Vickers (1967) estimated nearly
450 mm for a maximum 24-hour rainfall for a small watershed in southeastern Jamaica.
If these figures are representative, then episodic rainfall of very high magnitude recurs in
the Caribbean and adjoining areas at intervals of several years. Other areas of seasonal
humid tropics may carry similar figures. High-magnitude storm rainfall may be expected
43 Miscellaneous factors

Fig. 3.5 Generalised diagram of an easterly wave. Rain falls from the shaded area

at short intervals within the geographical location of 10 and 30 degrees of latitude, where
tropical cyclones are common in summer. Within such a latitudinal belt, if orographic bar-
riers run across the common paths of these depressions, as reported from the Philippines,
Taiwan, eastern Jamaica, Puerto Rico and Madagascar, rainfall can be especially high and
intense.

3.6╇ Miscellaneous factors

The general character of climate at a particular location in the tropics may alter for a brief
period of time due to the impact of factors such as the release of aerosols in the atmosphere
by volcanic eruptions or the El Niño Southern Oscillation. Large eruptions may release
a high amount of volcanic material into the atmosphere. The coarser fragments return to
Earth within a short time, but the finer parts stay much longer in the atmosphere as aero-
sols. Their presence influences various climatic properties during this period. If the erup-
tion is very large, the temperature of the next year may be reduced and the nature of the
rainfall changed. These may have some minor influence on geomorphological processes,
especially as the volcanic ash travels a long distance worldwide. The June 1991 eruption of
the Pinatubo volcano in Luzon, the Philippines, created a stratospheric aerosol cloud that
had travelled round the Earth by mid-July. The specific eruption that took place on 15 June
1991 was estimated to have released about 20–22 km3 of ash. The temperature of even the
temperate countries fell below average along with increased storminess in the year after the
huge 1815 eruption of Tambora in Sumbawa, eastern Indonesia.
The El Niño Southern Oscillation (ENSO) is a disturbance that occurs periodically, altering
the rainfall and storminess across the tropics. This is related to a change in pattern, known as the
Southern Oscillation, between sea-level barometric pressures over the eastern tropical Pacific
and the western tropical Pacific and Indian Oceans. In a majority of years, winds blow from
the high-pressure area off the west coast of South America across the Pacific Ocean towards
the low-pressure area off East Asia. Blowing across the ocean, this wind generates surface
44 Tropical hydrology

currents, which travel in the same direction as the wind. Surface water off the Peruvian coast
thus is pushed westward and deeper colder water comes up from below to fill the gap. The
process is known as upwelling. The upwelling water is not only cold but also rich in plankton
which attracts fish. The coast of Peru therefore has a cool and dry climate, while warm water
driven by wind across the equatorial Pacific piles up on the Asian side of the ocean.
The atmospheric pressure difference between the East and West Pacific disappears in
certain years. Then the east-to-west wind across the ocean weakens, and the surface water
that was piling up against the coast of East Asia tends to flow back east across the Pacific.
If this condition is fully developed, warm water reaches the coast of Peru, altering the
regional climatic condition (making it warmer and drier) and also reducing the fish catch.
Such a condition is locally known as El Niño. The phenomenon has been named after
infant Jesus by the fishers of Peru, as it tends to develop towards the end of December. The
years of El Niño strongly correlate with years of drought in Australia, parts of Southeast
and South Asia and Africa. In contrast, El Niño years could be times of heavy rains and
floods on the west coasts of the Americas. Such related environmental behaviours across
vast distances are known as teleconnections.
El Niños tend to recur after several years (the average time difference between two El
Niños being 3–7 years), and these conditions last for 1–2 years. El Niños vary in strength.
The El Niños of 1982–83 and 1997–98 were particularly strong, bringing floods, landslides,
forest fires and death and destruction across many parts of the tropics. The 1972–73 El Niño
was remembered for severe droughts in Africa, especially in the Sahel area around the south-
ern boundary of the Sahara Desert. It has been noticed recently that the atmospheric change
that promotes conditions for El Niños to form may occur in certain years at the Central Pacific
rather than off the coast of Peru. Such El Niños are known as the Central Pacific El Niños.
In certain years, the departure from the norm happens the opposite way and cold water
flows east across the Pacific. This is the year of La Niña. During La Niña, rains and floods
may threaten parts of Australia, Asia and Africa and the coastal waters off Peru could be
especially cold. All these cause parts of the tropics to undergo accelerated dry and wet
conditions periodically, which not only affect the economic and social conditions but also
the geomorphic processes by changing the hydrological norms.
Aalto et al. (2003) have shown that major depositional events related to the floodplains
of the Beni and Marmore Rivers in northern Bolivia, tributaries to the Madeira, have a
recurrence interval of about 8 years which corresponds to the La Niña years. Periodic slope
failures and flooding in parts of the tropics therefore occur, interrupting the geomorpho-
logical processes which are associated with the expected climate of a location. A strong
linkage exists between episodic meteorological phenomena and geomorphology.

3.7╇ Water balance

Water balance, a term first used by the climatologist C. W. Thornthwaite in 1944, deter-
mines an accounting relationship between the incoming and outgoing water for a soil pro-
file or a drainage basin. The incoming water could be from various forms of precipitation
45 Water balance

Fig. 3.6 Pattern of water moving through soil

and snowmelt. The outgoing water is the sum of evapotranspiration and overland flow (in
case of a soil profile) or streamflow (in case of a drainage basin). The positive difference
in volume between the incoming and outgoing water is usually stored in the soil; it raises
groundwater level or enhances streamflow (Fig. 3.6). The negative difference leads to dry
soil and low streamflow. It can be written as
P + Sn = AET + SF (or Q) + ∆SM + ∆GWS + GWR (3.1)
where P = Precipitation
Sn = Snowmelt (on high mountains in the tropics)
AET = Actual evapotranspiration
SF = Surface flow
Q = Discharge
ΔSM = Change in soil moisture
ΔGWS = Change in groundwater storage
GWR = Groundwater runoff
Water balance (also called water budget) is based on the two important criteria of precipi-
tation and actual evapotranspiration. If precipitation is greater than actual evapotranspir-
ation, the extra water first fills up the available void space between the solid grains in the
soil, then moves down the soil profile to recharge groundwater. If the incoming volume
of water continues to be more than actual evapotranspiration, as often happens during the
rainy season, the extra water flows out of the soil profile or drainage basin via streams.
Streams receive the excess water either as groundwater runoff or as overland flow running
on the surface. Computing water balance is an easy way of determining the behaviour of
soils and rivers. Instructions for computing water budgets are available in Thornthwaite
and Mather (1957) and Dunne and Leopold (1978).
46 Tropical hydrology

(a)

(b)

(c)

Fig. 3.7 Water balance. (a) Moisture-surplus equatorial location, Januarete, Brazil (0.5°N, 69°W); (b) Moisture-deficit arid
location, Hyderabad, Pakistan (25.5°N, 68.5°E); (c) Seasonal climate, Kericho, Kenya (0.2°S, 35°E). Data for (a) and
(b) from Strahler, 1975, from C. W. Thornthwaite Associates, 1962–65. (c) from Dunne. and Leopold, 1978. With
permission from T. Dunne

Three basic patterns of water balance can be recognised in the tropics. Near the equa-
tor where rainfall is high and a dry season is not conspicuous, precipitation is higher than
actual evapotranspiration (Fig. 3.7a). Here the actual and potential evapotranspirations are
the same because moisture is always available from high rainfall to meet the demand for
evapotranspiration. As a result, streams are seldom dry, the groundwater table is near the
surface except in special cases and the soil is moist. All this is reflected in the ambient
dense natural vegetation.
Arid areas near the subtropical high-pressure systems show the other extreme. In these
areas precipitation is low, but under a hot and cloudless sky potential evapotranspiration
is high. Actual evapotranspiration is less than the potential evapotranspiration due to the
47 Climate and geomorphology in the tropics

limited rainfall. There is not enough precipitation to meet the evaporation demand, and
the water balance is negative throughout the year. The soil is therefore dry, groundwater is
deep and streams are usually low or dry (Fig. 3.7b).
The annual pattern of water balance is more complex in places with strongly seasonal
rainfall. Dunne and Leopold (1978) provided the example of Kericho, Kenya (Fig. 3.7c),
which remains relatively dry during December to February but receives high rainfall in April
and May. Kericho is located at low latitude, but its climate is seasonal due to the location in
the interior and on the western slope of a mountain range. As Figure 3.7c illustrates, the end
of the dry season has a moisture deficit as water is lost by evapotranspiration and rainfall is
absent or low. With a rise in precipitation, soil moisture is recharged from March, becoming
saturated by April. Beyond April, the excess water flows to the river either as groundwater
(a process known as baseflow) or overland (also called surface runoff). Rivers are high so
long as surplus moisture is present, i.e. precipitation is higher than potential evapotranspir-
ation. Potential and actual evapotranspirations are the same at this phase. As precipitation
drops and temperature rises, the volume of precipitation falls below potential evapotranspir-
ation, the difference being met in the early stage of moisture shortage by the evaporation of
soil moisture. Potential and actual evapotranspiration continue to be equal. Some time into
the dry season, water is no longer available from the soil, potential evapotranspiration is no
longer reached, and the period of moisture deficit completes the annual cycle.
Rivers in the seasonal tropics thus fluctuate considerably between wet and dry seasons.
During the wet season the rivers are high. In many parts of the tropics, large storms (such
as tropical cyclones) arrive at this time to cause flooding. Rivers of the monsoon tropics
are not only seasonal in flow but they are also floodprone. Tropical slopes also tend to fail
seasonally, commonly when they carry too much moisture.

3.8╇ Climate and geomorphology in the tropics

Broadly speaking, geomorphological processes in the tropics are influenced by three kinds
of climatic variation. Near the equator, it is nearly always warm and humid with a posi-
tive water balance and, unless anthropogenically altered, the land is under well-developed
vegetation. As discussed in Chapter 5, this leads to advanced weathering and develop-
ment of deep soils in most places (Dudal, 2005; FAO, 1998). High-magnitude storms are
not common and most of the rain is from ITCZ convergence and thundershowers. Away
from the equator, the water balance carries a seasonal component with the associated vari-
ation in streamflow. The rivers are high during the wet season when most of their work is
done. Large tropical storms also are characteristic of seasonal tropics. Weathering will be
advanced but not as much as the equatorial tropics in most places. The arid tropics tend to
have a negative water balance and weathering and soil formation are restricted. Streams
tend to carry water only for a very short period after a rare rainstorm.
Such climatic variations give rise to different types of geomorphic processes and associ-
ated landforms. Episodic changes such as tropical cyclones or floods and droughts associ-
ated with ENSO are superimposed on the general pattern in certain parts of the tropics,
48 Tropical hydrology

periodically influencing the landforms in such regions from time to time. Even if we con-
sider only the climate, we should expect a good deal of variation in the tropical landscape.
The reality is even more complex, as the climatic variations operate in conjunction with
different types of tectonism, lithology and relief as introduced in Chapter 2. The tropics are
a mosaic of different landforms and a range of geomorphic processes, depending on the
regional climatic characteristics that operate on this wide range of landforms. The surface
of the humid tropics is protected by vegetation, but the vegetation cover varies, depending
on local climate and relief. In the next chapter we review tropical vegetation.

Questions

1. You are visiting an island with a high volcanic mountain in the tropical Pacific. What
would be the pattern of distribution of annual rainfall on this island?
2. Describe the nature of rainfall during the wet monsoon season. Does it rain uniformly
throughout the wet season?
3. You have two rivers flowing in opposite directions. In one case, moisture-laden winds
move upstream and in the other the winds travel downstream. If a large rainstorm is
embedded in the wind system, what would be its effect on these rivers?
4. Tropical cyclones have been described as geomorphic events in parts of the humid trop-
ics. Why? In which areas are tropical cyclones expected each summer?
5. Explain how climatic variations related to ENSO events influence geomorphic proc-
esses in the tropics. Can you find good evidence of such effects in literature or on the
web?
6. Describe a typical tropical rainstorm.
7. Figure 3.7c also shows runoff. Explain how this correlates with seasonal variations in
temperature and precipitation.
8. Construct a water budget (including surface runoff) for a station for which you have
access to the necessary data. See Thornthwaite and Mather (1957) and Dunne and
Leopold (1978) for the methodology required for computing water budgets.
9. Explain the quote from J. F. Griffiths at the beginning of the chapter.
4 Erosion and land cover in the tropics

We passed through extensive forests, along paths often up to our knees in mud, and were
much annoyed by the leeches for which this district is famous.
Alfred Russel Wallace

4.1╇ Erosion from tropical rainfall

Tropical rainfall can be strikingly erosive (Fig. 4.1) because it falls with high intensity and
often in large amounts. Erosion results from a combination of erosivity and erodibility.
Erosivity is the potential ability of a geomorphic agent to erode. Erosivity of a particular
rainfall event depends on the physical characteristics of the rain, such as intensity, duration,
amount and drop size. Erodibility is the measure of the level of vulnerability of the surface
material to erosion. It depends on the properties of the surface material, such as the texture
of the surface material, the slope of the land and the protection offered by the local vegeta-
tion cover. Intense rainfall, as from convectional thundershowers, erodes, especially when
falling on loosely structured soil.
The kinetic energy or momentum of a rainfall event determines its erosivity. The inten-
sity of rain is measured as the rate of rainfall, the amount of rain arriving over a fixed
period. It is generally measured in mm hr−1, but other time periods are also used, such
as millimetres per 30 minutes. Hudson (1981) mentioned that rainfall intensity seldom
exceeds 75 mm hr−1 in temperate countries, and that only in summer thunderstorms. In
comparison, intensities reaching double that amount are often experienced in many parts
of the tropics. Hudson recorded a maximum of 340 mm in an hour in Africa. Low-intensity
rain that lasts for long periods usually falls as small raindrops, whereas high-intensity rain
carries a range of sizes including big raindrops. The kinetic energy of the rain, on which
its erosivity depends, is determined by rainfall intensity and drop size. Hence, a tropical
thundershower on unprotected soil can be extremely erosive.
When raindrops with high kinetic energy impact the surface, they break it up and
splash particles in the air, a process known as splash erosion. This is the beginning
of erosion. Most of the splashed-up material comes to rest in the downslope direction.
With time and more rain, water starts to run downslope over the surface, removing the
detached particles. Runoff begins as a very thin sheet of water moving downhill, concen-
trated selectively in rills after travelling for some distance. More erosion and removal of
surface material happens progressively as rills get bigger and become gullies (Fig.€4.2).
Based on empirical evidence, Hudson (1981) identified the intensity of 25 mm hr−1 as
49
50 Erosion and land cover in the tropics

Fig. 4.1 Gully erosion at Baringo, Kenya. From Gupta and Asher, 1998. By permission of Wiley. Photograph: A. Gupta

a threshold that separates erosive and non-erosive rain, with an excellent correlation
between rain exceeding this threshold and soil loss. Erosion of the surface material there-
fore may depend on the proportion of the rain in a rainfall event that falls at intensities
greater than 25 mm hr−1 or a comparable value. Research in Nigeria indicated another
index, AIm, where A is the amount of rain and Im the maximum intensity over a 7.5 minute
period (Lal, 1976). A large volume of intense rainfall, as from convective thundershow-
ers, is highly erosive. Hudson (1981) noted that most of the annual erosion happens in
large storms with high-intensity rain such as convective thunderstorms, and not from
gentle low-intensity rain. He attributed this as a probable cause for intense soil erosion
in tropical and subtropical regions. Drops between 1.2 and 2.5 mm in size and with a
terminal velocity of 4.6–7.4 ms−1 have been measured in monsoon rainstorms in India
with hourly rainfall intensities of 8.5–55 mm (Pisahorty, 1991). Tropical rainfall often
falls with even higher intensity. Tropical rainfall thus has the potential to erode surfi-
cial material much more effectively than temperate precipitation, where gentle rain may
commonly fall over a long period.
The amount of annual rainfall and the extent of natural vegetation cover are directly
related. As a result, in spite of the prevalent intense tropical convective thunderstorms,
well-vegetated areas such as the tropical rain forest are usually lightly eroded. Areas most
vulnerable to erosion are those where intense rainfall falls on land with only a partial
coverage of vegetation. In a study that has become classic, Langbein and Schumm (1958)
plotted annual sediment yield against annual precipitation adjusted for temperature, i.e.
corrected for evapotranspiration. Their data were from small river basins in the United
States. The graph showed a single peak, the region with maximum erosion being under
short grass with about 300 mm of adjusted precipitation in temperate United States. This
peak in annual sediment yield was moved to about 450–500 mm by Dendy and Bolton
51 Erosion from tropical rainfall

Fig. 4.2 Rainfall giving rise to runoff and erosion on a slope

(1976), who analysed data from 800 drainage basins in the United States. The erosion peak
was still in grassland. In comparison, a protective cover of vegetation reduces sediment
yield in temperate continental areas with more rainfall. In drier areas, lack of running water
on the surface due to low rainfall limits erosion and sediment movement.
Wilson (1973) found double peaks in the global pattern, one for the semi-arid grassland
with about 380 mm of rain and the other for seasonal climates including the Mediterranean
and seasonal humid tropics. The data for the tropical wet–dry climate indicated sediment
yield peaking at a mean annual precipitation of 1780–1905 mm. With higher rainfall, the
natural ground surface is well protected by tropical rain forest and sediment yield drops
(Douglas, 1967). Fournier (1960) located the highest rate of erosion also in the seasonal
tropics and a double peak in the global pattern.
Sediment yield, however, does not depend solely on annual rainfall. It depends also on
a multiplicity of other factors that determine erodibility, such as ambient geology, local
slopes, properties of rainfall and vegetation cover. Natural vegetation has been destroyed
across large parts of the tropics, especially since the mid-twentieth century, for various
reasons:€ demand for tropical timber; expansion of agriculture; migrant settlements; and
urbanisation. Heavy and intense rain therefore falls, in many places, on bare ground or
a cultivated cover. Where the forest cover has been destroyed, the rate of sediment loss
increases tremendously as the land surface becomes exposed to high and intense rainfall.
For example, an annual sediment yield of 30 t km−2 from the Javanese rain forest rose to
1590 after logging (Anderson and Spencer, 1991). It is therefore necessary to examine both
the distribution of natural vegetation in the tropics and its destruction in order to under-
stand the prevalent pattern of erosion and sediment production.
52 Erosion and land cover in the tropics

Fig. 4.3 Map of tropical vegetation, pre-destruction stage

4.2╇ Distribution of natural vegetation in the tropics

Temperature and precipitation jointly determine the amount of moisture available to plants.
Rain forests occur in the tropics in areas with high rainfall and without a pronounced
dry season. Warm areas need higher precipitation to maintain a forest, as a large amount
of moisture is lost to the atmosphere by evaporation. The vegetation types in the tropics
vary on a gradient determined by moisture availability, seasonality of climate and eleva-
tion. With progressively less rainfall we find dry-season deciduous forest (also known as
tropical dry forest), grasslands and thorn scrubs, and deserts (Fig. 4.3). Altitudinal zoning
of vegetation occurs on high mountains, where with elevation and changing rainfall and
temperature, the rain forest of lowlands is altitudinally replaced by lower montane forest,
upper montane forest, subalpine forest and, on mountains that rise beyond 3500 m, a tree-
less community towards the top. On wet mountains such as in New Guinea the treeline may
rise to nearly 4000 m, provided the forest has not been disturbed. Ice covers the surface
on very high mountains such as the Andes and Himalaya and tops of high volcanic peaks
such as Mount Kenya or Kilimanjaro. The vegetation is denser on the wetter sides of the
mountains and the different altitudinal zones of forests and tree vegetation reach higher
elevations. Specialised forest types occur in specific environments. The major examples
are mangroves on tidal coasts, freshwater swamp forests on floodplains and deltas, heath
forests on relatively infertile lands, and peat swamp forests on deep ombrotrophic peat ris-
ing in a dome above the water table, as on the island of Borneo.

4.3╇ Tropical rain forests

Tropical rain forests occur in three large blocks within the tropics. The most extensive area
is found in South and Central America covering 4 million km2 in area. Most of this block
is located in the basins of the Amazon and Orinoco rivers. A much smaller and discon-
tinuous stretch lies on the Pacific side of the Andes in Ecuador and Colombia, extending
northwards through Central America to the southern border of Mexico. A third patch, now
almost entirely deforested, used to occur as a narrow strip on the Atlantic coast of Brazil.
53 Tropical rain forests

Fig. 4.4 Shorea curtisii, a tall tree of the tropical rain forest in Southeast Asia. Photograph: A. Gupa

The second largest area of the tropical rain forest is an Australasian block, extending dis-
continuously from the Queensland coast of Australia and Papua New Guinea and other
Pacific islands through the islands and peninsulas of Southeast Asia to the wet slopes of
the Western Ghats of south India overlooking the Arabian Sea. The smallest extent of the
rain forest is in Africa, centred on the basin of the Congo, with small outliers in East Africa.
Small patches are found on the wet east coast of Madagascar where rain forest used to be
continuous until recently. A strip detached from the Congo forest occurs along the coast in
West Africa.
The nature of the forest varies depending on elevation, rainfall, soil characteristics,
etc., but the well-developed variety found in the lowlands is mostly evergreen and high.
The general canopy is at 30–40 m above ground with scattered tall trees up to 50 m
high, known as the emergents, rising above this level. Smaller trees occur underneath
the canopy. The trees tend to have trunks with a wide girth and branching only near the
top (Fig. 4.4). Trees of the montane rain forest are smaller but still grow 15–33 m tall
on the lower slopes. Ground vegetation is usually sparse. A litter layer covers the for-
est floor. It varies in thickness from 1–2 cm of recently fallen litter to about 15 cm of
material on infertile soil (Kellman and Tackaberry, 1997). The most extensive tropical
forest type in the past was probably semi-evergreen rain forest with a largely deciduous
canopy and an evergreen understorey, which still occurs in Africa. With repeated burn-
ing, this type tends to change towards deciduous forest or savanna (R. Corlett, personal
communication).
54 Erosion and land cover in the tropics

Fig. 4.5 Downward passage of rainwater through a tropical rain forest

4.3.1╇ Passage of rainwater through tropical rain forest

The structure of the forest affects the falling of raindrops. Only a fraction impacts directly
on the ground. In a tropical rain forest, or indeed in any high forest with well-developed
canopy, rainwater, after reaching canopy level, moves as (Fig. 4.5)

• throughfall (raindrops pass through gaps in the canopy to reach the ground)
• drips from the canopy (rain intercepted at canopy level may drop later from leaves and
twigs)
• stemflow (water accumulated in leaves and twigs at the canopy level, slowly running
down straight, long and wide trunks to reach the ground).
Both throughfall and drips from canopy may cause splash erosion, throughfall because of
the high intensity of the tropical rain and canopy drips as the canopy in a rain forest is high
enough for the falling drops to acquire some velocity. In general, however, the potential for
erosion by rainfall in a tropical rain forest is low (Brandt, 1988). Vegetative litter on the
ground may also reduce rain splash.
After a period of intense rain, water on the ground may briefly flow on leaf-littered sur-
faces or flow through the subsurface via soil pores and pipes at shallow depth as described
by Sidle et al. (2000). A spell of intense convectional showers is often followed by water
emerging and cascading down from exposed ends of soil pipes midway up a vertical bank
55 Tropical deciduous forests, grasslands and deserts

or a road cut. A number of case studies have been carried out on rain forest slopes, for
example in eastern Puerto Rico, Southeast Asia and northern Australia. Bruijnzeel (1989)
generalised that 80–85 per cent of the rain reaching the ground surface arrives via through-
fall and 1–2 per cent as stemflow. The rest is retained by the canopy and evaporated back
to the atmosphere. Comparable figures came from a study in the Danum Valley of Sabah,
East Malaysia where the annual rainfall is around 3600 mm:
interception 17.4%
throughfall 80.7%
stemflow 1.9%.
According to this study, (Sinun et al., 1992), about 2.0–2.5 per cent of the rain on reaching
the ground travels as overland flow. The rest infiltrates and probably flows mostly through
soil pipes. Surface runoff and erosion therefore is limited in a tropical rain forest and wide-
spread overland flow is common only for short periods following heavy and intense rain.
The water, however, may react chemically with organisms and plant litter to turn acidic and
act on underlying soils and rocks as described in Chapter 5.

4.4╇ Tropical deciduous forests, grasslands and deserts

Trees of the rain forest shed their leaves individually and never in a synchronised fash-
ion species-wise, but the tropical deciduous forest displays a seasonal pattern, with trees
becoming bare in early summer. The leaves grow back with the arrival of the wet monsoon,
but the ground surface is prone to erosion in the first few weeks of the wet monsoon before
new leaves grow to provide some protection from the momentum of raindrops. Compared
to the rain forest, the canopy is less dense and at a lower height. Shrubby undergrowth
partially covers the ground surface and bamboos are common although not ubiquitous.
Erosion by sheetflow and gullies happens on the slopes under tropical deciduous forest.
This type of forest is found skirting the drier edges of the rain forest in East, Southeast and
South Asia, with the best example in South Asia. It occurs in patches in Africa, the Pacific
slopes of the Andes in tropical South America, Brazil, Central America, Mexico and the
Caribbean islands.
Tropical grasslands or savannas are plant communities dominated by grasses or grasslike
plants (graminoids), although not entirely without trees. Trees occur where the moisture
supply is better, e.g. in areas of locally high rainfall or along stream banks, a phenomenon
known as gallery forest (Fig. 4.6). The canopy is never closed. Savannas therefore run the
whole gamut from open forests with a graminoid understorey to grasslands without trees
(Fig. 4.7). Tropical grasslands and thorny scrubs cover enormous areas in tropical Africa.
Kellman and Tackaberry (1997) associated African savanna with subhumid areas receiving
between 500 and 1500 mm of annual rainfall as an intermediate region between the tropical
rain forest of Central Africa and the arid subtropical deserts. Savannas, however, also occur
where the soil is unsuitable for tree growth or in areas affected by anthropogenic alteration,
and thus are found outside the subhumid belt. For example, the savanna known as llanos
in South America is associated with 800–1500 mm of annual rainfall. These grasslands,
56 Erosion and land cover in the tropics

Fig. 4.6 Grassland with gallery forest, Masai Mara, Kenya. Photograph: A. Gupta

Fig. 4.7 Short grassland, Masai Mara, Kenya. Photograph: A. Gupta

however, have been reported to occur also in areas receiving much higher rainfall but with
a very long dry season. Tropical grasslands extend over huge areas south of the Amazon
Basin rain forest, and along the Andes foreland in the western Orinoco Basin. Scattered
areas of various types of grassland occur throughout tropical South America, Central
America and the Caribbean islands, not only where the annual rainfall is low (subhumid)
but if the geology or landform is also unsuitable for plant growth. Grasslands thus occur on
old crystalline rocks of the Brazilian Shield and Guinea Shield and the seasonally flooded
Pantanal swampland in southwestern Brazil. In Asia-Pacific, tropical grasslands are exten-
sive in northern Australia, in the interior of Papua New Guinea and the dry interior of
Myanmar and the Indian Peninsula.
57 Anthropogenic alteration of the tropical vegetation

Fig. 4.8 Vegetation clearance for oil palm plantation in eastern Sumatra, Southeast Asia. Note lines of young oil palms towards
the upper left corner, canals put in for drainage and burning of vegetation in the centre. IKONOS satellite image
© Center for Remote Imaging, Sensing and Processing, National University of Singapore (2006), reproduced with
permission. See also colour plate section

Tropical grasslands cover about 60 per cent of the uncultivated tropical land area
(Kellman and Tackaberry, 1997) and, together with thorny scrublands, indicate that a sea-
sonal climate with limited annual rainfall covers most of the tropics. The extensive occur-
rence of grasslands, thorny scrubs and subtropical deserts in the tropics (Fig. 4.3) is a
reminder that geomorphic processes vary across these latitudes, and a perhumid climate or
vegetation should not be expected everywhere.

4.5╇ Anthropogenic alteration of the tropical vegetation

The destruction of the natural vegetation started with the beginning of agriculture at the
beginning of the Holocene. Within a few thousand years, forests were being depleted not
only for agricultural expansion but also for construction material and urbanisation. Fertile
river valleys and coastal plains such as the Ganga Plains or volcanic slopes as in Java were
cleared thousands of years ago. From the mid-twentieth century the demands for timber
and for rubber and oil palm accelerated the destruction of the tropical rain forest, especially
in South and Southeast Asia (Fig. 4.8). Rain forests were also destroyed in large-scale
projects for resettling people from densely populated areas, as in Indonesia and Malaysia,
and by spreading of pasture in Central America. Tropical seasonal forests were depleted
58 Erosion and land cover in the tropics

for agricultural land and firewood. Details of deforestation have been reported by many
authors (Douglas et al., 1993; Gupta and Asher, 1998; Corlett, 2009). Vast areas of lowland
forest have disappeared since the 1970s in Malaysia and western Indonesia, in fact from
everywhere. Such widespread deforestation implies that:
1. the present forests are a fraction of their past extent
2. with the disappearance of the forest canopy and other natural protective devices, soil
erosion has increased and the present rates of erosion on slopes and deposition of sedi-
ment in rivers and coastal waters are not necessarily natural rates
3. a significant part of this destruction followed by changes in the physical environment
has happened very recently, in the last few decades, and change is continuing.
The anthropogenic component of these changes affecting the physical environment and
geomorphic processes in the tropics is examined in detail in Chapter 17. It is, however,
important to emphasise even at this stage that current landforms and operating processes
in the tropics are not entirely natural and include a strong anthropogenic component. This,
of course, is a global trend, but the tropical countries have recently experienced a particu-
larly rapid rate of population growth and developmental changes which are still continuing.
As a result, many tropical slopes are now exposed to wind and rain damage following the
removal of the protection offered by a vegetation cover. These slopes fail repeatedly with a
greater volume of sediment loading the streams and ultimately reaching the coastal waters.
The present landforms and geomorphic processes have been significantly altered from the
natural appearance and rates. Tropical geomorphology therefore requires an understand-
ing of both the distribution of natural vegetation that still survives and its progressive
anthropogenic alteration. This, given the nature of tropical rainfall, probably had a greater
impact in the humid and subhumid tropics than in other parts of the world. A natural ana-
logue of this anthropogenic destruction is the nature of post-hurricane (or typhoon) slopes
when the destruction of the vegetation in high winds is followed by torrential rain. Slopes
fail, rivers flood and coastal waters are polluted with sediment.

Questions

1. Why is tropical rainfall highly erosive?


2. What amount of annual rainfall has been associated with high sediment yields? Can you
explain such figures?
3. Why is erosion limited in an undisturbed tropical rain forest?
4. Is it correct to state that a seasonal climate, with limited annual rainfall and grassland
with scattered trees, best represents the tropics?
5. Distribution of the surviving natural vegetation and its progressive anthropogenic alter-
ation are both important factors in tropical geomorphology. Do you agree?
Part II

Process geomorphology in
the tropics
5 Weathering in the tropics

Hover through the fog and filthy air


William Shakespeare

5.1╇ Introduction

The process of weathering alters rocks at or near the surface of the Earth, transforming hard
rocks into soft material that can be removed easily (Fig. 5.1). Rocks, which are formed at
various depths inside the Earth’s crust under conditions of high pressure and temperature,
undergo a series of physical and chemical changes when they are exposed to the atmos-
phere at or near the surface of the Earth where much lower temperate and pressure prevail.
The term ‘weathering’ encompasses the full range of such changes. A formal definition can
be constructed as
Weathering is a process, at or near the interface between the crust and atmosphere, which
alters the physical and chemical nature of rocks in situ.

The process of weathering comprises a set of sub-processes that can be grouped broadly
into two classes:€ physical and chemical weathering. Physical weathering includes the
sub-processes that lead to the breaking up of rocks into smaller units without significant
chemical alterations. In chemical weathering, original minerals in the hard rock react with
ambient atmospheric conditions and with any water flowing across the rock surface to
change into new and softer minerals. Usually both types of weathering occur together,
although one may predominate. In the arid tropics, where atmospheric moisture is very
low, rocks tend to break up into smaller components without much chemical change. In the
humid tropics, chemical weathering is accelerated in the presence of high temperature and
moisture. The pattern and products of weathering are therefore different in various parts of
the tropics (Fig. 5.2). The nature of the final weathered product depends on

• mineral composition of the original rock


• prevailing climate
• ambient vegetation
• time.
A rock, when well weathered, becomes softer. It is the first step to erosion. The upper
part of the soft weathered rock is further transformed by soil-forming processes into soils.
The subsurface material therefore shows a vertical zonation:€ soils, weathered rock and
61
62 Weathering in the tropics

Fig. 5.1 Schematic diagram indicating the sequence of changes over time from hard rock to soft weathered rock and soil

(a)

(b)

Fig. 5.2 (a) Physical weathering in southern Israel:€note clasts of broken rock on the surface; (b) Chemical weathering in
Mauritius:€red fine-grained soft material developed over basalt; growth of sugarcane indicates a well-developed soil.
Photographs: A. Gupta

Fig. 5.3 The vertical zonation of soil, weathered rock and unweathered bedrock. Duricrusts (hard layers of precipitated
material) may occur in soil and weathered rock. Regolith describes the total material above unweathered rock.
Saprolite usually refers to weathered rock that still carries the rock structure inherited from the unweathered rock
63 Sub-processes of weathering:€a brief review

unweathered rock (Fig. 5.3). The structure of the original rock may still be recognised in
parts of the weathered rock layer below the soil. The total thickness of soil and weathered
rock increases with time and rate of weathering. This rate is extremely slow in the arid
tropics, giving rise to a thin layer of physically broken material on hard rock. In the humid
tropics, where accelerated chemical weathering changes minerals at a rapid pace, a thick
layer of soil and weathered material forms and is found over hard rocks. Chemical weath-
ering is therefore an important process in the humid tropics where slopes are underlain by
thick layers of soft erodible material, except where such material cannot accumulate as it is
quickly removed by erosion. Slopes affected by tectonics and storms therefore do not gen-
erally carry a thick weathered layer. The eroded material from the soft soil and weathered
rock ends up in rivers. Sediment carried by rivers is thus weathering-dependent.

5.2╇ Sub-processes of weathering:€a brief review

Weathering combines several sub-processes, which may act together. These, however, are
easier to comprehend when discussed individually. This section is a brief review of the
sub-processes.

5.2.1╇ Physical weathering

Physical weathering leads to mechanical disintegration of a rock into smaller particles;


the process therefore is also known as mechanical weathering. Apart from breaking down
a large piece of rock into small individual particles, each of which is easy to remove indi-
vidually, such disintegration also increases the specific surface of the rock. The specific
surface of a solid is defined as the sum of its surface areas. A cube with the volume of
1€cm3 has six sides, each with an area of 1 cm2. The specific surface of the cube therefore
is 6 cm2cm−3. On breaking this cube into 10 units of 0.1 cm3 each, the total specific surface
becomes 60 cm2cm−3. When broken to units of 0.001 cm−3, the size of a colloid, the total
specific surface becomes 60 000 cm2cm−3 (Ruhe, 1975). As chemical reactions take place
when water flows across the surface of minerals, an increase in surface area accelerates
chemical weathering. Furthermore, chemical bonds in the structure of minerals break down
when particles disintegrate to a very small size, leaving unbalanced charges at various
points on the surface of the mineral, which leads to chemical reactions. Physical weather-
ing thus accelerates chemical weathering.
Physical weathering happens when external stresses set up by ambient conditions are
stronger than the internal resistant forces of the rock, and the rock disintegrates mechan-
ically. These stresses are set up by pressure release, crystal growth, alternate wetting and
drying, and, in case of weak material, by biological agents.
The pressure on the surface of a rock such as granite is reduced when its overburden (the
material lying on top of the granite which was originally formed inside the Earth’s crust)
is removed by erosion. This may lead to a slight expansion of the rock and the develop-
ment of shear planes at right angles to the direction of stress release. The upper part of the
64 Weathering in the tropics

Fig. 5.4 Exfoliation in andesite, Lombok, Indonesia. Photograph: A. Gupta

rock therefore develops joints that are curved and run parallel to the surface of the rock.
Concentric slabs then come off the rock as various processes of weathering act on it. The
process of curved slab removal is called exfoliation (Fig. 5.4). Joints may also develop
in a rock as it cools from a hot magma or is subjected to tectonic forces. Such joints usu-
ally occur as parallel or right-angled sets. Where joints intersect, rocks may break up into
blocks as the first step to disintegration. Joints also allow water to enter a rock and promote
chemical action.
Growth of ice crystals from water that enters openings in rock and freezes also gen-
erates stress. The change of state from water to ice leads to a 9 per cent expansion in
volume and with ice filling the cracks the stress on the rock can be high, the maximum
recorded being over 2000 kgcm−2. This far exceeds the tensile strengths of rocks which
split up into small particles. Crystallisation of minerals such as halite, gypsum and cal-
cite from salt solutions filling openings in rock may lead to rock disintegration in a
similar fashion. Alternate wetting and drying leads to physical disintegration of soils and
the breakdown of certain rock minerals, such as phyllosilicates, that undergo volumetric
changes by absorption or loss of water. Direct heating of rock faces to high temperature,
as in bushfires, also leads to rock disintegration. Tree roots entering cracks in a rock
may expedite its disintegration, but possibly with simultaneous chemical weathering by
accompanying acid solutions as discussed below. Tree roots by themselves, however, are
only capable of disturbing relatively softer material such as soil or weathered rock. Ants
and termites also turn up huge volumes of soft unconsolidated material that lie over hard
rock (Fig. 5.5).
It follows that physical weathering in the tropics is significant in (a) relatively dry areas
where crystal growths and episodic bushfires occur and (b) in mountains and deserts where
the temperature may oscillate across the freezing point of water, falling below freezing at
night or in winter. But physical weathering does not happen in isolation even in deserts.
Weathering in the central Namib Desert is associated with salt weathering, and lichens and
65 Sub-processes of weathering:€a brief review

Fig. 5.5 Termite mounds, Kakadu, Northern Australia. Photographs: A. Gupta

Box 5.1 Phyllosilicates and clay minerals


Phyllosilicates are a group of minerals which are physically made of alternate layers of silicon tetrahedra
(SiO4) stacked vertically with either layers of minerals gibbsite (Al(OH)3) or brucite (Mg(OH)2). As a result,
these minerals tend to cleave easily in thin sheets. In some minerals, water can be stored in between the
sheets or stored water can be lost from there. As a result, these minerals can swell or shrink, a property
which makes them unstable and prone to disintegration. Minerals such as kaolinite, talc, montmorillonite,
smectite and various types of mica such as muscovite or biotite, fall in this group. A number of fine-grained
minerals that belong to this group are known as clay minerals. Kaolinite, for example, is a clay mineral. Clay
minerals are earthy masses, which are very soft and may hold some water. They are products of weathering
and are easily eroded.
The word 'clay' denotes grains smaller than 2 μm in diameter. Clay minerals are also fine-grained but they
have extra characteristics of being a group of hydrous silicates with a layered structure. Although a number of
clay minerals have the texture of clay, the two terms do not mean exactly the same. The clay fraction, being a
size term, may also include quartz, feldspar, oxides of iron and carbonates. In comparison, certain clay miner-
als may have a texture coarser than clay.
66 Weathering in the tropics

the moisture in fogs dissolving carbonates in the ground. Minor solution features on granite
and marble are quite common (Viles, 2005). Elsewhere in the tropics chemical weathering
dominates, although disintegration by physical weathering, if it happens, would be a sub-
process in support.

5.2.2╇ Chemical weathering with special reference to the humid tropics


The process and rate of chemical weathering depend on the combined effect of a number
of factors:
1. temperature
2. rainfall
3. vegetation
4. physical weathering
5. lithology
6. relief.
The first three factors are common in the humid tropics; others may be of local import-
ance. Chemical weathering is therefore extremely effective in the humid tropics. Chemical
weathering happens when a weak acidic solution passes across the rock face. This acidic
solution is derived in several ways. Rainwater may contain a limited quantity of carbonic
(H2CO3) and sulphuric (H2SO4) acids. Carbonic acid is also produced by microbes in the
soil by oxidising organic matter to CO2. Sulphuric acid is produced by bacteria from sul-
phide minerals, and a low concentration of this acid may also occur in rainwater. Water in
the subsurface (also called soil water) may carry organic acids (humic and fulvic acids) in
the tropics, released by fast partial decomposition of vegetative matter on the forest floor.
Schlesinger (1991) provided a list of mean residence time for litter in different kinds of
forest. The mean residence time for forest litter is 353 years in boreal forest, 4 years in
deciduous forest, 3.8 years in Mediterranean scrubland, but only 3 months in the tropical
rain forest. The prevailing high temperature in the tropics increases the decomposition rate
of soil organic matter (Berner and Berner, 1996).
The importance of biomass storage has also been demonstrated by Taylor and Velbel (1991),
working in the southern Appalachians, which are outside the tropics but their evidence can be
treated as an indicator. According to their calculations, stored biomass increases the weather-
ing rate of minerals by a factor of four. Biomass may be defined as the dead organic matter
resting on the surface of the ground and in the soil. The chemical constituents between the
rainwater and the soil water change because of the selective storing of elements (such as phos-
phorus and nitrogen) by vegetation. As only pure water is lost by transpiration, vegetation
increases the concentration of dissolved ions such as Cl− in soil water (Berner and Berner,
1996). Acids thus produced and carried in soil solution expedite rock weathering. The hydro-
gen ions in acids replace the cations in minerals as the acid solution flows across the rock face,
a process known as dissolution, which happens at a fast rate in the humid tropics. In addition,
organic acids colour soil solutions brown or red (a colour common in the tropics), disintegrate
minerals, and remove iron and aluminium by chelation (formation of a ring structure round the
67 Sub-processes of weathering:€a brief review

Table 5.1╇Common primary minerals and weathering


Mineral General composition Type of reaction

Olivine (Mg, Fe)2SiO4 Formation of iron oxides,


acid dissolution ©
Pyroxenes Ca(Mg, Fe)Si2O6/(Mg, Fe)SiO3 Formation of iron oxides,
acid dissolution ©
Amphiboles Ca2(Mg, Fe)5Si8O22(OH)2, Na and Formation of iron oxides,
Al may be present acid dissolution ©
Plagioclase feldspar Solid solution between albite Acid dissolution (i)
(NaAlSi3O8) and anorthite
(CaAl2Si2O8)
K-feldspar KAlSi3O8 Acid dissolution (i)
Biotite K(Mg,Fe)3(AlSi3O10)(OH)2 Formation of iron oxides,
acid dissolution (i)
Muscovite KAl2(AlSi3O10)(OH)2 Acid dissolution (i)
Volcanic glass Water and acid dissolution (i)
Quartz SiO2 Difficult to dissolve
Calcite CaCO3 Acid dissolution ©
Dolomite Ca,Mg(CO3)2 Acid dissolution ©
Gypsum CaSO4.2H2O Dissolved by water ©
Anhydrite CaSO4 Dissolved by water ©
Halite NaCl Dissolved by water ©
Pyrite FeS2 Formation of oxides of iron
and sulphur

Note:۩ congruent or simple dissolution; (i) incongruent or dissolution with reprecipitation


Source:€Berner and Berner, 1996

metallic ion). These acts are followed by a downward transfer to the subsurface, where iron
and aluminium are reprecipitated, usually at a depth of tens of centimetres.
Dissolution is one of a series of chemical reactions, probably the most important, that
together constitute chemical weathering. In dissolution, the primary mineral (Table€5.1)
is either dissolved completely by acidic solutions passing the rock face (congruent dis-
solution) or a fraction of the original mineral is reprecipitated to form one or more
secondary minerals (incongruent dissolution). Oxygen dissolved in water also reacts
with reduced iron and sulphur to form new minerals. Minerals such as halite, calcite
and gypsum easily dissolve in subsurface or rainwater, both seasonally abundant in the
humid tropics.
Acidic solutions pass downwards and laterally through the soil and already weath-
ered rock to reach primary minerals in unweathered material. Berner and Berner (1996)
described a selective beginning of dissolution on the face of the mineral crystals, where
solutions first attack rows of atoms slightly out of place in the crystal lattice. Reactions
with the flowing solution start in these locations, leading sequentially to selective etching,
formation of etch pits on the mineral surface, and coalescence of such pits, ending finally
in the disappearance of the original mineral.
68 Weathering in the tropics

Silicate minerals commonly dissolve incongruously. Congruent dissolution is limited to


only olivine, amphiboles and pyroxenes with low iron content, as the following chemical
reactions from Berner and Berner (1996) illustrate.
Mg2SiO4 (forsterite, an olivine with very little or no iron) + 4H2CO3 (carbonic acid) →
2Mg2+ + 4HCO– + H4SiO4 (dissolved silica)

CaMgSi2O6 (diopside, a pyroxene with very little or no iron) + 2H2O


+ 4H2CO3 → Ca2+ + Mg2+ + 4HCO– + 2H4SiO4
The iron-free olivine and pyroxene, therefore, can be completely dissolved and removed
in solution. This is a simple case, but most silicate minerals weather by incongruent dis-
solution, which is complex. The following equation summarises the decomposition of a
feldspar (Berner and Berner, 1996).
2NaAlSi3O8 (albite, a plagioclase feldspar) + 4H2C2O4 (oxalic acid) + 2O2
+ 7H2O → Al2Si2O5(OH)4 (kaolinite) + 2Na+ + 2HCO3– + 4 H4SiO4 + 6CO2
After incongruent dissolution of albite, kaolinite will be left at the place of weathering,
CO2 will dissipate as a gas and other products will be removed in solution. Feldspars occur
primarily in igneous rocks along with quartz (SiO2), and quartz remains in situ with kaolin-
ite from feldspar as an undissolved residue. Granite is commonly weathered in this fashion,
the hard rock being replaced by sand from quartz and clay from weathered feldspar and
other minor constituent minerals.
Loughnan (1969) showed the importance of acidity in water for dissolving minerals
and rock component. Silica is not dissolved unless the water is highly alkaline. Oxides of
iron or aluminium are dissolved only in highly acidic or alkaline water. Microorganisms,
so commonly present in the soils of the humid tropics, produce carbonic and sulphuric
acids that expedite dissolution. Oxalic acid is exuded by fungi in the forest litter and the
upper parts of the soil. As a result, iron and aluminium in the humid tropics are efficiently
removed from the minerals undergoing weathering and deposited elsewhere. In the forests
of the humid tropics, therefore, organic acids colour soil solutions (subsurface water circu-
lating through soil) red or brown, decompose minerals, and chelate and solubilise iron and
aluminium and transfer them to lower soil horizons.
Other chemical reactions may happen separately or simultaneously with dissol-
ution. Absorption of water without chemical change in a mineral is known as hydration.
Anhydrite (CaSO4) changes into gypsum (CaSO4.2H2O) by absorbing water and swelling
up. Hematite (Fe2O3) changes into limonite or goethite (HFeO2). Iron is transferred from
ferrous state to ferric by oxidation with the atmosphere. For example, an iron-bearing
mineral may oxidise to iron oxides. Such oxides may then change to goethite (FeO.OH).
On losing water, goethite changes back to hematite. When metallic cations on the surface
of a mineral are replaced by free hydrogen in solution, the process is called hydrolysis.
For example,
4KAlSi3O8 (potassium feldspar) + 4H+ + 2H2O → 4K+ + 2Al2Si2O5(OH)4 + 8SiO2
69 Products of weathering

Table 5.2╇Common secondary minerals formed by weathering


Mineral General composition

Hematite Fe2O3
Goethite FeO. OH
Gibbsite Al(OH)3
Kaolinite Al2Si2O5(OH)4
Smectite (Ca,Na)Al3MgSi8O20 (OH)4.nH2O
Vermiculite
Calcite CaCO3
Gypsum CaSO4.2H2O

Potassium feldspar thus changes into clay minerals and silica. Carbonate minerals wea-
ther by simple congruent dissolution, when attacked by carbonic acid.
CaCO3 (calcite) + H2CO3 → Ca2+ + 2HCO3−
CaMg(CO3)2 (dolomite) + 2H2CO3 → Ca2+ + Mg2+ + 4HCO3−
Limestones undergo selective dissolution along existing bedding planes, joints and other
openings. The openings are enlarged by removal of Ca2+ and HCO3− in solution and, given
enough time, caves are produced. The air in caves, especially those connected by an air
passage to the atmosphere, has a low partial pressure of CO2. When water flows into such
caves along subterraneous passages carrying high levels of CO2, the gas is lost to the cave
atmosphere. As a result the water becomes saturated with CaCO3, which is reprecipitated
as calcite in the caves. Even in dolomite caves, only calcite is deposited, as dolomite is
difficult to precipitate. Calcite is also deposited in soils from solutions due to similar degas-
sing. The more arid the climate, the shallower is the precipitated layer of calcite, an exten-
sive layer of which is known as caliche.

5.3╇ Products of weathering

When a primary mineral is attacked by a reaction agent such as an acidic solution, it under-
goes chemical changes. Such changes produce one or more secondary minerals (Table 5.2),
free cations and anions that travel with the solution, and remnants of the original mineral.
The final product is not always the same. It depends on the rate of flow and the pH-value
of the attacking solution.
In an earlier example, plagioclase feldspar changed into kaolinite, but two other alter-
natives are common:€smectite ((Ca,Na)Al3MgSi8O20(OH)4.nH2O) and gibbsite (Al(OH)3).
The formation of the clay mineral depends on the rate of flow of water. A slow passage
of water across the face of the plagioclase increases the time of contact of a unit vol-
ume of water with the mineral, and the solution becomes progressively more saturated
70 Weathering in the tropics

Fig. 5.6 Formation of different clay minerals depending on rainfall, following Sherman’s (1952) example from Hawaii.
Montmorillonite is a type of smectite

with the dissolved material. The proportion of both silica and cations builds up in the
solution and resultant chemical change forms smectite instead of kaolinite. In contrast,
when water travels fast, concentration remains low, and cations and silica are flushed out
of the system. This gives rise to gibbsite being formed. Kaolinite is formed when slow
flushing takes place but the water is not stagnant enough to allow a pronounced build-up
of cations and silica. Therefore gibbsite is found where high rainfall and good drainage
prevail and smectite where water stagnates, i.e. at a break in a slope or a relatively dry
area. An excellent and often used illustration comes from Sherman’s work in Hawaii
(Sherman, 1952), where the volcanic highlands create a wind and lee effect against the
moisture-bearing easterly trade winds and formation of smectite, kaolinite and bauxite
(gibbsite) from basalt is directly related to annual rainfall on slopes (Fig. 5.6). A second
example is the case of fast-flowing water down the steep sides of andesitic volcanoes in
Indonesia that slows down near the break in the slope at the base. Kaolinite is found on
the upper slopes and smectite in poorly drained or swampy depressions at the base (Mohr
and van Buren, 1954). In South India, weathering in humid conditions gives rise to the
formation of kaolinite and gibbsite, but where rainfall is low, smectite prevails (Bronger
and Bruhn, 1989). The presence of extra cations such as Na+ in the primary mineral also
favours the formation of smectite. Variable residence times of water in different layers
of soil may give rise to a vertical zonation of clay minerals with gibbsite towards the top
and gibbsite and kaolinite below. Entrapment of water in slightly weathered buried rock
produces smectite (Berner and Berner, 1996). The characteristic clay mineral over most
of the humid tropics, however, is kaolinite.
This discussion is based on plagioclase feldspar as the primary mineral undergoing
weathering. Other silicate minerals likewise weather to different clay minerals, the end
product depending on their chemical composition and rate of weathering. Generalisations
can be made regarding weathering of different classes of rock.
71 Products of weathering

Fig. 5.7 Weathering in volcaniclastic rock, Puerto Rico, producing both boulders and red clay. Photograph: A. Gupta

5.3.1╇ Weathering of common extrusive igneous rocks

Common extrusive igneous rocks are basalt and andesite. Basalt is fine-grained and con-
sists of olivine, pyroxene and calcium-rich plagioclase feldspar. Andesites are generally
made of plagioclase feldspar, amphibole and pyroxene. A small fraction of olivine may
be present. Both these rocks weather rapidly in the humid tropics, and smectite, kaolinite
or gibbsite is produced depending on the hydrologic conditions. Unless volcanic clasts or
quartz veins are present in the original rock, basalt and andesite would, over time, wea-
ther entirely to clay. If clasts or quartz veins are present, coarse gravel is found embedded
within the clay (Fig. 5.7). The soils on extrusive rocks are generally clay-rich, dark and
contain cations such as Ca, Mg or Fe. This pattern of weathering is seen in the humid
tropics in mid-ocean islands, over flood basalts on continents and in subduction-related
volcanic mountains. Stable slopes on extrusive rocks and volcanic mountains in the humid
tropics therefore tend to have a thick weathered layer near the surface (Fig. 5.2b). Rivers
draining such areas carry fine-grained material. If volcanic clasts or vein quartz is present,
boulders are found with the clay in stream channels. Over time the total thickness of soil
and weathered rock may extend to several tens of metres.

5.3.2╇ Weathering of common intrusive igneous rocks


Granite and granodiorite are common intrusive rocks. Granite consists of quartz, potassium
feldspar, plagioclase feldspar and small amounts of muscovite, biotite and amphibole. As
expected, amphibole, biotite and plagioclase feldspar weather early in a hot and humid
environment, followed by potassium feldspar and muscovite. All these minerals change
into clay minerals and quartz is left as individual grains of sand. Granite under advanced
72 Weathering in the tropics

Fig. 5.8 Schematic zones in weathered granite after Ruxton and Berry, 1957

chemical weathering therefore transforms to a mass of sand and clay, usually producing
more clay by volume than sand. Granodiorite with comparable mineralogical composition
changes to similar end products.
Like igneous extrusive rocks, granite weathers deep in a hot and humid climate. More
than 100 m has been recorded. Ruxton and Berry (1957) proposed a general vertical profile
(Fig. 5.8) for weathered granite based on their work in Hong Kong. The profile indicates
fresh bedrock at depths with joints and various openings in it and various states of weath-
ered rocks above it. Solutions penetrate along the openings in granite and weather the rock
both vertically and laterally, the depth of weathered rock increasing with time. For most
of the profile, isolated bits of granite (known as corestones) are separated from each other
by sandy clay. The final end product is found at the top of the profile, a mixture of clay
with sand. The profiles tend to develop better on gentle slopes and ridge tops and less well
on steeper slopes. Weathered profiles up to 85 m have been reported from Hong Kong,
although most profiles tend to be less than 20 m deep (Bennett, 1984).
Intensity of weathering increases with distance from the centre of a corestone. For
example, the centre of a corestone in zone 3 of the Ruxton and Berry model is likely to be
fresh, unweathered rock. Its margins, however, may be reddened indicating the beginning
of chemical weathering. Beyond the margins of this corestone, feldspars and biotite of the
original rock have undergone chemical changes leading to sand and clay minerals replacing
weathered rock away from the corestone. Further away clay will be present, reddened by
iron oxides. This is zone 2 with only a few corestones. With time the entire corestone will
disappear, as in zone 1, replaced by sand and clay minerals.
A granitic hill is commonly dome-shaped, reflecting its formation inside the crust where
the granitic magma commonly fills a large dome-shaped cavity and also forces a bulge to
form. Such hills are often seen standing in isolation in the middle of a plain which appears
73 Products of weathering

Fig. 5.9 Corestones, Southeast Asia

to have formed by erosion. Such hills are called inselbergs (island mountains) or bornhardts
(after an early German geologist). With time, the granite is weathered into the standard
bimodal type of granite regolith:€finer material and corestones. The fine material is com-
monly moved by slope processes and running water over time, leaving the slopes of very old
granitic hills covered by corestones (Fig. 5.9). Bare granitic domes are also seen protected
by a thick weathering layer on the surface (called a rind), cut with deep grooves.

5.3.3╇ Weathering of sedimentary rocks

Weathering of sedimentary rocks depends on their composition. Sandstone primarily con-


sists of sand-size grains, mostly quartz with some feldspar and fragments of earlier rocks,
held together either by cements such as calcite or silica, or by a clay matrix. Only feldspars
and rock fragments weather chemically and so does the cement. Weathering is usually
slow and limited, but over time a thin layer of loosened sand grains tends to overlie the
unweathered sandstone. The decomposition of the cement and the feldspar loosens the
quartz grains. Individual grains drop off, the rock becomes fragmented and sand grains are
removed over time. If water is present in the subsurface, then piping and further removal
of the rock mass occur. Water penetrates sandstones along joints and other openings, breaks
down the rock by dissolving its cement or loosening its matrix and then removes the indi-
vidual grains (Fig. 5.10).
74 Weathering in the tropics

Fig. 5.10 Subsurface piping in sandstone, Northern Territories, Australia

Fig. 5.11 Weathering in conglomerate, Kuta Tjuta, Australia

Water circulating inside sandstone along pipes is more effective than water running over
the sandstone surface (Young, et al., 2009). Water running on top may even protect the
surface by precipitating crusts of silica or iron oxides. This happens best in alternating wet
and dry conditions in a seasonal climate. Young et al. (2009) explain the decomposition of
the massive rock of Uluru in Australia by drawing attention to the presence of 80° dips in
75 Weathering and vertical zonation

rock which allows penetration of water throughout its mass. The surface of Uluru displays
a number of pipes emerging from inside the rock mass. The surface also shows a series of
accelerated spalling (breaking of the rock in thin layers starting from the surface) in arkosic
sandstone, which has been attributed to both tensional stresses set up during valley excava-
tions and hydration. The role of a protecting caprock is crucial, as the absence of it allows
water to penetrate the sandstone thereby weakening it (Young et al., 2009). In the case of
coarser conglomeratic sandstones, as the cement or matrix decomposes the rock breaks
down to pebbles, cobbles or boulders as in the case of the conglomeratic hills of Kota
Tujah in Australia (Fig. 5.11). Usually the weathering and erosion rates are slow in sand-
stone, but if the sandstone is part of a craton, the rock has time to gradually disintegrate
and be removed. The solubility of silica has been directly related to temperature, which
suggests that tropical streams may carry a relatively higher silica load (Meybeck, 1987).
This has not always been the case. Very low silica concentration, ≤ 1 mg/l, has been usually
recorded for waters draining quartz-rich areas in Venezuela (Chalcraft and Pye, 1984).
Weathering and subsequent erosion along joints may give rise to characteristic topography
in sandstone. These include cliffs with rounded tops, sandstone towers separated by joint-
controlled corridors, deep shafts, caves, and surface features ranging from small pits, hol-
lows, boxwork joint-driven patterns to larger hollows in rock called alveolis and tafonis.
Unlike sandstone, limestone undergoes rapid chemical change in the humid tropics. The
calcium carbonate in the rock goes into solution and then is transported out of the source
area. The calcium may be reprecipitated as the water circulates because of temperature and
pressure change in underground caves. The non-carbonates in the limestone are released,
frequently as precipitates, as the rock is dissolved. If iron is present, it colours the weath-
ered material and soil. Soils on limestone in the humid tropics are usually red. Even a small
amount of iron in the soil water will accomplish this. In the arid tropics where moisture
is limited, physical weathering tends to break up limestone first in blocks and then into
individual grains. The solubility of limestone and the resulting topography in the tropics is
discussed in Chapter 15.

5.3.4╇ Weathering of metamorphic rocks

Our knowledge about weathering of metamorphic rocks in the tropics is limited. Chemical
weathering of such rocks depends on mineralogical content and grain size.

5.4.╇ Weathering and vertical zonation

Weathering transforms a hard rock into a vertical sequence of layers. From the surface
downwards these are soil, weathered rock and unweathered rock (Fig. 5.3). These layers
differ from each other in texture, composition and erodibility (the capacity of a material
to be eroded). In general, weathered material erodes from the top. Other terms have been
used to describe the different components of the column shown in Figure 5.3. The term
saprolite is used, usually in North American literature, to describe a deep section of soft
weathered material consisting of clay, iron oxides and sand, some of which may still retain
76 Weathering in the tropics

Fig. 5.12 Diagrammatic sketch of soil horizons

the structure of the original rock, e.g. joint patterns. The term regolith is used to refer to
the part of the column consisting of soil and the weathered rock from which all marks of
rock structure have been obliterated. Regolith is sometime used loosely to indicate any soft
residual material.
Each of these layers can be subdivided. Soil is divided into horizons (Fig. 5.12).
Weathered rock can be subdivided into different layers, for example the generalisation of
Ruxton and Berry (1957) showing progressive penetration by weathering both downwards
and from the outside of a rock towards its centre. The cross-country vertical profile, how-
ever, differs in total thickness or thickness of individual layers. It is possible to trace the
complete profile of weathering from a hilltop to the neighbouring valley bottom but the
details of the profile, especially the thickness of different layers in the subsurface, would
vary even if the underlying lithology remains the same. This is primarily due to variations
in the flow rate of water through the subsurface material. Rock types, frequency of slope
failures, climate and vegetation are other determining factors. Given suitable conditions,
e.g. gentle slopes, igneous rocks, a location in the humid tropics and an absence of human
interference, weathered material may extend to 100 m or more in the subsurface. This is an
extreme figure, but depth and rates of weathering in the humid tropics are strikingly more
advanced than in cooler and drier areas.
Engineers denote the strength of the subsurface material and changes in permeability on
a scale of weathering divided into six grades (Table 5.3). It reflects the vertical sequence of
changes in the weathered material.
77 Pans and crusts

Table 5.3╇Mass weathered grades as used by engineers


Grade Class Summary description

VI Residual soil Soil with developed horizons


V Completely weathered Mostly soil; fragments of original rock or rock
structure may be present. Such fragments, if
present, are discoloured
IV Highly weathered Discoloured rock with openings. Not more
than 50 per cent of the rock mass is decom-
posed and can be extracted with a geo-
logical hammer. Scattered corestones may
be present
III Moderately weathered Mostly discoloured rock but less than half
decomposed; penetration of weathering
along openings, interlocked corestones
II Slightly weathered Discolouration along openings, rest of the rock
resemble fresh rock.
I Unweathered fresh rock Original rock, no weathering

Source:€Dearman, 1974, 1976

5.5╇ Pans and crusts

Soil and soft weathered rock are eroded more easily than the hard fresh rock underneath.
This is the general rule, but hard, usually thin, layers of reprecipitated material present
within the soil or weathered rock may complicate the pattern. Such layers are known as pans
or crusts and they may occur at any depth between the surface and fresh rock. The thickness
of the hard layer varies but a measure of 10 m and more has been reported from various
locations (Goudie, 1973).
Dissolved material such as calcium carbonate or silica is carried by solution to the sub-
surface and, under suitable conditions, reprecipitated as hard layers called crusts. These
crusts have three important geomorphological functions. First, being difficult to erode they
protect the underlying soft soil or weathered rock. Second, they hinder soil formation,
which requires vertical movement of material. Third, such crusts prevent water from pene-
trating to depths below. If the crust is near the surface and the rainfall is high, water tends
to saturate the material on top of the crust making the upper part of the subsurface unstable.
The effects are discussed in detail in Chapter 6.
The general term for such pans and crusts is duricrust, but self-explanatory and specific
terms are also used:€silcrete, calcrete, ferricrete, etc. As encrustment requires both dissol-
ution and reprecipitation, seasonality in climate helps, with crusts being very common in
subhumid and semi-arid tropics. Calcrete shows this well. For example, isolated nodules
of calcium carbonate (locally known as kankar) are common in India in a strong seasonal
climate with about 1000 mm annual rainfall. In more arid parts of India, such as Rajasthan,
precipitation of calcium carbonate becomes extensive, forming bigger nodules and ultim-
ately a continuous subsurface layer close to the surface in the driest areas. Such layers are
78 Weathering in the tropics

Fig. 5.13 A stone line in deeply weathered igneous material, Kenya

called caliche. In very dry tropics, caliche may occur on the surface itself. Ferricrete is
common in more humid areas. A particular type of ferricrete is laterite. Laterite is an old
term (Buchanan, 1807), and associated with a characteristic profile which is not shared with
all kinds of ferricrete. Laterite has a hard surface layer of iron oxide that could be nodular,
pisolitic or spongy. A mottled zone lies below this surface layer, a pallid zone underneath
the mottled zone and a ferrigenous bedrock underneath. Unlike common ferricrete, laterite
is soft unless exposed to the atmosphere, hence the hard layer at the surface.
In places, a thin layer of single or multiple clasts have been noticed within a deep weath-
ered layer. Such layers are known as stone lines (Fig. 5.13). Stone lines have been asso-
ciated with various origins. A stone line has been interpreted as lag gravel, remnant of
the pavement of an old surface, a climate change indicator and the result of bioturbation.
Whatever their origin, stone lines are not ubiquitous.

5.6╇ Effects of weathering

There are two major results of weathering. First, the land surface is mantled with soft
material that can be easily eroded and removed. Second, the soft weathered material can
be transformed into soils, the base for vegetation. To a large extent, the vertical zonation
of soil, weathered material and the underlying hard rock determine the nature of hillslopes
and the type of downhill processes that operate on it. The weathered material is eroded
from the slopes and ends up in streams as dissolved or solid load. Weathering is the first
step towards eroding a land surface.
Chemical weathering slows down when fresh rock is not exposed either to the atmos-
phere or to the circulating subsurface water. A thick cover of weathered material and soil
prevents the land surface from being eroded and the hard rock underneath remains in a
79 Effects of weathering

nearly unchanged state. If the cover is thin, then the processes of erosion directly impinge
on the hard rock, slowing down denudation. In between lies the optimal situation where
chemical weathering is active at the hard rock surface creating new weathered material,
but the regolith does not thicken significantly, as its upper part is regularly or episodically
removed by erosion.
Carson and Kirkby (1972) identified slopes with a thick cover of regolith as trans-
port-limited slopes. On these slopes, weathering produces material much faster than it is
removed. Transport-limited slopes are expected to be covered with a thick regolith that
buries underlying surface irregularities in the hard rock, promotes vegetative growth and
produces a gentle slope on the surface. On the other hand, material is transported out fast
from weathering-limited slopes, where the rate of erosion is higher than the rate of weath-
ering. Such slopes have a thin regolith, even fresh rock surfaces may be exposed, the rock
exposures tend to steepen the slopes, and vegetation is usually sparse due to thin soil and
active erosion. Streams draining transport-limited slopes often correlate with a high dis-
solved load in the local rivers whereas weathering-limited slopes tend to produce clastic
material as suspended and bed load in the streams. Advanced chemical weathering should
contribute solution loads to the streams of the humid tropics, but as the tropical slopes are
often unstable and prone to failures (as discussed in Chapter 6), tropical rivers carry con-
siderable amounts of solid load as well. As most of the sediment load in a river is frequently
contributed by the headwaters in tectonic mountains of variable lithology, the relationship
between weathering and river sediment correlates well for individual rivers but an over-
arching rule is difficult to construct.
Gibbs (1967, 1972) studied the geochemistry of the Amazon River system in the mid-
1960s. Both the Amazon and the Orinoco system were studied in detail by Stallard (1985)
and Stallard and Edmond (1983, 1987). Even for the 6300 km long Amazon, the sediment
load is derived primarily from the headwaters in the Andes. Gibbs (1967) indicated that,
in the Andes, the continuous availability of unweathered rock following removal of the
weathered material by active erosion leads to progressive chemical weathering. In con-
trast, the very thick cover of secondary minerals on fresh rock isolates the rock surface
from incoming rainwater and not much material gets to the river as suspended or bed load.
Stallard (1985) and Stallard and Edmond (1983, 1987) related the chemistry of the water
of the Amazon and Orinoco to lithology and relief. The amount and nature of the sediment
load in a river therefore depend significantly on weathering of the bedrock of the river
basin and, as most of the load comes from the headwaters, ultimately on mountain lith-
ology and relief.
High weathering rates and transport-limited slopes are often correlated with high dis-
solved loads of regional rivers. In contrast, rivers draining weathering-limited areas
usually carry suspended load and bed load. Stallard (1985) showed this for the basins
of the Amazon and the Orinoco. He divided the Amazon Basin into five major erosional
regions:€the Andes, the foothills, the elevated shields, the lowlands and the lowland shields.
The highest erosion rates and the biggest sediment concentration in river waters came from
the tectonically active areas, not the geologically ancient low-relief areas or the lowland
built by deposition. This pattern holds for a number of large river basins. Very little dis-
solved or solid load is found in the rivers flowing across only the lowland and the shields,
80 Weathering in the tropics

lithology permitting. Most solids are in the resistant mineral phase. Weathering reflects
both bedrock and erosional regime.
The nature of the sediment, however, changes along a long river due to episodic transfer
and long-term storage on floodplains (Chapters 7–9). Even the feldspar-rich sediment that
arrives in the Orinoco from the granites of the Guinea Shield is finally transformed into
the quartz sand of the Orinoco, given the long-distance transportation and long-term stor-
age (Johnsson et al., 1988). Weathering continues on detritus grains as they move from the
headwaters to the sea. The transformation of the weathered product to the river load in the
tropics is examined in detail in Chapters 7–9.
Rapid rates of erosion are indicated by the presence of chemically unstable and cation-
rich minerals in the suspended load and bed load. Geomorphologists now have techniques
based on isotopes to identify the main source areas of sediment even for very large rivers.
For example, the isotopic composition of the sediment of the Brahmaputra River in Assam,
India has revealed that 45 per cent of the river’s sediment comes from 4 per cent of the drain-
age area, the Eastern Syntaxis, where the Himalaya mountain bends and the Brahmaputra
flows through a 5 km deep gorge past the Namche Barwa Peak (Singh, 2007).
We need more case studies relating weathering to stream load, but a general picture has
started to emerge.

5.7╇ Tropics and weathering

A number of generalisations can be made regarding weathering in the tropics.


1. Given the ambient conditions of high temperature and humidity, weathering, especially
chemical weathering, is accelerated in the humid tropics, unless the underlying rock is
chemically inert.
2. The slopes of the humid tropics are usually underlain by regolith, consisting of soil and
weathered rock, which is several metres thick.
3. The weathered rock is normally clay-rich and kaolinite is the common clay mineral.
4. The weathered rock is not necessarily homogeneous and duricrust layers are common.
5. The impediment caused by such crusts to the downward flow of water leads to slope
failures, as discussed in Chapter 6.
6. Tropical streams may carry a significant amount of dissolved load, but these streams
mainly carry clasts of varying size as suspended and bed load. Most of the suspended
and bed load is derived from the mountains at the headwaters of the basin.
7. The river sediment tends to be episodically transferred with periodic deposition on the
floodplains, where the grains are weathered further.
8. The accelerated chemical weathering of the humid tropics influences subsequent slope
and channel processes.
9. Weathering in the arid tropics has a different pattern, physical disintegration being the
principal weathering process.
81 Questions

Questions

1. Name a mineral that exists both as primary and secondary types. How is it formed?
2. State any reasonable annual rainfall figures associated with the weathering of a plagio-
clase feldspar (albite) to
a) gibbsite:€mm
b) smectite mm
c) kaolinite mm.
Now name three locations in the tropics to go with your rainfall figures. Be specific
with your choice of location. Broad answers such as 'Africa' are unacceptable.
3. Imagine you are in Java (Indonesia) walking down a steep volcanic slope. What kind
of volcanic rock would you expect there? The area gets about 1800 mm of rain a year.
Do you expect to walk on the same clay mineral all the way? Justify your answer.
4. On which side of the Hawaiian Islands should you expect bauxite? Why?
5. Why should an increase in the specific surface of a rock fragment raise its potential for
chemical weathering?
6. Suppose you have hillslopes in the humid tropics underlain by the following rock
types. What size material should you expect on the slopes after advanced weathering?
Why? What type of sediment load should you expect in rivers draining such areas?
a) basalt
b) granite
c) sandstone
d) limestone.
7. You are in eastern Puerto Rico. The slopes are in volcanic rocks weathered to boulders
set in red clay. What would be the sediment load of a local stream if
a) a landslide occurs?
b) a large flood comes down the river?
c) headwaters of the river receive augmented drainage? As a result, the average dis-
charge increases significantly but not flood peaks.
8. Should there be any difference in appearance between a weathering-limited slope and
a transport-limited slope? If so, how would they differ? Provide brief descriptions and
two diagrammatic sketches.
9. How do pans and crusts form in the subsurface? Do you anticipate any significant role
played by these features on erosion of slopes?
10.╇ What happens after weathering?
6 Slopes:€forms and processes

I have climbed and ruminated upon too many great bornhardts, in company with the
leopard and the baboon, to believe that these most powerful of landforms, glorious in
the sun and rain alike, ever originated foetally within the dark body of the earth. The
leopard and the baboon don’t believe so either.
L. C. King

6.1╇ Properties of a slope

The word slope indicates both (1) an inclined unit on the surface of the Earth and (2) the
measured level of such inclination. The surface of the Earth is a combination of slopes of
various kinds. This assemblage of slopes mostly consists of hillcrests, valleyside slopes
and slopes along streams orthogonal to the termination of valleyside slopes. As Dunne and
Leopold (1978) have described it, hillslopes cover virtually the entire landscape. A slope
is probably visualised best as a profile drawn from the hillcrest to the valley bottom. Both
water and surficial sediment are transported under gravity downslope. Sediment commonly
takes several moves to traverse the entire length of a hillslope, and between such moves
it rests at progressively downslope locations. The surface of a slope therefore reflects this
pattern of storage and transfer of material.
The geometry of a slope is expressed by three measures:€gradient, length and width. The
gradient of a slope is measured in degrees, ratios or percentages, and its length and width
in linear measures such as metres. A complete description of the slope requires details of
its geometry and a hillcrest to valley bottom cross-section showing a vertical sequence of
soils, weathered rock and bedrock along the profile. Accumulation of soft regolith towards
the top of the slope and active erosion by water towards its base tend to give a slope pro-
file a convexo-concave appearance, but not always. For example, a weathering-limited
slope, as described in Chapter 5, tends to have a steep, almost vertical segment below the
crest of the slope. Material falling off this steep segment comes to rest below it, giving the
next downslope section a straight profile. In 1942, Wood proposed a four-fold division of
a slope profile. In the downward direction these are waxing slope (convex crest), free face
(vertical), constant slope (straight) and waning slope (concave). King proposed a similar
four-fold division in 1962 but with different terms:€crest, scarp, debris slope and pediment.
These units differ not only in form (appearance) but also in subsurface material and geo-
morphic processes that operate on the surface (Fig. 6.1). In a transport-limited slope, accu-
mulation of weathered material covers up the free face (scarp) and the profile of the slope
is modified to a convexo-concave appearance.
82
83 Properties of a slope

Fig. 6.1 The sub-units of Wood and King. Note that with time the straight debris slope (also called a talus slope or constant
slope) increases in area due to the accumulation of material arriving from the top. It then proceeds to bury the scarp
face and change the original appearance of the slope. The size of the talus or debris accumulation is reduced when
its constituent material is removed episodically by a slide or flow. Its appearance depends not only on the nature of
its constituent material, but also on the time elapsed since the last removal or significant addition of material from
the top

Any given point on a slope displays a balance between two forces over an interval of
time:€(1) the erosive force that removes material from the surface and moves it downslope,
and (2) the resisting force that opposes its removal. The erosive force usually depends
on two gravity-driven processes:€movement of material en masse and running water. The
first of these is collectively termed mass movement or mass wasting. Both erosive forces
increase if the slope is high and water is present. The upper parts of high mountains, such
as the Andes or the Himalaya, are also eroded by ice, a third operating process on selective
slopes. The high mountains are discussed later as a special case in Chapter 13. The resisting
84 Slopes:€forms and processes

force essentially depends on the slope material. Hard rock with little structural weakness
close to the surface is difficult to erode. Erosive processes in general are also limited where
an abundance of vegetation binds the regolith together. In contrast, loose regolith, tecton-
ically shattered rocks and bare slopes tend to accelerate slope erosion. In sum, slopes are
failure-prone when the following properties are high:

• gradient
• weathered or soft rock in the subsurface
• geological structures susceptible to erosion
• presence of water on the surface or in the subsurface
• absence of vegetation
• tectonic disturbance
• undercutting of the base of the slope by a river or the sea.
For example, tropical cyclones passing over hilly areas will lead to multiple slope failures.
The high winds will destroy the vegetative cover and then intense high rainfall will lead to
slope failures. Similarly, steep slopes in an area of weak rocks and tectonic disturbances,
e.g. in the vicinity of an active plate boundary, are hazardous. Anthropogenic activities may
also increase the potential for slope failures. Destruction of vegetation, local steepening of
slopes or construction of houses on the mid-slope increasing the weight on the slope may
increase slope failures.

6.2╇ Mass movement on hillslopes

Mass movement is the general term for movement of material downslope under gravity,
usually in a relatively dry state. Three variables characterise such movements. First, the
nature of movement is an important characteristic. The material may fall down a cliff, slide
down a surface, move as a hyperconcentrated flow when lubricated by water, overcome
physical barriers by riding a cushion of air mixed with fine dust particles, etc. Second, the
nature of material that moves is a crucial property. The moving mass may range in texture
from boulders to clay and it usually occurs in an unsorted mixed state. The third variable
is the speed of movement. The moving material may creep downslope very slowly. In con-
trast, certain types of mass movement are near-instantaneous. Various classifications have
been proposed to organise mass movements into classes based on these three variables.
Of these classifications, the better-known ones have been proposed by Sharpe (1938),
Varnes (1958; 1978), Hutchinson (1968), Carson and Kirkby (1972) and Nemcok et al.
(1972). These attempts vary in the criteria used for classification and in their complex-
ity. Some also require more careful field observations than others. We will use a simple
classification system, adequate for our purpose. This classification (Table 6.1) is based on
Dunne and Leopold (1978) and has been used before in Kale and Gupta (2001). It is based
primarily on the nature of movement and each type is further subdivided according to (1)
the property of the moving material, (2) the presence or absence of moisture in the material
and, in certain cases, (3) the speed of movement.
85 Mass movement on hillslopes

Table 6.1╇ Types of mass movement


Types Subtypes

Fall Rock fall


Debris fall
Topples
Slide Planar slide (often further categorised as rock slide, debris
slide, mudslide, etc., depending on the material)
Rotational slide, also known as slump (always in fine material)
Flow Debris avalanche
Debris flow
Earth flow
Mudflow
Lahar
Solifluction
Creep Soil creep

Modified from Dunne and Leopold, 1978

6.2.1╇ Falls

This term is used when a mass of material breaks off from a near-vertical or vertical slope
and falls through the air to the lower slopes (Fig. 6.2). If this falling material consists of
a single chunk or several pieces of rock, it is a rock fall. The falling mass may also be a
mixture that includes material from boulders to fine particles, in which case it is known
as a debris fall (Table 6.1). The movement is very fast, almost instantaneous. The coarser
components of a fallen mass (particularly the broken bits of rock) may roll or bound fur-
ther downslope after landing. Rock falls from a vertical cliff of hard rocks could be large
in size or may fall from a considerable height. Soft rocks or unindurated material can only
build a steep slope of limited height and therefore falls in such material are smaller in both
size and height.
Potential conditions for rock falls exist when the rock is well jointed and weakened over
time by several factors such as alternate freezing and thawing of the slope material, intru-
sion of tree roots or the development of stress-release joints as explained below. A trigger
action for starting a fall usually involves a well-jointed rock cliff being vigorously shaken
by earthquakes or undercut by a river or the sea. The undercutting of the lower part of
the slope makes it unstable. Falls commonly occur in seismic high mountains such as the
Himalaya or Andes. The potential for falls in these high mountains is also higher because
the present glaciers used to extend further down the valleys during ice ages. The retreat
of glaciers up-valley in the warmer Holocene removed ice masses from the lower valleys,
removing the lateral and downward pressure of ice against the rock of valleysides and
valley bottoms. This release of pressure fractured sidewalls and uparched valley floors,
and led to the development of new joints, splitting rocks into slabs and raising the poten-
tial of rock falls. Walls of mountain valleys therefore also fail as a delayed effect of past
glaciation.
86 Slopes:€forms and processes

Fig. 6.2 Different types of slope failure. After Dunne and Leopold, 1978

Topples is a special case. If joints are near-parallel to the rock face or bedding planes
are in a similar position due to the folding of the strata, rock slabs often fail along the
cross-joints and topple forward. Igneous rocks such as basalts or dolerites with columnar
joints and folded bedded sedimentary rocks or metamorphic slates or schists tend to topple
(Selby, 1993).
A mountain slope affected by rock or debris fall therefore would have a near-vertical
section with screes of fallen material below. This is the free face or scarp and constant or
debris slope mentioned earlier. The material on the scree slope (also known as talus) tends
87 Mass movement on hillslopes

to be unstable and fails again, to provide streams at the bottom of the valley with coarse bed
load. At times, boulders released by rock falls may also bound downslope to reach valley
bottom streams.

6.2.2╇ Slides

Slides are failures along a surface of separation, over which the failed mass slides downhill.
Translational (planar) slides indicate movement en masse over a surface separating two dif-
ferent types of material such as regolith and unweathered rock or shallow surficial material
and a subsurface encrustation (Fig. 6.2). Translational slides occur in a range of material
and hence it could be qualified by the property of the slid material. For example, a planar
slide can be identified as a rockslide (contains broken bits of rock), debris slide (a mixture
of rock and regolith) or mudslide (clay-size material, usually water is present). In contrast
to translational slides, rotational slides occur in deeply weathered, near-homogeneous fine
material (Fig. 6.2). The type and level of weathering of the slope material therefore would
be a determining factor, and the slid material, when it comes to rest, could be at any state
between nearly undeformed and well deformed.
Slides vary widely in size, distance travelled and rate of movement. Dunne and Leopold
(1978) mentioned that the slid material has been reported from cubic metres to cubic kilo-
metres in volume and the distance travelled between metres and kilometres. The rate of
movement has been found to range from several centimetres a year (very slow) to several
centimetres per second (fast). The slower rockslides are also called rock glides. The rate of
movement depends on three factors:

• gradient
• amount of moisture in the slid material
• presence of lubrication.
Antecedent rainfall and trigger rainfall (Box 6.1) in combination determine the last two
factors and lead to the crossing of the erosional threshold. An erosional threshold is the
condition, level or stage at which a striking change is produced, such as a slide, which
significantly modifies the landscape (Schumm, 1977). There are three types of erosional
threshold:€extrinsic, intrinsic and geomorphic. An extrinsic threshold is crossed when, for
example, an increase in velocity and depth of a river allows it to pick up sediment previ-
ously resting on the channel bed. An intrinsic threshold is crossed with ongoing changes in
the property of material. For example, progressive weathering of slope materials may even-
tually cause the slope to fail. Geomorphic thresholds are crossed when a landform changes
over time to a state of incipient instability. For example, undercutting of a rock cliff may
result in the collapse of part of the cliff face (Schumm, 1977). A combination of antecedent
and trigger rainfalls results in crossing of both intrinsic and extrinsic thresholds.
In translational slides the failed mass moves along a subsurface plane of failure that tends
to coincide with (1) a structural surface such as a bedding or foliation plane, (2) the contact
between regolith and unweathered rock with little cohesion between the two, or (3) a sub-
surface pan such as an iron or silica encrustment layer. Cracks and openings in the material
above the plane of failure allow rapid penetration of water which is interrupted or slowed
88 Slopes:€forms and processes

Box 6.1 Antecedent and trigger rainfall


Slope material fails when it is saturated with water and the increased weight due to the stored moisture is
more than the resisting force at the plane of failure, a weak zone to begin with. Moreover, the presence of
moisture lubricates the plane of failure. The moisture accumulates over time.
Antecedent rainfall refers to the total amount of rain that has fallen over a selected number of days prior
to the failure. The number of days selected depends primarily on location, and usually is between 5 and 30.
Moisture from antecedent rainfall saturates the soil to some extent, following which water from a large rain-
fall event (the trigger rainfall), supersaturates both soil and weathered rock, and a failure occurs. The trigger
rainfall has to be large if the antecedent water in the soil is low, but if the soil is near saturation, a small
amount of rain may act as the trigger. This explains why slopes fail in intense rain from a tropical cyclone
(measured as 102 mm in a day) without much antecedent moisture. On the other hand, slopes often tend to
fail late in the rainy season when the antecedent moisture is high and only a small amount of trigger rainfall is
required. Numerous landslides occur when both are high. Pitts (1992) has shown that slopes fail in a number
of places in western Singapore when rainfall thresholds are exceeded, but not earlier.
Gabet et al. (2004) studied the Himalaya Mountains in Nepal, which are prone to slope failures because
of steep slopes, weathered bedrock and intense monsoonal rainfall occurring in combination. Based on three
years of fieldwork in a small drainage basin, they have shown that a number of small landslides start on these
slopes from the beginning of the wet monsoon season and larger ones usually after the monsoon has been in
place for some time. They observed slope failures in the basin of the 136 km2 drainage basin of the Khudi Khola
on the southern flank of the Annapurna Himalaya in Nepal. It is a rugged tectonic catchment with a mean
elevation of 2565 m. Heavy seasonal rainfall (3000–5000 mm yr−1) occurs over the basin due to monsoon-
driven moisture impinging on the mountains. The average soil depth is only 50 cm but the bedrock is deeply
weathered and permeable. Thus the regolith can store water up to a certain amount (Fig. 6.3).
Gabet et al. (2004) determined that landslides are triggered after about 860 mm of rain has fallen from the
beginning of the monsoon to fill the regolith. This is the antecedent rainfall. As the regolith is gradually filled,

Fig. 6.3â•… Role of antecedent and trigger rainfall in landslides in the Nepal Himalayas. From Gabet et al., 2004.
With permission from Elsevier
89 Mass movement on hillslopes

Box 6.1 Continued


the slope begins to fail with trigger rainfalls, the threshold amount of which drops with time until the required
amount stabilises at 11 mm per day (Fig. 6.3). Thus two distinct rainfall thresholds, the seasonal accumula-
tion (antecedent rain) and a daily total (trigger rain), are required to initiate slope failures. Larsen and Simon
(1993) reached a similar conclusion regarding slope failures in the hills of Puerto Rico. There, landslides tended
to cluster towards the end of the hurricane season when intense rain falls on saturated ground. Gabet et al.
(2004) mentioned suspended sediment waves travelling at the rate of 2–3 ms−1 in a local river. In this kind of
hilly or mountainous environment with large storm rainfalls, the material coming off slopes is simultaneously
transferred out of the small drainage basins to major rivers or, as in the case of islands like Puerto Rico, to the
sea. Only the very coarse material may remain in the valley (Chapter 7).

down at the structural surface, contact surface or encrustment layer, leading to accumulation
of water in the mass overlying the potential plane of failure. The slope then fails due to both
the extra weight of water and lubrication. As gullies advance, slides in alluvial or colluvial
material may occur due to pressure release along the channel walls (Fig. 6.4)
Rotational slides, also known as slumps, tend to occur in deep homogeneous material
such as a thick layer of clayey silt formed by advanced chemical weathering. Slumps, there-
fore, have been associated with the humid tropics, but the weathered layer in the humid
tropics is not always homogeneous as it frequently includes impermeable layers such as
iron pans or silcrete. Shallow planar slides are therefore also common in the humid tropics.
In a slump, the failed material is deformed as the mass above the ruptured surface is rotated
backwards giving rise to tilting, formation of small faulted blocks at the back of the slide,
a hummocky surface, ponded drainage and a curving rampart at the toe (front end) of the
slide (Fig. 6.2). If the water ponded between the blocks at the back of the slide seeps through
the weakened failed material, subsequent failures may occur. Slumps tend to happen when
water from heavy rainfall enters the subsurface, usually through an opening into a thick, fine
homogeneous material, or when a river undercuts its bank in thick beds of fine material.
In steep mountains, very large slides may come down a mountain slope, cross the valley
bottom and ride up the opposite hillslope for a limited distance. The ground between the
source and the final location of the slid material surprisingly does not always show much
disruption along the track of the slide. Even nearly undisturbed snow, fallen prior to the
failure, has been recorded. Shreve (1965) studied the huge Blackhawk landslide in southern
California, and suggested that as the huge slide came downhill, it began to ride a cushion of
compressed air in the manner of a hovercraft, which allowed the material to travel a large dis-
tance. The cushioning effect has also been explained by the production of steam from water
vaporised by the heat of friction at the base of the sliding material (Habib, 1975). The concept
of sturtzstrom was proposed by Hsü (1975). According to this concept, the innumerable solid
units inside the moving mass of a cloud of dust exchange kinetic energy by collision. This
energy keeps the particles separate and thus the mass moves like a fluid (fluidisation) with
very little internal friction. An earlier version was provided by Heim in 1882 after studying a
large failure at Elm, Switzerland. There the failed material travelled for about 3 km at a vel-
ocity exceeding 160 km per hour. For this to occur, the flow needs to be cushioned.
90 Slopes:€forms and processes

Fig. 6.4 Slides along headwalls of a gully cut into the alluvial fill of an intermontane basin, Upper Araguaia Basin, eastern
Brazil. Photograph: A. Gupta

6.2.3╇ Flows

Unlike slides where the deformation of the failed mass is limited, the entire mass is
deformed internally in flows and the moving mass behaves like a viscous fluid. Flows also
are accompanied by higher volumes of moisture and fine-grained material. Certain clas-
sifications recognise very large slides such as sturtzstorms as dry flows and identify them
as rock avalanches if the primary constituent is broken rock. When the constituents are
mixed, such failures are called debris avalanches. The term avalanche recognises the very
high downslope velocity of these failures. Debris flows consist of fast-moving solid units
of varying texture including fine material and relatively more moisture and air mixed with
the solids in the moving mass. Debris flows are common in the mountains of the humid and
arid tropics, where high-intensity rain usually causes a series of debris flows to come down
the valleys of small steep streams, especially if sediment is already stored in the valley bot-
toms. For example, the hillslopes of the Caribbean islands are marked by numerous debris
flows after high rainfall or the passage of a hurricane (Fig. 6.5). Rapid snowmelt in the high
mountains, collapses of glacial ice or a moraine with a lake behind it, and overflows from
or collapse of volcanic crater lakes all have the potential for creating debris flows.
A debris flow commonly starts as a slide or a fall of debris at the head of a small valley
draining hilly areas. The addition of more moisture downslope, and probably dilatancy,
transforms the slide into a flow. Debris flows usually come down channels of small streams,
but, as Costa (1984) states, debris flows are quite capable of eroding their own channels
across sloping surfaces. They act as an erosive process, especially over steep slopes, and
may erode up to several metres in bedrock and more in loose material, breaking tree trunks
at ground level. The destruction along the track is related to the velocity of the flow, but such
destruction is seldom seen where the debris flow terminates. Morphologically, debris flows
91 Mass movement on hillslopes

Fig. 6.5 Debris flow after a heavy rainstorm in the Blue Mountains, Mahogany Vale, eastern Jamaica. Photograph: A. Gupta

terminate in a curved rampart of large boulders; flows behind the rampart are bounded by
levees of boulders, displaying hummocky tops. Individual flows measure between several
metres to about a hundred metres wide with a maximum reported depth of 12 m. Usually
the flows are shallow and less than 5 m deep (Selby, 1993; Costa, 1984).
Moving debris flows have been videotaped, displaying a mass like wet concrete flowing
over the ground surface. Pebbles, cobbles and boulders collide with each other with bang-
ing noises as they are tossed round within the mass and carried downslope. Debris flows
move in a series of surges. The surges arrive enriched with boulders and watery turbulent
slurries rich in suspended sediment, but only a few boulders arrive between the surges
(Costa, 1984). Observed velocities range from 0.5 to 20 ms−1 and they are fast over steep
slopes. Even over low gradients, debris flows are capable of travelling long distances,
especially those with a high clay content. Debris flows not only erode their tracks but also
bring down a huge amount of material, which often reaches valley flats at the base of the
slope, augmenting the sediment load of streams. They also build parts of alluvial fans at the
mouth of valleys emerging from mountains.
The solid content of a debris flow varies between 35 and 90 per cent and its water con-
tent between 10 and >30 per cent by weight. Unless the flow originates from a clay-rich
source, the clay content of debris flows is quite low, below 5 per cent. The increased bulk
density supports the transfer of boulders. Excellent general accounts of debris flows have
been provided by Costa (1984, 1988).
Slow, dry flows of clay-sized material are earthflows. These usually travel less than a
metre annually. Earthflows are derived from clay-rich rocks or clayey regolith. They usu-
ally start by a slope failure that releases clay which then spreads over a low gradient as a
toe-shaped feature. In contrast, mudflows are formed when a fine-grained failed material
incorporates a large volume of water. Mudflows may occur in the arid tropics when rapid
rainwash from cloudbursts sends a large volume of mud into river gorges. They can carry
92 Slopes:€forms and processes

Fig. 6.6 Lahar filling a stream channel on the southern slope of the Merapi Volcano, Java, Indonesia. The parapet of a bridge
which used to take the traffic over the river indicates the volume of the fill. The material is used in construction,
explaining the vehicle in the channel. Note the texture of the material. From Gupta and Ashar, 1988. By permission of
Wiley. Photograph: A. Gupta

material as coarse as boulders because of their thickness and viscosity. On reaching flatter
ground beyond the mountains, mudflows tend to spread out and terminate.
Lahars are mudflows of volcanic origin (see Chapter 14). The name is adopted from
Bahasa Indonesia, lahars being extremely common around Indonesian volcanoes. Lahars
start by the addition of a large volume of water to an accumulation of volcanic material.
Slopes of explosive volcanoes, as in the Andes Mountains and on the islands of Indonesia
and the Philippines, are usually covered with old pyroclastic material ranging in texture
from fine-grained ash to volcanic boulders. Thicker layers accumulate in river channels
radiating out of a central volcanic cone. Large and intense falls of rain, as expected in trop-
ical latitudes, create lahars from the stored pyroclastic sediment which then move down-
slope, especially along the river valleys. Pyroclastic flows travelling down a valley and
mixing with the water in the channel may also start lahars. In high mountains, such as the
Andes, lahars may be caused by snowmelt. Lahars travel fast and bring down huge vol-
umes of volcanic material to bury the pre-existing topography and settlements (Fig. 6.6).
They are extremely dangerous in terms of loss of life and property. The 1991 and 1992
eruptions of Mount Pinatubo in Luzon, the Philippines, deposited huge amounts of pyro-
clastic deposits and lahars (Newhall and Punongbayan, 1996; Nossin, 2005).

6.2.4╇ Creep
Creep is the slow downslope movement of the upper parts of the surficial material. It is
frequently described as soil creep because it is the soil part of a weathered profile that tends
to move imperceptively downslope. The shallow upper part of a weathered profile may
move due to three reasons:€slow plastic deformation of the moving mass under its weight;
93 Running water on hillslopes

Fig. 6.7 Running water on hillslopes

alternate wetting and drying; and alternate freezing and thawing. The last characteristic is
limited in the tropics to the high mountains. It is a very slow movement, usually unnoticed,
and continues for a very long time. Soil creep is common under the tropical rain forest,
especially if a thick weathered layer is present and, although imperceptible, it may move a
considerable volume of material over time to a stream at the bottom of the valley. Water, by
reducing the friction between grains, tends to speed up soil creep but even then it is a very
slow and imperceptible process. Its morphological manifestation breaks up a hillslope into
a series of terracettes, and piles up material against the base of an upstanding object such
as a tree trunk or post. Over time, such upstanding objects may also bend downslope in an
area of continuing soil creep. Thus the presence of a soil creep is identified. Soil creep is
not a hazard, except where the creeping material is unstable, e.g. a clay-rich plastic soil.

6.3╇ Running water on hillslopes

Splash erosion, as discussed in Chapter 4, is caused by intense rain with large raindrops
splashing on bare or near bare ground. Splash erosion also depends on the nature of the sur-
face material. Sand or silt is disturbed easily, but a soil that contains a high amount of any
binding material such as organics, clay and CaCO3 reduces splashing. Running water, how-
ever, erodes slope material more effectively. Water accumulates on the surface when rain
intensity is higher than infiltration (the steady rate at which water enters the subsurface). A
thin film of water thus accumulates on the surface before flowing downhill. This is the Horton
overland flow. If, however, all the empty space in the soil (i.e. pores between soil grains) is
already full of water, rainwater cannot enter the subsurface. Instead it emerges to flow on the
surface. This is the saturation overland flow (Fig. 6.7). Horton overland flow tends to occur
over most of the area covered by a hillslope, the saturation overland flow is common towards
the base of the slope and near streams that flow at the bottom of valleys. Exceptions may
occur to this general rule with unusual breaks in gradient or particular types of soil.
Erosion, i.e. removal of grains of material on slopes by running water, requires a min-
imum velocity which is dependent on the depth of flowing water. Thus a belt of no erosion
94 Slopes:€forms and processes

occurs near the crest below which the running water reaches sufficient depth and velocity
to remove surface material. The length of the belt of no erosion depends on the gradient
and soil properties. The running water downslope is known as sheetwash. Sheetwash is
about a centimetre or so in depth but not a uniform spread, as small threads of water with
deeper and faster flow move across the sheet. These threads are effective in moving surface
material, especially those loosened up by splash erosion. If the sheet of water is deeper
(up to 1 m) and faster because of high and intensive rainfall, it is called sheetflood. Small
channels (rills) are eroded where, over time, flow is concentrated and some of the small
channels deepened. Sheetwash is thus replaced by rills downslope. Rills tend to enlarge
into gullies, which are bigger rills. Dunne and Leopold (1978) have defined gullies to be
at least 30 cm (one foot) deep and wide. Bare slopes or slopes with limited vegetation tend
to develop a network of rills and gullies over time (Fig. 4.1). Dunne and Leopold (1978),
after reviewing a number of case studies, concluded that in many cases flowing sheets of
water remove more sediment than rills or gullies. The nature of a specific location (rain-
fall intensity and amount, gradient, soil properties, vegetation) determines which type of
erosion would be more effective at a particular location. Material eroded and transported
downslope ends up at the base where the gradient lessens and the velocity of the running
water drops. A fan-shaped accumulation of sediment tends to occur at the mouth of gul-
lies, which is removed by the valley bottom stream at intervals.

6.4╇ Storage and transfer of surficial material on tropical slopes

A range of slope processes occur in the tropics, of which some are more common. The
efficacy and frequency of the processes depend on the location. Rocks, relief and rainfall
are crucial background factors. Slope processes vary between the humid and arid tropics.
In the humid tropics, slopes are commonly under a thick layer of soil and weathered rock
due to effective chemical weathering. The slope processes that operate are commonly
moisture-driven as illustrated by Gabet et al. (2004) and slopes are expected to fail peri-
odically, especially in the wet season. In contrast, weathering is limited on the slopes of
the arid tropics which are usually under a thin layer of clasts of broken rock and sand.
Slope processes are rare and episodic, and tend to occur following cloudbursts when the
intense precipitation is heavy enough to cause debris flows, etc. without any antecedent
rainfall.
A number of case studies concerning the storage and transfer of surficial material on
tropical slopes have recently been published, allowing generalisations on slope processes
in certain environments. It is not possible to cover all types of slope, but we will discuss
two cases for which a reasonable amount of information is available:

• slopes of the tropical rain forests


• steep slopes of hills and low mountains.
Slopes of the arid tropics, volcanic landforms and anthropogenically disturbed areas are
discussed later, in Chapters 12, 14 and 17 respectively.
95 Storage and transfer of surficial material

6.4.1╇ Tropical rain forest slopes

Tropical rain forests are best developed on lower slopes of mountains and near the con-
tact between hills and plains. In general this is an environment of rolling hills with a thick
soil and a deeply weathered layer. The thickness and nature of surficial material, however,
depend on local lithology and steepness. A number of case studies have been carried out on
slopes of the rain forest, for example in eastern Puerto Rico, Southeast Asia and northern
Australia. In a forest, rainwater, after reaching the canopy level where part of it is inter-
cepted and evaporated back to the atmosphere, moves as (Fig. 4.4)

• throughfall
• fall from the canopy
• stemflow
• flow on leaf-littered surface, for a brief period after intense rain
• flow through the subsurface via soil pores and shallow pipes.
A description of such movement of water is in Sidle et al. (2000).
In the rain forests of the Danum Valley in Sabah, East Malaysia where the annual rainfall
approaches 3600 mm, rainwater ends up as
interception (17.5 per cent of the total rainfall)
throughfall (80.7 per cent)
stemflow (1.9 per cent).
According to this study (Sinun et al., 1992), between 2.0 and 2.5 per cent of the rain travels
as overland flow on reaching the ground. The rest infiltrates and probably flows through
soil pipes. Both throughfall and fall from the canopy may cause splash erosion, loosening
the upper soil and preparing it for removal in surface runoff. Given the brief and intense
rainfall, widespread overland flow lasting for short periods is common in the rain forests.
The brief duration of flow on a litter of leaves overlying humus-rich clays is not particu-
larly erosive. The movement of slope material under gravity is also limited, and commonly
occurs by surface creep. Apart from slow but continuous creep, sediment moves down-
slope for brief periods in surface runoff (sheetflow) and through subsurface pipes emerging
at the face of vertical banks cut in the regolith. Under well-developed forests, the amount of
sediment that is eroded and removed is limited, usually not exceeding 102 tkm−2yr−1. Such a
rate is unlikely to either form erosional scars or build extensive depositional features. The
sediment commonly consists of sand, clay and oxides and hydroxides of iron and alumin-
ium, particularly on igneous rocks. Gravels (up to boulder size) may occur on slopes as lag
or exposed corestones. Rills and gullies, if formed, are few in number until the vegetation
cover is destroyed. Such is the nature of the slopes of the forested hills of the tropics.
With the destruction of vegetation, however, either in storms such as tropical cyclones
or by anthropogenic clearing, sediment loss can be extreme. Such cases are discussed later
(Chapter 17), but just to illustrate, Anderson and Spencer (1991) in a review of tropical rain
forest conditions refer to a case study in Java that reported an increase in sediment yield
from 30 to 1590 tkm−2yr−1, an increase of 53 times. Such a figure is comparable to erosion
in natural disasters, such as tropical cyclones, where destruction of the vegetation cover in
96 Slopes:€forms and processes

high winds is followed by heavy rainfall. In all these cases, the slopes become dissected by
landslide scars and gullies. Elsewhere, the rain forest is a quiet environment, except in tec-
tonically unstable areas. Soil creep is the dominant mode of transportation down slope and,
unless lithology intervenes, a convex–concave slope tends to develop over a thick layer of
soil and weathered rock.

6.4.2╇ Steep slopes of hills and low mountains

Traditionally, given the thick layer of soil and weathered rock in the humid tropics, rota-
tional slides (slumps) were considered as the common type of failure on hillslopes. The
failures, however, are just as often translational or planar. The plane of failure frequently
turns out to be a relict structural surface or an encrustment layer that impedes the down-
ward passage of subsurface water, as explained in section 6.2.2. Usually, when slopes fail
in this fashion, a number of small and shallow translational slides are seen on a hillslope,
often at a comparable elevation. Antecedent and trigger rainfalls (Gabet et al., 2004) are
important for both translational and rotational slides (Box 6.1).
Simon et al. (1990) reviewed failures and developments of hillslopes in the low moun-
tains of eastern Puerto Rico. They came to certain conclusions which may be extended to
cover most of the hillslopes of the humid tropics. Shallow translational failures are prob-
ably the primary mechanism of hillslope denudation. This happens because of (1) deeply
weathered rocks and thick soils, (2) occurrence of discontinuous zones of contrasting dens-
ity and permeability throughout such layers, and (3) presence of relict features and tensile
stresses. Both pathways and impedances for the movement of surface water therefore occur
in the regolith. Deep failures do happen, but shallow failures are common. Simon et al.
(1990) measured the depth of failure in eastern Puerto Rico as between 3 and 7 m. Slopes
around the heads of channels (cove slopes) tend to fail when the saturation level is high
and subsurface water converges towards the channel. Although elevations in eastern Puerto
Rico rise to beyond 1000 m, slopes usually fail at between 600 and 800 m, probably due to
water accumulated from the upper slopes crossing the threshold of failure at this elevation
band. Such elevation-selective failures may happen on slopes in other areas. These events
combine to determine the characteristic form of the local hills.
Large-scale failures also occur and these are normally associated with tectonically active
areas or steep slopes within the tropical cyclone belt. The worst scenario is a combination,
such as a tropical cyclone in action over steep and tectonically unstable hills. Simonett
(1967), working in Papua New Guinea, estimated that the denudation rate of the earthquake-
prone Toricelli Mountains where landslides are common is about 100 cm in 1000 years.
In comparison, the denudation rate in the neighbouring Betwani Mountains, less affected
by earthquakes, is 22 cm per 1000 years. Steep tectonic mountains are prone to slope
failures that contribute a large proportion of the sediment that travels down major rivers
whose headwaters originate in mountains associated with active plate margins. Common
examples of failures from steep slopes in the literature are from the Himalaya, Andes,
mountainous Papua New Guinea, Hong Kong, the escarpment zone of eastern Brazil, the
Caribbean and Taiwan. Undoubtedly steep slopes fail almost everywhere in the tropics, but
these areas have been better studied.
97 Storage and transfer of surficial material

Fig. 6.8 Landslide in the Blue Mountains, eastern Jamaica overlooking the Yallahs River at Mavis Bank. A number of factors
control the slopes of the Blue Mountains. A regional fault runs along the hillside; the slopes are in alternate beds of
greywacke and shale; and the area is periodically affected by storm rainfall, some of which is caused by hurricanes.
Note also that, as a result, large quantities of very coarse material fall into the Yallahs, which can remove and
redistribute most of it downstream if it is in flood from hurricane rain simultaneously with slope failures. Photograph:
A. Gupta

Slope failures in the tropical mountains (Fig. 6.8) tend to contribute most of the
river sediment, irrespective of the scale of the basin. Larsen and Torres Sánchez (1992)
calculated that 81 per cent of the total sediment transported out of the Mameyes River
basin (17.8 km2) in eastern Puerto Rico comes via mass movements. In this paper they
also report the effect of a 1989 hurricane (Hurricane Hugo) which had a sustained
wind speed exceeding 46 ms−1 and a 24-hour rainfall total ranging between 100 and
339€mm. Average rainfall intensity varied between 34 and 39 mmhr−1. Up to 97 per cent
of the trees were defoliated in certain river basins; a very high proportion of these were
destroyed. This was followed by hurricane rain falling on exposed ground. More than
400 landslides occurred on the slopes facing the approaching hurricane. Two hundred
and eighty-five landslides were mapped within 64 km2 from aerial photographs. Field
mapping after similar storms has indicated even higher figures in Jamaica, approaching
60 mass movements per km2, more than half of which being reactivated slope failures
(Maharaj, 1993). The size of such slope failures varies. The biggest one mapped in
Puerto Rico was a debris slide in which about 30 000 m3 of material was moved into the
Rio Cubuy, exposing unweathered bedrock along the erosional scar. It happened three
days after the hurricane.
Most of the landslides occurred within the area of heavy rain, bounded by 200 mm iso-
hyets in this case (Larsen and Torres Sánchez, 1992). Seventy-five per cent of landslides
occurred on forested hillslopes, which indicates that vegetation cover may not be a restrict-
ing factor in large storms. Shallow soil slips and debris flows were identified as the primary
types of failure. The failure plane of soil slips tends to be at the contact between soil and
98 Slopes:€forms and processes

Fig. 6.9 Debris flows and coarse channel sediment following the December 1999 flood in northern Venezuela. Photograph:
M. C. Larsen, US Geological Survey

weathered rock. Larsen and Torres Sánchez calculated an approximate denudation rate of
164 mm per 1000 years, taking into account the estimated volume eroded by landslides and
the 10-year recurrence interval rainfall in the hurricane. High-magnitude formative events
like this are frequent (once in 5–10 years) in certain humid tropical regions with steep
slopes, indicating the effectiveness of climate in river basin geomorphology of these areas.
Short-duration, high-intensity rainfalls regularly cause shallow slips and debris flows in
humid tropical mountains. Bigger events such as tropical storms add larger and deeper
slides to it at intervals of 5–10-years. These high-magnitude events erode huge parts of the
hillslopes, contributing a large and coarse sediment load to the streams and, for islands or
coastal mountains, to the coast and offshore regions. The effect of the excessive water and
sediment loading on streams is discussed in Chapter 7.
Larsen et al. (2001a, 2001b) described a sequence of rainfall-triggered debris flows and
flash floods in northern Venezuela. The region received over 1200 mm of rain in the first
16€days of December 1999, including 911 mm between 14 and 16 December when the
slopes were already saturated. This was a catastrophe with lives lost (estimated between
15 000 and 30 000) and property damaged, but the event also illustrated the role these
episodic events play in modifying mountain slopes, stream channels, coastal features and
river-mouth fans (Fig. 6.9).
Creeps, planar slides and debris flows seem to be the dominant processes on tropical
slopes, although all the types discussed earlier do occur in appropriate environments. For
example, the type of mudflows known as lahars erodes and reforms volcanic hillslopes.
Ice, steep slopes and tectonically shattered rocks cause a range of slope failures on high
mountains such as the Himalaya or Andes, which is discussed in a separate chapter on trop-
ical mountains (Chapter 13). It is, however, worthwhile to recall that the sediment load of
rivers essentially owes its origin and characteristics to the tectonics, volcanism, weathering
and mass movements on mountain slopes, often hundreds or thousands of kilometres away
from the sea.
99 Questions

6.5╇ A general description of tropical slopes

Slopes in the tropics, as everywhere else, are determined simultaneously by lithology, tec-
tonics and a combination of geomorphic processes. The processes are climate-dependent
to a large extent and vary across the tropics. Thus slopes on convergent plate margins tend
to undergo rapid changes, especially if periodically visited by tropical storms. A large vol-
ume of sediment is derived from such mountains and volcanic highlands at a rapid rate
and stored in wide valleys at low elevations. For example, the alluvium of the Ganga val-
ley is mainly derived from sediment eroded by slope failures and running water from the
Himalaya.
Parts of the ancient Gondwana constitute stable areas of the tropics in every continent.
Extensive low-relief features with scattered isolated hills or plateaux are common features
of such a landscape. Slopes are either very low in the plains or near vertical as the sides of
the plateaux often rise in steep escarpments. The rocks, however, are cratonic with a thin
regolith on top. These areas, irrespective of erosional slope processes in operation, do not
produce much sediment. Slopes, processes and sediment supply therefore vary across the
tropics, and it is difficult to generalise about tropical slopes. It is necessary to qualify first
the regional geological and climatic backgrounds. Many of the succeeding chapters pro-
vide examples of different types of tropical slope, and highlight the variability in tropical
landforms due to geological history, climate and anthropogenic alterations. The quote at
the beginning of this chapter highlights the controversy that once reigned regarding the
origin of tropical landforms.

Questions

1. A number of geomorphologists believe that slump is the primary type of slope failure in
the humid tropics because of the thick mantle of weathered material that covers the hard
rocks. Do you agree? Justify your viewpoint.
2. How would you expect hillslopes to look in
a) an area frequently visited by tropical cyclones?
b) a granitic area in the humid tropics before and after deforestation?
3. What is your expectation of the sediment load of a river that
(a) drains a steep, tectonically active mountain area?
(b) drains a forested area of low granitic hills in the humid tropics?
4. Chapter 6 includes a list of slope properties that induce failures. List the properties
which stabilise slopes. Give examples in support of your chosen properties.
5. Why is the amount of moisture retained by the surfical material so important for slope
failures?
6. Fig. 6.3 is from Gabet et al. (2004) where the data are from a drainage basin in the
Himalaya Mountains in Nepal. Does it demonstrate that landslides are more frequent
towards the end of the wet monsoon?
100 Slopes:€forms and processes

7. Pitts (1992) plotted a similar diagram for western Singapore. Antecedent rainfall was
plotted against trigger rainfall. It was possible to identify the threshold beyond which
landslides occur, but the data showed a wide scatter for smaller and fewer events that
had occurred in urbanised western Singapore following a limited amount of rainfall.
Why the scatter?
8. It has been said that most of the sediment of major streams is derived from headwater
mountains. Is this solely related to failures in steep slopes?
9. How would you determine whether the surface material under a tropical rain forest is
creeping downslope? Design a methodology for demonstrating this phenomenon and
measuring the rate of creep.
10.╇Larsen and Torres Sánchez (1992) reported on slope failures following the arrival of
Hurricane Hugo over eastern Puerto Rico in 1989. There were many failures, but the
biggest one happened three days after the hurricane. Why the delay?
11.╇Do you agree with the summary statement that creep, planar slides and debris flows
seem to be the dominant processes on tropical slopes? Justify your decision.
7 Rivers in the tropics

The moisture-laden wind blows at speed


Awakening the mad river
Rabindranath Tagore,
translated from Bengali

7.1╇ Components of a river system

Rivers are conduits for transferring water and sediment from the land to the ocean. Among
the multitude of factors that control their dimensions, appearance and behaviour, four are
of high importance:€water, sediment, sediment texture (grain size) and channel slope. If any
of these changes, the river can be expected to accommodate the change with a proportional
shift in its form and function. To illustrate, other factors being equal an increase in the
supply of sediment would lead to deposition in the channel. However, if the increased sedi-
ment simultaneously becomes finer, the river may be able to transfer most of the fine sedi-
ment downstream. In contrast, an increase in stream flow due to rainfall from a large storm
or tectonic tilting of the channel slope may lead to erosion. This relationship (Fig.€7.1) was
expressed in a simple diagram by W. M. Borlaug based on an equation derived by E. W.
Lane (1955). We need to explore these factors in order to understand rivers.

7.2╇ Water in river channels

The basic measure of water in rivers is discharge. Discharge at a point on the river is
defined as the volume of water passing that point in unit time. Discharge can therefore be
computed by multiplying the cross-section of the channel (also known as channel area)
with the velocity of the stream. It is written as
Q = A.v = w.d.v (7.1)
where Q = discharge
A = channel area
v = velocity of the flowing water
w = width of the surface of the channel
d = mean depth of the channel.
Channel width and depth are both measured in m, velocity in ms−1, and therefore the
Â�discharge is in m3s−1.
101
102 Rivers in the tropics

Fig. 7.1 The river balance. Generalised from the original diagram proposed by W. M. Borlaug and E. W. Lane. From Gupta,
2011. By permission of Royal Academy for Overseas Sciences, Brussels

Velocity in a river is dependent directly on the component of gravity that pulls a volume
of water downslope, and indirectly on the friction that has to be overcome before water
can flow in the channel. The friction (also known as the resistance) is the summation of
several characteristics of the channel:€unevenness of the surface of the bed and banks of
the river, the bends in the channel, channel vegetation, nature of the sediment, water tem-
perature, fish in the river, etc. It is a basket term that includes everything that retards the
flow of water. It written as n (called Manning’s n) or ff (called friction factor). Various
general equations of river velocity have been derived from these two forces (gravity and
friction).
Manning’s equation:€v = (R2/3s1/2)/n = (d2/3s1/2)/ n (7.2)
Chézy equation:€v = C√(ds) (7.3)
d’Arcy–Weisbach equation:€v = √(8gds/ff) (7.4)
where v, d and n have the same attributes as before and
R = hydraulic radius of the channel defined as wetted perimeter/surface width, which
can be approximated for wide channels by mean channel depth (this approximation
works for most rivers)
s = slope
C = a constant that represents resistance to flow and can be derived as √(8g)/ff
ff = friction factor used in the d’Arcy–Weisbach equation
g = acceleration due to gravity.
Both the Chézy and d’Arcy–Weisbach equations are dimensionless and theoretically pref-
erable to Manning’s, which is not. But Manning’s equation remains popular, as it is easy
to compute and physically perceive. It is so much in use that a manual exists for estimat-
ing Manning’s n of a river by comparing it with photographs of a number of streams with
103 Water in river channels

Table 7.1╇ Manning’s roughness coefficients (n) for various types of channels
Channel type Manning’s n

Canal with smooth concrete 0.012


Canal with ordinary concrete 0.013
Straight unlined earth canals 0.020
Rivers in fair condition, some vegetation 0.025
Winding natural streams 0.035
Mountain streams with rocky beds and bank vegetation 0.040–0.050

Values of Manning’s n from Chow 1964

known values of n (Barnes, 1987). Tables are also available which provide a physical
description of the river with a value of n (Table 7.1).
We see from the velocity equations that stream velocity is related directly to channel
depth and slope and indirectly to all types of friction summed together. The fastest vel-
ocity in a river occurs slightly below the surface due to friction with air which slightly
slows down the flow at the surface. Velocity is low near the channel bottom as expected.
The velocity that occurs at 0.6 depth of the channel is acceptable as the average vertical
velocity but the mean of the velocities at 0.2 and 0.8 depths gives a better approximation.
Cross-channel velocity is higher over deeper parts, and isovels (lines of equal velocity) in
the channel tend to follow the channel shape (Fig. 7.2).
The velocity of a river increases downstream. Its slope decreases downstream but its depth
increases, and depth carries a higher power than slope in the Manning’s equation. Rivers
tend to flow over sand, silt or clay in their lower courses and thus resistance to flow is also
reduced.
Flowing water exerts force on the bed and banks of a channel. This is the force that
erodes. The force per unit area of the channel is called the boundary shear stress and can
be written as
τ = γds (7.5)
where τ = boundary shear stress, determined in Newtons per square m (Nm−2)
γ = specific weight of water (9807 Nm−3 for clear water)
d = depth (or the hydraulic radius) of the channel in m
s = slope of the channel.
The form of the equation indicates that rivers flowing with high velocity will exert greater
stress on the bed and bank of the channel and hence erode. A river in flood is erosive.
Erosion or deposition in a stream is also determined by its unit stream power or the
power that a stream exerts over a unit area of the channel (Bagnold, 1966). This is meas-
ured in watts per square metre (Wm−2), and expressed as
ω = γQs/w = γwdvs/w = τv (7.6)
where ω = unit stream power and the other symbols are as before.
104 Rivers in the tropics

(a)

(b)

(c)

Fig. 7.2 Idealised cross-channel distribution of velocity in a river:€(a) location of approximate mean velocity; (b) vertical
velocity distribution; (c) cross-channel velocity distribution with shaded area indicating location of fast flow in
channel

7.3╇ Sediment in river channels

Sediment is commonly perceived as the load that the water of the river carries, which can
be divided into two major types:€solution and solid (also known as particulate) load. The
solid load is further differentiated by its mode of conveyance. If the immersed weight of
a particle is carried by the water, it is classified as suspended load. If the weight is carried
mostly by the solid bed of the river, it is bed load.
Commonly, the solution load of a river is only a very small fraction of its total load.
However, the solution load of the river may be significant if the upstream drainage basin
includes a large area of rocks that are easily dissolved, such as evaporites or carbonates.
105 Sediment in river channels

The major chemical elements transported in solution are calcium, silica, iron and alumin-
ium. Stallard (1985) has demonstrated that advanced chemical weathering tends to increase
the solution load in parts of the Amazon and Orinoco Basins due to the presence of elem-
ents from highly weathered rocks in their solution load. The total volume of solution load
increases with discharge, although the increase in discharge also tends to dilute the concen-
tration at the same time unless a very large amount of soluble material is available to the
river. Thus, a higher concentration is usually found in low flows.
It is obvious from the equations discussed earlier that the power of a river and the stress
it imposes on unit areas of the channel bed and banks are related to its velocity. As the
velocity rises, the river tends to remove material from the channel perimeter and transport
it. This is erosion by rivers. Very coarse material such as cobbles and boulders are always
carried as bed load, except in extreme floods. Such catastrophic floods occurred in glacial
lake bursts towards the end of the Pleistocene (Baker, 1981, 2007). Very large floods, such
as those caused by tropical cyclones, have been known to suspend pebbles and cobbles for
short distances. But this happens rarely, and pebbles and cobbles, just like boulders, are
usually carried as bed load. Sand usually travels as bed load, but in high floods it is fre-
quently carried suspended by eddies in a river. Silt and clay are transported as suspended
load, unless they are in aggregates.
The removal (erosion) from the channel of individual particles (clay to boulders) is a
function of stream velocity. The critical erosion velocity of a particle is the velocity at
which it starts to move. Grains of clay and fine silt, although of finer texture, tend to stick
to each other in clumps and therefore require higher velocity for erosion than expected.
Sand is removed at a relatively lower velocity than silt or clay, as each grain is independ-
ently subjected to stress. Once sediment grains are entrained, they can be transported with
a velocity lower than the one for entrainment.
A sediment grain being carried either on the bed or in water comes to rest (a phenomenon
known as deposition) when the stream velocity falls below the level needed to carry it. This
is the fall velocity, which is related solely to grain size. Hjulstrom mapped both velocities
against grain size (Fig. 7.3) to explain the work of a river at a particular velocity. According
to the diagram, sand is the easiest to erode from the perimeter of the channel. Anything
coarser (pebbles, cobbles, boulders) requires a higher critical erosion velocity and, given
their size and weight, they are only transported for a short time, as their fall velocity is also
high. Sand is carried longer. The critical erosion velocity of silt and clay is higher than sand,
but once suspended in water, they are transported for a long time as their fall velocity is very
low. Thus river transport leads to a separation of material downstream, finer material travel-
ling longer. This grouping of particles of different size over distance is called sorting.
The channel of a river flowing from mountains to the sea therefore contains pebbles, cobbles
and boulders near the mountains and silt and clay near the sea. Sand is ubiquitous, and is found
just about everywhere although the modal size of the sand grains decreases downstream.
Certain generalisations can be reached regarding the nature of the suspended load in a
stream:

• The suspended load increases directly with discharge.


• The vertical distribution of the suspended load in a stream usually but not always paral-
lels the vertical velocity distribution.
106 Rivers in the tropics

Fig. 7.3 Hjulstrom diagram of velocity required to remove and transport sediment load of a given texture. The velocity
required to move a sediment grain varies and hence a band of erosion velocity has been shown. Sand (0.06–2 mm)
is eroded at a relatively lower velocity than either clay or silt (< 0.0.6 mm) or coarser material (pebbles (2–64 mm),
cobbles (64–256 mm) and boulders (> 256 mm). The line marked ‘fall velocity’ indicates the velocity required to keep
a particular grain of sediment moving. Erosion and fall velocities together indicate that material coarser than sand is
removed only at high velocity (floods) and deposited soon. Sand is picked up first and transported over a wide range
of river velocities. Silt and clay are erodible at a velocity higher than sand but once in motion they are carried for a
long distance. This leads to sorting of sediment in the downstream direction and also determines that rivers that flood
will have a diagnostic appearance as discussed in the text. After Hjulström, 1939

• The peak concentration of the suspended load arrives before the peak in the hydrograph,
unless the supply of sediment is abundant
• The suspended load may exhibit hysteresis; for the same discharge, the rising limb of the
hydrograph may carry more sediment than the falling limb.
All these conclusions are diagrammatically illustrated in Figure 7.4.
The amount of suspended load being carried can either be computed or directly meas-
ured from samples of river water. The numerical figure is called sediment discharge and
is determined as milligrams per litre or tonnes per day. Bed load is difficult to measure
and is often computed from theoretical bed load functions proposed by Shields, DuBoys,
Meyer-Peter and Muller, Einstein and others. Data on the bed load of streams is therefore
not as easily available as suspended load. However, the bed load generally constitutes only
a small fraction of the total sediment load of most rivers and its amount may fall within the
error term for suspended load measurement. It is therefore not always measured. Bed load,
however, is important for certain rivers like the Brahmaputra, which are special cases.
Material on the channel bed moves in three ways:

• rounded or subrounded grains can be rolled


• any shape of grain can be dragged by slipping and sliding along the river bed
107 Sediment in river channels

(a) (b)

(c) (d)

Fig. 7.4 Idealised pattern of suspended load in a stream channel; Q = discharge of water, Qs = discharge of suspended load

• grains may saltate, i.e. move in progressively decreasing hops after being hit by another
moving grain.
Two terms are commonly used to describe the transporting ability of a river. The compe-
tence of a stream at a given flow is the size of the largest grain that can be moved as bed
load. The capacity of a stream at a given flow is the total sediment load.
The ability of a stream to erode and transport sediment is related to its bed shear stress
(τ) or unit stream power (ω). Bagnold (1966) has shown that stream power determines
the capacity of a stream to transport sediment. The higher the values of shear stress and
power, the more capable is a stream of carrying out erosion and sediment transport.
Extremely high erosion is therefore carried out in floods or steep river gorges, especially
where high values of depth and slope are periodically attained given that the drainage
area upstream of the gorge is capable of producing large flood discharges (Baker and
Costa, 1987). For example, Williams (1983), after examining a large database, stated that
within his data set at least 1063 Wm−2 unit power, 225 Nm−2 shear stress and 2.52 ms−1
mean velocity were needed to move a 1.5 m boulder. Such figures are exceeded in many
gorges and steep floodprone reaches. Coarse material is therefore transported through
the gorges and steep reaches only to be deposited in flatter alluvial basins beyond. There
are instances of floods with shear stresses in the region of 103 Nm−2, as happens in hilly
areas during tropical cyclones, which can move boulders a few metres in diameter as bed
load (Fig. 7.5).
108 Rivers in the tropics

Fig. 7.5. Buff Bay River, eastern Jamaica, draining the northern slopes of the Blue Mountains. The coarse bed load material
indicates the high competence of this river in flood. The very large rounded boulder is 5 m in diameter. From Gupta,
2000, with permission from IAHS. Photograph: A. Gupta

7.4╇ Channel geometry

The dimensions of a stream channel are determined mainly by a number of background


factors:€discharge, amount of sediment, texture of sediment and channel slope. These fac-
tors are very effective when the river is entirely in alluvium. If the river is cut into rock
or flowing over a thin layer of alluvium on rock, the underlying lithology and geological
structure also become a determining factor for the geometry of the stream channel.
An alluvial river tends to be sensitive to changes in water and sediment and adjusts to
changes easily. Richards (1982) has called them self-formed rivers. Alluvial rivers there-
fore tend to change their channel geometry frequently. On the other hand, a river flowing
through bedrock requires a high amount of energy to erode the bed and adjust its morph-
ology. Rivers on rock therefore only change their channel geometry under conditions of
high energy, e.g. during a large flood or an increase in slope due to a regional tectonic activ-
ity. Such changes are separated by long time intervals. An extreme example is the Upper
Indus, which has stayed pinned into a tectonics-driven suture zone. A river flowing through
a rock canyon cannot change its course within the gorge without catastrophic events taking
place. Temporary changes may, however, happen to the alluvial features deposited within
the rock-bound channel (Chapter 8).
It should be stressed that rivers in alluvium and rivers on rock tend to appear and behave
differently and models or statistically generated parameters cannot be transformed directly
from one type to another. The situation becomes complicated when rivers have alluvial
banks but their bed is on rock or a thin layer of alluvium on rock. A long river usually
flows over rock in its upper course in the mountains and in an alluvial valley downstream.
Certain structure-controlled rivers, such as the Narmada, Irrawaddy and Mekong, flow
109 Channel geometry

alternatively in rock gorges and alluvial basins, thereby alternating their form along their
course.
Channel geometry for alluvial rivers is better known. In this section we review channel
geometry for both types but with more details for alluvial rivers.
The geometry of a river channel has four major characteristics:

• channel slope
• channel size
• channel shape
• channel pattern.
These four are interrelated and the river is the sum of all four characteristics. It is, however,
easier to examine each characteristic separately and then assemble the complete picture.

7.4.1╇ Slope of the river

The slope (also known as the gradient) of a river is determined by plotting the elevation
of its bed against distance from its source to its mouth (Fig. 7.6). The plotted line is called
the longitudinal profile (long profile). As the headwaters of most rivers start in mountains
or uplands, the upper part of a longitudinal profile is steep, giving rise to a concave-to-��the-
sky appearance. The degree of concavity is important as the gradient of a river controls
the velocity and power of a river to a large extent (see equations 7.1–7.6). When plotted
on semi-logarithmic paper, long profiles become straight lines or a combination of straight
lines. The standard form of the equation is
h = alb, (7.7)
where h = elevation
l = distance from the source of the stream and
a and b are numerical coefficients for individual rivers.
Rivers like the Ganga or Brahmaputra tend to erode and pick up sediment as they des-
cend the steep southern slopes of the Himalaya Mountains. The sediment is deposited in
the channel or overbank in the lower, flatter part of the long profile. Similarly, 90 per cent
of the sediment of the Amazon is derived from the Andes (Meade, 2007). The gradient of
a river tends to lessen progressively along its length. The term 'base level' is used to indi-
cate the level of the water surface at the mouth of a river. The base level for a stream could
therefore be the level of the sea or of a lake or of a bigger river at the confluence, and this
level is recognised as the limit to which the river can deepen its channel.
Mackin (1948) visualised an equilibrium condition in rivers where the gradient at each
point is just right for passing a quantum of sediment from upstream to downstream with-
out any significant erosion or deposition happening. This notional longitudinal profile is
known as the profile of equilibrium or a graded profile. A graded profile is not necessarily
smooth but can have sections that vary in slope in order to adjust to local rock types or
changes in water and sediment delivered to the stream. In reality, the profile of a river at
grade tends to fluctuate insignificantly on both sides of the expected profile. The ultimate
profile is a function of a number of variables:
110 Rivers in the tropics

Fig. 7.6 Generalised long profile of the Nile from the White Nile headwaters to the sea. Note variations in slope. From
Woodward et al., 2007. From Said, 1994, by permission of Wiley.

• regional lithology
• discharge
• amount and texture of channel sediment
• resistance to flow
• stream velocity
• width of the channel
• depth of the channel
• regional slope.
The major channel properties (profile, size, shape and pattern) are mutually dependent. The
design of alluvial canals in Egypt and India reflects this behaviour. After the initial scour
and fill and finding of the appropriate width, depth and gradient, these canals become sta-
ble and are said to be in regime. The concept is comparable to that of graded rivers. A river
maintains its channel dimension and form for a long time. It may undergo minor short-term
fluctuations but usually returns to its long-term characteristics. A major disturbance in the
drainage system is required to significantly change a graded river.
It has been shown by Langbein and Leopold (1964) that channel forms adjust until
the energy used per unit bed area for the river to flow is uniform along the channel. This
indicates that power expenditure per unit bed area is constant (Richards, 1982). As the dis-
charge increases downstream, it becomes necessary for the slope to decrease and the river
develops the concave-to-the sky longitudinal profile.
This implies that a river needs to adjust to the material lying on its bed. A river flow-
ing across several rock types may display a composite profile formed of segments at
varying slopes on different lithologies due to both differences in the erodibility of dif-
ferent rocks and the variation in size of the resultant bed load. Each segment will have
a separate curve with different slopes and concavity. Coarser material will cause the
gradient to steepen. A graded profile need not be a smooth profile but may have sev-
eral protuberances, as demonstrated by Wolman (1955) for the Brandywine Creek, PA,
USA.
111 Channel geometry

Fig. 7.7 The Narmada at Dardi, central India. Note the waterfall at the knickpoint in the background with a gorge with rock
shoulders downstream of the falls. Imbricated boulders 2 m in size indicate the competence of the Narmada in flood.
From Gupta et al., 1999. By permission of Wiley. Photograph: A. Gupta

The concave profile is a generality that is not found everywhere. Rivers in arid envir-
onments may have a straighter or a convex profile, as discharges may not reach the mouth
of the stream, being lost by evaporation and seepage through channel sediment. A concave
profile is interrupted in places where rivers cross harder rocks or where uplift has taken
place. Here the gradient is steep, so the river tends to flow over it with higher velocity and
power. A break in the profile is called a knickpoint. Knickpoints frequently, though not
always, give rise to waterfalls or cataracts (Fig. 7.7). The high energy of the river at a steep
knickpoint leads to erosion and the knickpoint tends to retreat slowly and finally disappear.
A retreating knickpoint may leave a gorge section downstream that starts from the original
location of the knickpoint.
Large rivers like the Mekong have deep scour pools in rock-bound sections, the base of
which could be at elevations below the sea level, even a long distance from the sea (Fig.
7.8). More than 90 deep troughs were found eroded on the bed of the Changjiang (Yangtze),
during the construction of the Three Gorges Dam, when the river was diverted for dam con-
struction and its former bed exposed. The locations of such troughs were associated with
tectonic zones of weakness and the deepest of these were about 80 m deep (Yang et al.,
2001). The bed of a river can be either smooth or show significant relief. The longitudinal
profile should not be expected to be smooth everywhere.

7.4.2╇ Size of the river channel

Intuitively it seems that the cross-section of a river (also referred to as channel size or chan-
nel capacity) depends on the discharge it carries when full. This discharge is referred to as
112 Rivers in the tropics

Fig. 7.8 The Mekong eroding into quartzitic rocks along the border of Thailand with Lao PDR, upstream of Pakse, Lao, PDR.
Scale indicated by trees at the bottom left. IKONOS satellite image © Centre for Remote Imaging, Sensing and
Processing, National University of Singapore (2003), reproduced with permission. See also colour plate section

the channel-forming discharge or dominant discharge. A channel-forming discharge should


have the following properties.
1. It is capable of maintaining the channel at the existing cross-section without significant
erosion or deposition.
2. It is not exceeded frequently enough to deposit sediment overbank to elevate the
floodplain.
Bankfull, discharge is the discharge that fills the channel when any addition of water will
cause the river to flow overbank. As the two banks of a river are usually not at the same
elevation, bankfull discharge is identified from the width–depth ratio of the channel. If this
ratio at different heights above the bed is plotted on a graph, the lowest width–depth ratio
occurs at the bankfull depth. It has been statistically shown that width–depth ratio occurs at
an interval of approximately 1.5 years on the annual maximum series. The annual �maximum
series is a listing of the highest discharge of the year for a number of years. Therefore a dis-
charge of this recurrence interval may be taken to represent the bankfull discharge. Studies
in the United States and Western Europe have shown that the bankfull discharge is statis-
tically related to the channel dimensions. Measures of channel dimensions such as channel
width, channel area and bend amplitude for various rivers show this relationship when plot-
ted against their bankfull discharges (Leopold et al., 1964).
This statistical relationship, however, does not hold without caveats for certain types of
river. These are rivers deeply incised in rocks or alluvium, rivers periodically affected by
high-magnitude floods or rivers in a strongly seasonal climate. Pickup and Rieger (1979)
113 Channel geometry

have shown that the size of many tropical rivers may depend on more than one discharge
where floods and seasonality are common in both the arid and humid tropics. They attrib-
ute the task of channel forming to a set of discharges, termed by them as nested discharges,
all of which to some extent control the channel size. This characteristic of tropical rivers is
discussed in detail in section 7.6.

7.4.3╇ Shape of river channels


Two rivers of the same cross-sectional area may have different shapes:€one deep and nar-
row, the other shallow and wide. Channel shape is calculated as the dimensionless width–
depth ratio (also known as form ratio), which is the ratio between the water surface width
and the corresponding mean depth of the river. This ratio changes as the water level rises or
falls in the channel, reaching its minimum value at the top of the banks. The rate of change
depends on the slope of the banks; a channel with near-vertical banks shows a different rate
of change than the case where the banks (one or both) are sloping.
The shape of a river channel at a given cross-section is a function of three variables:
1. bed and bank material
2. volume and texture of the sediment being transported by the river
3. river discharge and its variations.
Channels with cohesive banks in silt and clay or fine sand tend to be steep. Rivers with
a high proportion of coarse material in their banks tend to have sloping banks (Schumm,
1960). Finer material encourages vegetative growth in the channel and banks, which further
stabilises them. Absence of vegetation on coarse bank material and lack of cohesiveness
allow the active erosion of banks, widening the channel. Rivers that carry a high proportion
of coarse bed load tend to have high form ratios but rivers with a high suspended load (usu-
ally fine sand, silt and clay) have relatively low ones (Lacey, 1930). Rivers with marked
seasonality in discharge and those that flood at short intervals tend to have high form ratios
(section 7.6). All these background factors are interrelated. A river that floods will bring in
coarse material and a large bed load, thereby maintaining a wide channel.
Leopold et al (1964) listed four generalisations regarding channel shape.
1. Most rivers have approximately parabolic cross-sections.
2. River channels become asymmetric at bends.
3. The width–depth ratio of river channels increases in the downstream direction.
4. Large rivers have high width–depth ratios.
The shape of the channel is often referred to as the form of the channel. It is an important
measure because it relates to two other properties of the river:€the channel pattern and the
texture of channel sediment.
The shape of each channel changes (1) as it is gradually filled with water or (2) in the
downstream direction. Leopold and Maddock (1953) described a technique for measuring
this change, called hydraulic geometry. It is determined by plotting width (w), mean depth
(y) and velocity (v) of a channel against corresponding discharge values on logarithmic
graph papers. The plotting can be done for two different sets of data. In the first set, all
114 Rivers in the tropics

b + f + m = 0.28 + 0.42 +0.30 = 1.00

Fig. 7.9 A classic example of at-a-station hydraulic geometry determination, Powder River at Locate, Montana, USA. From
Leopold and Maddock, 1953. Courtesy of USGS

measurements are from the same cross-section on the river. It is at-a-station hydraulic geom-
etry (Fig. 7.9). In the second case, measurements are taken at various points along a river for
the same type of flow such as half-bankfull. This is downstream hydraulic geometry.
The relationships of width, mean depth and velocity against discharge can be expressed as
w = aQb (7.8)
y = cQf (7.9)
v = kQm, (7.10)
where Q, w, y and v are the same as before and b, f and m and a, c, and k are numerical expo-
nents. It can be shown that b + f + m = 1 and a.c.f. = 1. This allows a comparison between riv-
ers as the relative values of b, f and m will vary although they would sum approximately to 1.
A wide channel with banks in coarse material will have a high b and low f, as it gets relatively
wider with increasing discharge. A narrow river with sloping banks will show the opposite. A
115 Channel geometry

Table 7.2╇ Values of at-a-station hydraulic geometry


River and location b f m Source

Yallahs at Llandewey, Jamaica 0.51 0.29 0.23 Gupta (1975)


Godavari at Shagad, India 0.32 0.33 0.36 Deodhar and Kale
(1999)
Godavari at Nanded, India 0.24 0.63 0.12 Deodhar and Kale
(1999)
Tapi at Burhanpur, India 0.24 0.51 0.26 Kale et al. (1994)
Sina at Kolegaon, India 0.19 0.47 0.35 Deodhar and Kale
(1999)
Bhima at Takali, India 0.13 0.43 0.44 Deodhar and Kale
(1999)
Buff Bay at Balcarres, Jamaica 0.07 0.47 0.51 Gupta (1975)
Narmada at Rajghat, India 0.04 0.50 0.46 Kale et al. (1994)

Fig. 7.10 Idealised sketch of four channel patterns

selection of at-a-station hydraulic geometry values are shown in Table 7.2. Hydraulic geom-
etry implies that channel dimensions are adjusted to flow conditions. This is comparable to
the concept of stable alluvial channels designed by engineers in India mentioned earlier. It is
an excellent non-subjective way of describing the shape of a river channel.

7.4.4╇ Channel pattern

Rivers not only adjust their cross-sections: they also adapt a planform that is suited to their
discharge, sediment and regional slope. The planform or channel pattern has been defined
as the configuration of a single river as seen from the sky. Rivers are commonly classi-
fied as one of four types:€straight, meandering, braided and anastomosing or anabranching
(Fig.€7.10). These classes can be subdivided, based on channel measurements. The major-
ity of rivers are either meandering or braided. Many rivers display the same planform
throughout their length, although exceptions do occur. For example, a braided river may
116 Rivers in the tropics

change to a meandering pattern with altered gradient and sediment load, or a seasonal river
which braids during the dry season may take on a meandering appearance with the rising
of the water level in the wet season.
All natural rivers have an inherent tendency to follow a sinuous flow. The sinuosity of a
river channel is calculated as a ratio called the Sinuosity Index.
Sinuosity Index = Distance between two points along the channel/distance between the
same two points along the valley.

Straight rivers are rivers whose channels run straight for several times the channel width or
for which the computed sinuosity is very close to or exactly 1.
A straight flow between two points obviously gives the channel the highest possible gra-
dient, and hence these rivers flow on slopes steeper than that of other channel types in the
region. The straightness in many cases is derived from geological controls such as faults,
which determine the gradient of the channel and its ability to erode actively. For example,
the Irrawaddy in upper Myanmar changes its planform from a meandering pattern through
a wide alluvial valley to a deep straight gorge on reaching the Sagaing Fault (Fig. 7. 11).
The banks of a straight channel may run straight but the thalweg (the line joining the
deepest part of the channel) may not. It may follow a winding path between alternate bars
on either side (Fig. 7.10). Pools (deep reaches with slow flow) alternate with riffles (shal-
low reaches with fast flow) along a straight channel. Fine sediment tends to accumulate
in pools, coarse material in riffles. Measurements along straight channels have shown that
pools and riffles tend to be spaced at regular intervals, about 5–7 times the mean width of
the channel.
A meandering river is expected to have a sinuosity index of 1.5 or above, but more
importantly it should have a wandering channel with the thalweg (the deepest part of the
channel) regularly winding from one bank to another, approaching different banks alter-
nately (Fig. 7.12). Each bend is asymmetrical in cross-section with a deep pool on one side
and a crescent-shaped bar (point bar) on the other. Halfway between two successive bends
the channel is symmetrical with the thalweg in midchannel. A meandering channel tends
to migrate at bends, eroding the meander cliff side but extending the point bar by depos-
ition. Thus its original dimensions are maintained. A meandering stream thereby wanders
across its valley floor. Brice (1983) recognised two different types of meandering channel.
The first type carries approximately equal widths for each cross-section and is considered
passive, i.e. does not move much across its valley floor. The second type is mobile and
its cross-sections are wider at bends with well-developed point bars on the concave side.
Schumm (1960) has shown that a significant proportion of clay and silt in the bed material
is associated with high sinuosity of the channel. Streams with coarse bed load, however,
also meander. Schumm (1968) suggested that the morphology of a meandering channel
depends on the type of load − suspended, mixed or bed load − that it carries. The dimen-
sions of a meandering river (Fig. 7.12)€– the wavelength of the meanders, channel width
and the meander radius€– are statistically related to each other and to its bankfull discharge
(Leopold and Wolman, 1957).
Braided rivers display a number of sub-channels between the two banks of the entire river
with mid-channel bars separating such channels (Fig. 7.10). These channels are characterised
117 Channel geometry

Fig. 7.11 Diagrammatic sketch of the Irrawaddy River changing its course from a free meandering river in an alluvial floodplain
to a straight restricted gorge section as it enters the Sagaing Fault in north Myanmar. Sketched from a SPOT image on
the web and checked with topographical maps. From Gupta, 2005b. By permission of Oxford University Press

by their large width, high width–depth ratio and the frequently changing position of the
smaller channels and mid-channel bars. Leopold and Wolman (1957) concluded from field
data that for a given discharge, braided channels have steeper slopes than meandering riv-
ers, an observation that has been supported by the experimental data of Schumm and Khan
(1972). Braiding is associated with a high coarse sediment load that is transported and
deposited repeatedly, following fluctuations of discharge. They are usually associated with
steep slopes, abundant supply of coarse material and a fluctuating discharge. In such rivers,
the pattern stays the same but the locations of channels and bars shift following individual
high flows. It is therefore not surprising that braided channels have been reported from

• semi-arid regions of low relief that receive drainage from mountainous areas
• areas with seasonal rainfall
• highland areas in a variety of climates
118 Rivers in the tropics

Fig. 7.12 The geometry of a meandering river

• breaks of slope near the foot of a mountain


• glacial outwash areas
• periglacial areas, especially over permafrost.
In all these areas an abundance of coarse bed load and a fluctuating flow regime are
found.
Anastomosing (also known as anabranching) rivers (Fig. 7.10) display a multichannel
pattern where, unlike braided river channels, sub-channels and bars/islands are fixed in
position. The stable islands of an anabranching river are usually in clay or silt, reflecting
the load carried by an anabranching river. Such islands, being stable, are generally under
119 Channel network and nodes

vegetation. The individual channels of an anabranching system could be straight, meander-


ing or braided (Schumm, 1985). The channels are not only stationary; they are also deep.
Anabranching rivers are usually found in intermontane valleys, on low gradient fans, and
in lacustrine and alluvial plains, where floods cause channels to avulse through erosion-
resistance banks. Nothing striking happens between floods. A similar pattern is sometimes
seen in rock as described for the Sabie in South Africa by Niekerk et al. (1999) and the
Mekong by Gupta and Liew (2007).
The four characteristics of a river channel are interrelated and determined by the com-
bination of channel properties such as discharge, sediment, bed and bank material. For
example, braided rivers are associated with a high width–depth ratio, fluctuating discharge,
coarse and abundant sediment, and steep slopes. Variations do occur from such generalisa-
tions but only as exceptions. Schumm (2005) has summed up the discussion by dividing
rivers into two basic classes:€regime and non-regime. Rivers in regime are those described
above which tend to be stable and demonstrate the expected characteristics. Non-regime
rivers are either unstable or controlled by geological factors such as bedrock, the effect of
which surpasses all other factors.

7.5╇ Channel network and nodes

Individual channels join up to form a network that transfers water and sediment from an
area of land surface. This area is the drainage basin which is named after the main channel
in the network. The density and pattern of the network in the basin reflect various factors
such as geology, relief, hydrology and land use in the basin. In a classical paper written
almost 70 years ago, R. E. Horton explained certain rules behind the arrangement of this
network (Horton, 1945). Horton proposed an ordering system for the network which has
been modified several times since then. Figure 7.13 demonstrates Horton’s original system
and two major proposed modifications. Horton also demonstrated that the meeting of the
tributaries (nodes) and aspects of their dimensions follows certain relationships. Such rela-
tionships could be demonstrated using any of these systems.
Horton called the fingertip tributaries the first-order streams. Two first-order streams come
together to form a second-order stream, two second-order streams join to form a third-order
river and so on (Fig. 7.13a). It is a wonderful system but it requires extending the ordering of
the main stream all the way back to its source. This could be problematical as it is not always
clear which one is the major stream upstream of a confluence, especially from a map. In the
simplified system proposed by Strahler (1952a) both streams of the same order carry the
same order up to the confluence (Fig. 7.13b). Shreve (1967) added the order of individual
tributaries to the main stream, (Fig. 7.13c) so that the magnitude of the major streams can
be adequately expressed. As Horton originally demonstrated, geometric relationships exist
between the number of the tributaries, the lengths of the tributaries and the area of the basins
drained by tributaries of different orders. These relationships have been clearly discussed
by many geomorphologists and will not be reviewed here. For example, excellent reviews
have been provided by Strahler (1964) and Abrahams (1984). Primarily the idea of network
120 Rivers in the tropics

(a) (b) (c)

Fig. 7.13 The same channel network according to the ordering systems of (a) Horton, (b) Strahler and (c) Shreve

and ordering is very helpful in understanding the discharge of water and sediment along a
drainage basin. Horton’s rules work well for river basins in the tropics.
The comparative efficacies of two joining streams are important and determine the nature
of sedimentation and erosion at their confluences. Similarly the geomorphic coupling at the
contact between hillslopes and drainage channels to a certain extent determines the nature of
the landform (Harvey, 2002). Deposition of material accumulates at confluences and on foot
slopes if the main stream is not effective at the time the sediment arrives. In many cases, a
time-based accumulation of material is cleared in an episodic large event like a flood (Fig.
6.8). The landscape is a reflection of many factors including elapsed time since the last such
clearing (Fig. 7.14). The characteristics of the tropical climate act as a control factor.

7.6╇ River systems of the humid tropics

Do tropical rivers look and behave differently to those of temperate regions? Governed by
a different hydrologic environment many of them do. The form and function of tropical
streams, however, are also determined by other aspects of the drainage basins:€geology,
relief and land use. Such factors may locally override hydrology. We will examine the
characteristics of streams in the humid tropics in this section.
Rivers of the temperate regions demonstrate a relationship between bankfull discharge
and channel size and shape (Leopold and Maddock, 1953; Leopold and Wolman, 1957;
Leopold et al., 1964). This relationship can be explained by the magnitude–frequency con-
cept of Wolman and Miller (1960). The most effective force on these rivers is bankfull dis-
charge, which is powerful enough to determine the size and shape of the river and frequent
enough to maintain their dimensions. A large flood may temporarily enlarge the channel
121 River systems of the humid tropics

Fig. 7.14 Accumulation of coarse sediment at a coupling point, Tiger Leaping Gorge, Jinsha Jiang (Upper Yangtze),
Yunnan, China. From Gupta, 2011. By permission of Royal Academy for Academy for Overseas Sciences, Brussels.
Photograph: A. Gupta

but within months or years it would revert back to its original size. The river returns to
its former size with in-channel deposition from common post-flood flows carrying a sus-
pended load of silt and clay and also transporting a substantive amount of sand and small
pebbles. This has been repeatedly demonstrated in the field (Costa, 1974; Gupta and Fox,
1974). In contrast, the size of a river channel in arid areas is controlled by floods that occur
at long intervals of time (Wolman and Gerson, 1978) and in between the channel is usually
dry. The channels of arid areas are essentially flood-controlled (see Chapter 12). The rivers
in the seasonal tropics fall between these two cases.
Extreme floods commonly occur on rivers whose basins are located between about 10°
and 30° of latitude (Gupta, 2000). Such floods usually result from tropical storms occurring
in the wet season; a number of these storms are strong enough to be recognised as trop-
ical cyclones. Accounts of the resulting high shear stress and unit stream power, enhanced
stream competence, sediment transport and storage, and channel forms are now available for
a limited number of streams in Australia, South Asia and the Caribbean, and generalisations
are possible. In such floods, high stream power operating over a moderate to high duration
(Costa and O’Connor, 1995) are able to erode the bed and banks of the channel and transfer
sediment for the time period. Flood effects are preserved if the sediment is coarse and the
relief is high, as in such environments flood effects are difficult to undo (Gupta, 1988).
The geographical locations where tropical cyclone-driven floods (and smaller storms)
could be expected to affect river systems are

• North and Northeast Australia


• Parts of Southeast Asia affected by tropical cyclones, e.g. Myanmar, Viet Nam and the
Philippines
• Parts of the Indian subcontinent
• Madagascar and neighbouring coastal areas of East Africa
122 Rivers in the tropics

Fig. 7.15 Annual hydrographs of four stations on the Mekong River (Total drainage area 795 000 km2):€Chiang Saen, Thailand
(14% of the basin area) Chiang Khan, Thailand (36%), Pakse, Lao PDR (69%) and Stung Treng, Cambodia (80%). Data
from Mekong River Commission, Lower Mekong Hydrologic Yearbook, 1997. From Gupta, 2007b. By permission of Wiley

• The Caribbean Islands, the coastal areas of the Gulf of Mexico, and tropical and subtrop-
ical North America.
The effects are especially strong where storm tracks are intercepted by mountain ranges.
Taiwan in the subtropics is an excellent example.
Strong seasonality is associated with much of the humid and subhumid tropics (see dis-
cussion on monsoon rainfall in Chapter 3) irrespective of the annual total rainfall, which
may range from below 1000 to above 4000 mm. A very high percentage of the annual
rainfall (≥75 per cent) arrives in the wet season which may extend for 4–5 months. The
hydrographs of the rivers of the seasonal tropics therefore demonstrate a seasonal pattern
(Fig. 7.15). The wet-season discharge is much higher than that of the dry season. Almost all
the work of the river (erosion, sediment transport, etc.) is carried out during the wet season.
Large tropical storms arrive over these river basins episodically, usually in the wet season
when the river is already high and the soil saturated.
The effect of seasonality is also reflected in sediment transport in many tropical rivers.
In a recent publication, Billi and el Badri Ali (2010) reviewed suspended sediment trans-
fer in the Blue Nile for three years:€1967–69. Although this is a short period, they could
conclude that any significant transfer of sediment only happens during the wet monsoon
period, July to October. Their curves for suspended sediment plotted against time showed
considerable variation even within the wet period. During the three years of record, the
maximum and minimum suspended daily sediment discharges at Khartoum where the Blue
Nile meets the White Nile was 2 242 588 and 2484 t. A good correlation (R2 = 0.84) exists
123 River systems of the humid tropics

Fig. 7.16 Cross-section of the Auranga showing the 3-step physiography of the channel. From Gupta and Dutt, 1989.
(www.borntraeger-cramer.de)

between water and sediment discharges but not with sediment concentration. Peaks in sedi-
ment concentration appear to precede that in discharge. Most of the sediment is fine sand
with an unusual proportion of coarse sand being present. The clay fraction is also high,
but unlike fine sand, it tends to decrease with a rise in low discharge. Overall suspended
sediment concentration increases with flow velocity. This is a description which probably
holds true for many seasonal rivers in the monsoon tropics.
Rivers in the seasonal humid tropics use a set of channels of different size, progressively
smaller ones nesting inside the next bigger one in order to adjust to floods and seasonality.
Working in Australia, Pickup and Rieger proposed that channel forms result from a series of
discharges rather than a single dominant one (Pickup and Rieger, 1979). This nested pattern
has been reported for other rivers of the seasonal tropics. The Auranga, a fifth-order stream
that flows across the northeastern corner of the Indian Peninsula is an excellent example
(Fig. 7.16). It is a seasonal stream carrying sand and fine pebbles, dependent on the mon-
soon system of South Asia. During the dry season, the Auranga is a braided channel in sand,
and water flows only through the lowest part of its channel. Well-developed point bars occur
above the level of the dry-season braided channel. At some of the bends, even higher bars,
termed flood bars, exist, displaying an even wider channel. The dry-season braided channel,
point bars and flood bars are all inside the riverbanks marking the entire channel. During
the wet season, the river rises, fills part of the channel and meanders around point bars. The
discharge starts to fall at the end of the wet season, and the river becomes braided with small
sand bars separating channels that are only several centimetres deep. This kind of braiding
has been described in detail for the Platte River in the Great Plains of the subhumid western
United States (Smith, 1971). When high-magnitude floods associated with tropical storms
arrive in certain years, the river needs the entire cross-section that consists of the flood bars,
124 Rivers in the tropics

point bars and low channel to carry floodwaters. The channel is cleaned of sediment dur-
ing the floods but as the floods recede, flood bars are deposited at some of the bends with
the slowing down of the river velocity. Subsequently, new point bars are rebuilt in the old
places as determined by channel bends and in the dry season the braided low-flow channel
reappears below the level of point bars (Fig. 1.2c). The Auranga therefore flows through a
nested channel that is adjusted to the low flows of the dry season, the high flows of the wet
season and occasional very high floods (Gupta and Dutt, 1989).
A nested set of channel-forming discharges explains the size, shape and sedimentary
characteristics of stream channels in the seasonal tropics. These rivers have been reported
from different areas:€the Narmada in India (Gupta, et al., 1999), Yallahs in eastern Jamaica
(Gupta, 1975), the Burdekin in northeastern Australia (Wohl, 1992), and the Mekong in
Southeast Asia (Gupta and Liew, 2007). Both bed shear stress and unit stream power are
strikingly different for the low flows of the dry season, the high flows of the wet season
and high-magnitude floods.
Wohl (1992) examined the effect of large floods and highly variable discharge regime in
the seasonal tropics of northern Australia. During the dry season of April to November, the
Burdekin carries a discharge of 60 m3s−1 and transports very fine silt and clay. The discharge
rises to an average of 1260 m3s−1 during the wet season, and the sand and silt transport increases
significantly. The sand comes from the weathering of granitic rocks on the hills and plateaux
of the upper catchment. High-magnitude floods with a discharge of 15 000–30 000 m3s−1 are
related to certain aspects of the morphology of the rocky gorge of the Burdekin, e.g. location
of boulder bars and flood levees. These features form where the shear stress and unit stream
power of the Burdekin are reduced due to channel widening and the passing high-magnitude
floods tend to deposit some of the sediment. Only such huge floods have the competence
to move the coarse material of the boulder bars of the Burdekin. Fairly frequent wet-season
floods are capable of transporting a number of boulders found on bars, but extreme floods con-
trol the location of boulder bars which lesser floods may only rework slightly.
A channel-in-channel topography is common in these rivers and the overall channel
shape is box-like with high banks. The width–depth ratio of the channel decreases strik-
ingly with the arrival of high-magnitude floods that use the entire channel. Quantities of
very coarse sediment are moved during high-magnitude floods which are capable of dra-
matic erosion. The Narmada at Dardi, central India (Fig. 7.7) excavates and transports up to
2-metre cuboid blocks of orthoquartzite, stacking them like upstream-imbricated cushions
(Gupta et al., 1999). In flood, the river probably clears out all the sediment in the channel
except such huge boulders. Subsequent high flows of the wet season rebuild an inner chan-
nel by depositing sand and pebbles around boulders left by the high-magnitude floods at
appropriate locations in the big channel. This leads to the channel forms, even floodplains,
being inset within the bordering high banks that bound the high-magnitude floods. This
happens both in rock gorges and alluvial sections of the Narmada; except in rock gorges,
the deposited material is seldom smaller than pebbles, whereas in wider alluvial sections
sand builds up kilometre-scale point bars with megaripples on top (Fig. 7.17). Clearing out
the channel of all but very coarse sediment in hurricane floods and rebuilding depositional
features after the passage of the flood peak also occurs in the Caribbean mountain streams
(Gupta, 1975).
125 River systems of the humid tropics

Fig. 7.17 Sequence of channel form construction in the Narmada River. From Gupta et al., 1999. By permission of Wiley

In the case of the Narmada, which flows inside 10–15 m high banks, no floodplain
exists above the channel. A discontinuous and temporary floodplain-like feature, built of
silt and clay, is seen inside the banks in places. In the case of large rivers like the Ganga,
as described in Chapter 9, a shallow wide braided channel is flanked by riverbanks that are
filled during the wet monsoon, with a higher floodplain surface and flood banks beyond.
The entire cross-section is submerged during high-magnitude floods. Wolman and Miller
(1960), while opting for bankfull discharge, stated that as the threshold stress increases
catastrophic events become more important in shaping a channel.
Variations from the general picture occur if seasonality in flow is not pronounced or
the river carries only fine-textured sediment that can also be moved by flows of high fre-
quency. Local geological characteristics too may override the described pattern by forcing
the river to flow through sagged depressions or fault-guided gorges or across hard rock
barriers. Bedrock rivers also are different. They are modified only during floods, although
as bedrock sections are usually steep, these sections are not associated with significant
126 Rivers in the tropics

Fig. 7.18 Kaveri River, Srirangapatnam, Karnataka, India. Note structural control of the riverbed topography. Photograph A. Gupta

sediment deposition, almost all sediment being flushed out downstream. Bedrock reaches
may have an inner gorge and shoulders in bare rock (Fig. 7.7) but the size of the bars will be
small and they tend to be in material coarser than sand. Sediment may also build up in the
channel behind cross-channel barriers in bedrock. Thus the effect of geological structures
and tectonics may be locally important, overriding the hydrological component. The river
may be fault-guided like the Irrawaddy passing through the Sagaing Gorge (Fig. 7.11), but
even inside the channel, smaller faults and joints control channel relief, well illustrated
during low flow conditions (Fig. 7.18). Rivers in the seasonal tropics have their own style
of adjusting to the magnitude–frequency concept.
With the onset of aridity, rivers in basins of the subhumid and arid tropics tend to have
longer periods of low flow or no flow. Most of the work is carried out during a brief wet season
or the periodic passage of short-duration floods. Not many of these rivers have been studied.
In the arid tropics, the form and behaviour of rivers depend solely on the passage of floods of
varying magnitude and frequency. An account of such rivers is given in Chapter 12.
A number of generalisations can be made regarding the rivers in the seasonal tropics:

• A very high percentage of annual rainfall (75 per cent or more) occurs in the wet season,
which may extend up to 4–5 months.
• Large storms, when they occur, usually arrive in the wet season when the river is already
high and the soil saturated.
• It is difficult to measure discharge and sediment transport in high-magnitude floods.
• Calculated bed shear stress and unit stream power show striking differences among (a)
low flows during the dry season, (b) high flows during the wet season and (c) high-
magnitude floods.
• A large number of rivers in the seasonal tropics tend to braid during the dry season, but
not so in the wet period when they may have a meandering appearance.
• A channel-in-channel topography is common.
• The overall channel shape is box-like with high banks.
127 Questions

• The width–depth ratio of the channel decreases strikingly with the arrival of high-mag-
nitude floods when the entire channel is in use.
• Quantities of very coarse sediment can be moved during the high-magnitude floods.
• Readjustment of part of the sediment occurs during the high flows of the wet season.
• Channel forms are usually inset within the bordering high banks associated with high-
magnitude floods.
• Variations from the general trend may occur depending on the strength of seasonality
and the texture of sediment being transported. This in turn depends on basin location and
modification.
• Bedrock channels and alluvial channels look different and behave differently.
• The effect of geological structures and neotectonics may be locally important.
Rivers in the seasonal tropics have their own style of adjusting to the magnitude–frequency
concept. Their forms and behaviour should also be remembered in conceptualising and
managing the tropical environment.

Questions

1. Compute the discharge of a stream with 100 m width, 1.5 m mean depth and 0.0002
slope, given that n is (a) 0.02, (b) 0.035, and (c) 0.05.
What effect does a changing n have on velocity and discharge?
2. What kind of a river would you expect to go with each of the n values?
3. A river flows through a 2 km gorge. After a large flood on the river, the elevation of the
flood levels marked on the rock walls was 500 m at the beginning of the gorge and 480
m at the end of it. The average depth of the flood was 8 m in the gorge. Calculate the
shear stress and unit stream power for the flood in the gorge. Assume n = 0.03 and use
γ = 9807 Nm−3.
4. When you finally came to the gorge, the flood has abated and the water has gone down
to an average depth of 1.5 m. What would be the shear stress and unit stream power?
You have to guess a new n.
5. When the suspended load in a flood is plotted against discharge as in Fig. 7.3, the ris-
ing limb of the graph shows more sediment concentration than the falling limb. Why?
Under what conditions could the opposite be found?
6. Bed load is usually not measured for most streams. Why?
Explain under what conditions the bed load of a river would contribute a significant
proportion to the total load.
7. (a) The river Mekong passes through a rocky gorge upstream of the Lao capital of
Vientiane. Calculate the shear stress and unit stream power for a flood of 22 900 m3s−1
in the gorge.
γ = 9807 Nm−3, depth = 10 m, slope = 0.002 and width = 335m.
8. What type of sediment can the Mekong move in the gorge? Williams (1983) found that a
1.5 m boulder can be moved under conditions of at least 1063 Wm−2 unit stream power, 255
Nm−2 shear stress and 2.52 ms−1 mean velocity. Use these figures as general indicators.
128 Rivers in the tropics

9. The Mekong is wider and shallower at Vientiane, where the river flows through an
alluvial reach. Would the character of the sediment change between the gorge and
Vientiane?
If no tributary flows into the river between the gorge and Vientiane, and the height of
the flood at Vientiane is 3 m and the slope 0.0015, what is the width of the flood? You
may need to assume n from a table.
Is flooding extensive in the alluvial reach?
10. You are standing in the midstream of a meandering river looking at the cutbanks. The
river mostly carries silt, clay and fine sand as suspended load. In flood it also transports
large quantities of coarse sand and pebbles. What kind of stratigraphy would you see
in the cutbanks? Try the most likely scenario.
11. Most of the sediment in a river comes from its headwaters in the mountains. The val-
leys in the mountains are also areas where dams are constructed. How would a river
adjust itself after a dam is built across its main channel in the mountains? Answer this
question before you read Chapter 17.
12. A graded profile of a river is a smooth profile.
Is this statement correct? Justify your answer.
13. You have two rivers with the following at-a-station hydraulic geometry measures.
(a)╇ b = 0.44, f = 0.29, m = 0.27
(b)╇ b = 0.14, f = 0.45, m = 0.41
What would the two rivers look like? What else can you say about them?
14. Figure 7.11 shows the Irrawaddy in upper Myanmar entering a fault-guided gorge.
What transformation does the river undergo at the head of the gorge? What transform-
ation should you expect as the river emerges from the gorge to an alluvial valley again?
This transformation is not in the figure.
15. Which channel pattern would have islands with better-established vegetation and
why?
(a) a braided stream (b) an anabranching stream
16. How are the four characteristics of a river channel (slope, size, shape and channel pat-
tern) interrelated?
17. Find a topographical sheet of your area with contours and at a scale of at least 1:63â•›360.
Trace the drainage lines of a river basin from the map to a sheet of tracing paper. Mark
the drainage lines according to Strahler. Tabulate the number of streams for each order.
Divide the total number of streams of each order by the total number of the next higher
order. This is the bifurcation ratio. What range of ratio do you get for the basin? What
is the average of all such ratios?
18. (a) In which part of the humid tropics would you expect nested channels?
(b) Under what environmental conditions in these areas would nested channels not
occur?
19. How do the rivers of seasonal humid tropics adjust to the magnitude–frequency con-
cept of Wolman and Miller (1960)?
8 Alluvial valleys

Of golden sands, and crystal brooks


John Donne

8.1╇ Fluvial depositional environment

A river tends to deposit sediment in certain locations in a valley. Intuitively we expect sedi-
ment to be deposited as stream velocity drops, starting with the coarsest fraction. When that
happens, the preferred locations for depositing sediment are

• on floors of valleys in the mountains, frequently near confluences or barriers in the


channel
• at the foot of the mountains or plateaux (alluvial fans)
• along an alluvial valley in the plains (channel bars, floodplains and terraces)
• deltas at river mouths.
The valley floors in the mountains and alluvial fans are discussed later in Chapter 13 and
the deltas in Chapter 11. Here we review alluvial deposition in plains. The current alluvia-
tion in a river valley happens on top of an earlier sequence, which is mostly bedrock in the
mountains and the upper course of the river but could be an older layer of alluvium in lower
course. Meade described alluviation in the largest river basin in the world, the Amazon, as
follows:
As they leave their narrow mountain gorges, the streams coming off the Andes flow
into river beds that are self-made. Massive amounts of sediment brought down from
the mountains become the substrates over which and through which the flowing
waters, with their accompanying load of even more sediment, must make their ways.
Once off the Andean slopes and out of the confinement of mountain-girt channels,
the riverborne sediment is likely to endure many episodes of confinement and occupy
many rest stops and storage compartments before it reaches its ultimate destination.
(Meade, 2007:€46)

This is for the largest river in the world, but the pattern remains the same for most rivers
when scaled down appropriately.
The steep and possibly tectonic upper parts of a river system carry high stream powers,
and erosion and transfer of sediment is common (Blum, 2007). The depositional forms
in the highlands are fragmentary and sediment grains tend to have a brief residence

129
130 Alluvial valleys

Fig. 8.1 Source-to-sink passage of a river

time. The situation changes in the upper plains where the stream gradient is low and
valleys are usually described as mixed bedrock alluvial valleys. The alluvial forms such
as floodplains become continuous, sediment grains have longer residence time, and the
entire valley acts as a conveyor belt of sediment being transported to the sea. Alluvial
valleys and channels entirely in alluvium occur further downstream near coastal plains
and deltas (Fig. 8.1).
We examine the material, forms and sediment transfer in alluvial valleys in the follow-
ing sections.

8.2╇ The alluvial valley

In the upper part of their course in the plains beyond the mountains, rivers flow in a
mixed bedrock alluvial valley. Downstream, alluvium in the valley thickens and bed-
rock is rarely visible. The valley alluvium is a three-dimensional body which fills a
linear depression. This depression could be erosional, cut into rock below the surface
alluvium, or a tectonic downwarp. It is difficult to visualise the base of the alluvium
unless a number of boreholes are sunk or a geophysical survey is carried out which
allows mapping of the rock floor (Schumm, 1977). It seems likely, from the limited
amount of data available, that (1) the rock base is usually uneven and (2) a series of
alluvial deposits rest on the bedrock, not all of which have necessarily been deposited
by the current river.
Figure 8.2 is a generalised diagram of a valley cross-section. It indicates that the upper
level of the alluvium may not reflect all the morphology at the base. The present river
may not be located directly over the deepest part of the valley. There may be an older
alluvium below the alluvium of the present river. Past hydrology and river geomorphol-
ogy can be derived from the texture and sedimentary structure of the past alluvium.
Its nature and organic or carbonate contents may provide information about the past
131 The alluvial valley

Fig. 8.2 Generalised diagram of a valley cross-section

Fig. 8.3 Rock ribs in the Mekong, downstream of Chiang Saen, northern Thailand. Photograph: A. Gupta

climate. Such indicators are known as climate proxies, and they have been heavily used
for studying palaeoclimate during the Quaternary (see Chapter 16). The alluvium of the
current river lies on top and is reworked from time to time in high flows. The depth to
the rock base and its geological structure may still influence the form and function of
the present river. This has been demonstrated through the thick alluvium of the Amazon
River (Mertes and Dunne, 2007). Several structural highs cross the Amazon valley below
the thick alluvium. Where the river flows over such rock arches, the river narrows in
spite of the thick alluvium, its gradient steepens, and there is less deposition in the chan-
nel. After crossing a structural high, the Amazon widens and sediment is deposited in the
channel (Dunne et€al., 1998).
Where alluvium is thin or nearly absent, the older layers and the bedrock are visible.
This is common for rivers near the mountain front. Given the regional geology, bedrock at
shallow depth controls the form of the channel or even outcrops to form rapids and a line
of rocks across the channel, known as rock ribs (Fig. 8.3). This is common even in a large
river like the Mekong for the first 4000 km of its course (Gupta and Liew, 2007).
Alluvial morphological features that fill a river valley can be listed (Table 8.1), although
not every feature is present in all valleys.
132 Alluvial valleys

Table 8.1╇ Alluvial valley forms and sediment characteristics


Location of deposition Terms Characteristics

In channel Alluvium resting on Bed load in storage between trans-


bed port; period of storage usually
directly related to texture
Fills in abandoned or aggraded
channels
In channel but higher forms Point bars Point bars and marginal bars are
Marginal bars usually formed by channel
Midchannel bars shifting or found below bank col-
lapses. Midchannel bars of vari-
ous size are formed in braided
and anabranching streams
Overbank Floodplain Formed by river shifting (lateral
accretion) or overbank floods
(vertical accretion); in braided
streams by merging of midchan-
nel bars and islands
Back of floodplain Terrace Abandoned floodplain at a higher
level
Valley margins Alluvial fans and screes Sloping triangular bodies formed
by deposition by headwaters,
�tributaries or mass movements

8.3╇ The channel alluvium

The bed of a river is commonly floored with coarse alluvium. This alluvium is referred
to as bed material, the gravelly (pebbles, cobbles, boulders) part of it being carried as bed
load and the finer part (sand) as both bed load and suspended load. If the bed material is
coarse, gravel accumulates at intervals forming shallow riffles and rapids. They are sepa-
rated by deep pools where finer material accumulates. In tropical rivers that experience
high-�magnitude floods, such accumulations of coarse gravels probably start by deposition
of boulders in places dictated by channel geometry or exposure of bedrocks at the fall-
ing stage of the flood hydrograph. Cobbles, pebbles and sand then accumulate round the
boulders and rock features to complete the riffles. Sand is deposited in the pools later and
finer suspended material is transported downstream.
This type of deposition, which requires availability and transportation of very coarse
material, is usually seen in the upper river courses near the mountains, downstream of a
rocky gorge between two alluvial reaches, and where a tributary contributes coarse mater-
ial to the main stream. Sand is probably the commonest material deposited in an alluvial
channel. The deposited sand may mask the channel base or form bars of different types.
133 The channel alluvium

Box 8.1 Critical, subcritical and supercritical flows


The velocity of a surface wave moving over water is expressed as √(gy) where g is acceleration due to gravity
and y is the vertical distance to the water surface. At critical flow, water moves at the same velocity as that
of a surface wave moving relative to the flowing water. When the velocity of the flow is less than critical, it is
considered as subcritical flow. Similarly, supercritical flow occurs when the channel velocity is higher than the
critical velocity. The velocity of a river almost always stays subcritical.
Imagine a surface wave being created by a piece of rock falling into the river. At subcritical flow, waves from
this disturbance would move both upstream and downstream from the point of impact. We are familiar with
this scenario. If the flow of the river is critical, the upstream point of the wave would appear to be stationary
to an observer standing on the bank who would also see the downstream front moving at twice the velocity
of the river.
If the river flows very fast and the flow is supercritical, then the entire disturbance from the falling rock
will move downstream and no wave front will propagate upstream of the point of impact. This happens rarely,
only during very high floods or at waterfalls. An analogue can be drawn with subsonic and supersonic effects
with the velocity of sound replacing critical flow.
The ratio of stream velocity to wave velocity,√(gy), is the Froude Number (Fr). If Fr < 1, flow is subcritical;
if Fr = 1, flow is critical; and if Fr > 1, flow is supercritical.
A river is almost always at a subcritical stage when sandy ripples and dunes are created on the bed of the
channel. Plane beds are formed when the velocity of the river approaches the critical mark. Antidunes form
rarely, only at the supercritical stage.

Sand is easily eroded and transferred (Chapter 7), and the surface features and the internal
structure of a sand body deposited in the channel reflect the nature of the depositing flow.
Sediment accumulates on the channel bed as discrete grains or aggregates. The coarser
variety (sand-boulder) behaves as discrete grains. Sand and pebbles may form bars or a
layer of sediment on the channel bed. Pebbles tend to imbricate against each other with
individual pebbles sloping upstream. In extreme floods, as described for the Narmada in
Chapter 7, such a characteristic is extended even to metre-scale boulders (Fig. 7.7). Pebbles
and coarser material may lie buried under sand.
Aggregates of sand and silt create bedforms under flowing water. Such bedforms appear
as asymmetrical or symmetrical waves and vary in size from ripples (in centimetres) to
large dunes (tens of metres). Materials coarser than sand are usually found in bars or
forming riffles as described in Chapter 7. Three factors determine the types and size of
bedforms:

• flow velocity
• flow viscosity
• size distribution of bed material.
Bedforms of different types depend on the velocity of flow at which they are built.
They are grouped into lower and upper regime forms and are associated with critical
and supercritical flows (Box 8.1). Each type is associated with a characteristic mode
134 Alluvial valleys

Table 8.2╇ Types of bedform in sand and silt


Relative sediment
Type of flow Type of bedform Manning’s n transport

Lower flow regime Ripples 0.018–0.030 Small


Subcritical flow, Fr < 1 Dunes with superposed
ripples
Dunes 0.020–0.040
Upper flow regime Plane bed 0.010–0.013 Large
Critical flow, Fr = 1
Upper flow regime Antidunes:€Standing 0.010–0.012 Large
Supercritical flow, Fr > 1 waves
Antidunes:€Breaking 0.012–0.018
waves
Chutes and pools 0.018–0.030

Note:€Values of Manning’s n from Chow, 1964

of (a) sediment transport, (b) flow resistance and (c) amount of material transported.
Different bedforms may occur next to each other in the same cross-section of the channel
or replace each other over time as the discharge and velocity of the river change cross-
channel (Table 8.2).
In the lower flow regime, ripples and dunes cause high resistance to flow primarily by
form roughness. The sediment transfer is carried out by individual grains moving up the
back of ripples or dunes and sliding down the steep front face. The surface of water is out of
synchronisation with the lower regime bedforms. This creates a zone of separation down-
stream of the crests resulting in surface eddies (Fig. 8.4).
In contrast, resistance to flow is small and sediment transport large in the upper flow
regime, which is commonly seen when rivers are in flood. As the water velocity rises,
dunes are replaced by plane beds, although both dunes and plane beds may occur simultan-
eously in a channel. Antidunes are seen in extreme cases with very high flood velocities.
Chutes and pools are rare and only seen in waterfalls and rapids. The resistance to flow is
due not to form roughness but grain roughness. In plane beds, grains form a sheet which
has the thickness of a couple of grains, and continuously roll downstream. In antidunes,
in spite of sediment grains moving downstream, the forms move upstream. When break-
ing antidunes occur, much bed material is briefly suspended in water and then is partially
stored on the bed. The surface of the water stays in phase with the bedforms (Simons and
Richardson, 1971).
Such bedforms are rarely seen in gravel, except as poorly developed dunes and ripples.
The commonest form is a flat gravel bar, which is often seen after the passage of a flood
which has winnowed out all the fine material. Individual pebbles, cobbles or boulders on
this bar may be imbricated upstream.
135 Bars

Fig. 8.4 Alluvial bedforms and the nature of the flow in the river. After Simons and Richardson, 1971

8.4╇ Bars

Bars are transitional channel-features built by the deposition of sediment carried by the stream.
They are located at a level higher than the bed but lower than the bank top, and are exposed
during periods of low flow. The tops of the bars are therefore visible most of the time.
136 Alluvial valleys

The location, size, shape and structure of bars depend on a range of channel proper-
ties including channel geometry, gradient, variation in flow, and volume and texture of
sediment. Different types of bar therefore tend to be associated with variations in chan-
nel pattern. Alternate bar are seen in straight channels, located opposite channel pools.
Meandering rivers display point bars (as discussed in Chapter 7) on alternate sides, fol-
lowing meander bends. Braided streams mainly have bars in mid-channel but, in many
locations, a scouring of the banks and widening of the channel also lead to the deposition of
bars attached to a bank. The location of the bars in a rock-cut channel may also be associ-
ated with a rock protrusion on the floor or the accumulation of a pile of boulders forming
the core of the bar and accelerated sedimentation in finer material round it.
Bars are usually formed at the falling stage of a hydrograph and carry the history of
their formation in the texture and structure of their sediment. They also reflect post-
�depositional changes in the form of shallow flood channels or sand ripples on their top
surface. As expected, bars are common in rivers where the discharge fluctuates and sedi-
ment is available, often from the channel and the floodplain. During the dry season, the
beds of many seasonal rivers are an assemblage of barforms with one or several narrow
channels separating them. Such bars may disappear entirely during the wet season to
reappear in a renovated form in the next dry season.

8.5╇ Floodplain

A floodplain is a strip of land with very little relief that borders a river and is inundated at
times of high water. As the extent of the inundation depends on the size of the flood, flood-
plains are described according to flood probability, for example, an n-year floodplain. The
rarer the flood, the more extensive is the floodplain.
Floodplains comprise a number of distinctive features:

• the river channel


• point bars
• scroll bars
• depressions and risers on bar tops
• oxbow lakes (cut off sections of a meander bend)
• natural levees
• crevasse splays
• backswamps.
The first four of these features generally occur inside the channel of the river. Abandoned
and degraded portions of the former channel with these features may occur away from the
river’s present position to form part of the floodplain. Levees are raised linear features that
run parallel to rivers. As a river overflows, hydraulic conditions change at the top of the
banks and active sedimentation in fine sand, silt and clay takes place, building levees along
its channel. Levees are generally the highest feature in a floodplain, disrupting drainage to
and from the river, and thus swampy conditions form over the rest of the floodplain. These
are the backswamps that cover most of the floodplain. They are bounded on the side close
137 Floodplain

to the river by levees and on the other side by hillslopes and small screes or alluvial fans
of tributary streams. The smaller tributary streams could be blocked by a levee, forcing
them to flow parallel to the main stream for some distance until they could merge. Such
tributaries are known as yazoo streams, the term originating from the valley of the Lower
Mississippi. Such streams are common on the wide floodplains of large rivers, as on the
north bank of the Ganga between its confluences with the Gandak and Kosi.
The presence of levees on both banks also causes the main river to deposit sediment inside
the channel. Given high sediment availability and an efficient levee system, this may cause
the bed of the river to rise and, in extreme cases, to reach a level higher than the backswamps.
A collapse of part of the levee in a flood at this stage is a disaster leading to widespread and
fast inundation of the backswamps. Such a widespread levee collapse happens rarely. Instead
cracks in the levee may be enlarged in floods and sediment-laden floodwater may erode its
way through. As a result, a cone of sediment often breaks through the levee to spread down
the slope towards the backswamps. The cracks are called crevasses and the cones are called
crevasse splays. A number of crevasse splays occur along the levee of a large river. They are
identifiable by their relative coarseness, as most of the crevasse splay material is sand which
spreads over the silty and clayey floodplain. Coleman (1969) described the crevasse splays
of the Brahmaputra (locally known as the Jamuna) in Bangladesh.
Most of the floodplain, therefore, is dissected by tributaries and abandoned and overflow
channels of the main river. The low gradient across the floodplain creates an environment
of stagnant or slow-flowing water and deposition of fine material. A number of these aban-
doned channels, especially the cutoffs, tend to be filled with clay.
Floodplains can be formed in three ways:

• lateral accretion from movement of meandering rivers


• vertical accretion from overland flows
• island formation and their consolidation and channel abandonment by braided rivers.
The classical model of floodplain formation by lateral movement and building of point bars
was illustrated by Leopold and Wolman (1957) and Leopold (1973), who for years moni-
tored surveyed sections across a small stream, the Watts Branch in Maryland, USA. In this
model, a meandering stream moves laterally across its valleyflat, eroding on the convex
side and building point bars on the concave bank. Thus over time the entire floodplain is
built and the river maintains its dimensions as it shifts.
An account of vertical accretion, the second process, is provided by Schumm and Lichty
(1963), who recorded the reconstruction of a floodplain previously eroded in a large flood
on the Cimarron River, Kansas, USA.
Many floodplains are perhaps formed by a combination of both processes, although the
lateral movement of a meandering channel usually forms most of them. This is reflected
in the sedimentary structure of the floodplain, which has a layer of gravel at the base (bed-
load), a layer of sand (some cross-bedded) on the gravel (indicating lateral movement and
point bar formation) and a thin layer of fine sediment on top (overbank deposition). This,
however, is a simplified description. Floodplain sediment also includes deposition from
crevasse splays or abandoned channels, etc. on a minor scale. Braided rivers build flood-
plains by consolidating the mid-channel islands by filling the intervening channels as the
stream shifts.
138 Alluvial valleys

Fig. 8.5 Narmada floodplain inside the high banks of the large flood channel in alluvium. Note mudcracks in the clay.
Photograph: A. Gupta

Dietrich et al. (1999) summarised the floodplain deposition of the Fly River in the wet
equatorial hilly environment of Papua New Guinea as due to (1) advection of sediment
with overbank flows, (2) lateral diffusion from sediment-rich water of the Fly River and
(3) sediment-laden water travelling upstream into tributaries or small floodplain channels.
In general, the Fly has a low rate of channel migration and a limited current volume of
sediment deposition in the floodplain. Dunne et al. (1998) also mentioned the transfer of
sediment-laden water up secondary channels into the floodplains filling water bodies.
A floodplain is a place for storage. A grain of sediment is stored in the floodplain for
years until the main channel or a tributary comes across, erodes the alluvial fill and picks
up the same grain of sediment. In certain rivers, lateral movement of sediment may be com-
parable in volume and speed to downstream transfers. Descriptions of such movements
have been provided for the Amazon by Dunne, et al. (1998) and Mertes and Dunne (2007).
Sediment grains move downstream in a number of steps, interrupted by long periods of
storage. It may take a sediment grain centuries, even millennia, to reach the end of the jour-
ney at the mouth of the river (Leopold et al., 1964; Madej and Ozaki, 1996). In the humid
tropics, this long storage results in chemical weathering of the stored grains of sediment,
ending in the formation of clay minerals and a proportional increase in quartz. The sedi-
ment that reaches the sea is compositionally not the same sediment that came off the moun-
tains. Meade (2007) discussed this for the Amazon and Singh (2007a) for the Ganga.
Rivers in the seasonal humid tropics that are also affected by large floods have been
described in Chapter 7. Such rivers have a nested appearance with high banks enclosing
the entire channel required to transfer the discharge of high-magnitude floods. As the river
cannot normally climb out of the high banks, deposition related to annual high flows occur
between them. Figure 8.5 illustrates a floodplain built by clay and silt within the high banks
of the Narmada River. A similar pattern of deposition, except in coarse material, is shown
for the Yallahs River at Mavis Bank in the Blue Mountains of Jamaica (Fig. 8.6). The
139 Terrace

Fig. 8.6 Valley of the Yallahs River, Mahogany Vale, Blue Mountains, Jamaica. Note deposition of coarse sediment derived from
upper basin. Photograph: A. Gupta

photograph was taken in 1971:€seven and a half years after Hurricane Flora (1963) caused
tremendous flooding in these rivers. The channel was cleared of all except perhaps very
large boulders. Some of the coarse material was deposited to form a high terrace (under
young trees). The Yallahs then rebuilt its floodplain and a smaller channel as required by
its common pattern of discharge within the big channel. Its valley now demonstrates a river
channel, a floodplain and a hurricane terrace (or a hurricane floodplain) in coarse material.
This is the channel-in-channel physiography.

8.6╇ Terrace

Terraces are fragments of former floodplains which are no longer flooded as they are cur-
rently at a level higher than the present floodplains. Terraces are generally located away
from the river, behind floodplains, rising sharply from the floodplain level. The rise (also
called the scarp) and the flat top (termed tread) together constitute a terrace. Terraces are
formed when a river downcuts, lowering its bed to a level from which the old floodplain
cannot be inundated. The old floodplain becomes dissected and fragments of it survive as
a terrace; a new floodplain is deposited below it and above the riverbed.
A river lowers its bed elevation due to several causes. A drop in sea level, or a lake level
which was operating as the base level of a river, would expedite downcutting. This leads to
terrace formation along the lower course of a river. These are eustatic terraces. In tectonic-
ally active areas, almost always in the mountains, tectonic uplifts steepen the river gradient,
contributing higher energy for deepening its channel and forming tectonic terraces. Flights
of such terraces are seen where a river emerges from a gorge section in a tectonically active
140 Alluvial valleys

Fig. 8.7 Terraces along the Jinsha Jiang, southwestern China; the river is emerging from the Tiger Leaping Gorge.
Photograph: A. Gupta

area (Fig. 8.7). Climate changes during the Quaternary have been suggested to explain the
formation of terraces in alluvial material by the downcutting of the Upper Godavari and
Krishna Rivers in southwestern India (Kale and Rajaguru, 1987).
Extensive low terraces may occur in the valley of large rivers. Singh (2007a) has
described two large alluvial terraces above the floodplain of the Ganga in its middle course.
The first terrace (termed T1) is at a level 5–10 m above the floodplain and the next one (T2)
is 5–10 m higher. Abandoned channels are common on terrace surfaces.
Multiple terraces are seen following repeated tectonic movements or negative base level
changes. Such terraces will occur on both sides of a river and could be paired or unpaired
and at alternate heights. Terraces at approximately the same level on both sides of a river are
known as paired terraces. These are formed when the river downcuts rapidly. If the terraces
are not at the same level on either side, they are called unpaired terraces. These are formed
with a slow downcutting, giving the river time to meander and destroy the old floodplain on
both sides alternately (Fig. 8.8). Multiple terraces, between four and six in number, occur in
the tectonically active valleys of the Himalayan headwaters of the Ganga River (Khan et al.,
1982). Such terraces have also been reported from the Spiti Valley, Himachal Pradesh, India
(Sah and Virdi, 1997). Multiple terraces are quite common in tectonic mountains.
Terraces can be entirely alluvial or underlain by hard rocks, leading to their identifica-
tion as alluvial or bedrock (strath) terraces.

8.7╇ Valley margins

The edge of an alluvial valley is usually marked by hillslopes, small screes from mass
movements and alluvial fans (Chapter 13) built by tributary streams. Unless the hills are
141 Questions

Fig. 8.8 Paired and unpaired terraces

tectonically active and/or drained by an eroding tributary, not much sediment is transferred
across the valley to the main channel. Instead, the transfer of sediment happens down-
stream along the main stream and also laterally into its floodplain from the main channel.

8.8╇ Sediment transfer along the valley axis

Changes in sediment deposited along a valley axis are related to the size of the river. A long
river will show the difference between the steep upper reaches, the low-gradient mixed
bedrock-alluvial valley and the true alluvial valley near the coast. However, as large riv-
ers (a) tend to be polyzonal (different parts of the basin contributing water and sediment
via tributaries at different volumes and rates); (b) flow across varying geology; and (c) are
long-lived, they carry a history of adjustment to climate changes from the past, the final
pattern may be complicated. There should be an overall trend but local deviations from it
should be expected. The sediment in channel bars, floodplains and terraces reflect all this
(Blum and Törnqvist, 2000; Blum, 2007).

Questions

1. Sand is a very common material deposited in an alluvial channel.


Do you agree? Why?
142 Alluvial valleys

Fig. 8.9 Flood probability analysis of Baitarani River at Akhuapada, Orissa, India. Plotted from data in Ahuja and Majumdar,
(1959). From Gupta, 1988, by permission of Wiley.

2. What are the effects of the rock ribs in the Mekong as shown in Figure 8.3 on sediment
accumulation in the channel?
3. Can you explain the sedimentary structures shown on the side of the bar in the Jamuna
(Fig. 9.11)?
4. Define a 50-year floodplain. Can two 50-year floods occur in successive years?
5. Why is there a progressive increase downstream in the proportion of quartz grains in
river sand?
6. Identify the alluvial valley features in Figure 8.6. Rank such features in the order of
length of survival.
7. What is a polyzonal river system? Provide several examples.
8. Figure 8.9 displays the flood potential of the Baitarani River at Akhuapada, Orissa,
India. The data run from 1874 to 1957 (Ahuja and Majumdar, 1959). Gumbel Extreme
Value Distribution has been used to plot the regression line.
(a)╇Using the fitted line, find the flood discharge for the recurrence intervals:€1.1, 1.5,
2.33, 5, 10, 50 years.
(b)╇Would it be acceptable to use this diagram to determine the 100-year flood? The
500-year flood? Justify your answer.
(c)╇Calculate the unit discharges for the same floods. Unit discharges are calculated
by dividing the total discharge by the drainage area. Upstream drainage area at this
station is 11 360 km2. Can you identify any trend in the unit discharge figures?
9 Large rivers in the tropics

A huge river is the Padma, just like the sea


Buddhadev Bose,
translated from Bengali

9.1╇ Introduction

A number of large rivers drain the tropics, acting as huge conduits for transferring water
and sediment to the oceans. These include the Amazon, Orinoco, Magdalena and Paraná
in tropical South America; the Nile, Niger, Congo and Zambezi in Africa; and the Indus,
Ganga, Brahmaputra, Irrawaddy and Mekong in tropical South and Southeast Asia. These
rivers play a very important role in tropical geomorphology. We review a selected number
of large rivers in this chapter to illustrate their importance.

9.2╇ Characteristics of a large river

A proper definition of a large river is elusive. Potter (1978b) listed four essential properties
for large rivers:€size of the drainage basin, length of the main river, magnitude of water
discharge and volume of sediment transported. He then listed 50 largest rivers that col-
lectively drain about 47 per cent of the land surface, excluding Greenland and Antarctica.
All except one of these rivers are more than 103 km long and the smallest drainage basin
is 105 km2. Good data on the other two properties, discharges of water and sediment, are
not available for all major rivers, especially regarding their bed load. It is very difficult to
measure the bed load of a large river. Table 9.1 lists selected characteristics of the rivers of
the tropics that are longer than 2000 km.
Meade (1996) provided two lists, one of which ranked the top 25 rivers with the biggest
discharge and the other included the top 25 with the largest suspended sediment load. The
two lists do not coincide, as certain rivers such as the Zambezi carry a large water discharge
but a low sediment load. Sediment also accumulates behind impoundments in several riv-
ers with only a fraction reaching the sea. Such a river no longer functions as a natural
conduit controlling a continental-scale drainage basin. The Nile below the Aswan Dam is a
good example. It is expected, however, that a large river is long, drains a large basin, carries
a very big discharge and commonly, but not always, transports a large volume of sediment.
The great length of the rivers allows them to flow across a range of environments. For
143
144 Large rivers in the tropics

Table 9.1╇ Selected properties of large rivers in the tropics (ranked according to discharge)
Annual average Annual average
water discharge Length Drainage area suspended sediment
River (109 m3) (km) (106 km2) load (106 t)

Amazon 6300 6000 5.9 1000–1300


Congo 1250 4370 3.75 43
Orinoco 1200 770 1.1 150
Ganga– 870 B-2900 G-1.06 900–1200
Brahmaputra G-2525 B-0.63
Mekong 470 4880 0.79 150–170
Paraná-Uruguay 470 3965 2.6 100
Irrawaddy 430 2010 0.41 260
Zhujiang 300 2197 0.41 80
Salween 300 2820 0.27 about 100
Indus 240 3000 0.97 50
Magdalena 240 1540 0.26 220
Zambezi 220 2575 1.32 20
Niger 190 4100 2.27 40

Note:€Drainage areas are rounded off to 106 km2 to avoid discrepancies between various
sources. Figures from Gupta 2007a:€Table 1.1. For primary sources see references therein. The
table does not include the Nile because, in spite of a length of 6500 km, its current water and
sediment discharges are not comparable with figures in this table.
Source:€Gupta, 2007a and references therein

example, the Mekong flows in and out of narrow rocky valleys and wider alluvial basins.
Hydrological conditions also vary along their courses. Different parts of their large basins
contribute water and sediment in different fashions to the mainstream. The end part of the
river has to adjust to all such variations including changes in sea level. Large rivers are
therefore recognised as polyzonal.
The origin and sustenance of big rivers are usually associated with orogenic belts, as the
Amazon is with the Andes and the Ganga with the Himalaya. Part of their upper course may
follow a geological lineation like a long section of the Upper Indus that remains pinned to a
suture zone. Certain large rivers, however, start in rift valleys and cratonic settings. Neither
the Congo nor the Zambezi, which rise on the opposite sides of a common divide, originates
in an orogenic belt. Structural controls on a continental scale, like a downwarp, may deter-
mine the location of the river even in lowlands. The Ganga, for example, flows through a
downwarped foreland parallel to the Himalaya which is related to the tectonic origin of the
mountains. The Amazon connects the Andes to the South Atlantic Ocean through a tectonic
downwarp. Even away from the steep headwaters, smaller crustal deformations and trans-
verse belts of resistant lithology may regionally alter the low gradient and valley width of
the rivers, resulting in repeated changes in channel form and behaviour at the scale of tens
and hundreds of kilometres. Rivers flow over a steeper gradient in a narrow valley in such
145 Characteristics of a large river

Table 9.2╇ Large rivers:€tectonic settings and morphological variability


Tectonic setting Description Tropical example

Continental collision Rivers with headwaters in the Ganga, Paraná


belt:€rivers: mountain range at converging plate
Longitudinal foreland boundary and the main stream
basin flows axial or longitudinal to the
mountain range
Continental collision belt Rivers with headwaters in the Amazon, Orinoco
rivers: mountain range at converging
Transverse systems plate boundary but the main stream
flows transverse to the mountain
range across a stable platform or
craton
Continental collision belt Rivers that flow within mountain Brahmaputra,
rivers: ranges parallel to major structural Irrawaddy, Mekong
Intrabasinal setting trends almost to their mouths
Rift systems River basins formed by creation of
new relief in rift and rift-shoulder
settings
Cratonic settings River systems formed by cratonic tec- Lower Paraná, Orange
tonic events unconnected to plate
tectonics, e.g. large-scale igneous
doming

After Tandon and Sinha 2007

locations, and are likely to deposit less and transfer more sediment. Tectonics, climate and
eustasy combine to control a long river in its entirety, but any of these factors may domin-
ate part of its course. In sum, large rivers tend to have their headwaters in the uplifted large
mountains, but not always. They tend to drain continental-scale cratonic areas (Tandon and
Sinha, 2007) usually through wide low-gradient alluvial plains. Tandon and Sinha (2007)
summarised the tectonic setting and morphology of large rivers (Table 9.2).
The age of large rivers varies but some have existed for a long time. The extra-tropical
Mississippi is at least as old as the late Jurassic. The entire present course of a river, how-
ever, may not have come together at the same time. It could be an assemblage of sev-
eral subunits of different age and appearance (Gupta, 2007a). The nature of the river also
changes over time. Goodbred (2003) has demonstrated the changing pattern of the Ganga
over the last 50 ka in response to climate and sea-level changes.
Rivers of this size are supported by huge amounts of precipitation falling on their basins.
This requirement is met by heavy rainfall of the equatorial and monsoon tropics. Snowmelt
forms a substantive part of the baseflow for rivers originating in the high mountains. The
mainstem discharge is augmented cumulatively by a number of tributaries draining dif-
ferent parts of the basin, but the sediment supply often comes predominately from the
source-mountains of high relief and the tectonically fractured rocks and oversteepened
146 Large rivers in the tropics

(a)

(b)

Fig. 9.1 Average discharges of (a) suspended sediment and (b) fresh water in the Orinoco River and its tributaries. From
Meade, 2007. By Permission of Wiley

slopes (Milliman and Syvitski, 1992). The rest of the basin usually provides little sediment
but augments the discharge of the main river cumulatively at tributary junctions. This is
illustrated by Meade (2007) for the Orinoco, which receives its sediment from the Andes
and its foreland, but water from all parts of the basin (Fig. 9.1).
147 The Amazon

Large rivers with headwaters and tributaries in the high mountains contribute large vol-
umes of sediment to the oceans (Table 9.1) unless inhibited by impoundments, as in the
case of the Indus. The transfer of this sediment along a long route occurs in pulses and is
characterised by short transfer times and long storage periods, especially when sediment
could be stored in floodplains. A number of rivers − the Amazon, Ganga, Brahmaputra −
went through large shifts in discharge following climate change and meltwater floods at the
end of the Pleistocene and in the early Holocene. Sediment of this time may still exist in the
channel or floodplain of the river. Over the last several millennia, anthropogenic activities
such as deforestation, impoundments and land use changes have altered the characteristics
of many river basins and channels.
Several case studies in this chapter illustrate the way water and sediment reach the
mainstream in large basins and are carried out to the coastal waters and beyond, build-
ing deltas in the process. A very large part of the tropical continents is drained by large
rivers and the dynamics of their drainage basins is an essential component of geomor-
phology in the tropics. These basins integrate a variety of structure and processes on a
sub-�continental scale.

9.3╇ The Amazon

9.3.1╇ The physical setting

The Amazon is well studied, in spite of its size (Fig. 9.2). It still remains in natural condi-
tions and is not modified by anthropogenic activities. Plate tectonics control the location
and lithologic and topographic frameworks of the Amazon (Potter, 1978b). The basin was
delineated following the Miocene uplift of the Andes due to the subduction of the Nazca
plate below the South American. It is generally thought that the rise of the Andes cre-
ated this huge eastward drainage system. The tectonically and volcanically active Andes
Mountains slope down towards the east to a foreland basin. The Amazon flows from the
Andes across the foreland eastward through an axial trough, which is bounded to the north
and south by the crystalline rocks of the Guyana and Brazilian Shields. A fair amount
of knowledge exists regarding the underlying structure along the channel and floodplain
because of exploration for petroleum (Mertes and Dunne, 2007). Evidence from deep cores
indicates a 6000-m deep east–west crustal sag below the basin axis that ends below the
mouth of the Amazon in a graben, the Marajó Rift.
Precipitation over the Amazon Basin is influenced by the annual shifting of the Inter
Tropical Convergence Zone (ITCZ) and the South Atlantic Convergence Zone (SACZ)
over the Andean region. Annual average precipitation is about 2000–2500 mm, approxi-
mately uniformly distributed over most of the basin with the maximum of 7000–8000 mm
on the lower eastern slopes of the Peruvian Andes. The extreme northeastern and southern
parts of the basin are drier. The rains arrive first over the southern basin in November to
December, and then move north. The annual hydrograph of the river (Fig. 9.3) is uni-
modal, damped and delayed by the sheer size of the basin, length of the drainage network
148 Large rivers in the tropics

Fig. 9.2 The Amazon. Top diagram:€generalised map of lithological zones and structural features. Note the river crossing four
intercratonic arches and the Monte Alegro ‘intrusion’. Top: from Dunne et al., 1998. Bottom: from Meade, 2007. By
Permission of Wiley

and storage of water in the enormous floodplains, the cumulative size of which exceeds
100€000 km2. Both precipitation and river discharge are affected in certain years by the
El€ Nino Southern Oscillation (ENSO), low flows of the Amazon being associated with
the El Nino years (Mertes and Dunne, 2007). Such climatic fluctuations affect flooding
and sedimentation (Aalto et al., 2003) but the effect on the morphology of the rivers of the
Amazon Basin is yet to be understood.
149 The Amazon

(a) (b)

(c)

Fig. 9.3 Monthly rainfall and river discharge (1972–1996) at three locations on the Amazon River. From Mertes and Dunne,
2007. By permission of Wiley

The axial trough of the central Amazon Basin exhibits a remarkable suite of fluvial
landforms (section 9.3.3) under an extensive forest cover. The forest is interspersed with
savanna and recently deforestation has occurred towards the southern and eastern margins
of the basin.

9.3.2╇ Transfer of water and sediment along the Amazon

About 80 per cent of the Amazon’s sediment is derived from the Andes and its foothills,
although they cover only 800 000 km2 of this 6 million km2 basin. The rocks in the Andes are
primarily volcaniclastic and marine sedimentary rocks, metamorphosed to various degrees.
These mountains have been supplying the Amazon with 500–600 million tonnes of sedi-
ment throughout the late Cenozoic. Almost the entire sediment arrives from the Peruvian
Andes by the Amazon mainstem or from the Bolivian Andes via the Madeira (Meade,
2007). Rivers in the Andes Mountains are steep and flow on bedrock or gravel. In con-
trast, the river gradient across the subsiding Andes foreland is only a few centimetres per
kilometre. The Amazon has very low stream power, about 12 Wm−2, even at average peak
discharge, and only sediment finer than about 0.5 mm (sand, silt and clay) is carried into
the lowland Amazon Basin. Rivers start to meander in the downwarped foreland zone, and
150 Large rivers in the tropics

high sediment deposition leads to the building of bars and shifting of channels. Tributaries
draining land below the Andes and the Guyana and Brazilian Shields bring in very little
sediment but contribute high volumes of discharge. In fact, two of the biggest tributar-
ies of the Amazon, the Madeira and Negro, contribute more water to the Amazon than is
discharged by any other river except the Congo, Orinoco and the Changjiang. Meade has
further observed that the freshwater discharge at the mouth of the Amazon is five times that
of the Congo, six times that of the Orinoco and 12 times that of the Mississippi (Meade,
2007). All these are huge rivers and the comparison indicates the scale of the Amazon and
its importance in global hydrology.
The average annual sediment of the Amazon at Óbidos is about 1200 million t, at
present exceeded only by the combined flow of the Ganga–Brahmaputra. Far higher
sediment, however, is transferred laterally in the Amazon between the channel and the
floodplain than downstream along the channel as calculated by Dunne et al. (1998), who
studied sediment transport through the 2010 km of the lowland Amazon in Brazil. The
lateral exchange involves bank erosion, bar deposition in the main channel, settling from
overland flow on the floodplain and sedimentation inside channels within the floodplain.
Much of the sediment that leaves the channel in suspension during floods and enters the
floodplain is deposited there before clearer waters return to the Amazon during the falling
stage of the annual hydrograph. This is particularly demonstrated in the lower Amazon
where the large floodplain lakes are silting slowly but progressively. Sediment grains are
stored in the floodplain of the Amazon for centuries and millennia. Mertes et al (1996)
�estimated the mean recycling time to be between 1000 and 2000 years for the floodplain
sediment in the reach between the confluences of the Jutaí and Madeira. This allows
chemical weathering of lithic fragments, production of clay minerals and a progressively
quartz-rich �sediment. Ecologically, with annual inundation alternating with exposure, the
floodplain of the Amazon is a remarkable landscape (Junk, 1997).
The pattern of sediment transport is influenced by (1) the hydrology of the Amazon and
(2) geomorphic characteristics of the channel and floodplain. Regional and local tectonics
determine the second to a considerable extent. The pattern that persists is the deposition of
floodplain sediment and its subsequent removal by the shifting river (Meade, 2007).
The tidal limit is approximately up to Óbidos, almost 1000 km from the mouth of the
Amazon, due to the very low gradient of the river. It is difficult to measure sediment fluxes
in the end part of a river, but a large amount of the river sediment is apparently being
deposited on the floodplain, mobilised inside the channel of the Amazon, and also entering
the Atlantic Ocean. About half of the sediment that passes Óbidos settles on the continental
shelf beyond the mouth of the Amazon (Kuehl et al., 1986). A substantial part of the sedi-
ment is carried northwards and deposited along the coast up to the edge of the Orinoco
Delta. Meade (2007) suggested that the amount reaching this delta, about 1600 km from
the mouth of the Amazon, could be as high as 100 million tonnes. This is not a precise
number, but anything in that range indicates the size of the Amazon and its importance in
the geomorphology of the American tropics.
A different pattern of sedimentation seems to have operated on the continental shelf
near the mouth of the Amazon during the Last Glacial Maximum. It has been proposed
that the sea level dropped 120 m during the last regression and the Amazon incised to
12–25€m below its present bed and flowed on a steeper gradient. This allowed the transfer
151 The Amazon

of sediment offshore to build the Amazon Cone (Fig. 9.2). Subsequently, as the sea level
rose and accommodation space became available, considerable sedimentation happened in
the lower river and its floodplain. Less material reached the continental shelf. The reduc-
tion in the number of lakes between the middle and lower courses of the Amazon supports
this interpretation.

9.3.3╇ Morphology of the Amazon River


Beyond the Andes foreland, the Holocene Amazon floodplain is incised more than 10 m
below a landscape of low hills under thick forest cover. The hills had been dissected out of
Tertiary and Quaternary lacustrine and alluvial deposits that fill a central depression. The
Holocene Amazon has incised its channel and floodplain in the middle of this depression
and the surface of the floodplain displays a complex pattern of channels of different size,
scroll bars, levees and various kinds of depressions such as cut-off lakes and backswamps.
The floodplain is commonly flanked by discontinuous terraces, 5–15 m above the usual
inundation level.
The Amazon in Brazil has a remarkably straight and anastomosing channel within its
floodplain. Avulsions related to flow switching are common, and where the channel bends
it usually undercuts the cohesive terrace material above its Holocene floodplain. Over a
measured length of 2000 km in the downstream direction, the low water width of the river
increases from 2 to 4 km and the corresponding depth from 10 to 20 m. The floodplain con-
tains a dense network of channels, some of which are as big as large tributaries. The large
channels stay connected to the Amazon all year but smaller ones may become detached dur-
ing the dry season, although they continue to hold water. The main channel, anabranches
and floodplain channels all migrate and, as a result, numerous scroll bars and depressions
occur on the surface of the floodplain. Mertes and Dunne summarised the Amazon as an
entrenched river that is confined to its valley, remains straight and is relatively immobile
over a distance of hundreds of kilometres (Mertes and Dunne, 2007).
In general, the low-gradient Amazon flows through a 40–50 km wide floodplain
with considerable complexity. Anabranches, levees, scroll bars and large-scale chan-
nel migration or avulsion are common. A complex mosaic of lakes, lake deposits and
overbank sediments are also characteristic features in a wide floodplain. The smaller
of these lakes could be due to small low-sediment tributaries being dammed by rapid
alluviation on the floodplain of the Amazon. Larger lakes, up to 65 km across, occur
downstream.
The Amazon, however, in spite of flowing on top of thick sedimentary layers along
its central valley axis, is affected by a set of buried transverse structures (Fig. 9.2). Four
major structural arches (Iquitos, Jutaí, Purús and Gurupa) and the Monte Alegro Intrusion,
further to the east, modify the river in their vicinity. The Amazon straightens its course
as it crosses the structural highs. On a slightly steeper gradient of the water surface, the
floodplain narrows, the river tends to be at the foot of the terraces, scroll bars are limited
to channel margins and channel migration becomes limited. Gravity measurements for the
lower Amazon region relate direction change and form of the river with gravity anomalies
(Nunn and Aires, 1988), indicating that the Amazon crosses its floodplain only in specific
places. Even the biggest river of the world is structurally controlled.
152 Large rivers in the tropics

Other large-scale fracture patterns control the tributary network of the Amazon. For
example, Franzinelli and Igreja (2002) associated the alignment of the lower Negro
River with a NW–SE trending tectonic lineament. Here intersecting sets of faults have
created sunken crustal blocks and depressions along a half-graben that is inundated to a
depth of 20 m across a 20-km wide reach of the river. The Negro carries very little sedi-
ment and this small amount is flushed through the fault-lined depression with accumu-
lation of only little sand bodies along the margins of the river. The banks of the Negro
are in cohesive cliffs of lacustrine deposits and bedrock is exposed in the river chan-
nel due to a lack of sediment to cover it up (Franzinelli and Igreja, 2002). Latrubesse
and Franzinelli (2002) related straight reaches of the Solimões (middle Amazon) to
recent activities along a set of NW–SE and SW–NE trending fractures. These frac-
tures were originally mapped in 1952 from river and lake alignments by Sternberg and
Russell. Latrubesse and Franzinelli (2002) interpreted the confluence of the Purús with
the Amazon also as tectonically controlled. It coincides with a wide V-shaped sunken
tectonic block.
Three factors interacted through the late Cenozoic and particularly the late Quaternary to
determine the form of the Amazon channel and its floodplain. These are the basin tectonic
setting, climate and sea-level fluctuations. Together these factors controlled erosion, sedi-
ment transport and deposition in the Amazon valley.

9.4╇ The Zambezi

9.4.1╇ The physical setting

Unlike other large rivers described in this chapter (the Amazon, Ganga–Brahmaputra and
Mekong), the Zambezi does not rise from a tectonic mountain system. In contrast, it rises
from a small spring on a relatively low divide that separates it from the northwest-flowing
headwaters of the Congo. Its extreme upper course is characterised by several directional
changes and supported by steady drainage from a series of small pans (swamps) within
broad lowlands. None of these pans extends for more than 4 km in diameter, and they are
either perennially waterlogged or fed by marginal springs. Discharge to the upper Zambezi
is also augmented by base flows from shallow grassy valleys with high water tables and
anaerobic soils, locally known as dambos. Moore et al. (2007) have described these dam-
bos as sponges that support the Zambezi.
The river flows through a number of gorges and wide basins, a pattern which has been
attributed to an evolutionary history of river capture and course changes that started before
the disruption of Gondwana about 120 ma ago. Moore et al (2007) have divided this
2575 km long river that drains a 1.32 million km2 basin into three major sections, each
with characteristic geomorphology. The upper Zambezi extends from the source to the
Victoria Falls, the middle river from the Falls to Cahora Bassa Gorge at the edge of the
Mozambique coastal plain and the lower Zambezi flows over the coastal plain to the Indian
Ocean (Fig.€9.4).
153 The Zambezi

Fig. 9.4 The Zambezi drainage system. Rift basins:€G, Gwembe Trough; MP, Mana Pools basin; C, Chicoa or Cahora Bassa Basin.
Gorges:€B, Batoka; K, Kariba; M, Mupata; CB, Chora Bassa. Rapids and Falls:€1, Chavuma; 2, Ngonye; 3, Katima Mulilo;
4, Mambova; 5, Katombora; 6, Victoria. Floodplains:€Bar, Barotse; MM, Mulonga-Matable; OK, Okavango Delta.
Pans:€MP, Makgadikgadi. From Moore et al., 2007. By permission of Wiley

The upper Zambezi starts as a small channel cut into Precambrian basement rocks before
opening out into a low-gradient shallow stretch through the Barotse floodplain that ends
at the Ngonye Falls. Downstream of the Ngonye Falls, the gradient of the river begins to
steepen and the river downcuts into a set of rapids to flow over the 1700 m long and more
than 100 m high Victoria Falls, also known as Mosi-a-tunya (the smoke that thunders).
The middle river starts in a zigzag gorge banked by low escarpments before turning
abruptly east into the 100 km long Batoka Gorge inside which the Zambezi narrows to about
25–30 m in the 6 m high Chimamba Rapids over basalt. The difference between the upper
and the middle Zambezi is perhaps best appreciated by its change in average gradient, from
0.00024 above the Victoria Falls to 0.0026 below it. Downstream of the Batoka Gorge, the
Zambezi flows through a succession of broad basins separated by narrow defiles, altering
its form and behaviour at every change. These basins, often seen in popular television docu-
mentaries as important wildlife habitats (such as Mana Pools), have been designated world
heritage sites and the gorges (such as Kariba) have been used for power generation with
large lakes inundating the canyons. The middle Zambezi flows into the last of the gorges
(Cahora Bassa) to emerge onto the coastal plain as the lower Zambezi.
The lower Zambezi is a different river. It forms a 100-km long floodplain–delta system
of oxbows and swamps and meander bends. Old distributary channels are found along a
280-km stretch of the Indian Ocean coast between the cities of Quelimane and Beira. A
154 Large rivers in the tropics

6.4 m high tidal range, a 40–50 km incursion of tides up the river and coast-parallel low
ridges found up to 30 km inland and indicating former high sea levels mark this delta-face
(Moore et al., 2007).
A number of attempts have been made to reconstruct the history of the river. Such recon-
structions are based on a multitude of characteristics:

• alternating pattern of gorges and wide basins


• changes in flow directions of the tributaries (several of which are big rivers themselves)
forming elbow-bends
• sedimentary sequences on the coastal plain and offshore
• palaeocurrent directions in fluvial sedimentary rocks dating back to the upper
Cretaceous
• alignment of buried valleys under fluvial sediments
• distribution pattern of fish, fauna and antelopes in the basin.
In sum, the process of drainage evolution started in the early Cretaceous following the
disintegration of Gondwana and the opening of the Atlantic Ocean. The drainage system
that currently terminates in the inland Okavango Delta, different sections of the present
Zambezi that probably flowed as separate systems earlier, and the proto-Limpopo River
south of the current Zambezi basin, all were involved at a certain stage of its history. The
current Zambezi system evolved over time following headward erosion of sections of the
proto-Zambezi, a succession of river captures, updoming of parts of the continental sur-
face, the presence of low relief over much of the area, and tectonic activities along faults
with a NE–SW orientation. This history of drainage development has been discussed in
detail by de Wit (1999), Moore and Larkin (2001), Moore and Blenkinsop (2002) and
Moore et€al. (2007). This explains the changing characteristics of the Zambezi along its
course, the elbow bends in tributary rivers, the occurrence of past river sediments, the pres-
ence of huge palaeo-lakes and the pattern of current speciation of plants and animals.

9.4.2╇ Water and sediment

The average annual discharge of the Zambezi is 220 m3, befitting a major river, but it car-
ries a rather low sediment load, on average only 20 million tonnes of suspended sediment
(Meade, 1996). A number of dams and reservoirs (Fig. 9.4) undoubtedly impound a sig-
nificant proportion of its sediment load, but basically it lacks the presence of an unstable
tectonic mountain system within its basin, unlike the other large rivers reviewed in this
chapter. The pattern of discharge is seasonal (Fig. 9.5), with most of the rain arriving in
the southern hemispheric summer due to the meeting of the northeast monsoon and south-
west trade winds across the Inter Tropical Convergence Zone over the middle and lower
Zambezi, and that of the southwest monsoon and southeast trades over the upper basin
across what is known as the Congo Air Boundary. The two air boundaries move north in
winter resulting in drier months. Most of the annual rainfall of 600–1300 mm falls over the
northern basin, which contributes a higher runoff to the Zambezi (Moore et al., 2007). Like
its sediment load, the downstream passage of the Zambezi’s water is modified by the num-
ber of dams and reservoirs, such as the Kariba and Cahora Bassa. This certainly influences
155 The Ganga-Brahmaputra system

Fig. 9.5 Mean monthly discharge of the Upper Zambezi showing seasonality in flow. Drawn from data in Table 15.2 in Moore,
et al., 2007

the floodplains and marshes along the Zambezi and presumably has altered their ecological
conditions. These shallow-water, low-gradient, natural sediment traps are important wild-
life refuges, including fish, avifauna and mammals that require a wetland habitat such as
hippopotamus, and lechwe and sitatunga antelopes (Moore et al., 2007). These also could
be important for farming and settlements as in the Barotse floodplain.
Large floods periodically occur on the Zambezi. Moore et al. (2007) referred to a 100-
year flood in 1958 with approximately 9000 m3s−1 of water passing over the Victoria Falls.
This is three times the volume of the biggest monthly discharge (April) over the falls. Large
floods have occurred in this century in 2000, 2003 and 2009, and evidently are expected
to occur periodically on this seasonal river. A cyclical pattern of wet and dry years on the
Zambezi has also been suggested by McCarthy et al. (2000).

9.5╇ The Ganga–Brahmaputra system

The two rivers rise on the opposite sides of the Himalaya Mountains and travel a couple
of thousand kilometres before they meet and build a huge delta into the Bay of Bengal.
156 Large rivers in the tropics

Fig. 9.6 The Ganga Basin:€physiographic divisions and major rivers. Megafans of the large tributaries shaded. (b) after Blum, 2007

Together they drain over a million square kilometres, and discharge the fourth highest
Â�volume of water (970 × 108 km3) and second highest average annual suspended sediment
discharge (about 1000 × 106 t) to the sea. These numbers are reached because the rivers
drain a high and tectonically active mountain and intense monsoon rain falls over their
basins. The rivers are separately discussed first, followed by an account of their delta
�conjoint with that of the Meghna River (Fig. 9.6).

9.5.1╇ The physical setting of the Ganga River

The main headwater of the Ganga, the Bhagirathi, starts from the Gangotri Glacier
at Gaumukh at 3800 m. The river is called the Ganga after its confluence with the
Alaknanada at Devprayag. The Ganga descends to 290 m in 300 km at Haridwar, where
it leaves the Himalaya to enter the alluvial plain (Figs. 2.6 and 9.6). After descending
the steep Himalayan slopes, the Ganga and its tributaries flow through the low-gradient
extensive alluvial plains of the Himalayan foreland, a subsiding continental interior fore-
land setting.
157 The Ganga-Brahmaputra system

The south-flowing Ganga turns first southeast downstream from Haridwar and then east
to flow down a wide alluvial plain built by the river and its tributaries. All along its course,
it is joined from the north by large Himalayan tributaries with high discharge at intervals
of hundreds of kilometres. These rivers and the Ganga have built huge fan-shaped alluvial
deposits at the highland–lowland contact, known as megafans (see Chapter 13). Several of
the Himalayan megafans have been described in detail, e.g. the Kosi (Gole and Chitale,
1966; Wells and Dorr, 1987; Singh et al., 1993).
The tributaries from the south drain the old rocks of the northern edge of the Indian
Peninsula. A number of these tributary streams flow into the Yamuna, the largest tributary
of the Ganga. The Yamuna is also a Himalayan river that joins the Ganga at Allahabad.
Nearly 59 per cent of the combined discharge of 130 × 109 m3 at the confluence is from the
Yamuna (Das Gupta, 1984). At the eastern margin of the large alluvial plain of the Ganga
valley, the river passes through a gap in the basaltic low hills of Rajmahal and within a short
distance enters its delta. One of its two major distributaries, the Bhagirathi or the Hughli,
flows directly south within India, collecting drainage from the northeastern corner of the
Indian Peninsula. The other major distributary, the Padma, carries most of the discharge
into Bangladesh where it meets with the Brahmaputra, and then the Meghna, another major
river. The three major rivers have built the combined Ganga–Brahmaputra delta.

9.5.2╇ Hydrology

The Ganga is a rainfed seasonal river, although it does receive some summer snowmelt
from the Himalaya. More than 70 per cent of the annual rainfall, in certain locations about
80€per cent, comes from the wet monsoon between July and early October. The annual total
decreases from east to west, from 1600 to 500 mm. The southwestern basin is also drier with
an annual total between 500 and 700 mm. As expected, the rainfall rises on the Himalayan
slopes, reaching 1500–2300 mm (Singh, I., 2007). The rain often falls intensely, and trop-
ical storms, some reaching cyclonic status, periodically affect the lower basin leading to
floods.
The flow of the Ganga reflects both the seasonality of rainfall and the stepwise increase
in discharge where the major tributaries such as the Yamuna, Gomati, Ghaghara, Gandak,
Son and Kosi come in. Some increment is due to the baseflow that seeps in through the
deep alluvial cover. According to Das Gupta (1984), 50 per cent of the annual rainfall
enters the river as surface runoff, 30 per cent is lost by evaporation and 20 per cent seeps
to the subsurface. During the dry season, part of this subsurface water flows through the
high banks of the Ganga into its channel as baseflow. The mean discharge of the Ganga at
Farakka before it divides into deltaic distributaries is 70 547 m3 s−1. About 60 per cent of
this arrives from the Himalaya and the northern plains (Das Gupta, 1984).

9.5.3╇ Sediment

As expected, the sediment load of the Ganga comes mostly from the tectonic Himalayan
Mountains. Chemical weathering is not important in the Himalaya and the solution load
of the river is low, being diluted even further when the seasonal discharge is high. The
158 Large rivers in the tropics

Fig. 9.7 Changes in grain size of the bed material of the Ganga from Haridwar to Ganga Sagar. The middle reach coarsening is due
to the contribution from the southern rivers draining the Indian Peninsula. From Singh, 2007a. By permission of Wiley

suspended and bed load of the Ganga are very high; the suspended load being the second
highest after the Amazon. Milliman and Syvitski (1992) have estimated the annual sus-
pended load of the Ganga as 520 × 106 t. About 90 per cent of the sediment travels during
the months of the wet monsoon (Singh, I., 2007). The bed load of a large river is difficult
to measure but Wasson (2003) estimated that 600–2500 million tonnes of bed load reaches
the delta each year. Most of the sediment comes from the Himalaya via the large tributaries.
The source of the tributaries varies, as they may originate from the mountains, the foothills
or the plains below the hills (Sinha and Friend, 1994). Rivers starting in the northern moun-
tains and foothills contribute most of the sediment. The southern tributaries drain a craton
and contribute coarse sediment.
Figure 9.7 demonstrates the downstream change in bed material of the river from
Haridwar at the foot of the Himalaya to Ganga Sagar where one of the distributary chan-
nels, the Hughli, flows into Bay of Bengal. Measured from bar samples, it indicates a
general downstream fining characteristic interrupted by periodic coarsening of the bed
from contributions by large tributaries (Singh, M., 1996). The bar sediment of the Ganga
is dominated by sand, the mineralogy of which is essentially quartz with minor amounts of
feldspars, micas and rock fragments. Material from the weathered source rocks undergoes
further alteration when the grains form part of the floodplain alluvium and remain in stor-
age between being transported in high flows.

9.5.4╇ The valley morphology

The Ganga and its headwaters in the Himalaya flow in narrow, deep, almost gorge-like
�valleys flanked by small discontinuous patches of floodplains and terraces. The rivers in
the wide alluvial plain between the Himalaya and the peninsular craton are entrenched
below the surface of the plain. The channel of the Ganga remains within a 10–25 km
wide elongated lowland, which is bounded by several metres high alluvial cliffs, the
lowland being referred to as the Ganga River Valley (Singh, I., 2007). The cliffs enclose
159 The Ganga-Brahmaputra system

the channel with braid bars and meander scars, the floodplain, terrace-like features and
wetlands. The channel and the floodplain together could measure up to 3 km in width
and large sand bars, kilometres in dimension, are common in the channel. The Ganga
is confined within the valley and even the large floods rarely overtop the cliffs. The
floodplain and terrace-like features, however, are periodically inundated, the frequency
of which depends on their height above the channel. Gullies and small ravines are com-
mon, and abandoned channels and wetlands appear on top of the cliffs on the alluvial
plain.
Like the seasonal rivers discussed in Chapter 7, the channel of the Ganga becomes
braided with multiple channels and huge kilometre-scale bars consisting of braid bars,
lateral bars and point bars at low flow. The low flow effect is enhanced by the large-scale
transfer of water into irrigation canals or to meet other demands. Point bars develop
where the river displays a meandering pattern with local narrowing of the channel due
to extensions of peninsular lineaments under alluvium. The meandering pattern also
appears in high flows when nearly the entire channel is under water, submerging the
mid-channel bars. The bars occur at several levels related to the frequency of their
inundation. The higher bars are vegetated and usually farmed. During the dry period,
sediment transfer is confined to the deeper sub-channels. During the wet monsoon, sedi-
ment travels across the entire channel width and occasionally also over the floodplain.
At high flow, several metres of sediment is scoured from temporary storages on top of
the floodplain (Shukla et€al., 1999). The general channel pattern remains the same, but
the location and geometry of the bars vary over time. The river was mobile in the past,
but currently its lateral shifting is limited to a scale of several kilometres within the
high cliffs.

9.5.5╇ The setting of the Brahmaputra River

The Brahmaputra rises from the southern slopes of the Kailash Mountain north of the
main Himalayan Range, and flows eastward along the Indus–Tsangpo suture through the
Tibetan Plateau on a relatively gentle gradient, about 0.001. In Tibet, it is known under
various names, the most common of which is Tsangpo. After traversing a distance of
1200€ km on the Tibetan Plateau, the Tsangpo takes a large U-turn at 95°E in a 5075
m deep gorge around the Namche Barwa Peak of the eastern Himalaya, a part of the
Himalaya known as the Eastern Syntaxis. This sudden bend and the deep gorge are fas-
cinating but yet unresolved features of the regional geomorphology. The river then turns
south to enter the Arunachal Pradesh of India as the Sihang or Dihang. The huge river
in the gorge flows over a very steep gradient (0.03) with a very high velocity and is
extremely turbulent. Further downstream it reaches the Assam Plains and, after merging
with the Dibang and Lohit rivers flowing from the northeast, becomes the Brahmaputra,
a wide, deep braided stream flowing west–west–south in the reversed direction of the
Tsangpo. The Brahmaputra maintains this course until near Dhubri, where it turns south
to enter Bangladesh. In Bangladesh, it is known as the Jamuna, and joins the Ganga at
Arichaghat. From its source to the confluence with the Ganga, it is 2900 km long. Figure
9.8 maps part of its basin and the major tributaries.
160 Large rivers in the tropics

Fig. 9.8 The Brahmaputra River and its main tributaries upstream of the confluence with the Ganga. From Singh et€al., 2005.
With permission from Elsevier

9.5.6╇ Hydrology

After flowing through the dry Tibetan Plateau in the rain shadow of the Himalaya, the
Brahmaputra passes through the extremely wet eastern Himalaya Mountains, collecting
drainage from the wet southern slopes of the Himalaya and the hills of eastern India and
India–Myanmar border. The annual precipitation in the upper basin ranges from about 300
mm in Tibet to about 5000 mm in the Eastern Syntaxis. Like the Ganga, the Brahmaputra’s
discharge regime is seasonal, dependent on the wet southwest monsoon. In Tibet, snow-
melt, groundwater and rainfall are about equal contributors to the Brahmaputra’s discharge.
In India, the monsoon rainfall predominates, with about 70–80 per cent of the annual dis-
charge of the Brahmaputra in India arriving between June and September. The bankfull
discharge of the Brahmaputra is about 35 000 m3s−1 at Pandu near Guwahati (Fig. 9.9). The
mean annual flood is about 50 000 m3s−1, and the 25-year flood about 60 000 m3s−1. Large
discharges arrive in the Brahmaputra via the Himalayan tributaries to fill the river and inun-
date its floodplain.
Annual floods are expected on the Brahmaputra, given the seasonal rainfall, rapid change
in relief to a very flat gradient, obstruction of the channel by many sand bars and the local
161 The Ganga-Brahmaputra system

Fig. 9.9 The annual peak discharge of the Brahmaputra River at Pandu, near Guwahati, India. Overbank flooding is common.
From Singh, 2007b. By permission of Wiley

Fig. 9.10 MODIS/NASA image of the Brahmaputra Valley and the Ganga–Brahmaputra Delta. See also colour plate section

narrowing of the Assam Plains between the Himalaya Mountains and the Shillong Plateau
(Fig. 9.10). Floods on the Brahmaputra also occur due to seismic events. Massive earth-
quakes of 1897 and 1950, both with magnitudes of 8.7, partly blocked the river, extensively
flooding the Assam plains.
162 Large rivers in the tropics

9.5.7╇ Sediment

A huge amount of sediment and solute is discharged by the Brahmaputra:€1000 million


tonnes of clastic sediment and 100 million tonnes of dissolved matter (Singh, S. K., 2007
and references therein). This, however, is an estimate, as accurately determining the sedi-
ment discharge in the high flow of a large turbulent river like the Brahmaputra is near-
impossible. Bed load could be an important component of sediment transfer in a river like
the Brahmaputra (Galy and France-Lanord, 2001). S. K. Singh (2007) has commented that
given the scale of the runoff and the lithology of the eastern Himalaya, both physical and
chemical erosion rates should be higher for the Brahmaputra Basin than that of the Ganga.
Galy and France-Lanord (2001) determined that the total erosion in the Brahmaputra Basin
is about 1.5–2 times higher than that in the Ganga.
Isotopic compositions of the sediment of the Brahmaputra indicate that parts of the basin
in Tibet, the Eastern Syntaxis (around the Namche Barwa Peak), the hills at the eastern
end of the basin and the Himalayan slopes contribute 5, 45, 10 and 40 per cent of the total
sediment, respectively (Singh and France-Lanord, 2002). The Eastern Syntaxis Zone only
occupies 4 per cent of the basin, but according to S. K. Singh (2007) the erosion rate in
this zone is 14 mm per year, among the highest in the world. This is an area of tectonics,
high relief and intense rainfall. It has been surmised that such a rate of erosion leads to high
rates of isostatic rebound, which is responsible for the very high peaks, like Namche Barwa
(7750 m) and Gyala Peri (7150 m), in this area.

9.5.8╇ The channel of the Brahmaputra

The Brahmaputra displays a surprisingly wide range of variations in its channel and bed-
forms. Unlike most rivers, its steepest slope is not at the headwaters in Tibet but in the
middle of its 2900 km length where it crosses the Eastern Syntaxis. The gradient of the
river is variable on the Tibetan Plateau but averages round 0.001, interrupted by a number
of knickpoints. Downstream the river has cut a 5000 m gorge through the mountains of the
Eastern Syntaxis where the gradient steepens to 0.03. Its slope drops to 0.0001 in the mid-
dle of the Assam Plains (Singh, S. K., 2007). The average channel of the river also widens
to 8 km, but varies between 1 and 20 km in the tectonics-controlled lowland of the Assam
with associated changes in gradient and bedforms.
The upper Tsangpo in Tibet changes from a freely meandering channel to a braided
one. It then flows in a single straight channel that becomes a winding channel flowing
through gorges, before curving round the peak of Namche Barwa in a very deep gorge
and emerging on the Assam Plains as a wide braided river. The highly braided channel of
the Brahmaputra, which is full of sandbanks and high islands, has been described in detail
by Goswami (1985) in Assam, and Coleman (1969) and Best et al. (2007) in Bangladesh.
It is a wide mobile channel between narrow stable nodes (Coleman, 1969). Avulsions
have been recorded. Huge braid bars, lateral bars and bars formed by backwaters at the
mouths of tributaries are built by the material eroded from the mountains and at various
levels within the channel. The Majuli Island in Assam, which is between two channels of
163 The Ganga-Brahmaputra system

Fig. 9.11 A bar in the low season Brahmaputra (locally known as the Jamuna) at Sirajgunj, Bangladesh. Note the high volume
of sediment in episodic transport and storage and the sedimentary structures. Photograph: A. Gupta

the Brahmaputra, is 600 km2 in area, the largest river island in the world. The course of
the river has changed in the past, following earthquakes in this tectonically active basin
(Goswami, 1985).
Further downstream in Bangladesh, the river, known locally as the Jamuna, receives
the large Himalayan tributary of Tista from the west. The Brahmaputra in Bangladesh has
changed course in the past by avulsion, possibly following earthquakes, indicating the
mobility of large rivers in low flat terrain. The braidplain of the Jamuna is 11 km wide on
average, the flow depth is about 5 m and its Brice Braiding Index is 4–6. The braidplain
may widen to 15 km wide in floods, when scour depths of 40 m have been measured
(Best et al., 2007 and references therein). It is a huge river flowing through a dynamic and
highly variable alluvial environment (Fig. 9.11). Goswami (1985) gave the mean depth of
the 31€July 1977 high flow at Pandu, Assam, India as 13.4 m. The river would be several
kilometres wide at this stage and that measure of mean depth indicates several very deep
channels for this wide braided river. Coleman (1969) described the short-term channel
migration of the Brahmaputra in Bangladesh as drastic. The movement of major channels,
modification of bed and bar forms, and bank slumping are conspicuous during high flows,
and especially in floods.
The huge combined delta of the Brahmaputra and the Ganga is the final product of the
sediment eroded and transported from the Himalaya. Goodbred (2003) has described
the history of the delta evolution and the changing nature of the source-to-sink sedi-
ment transfer and discharge variation for the Quaternary, including the passage of the
sediment to the submerged Ganga Fan on the floor of Bay of Bengal as summarised in
Chapter 16.
164 Large rivers in the tropics

Fig. 9.12 The Mekong and its basin. From Gupta, 2007. With permission from Elsevier

9.6╇ The Mekong

The 4880 km long Mekong River drains a northwest–southeast trending pan-shaped basin in
Southeast Asia to the South China Sea (Fig. 9.12). The handle of the pan is in China where the
river starts at an elevation of 5000 m on the Tibetan Plateau. The river then flows in a partly
confined channel through the mountainous terrain of southwestern China, Myanmar, Lao
PDR and Thailand to the alluvial lowlands of Cambodia. Its delta stretches from Cambodia
165 The Mekong

to Viet Nam. The Mekong is a remarkably structure-guided river for most of its course; it is
strongly seasonal in nature. The 785 000 km2 basin is primarily rural with a generally low
density of population. It has been significantly affected anthropogenically only since 1993,
when the first of a planned cascade of dams was closed across its upper course in China.

9.6.1╇ The physical setting


The river runs on rock through narrow valleys in mountainous regions for the first
3000€km of its course. The next 1000 km are on mixed rock and alluvium, although the
valley is almost always enclosed by neighbouring mountains. Only for the last 600 km of
its course, can this huge river move freely through a 500-km wide alluvial lowland that
ends in a large delta.
The geology of the Mekong Basin is patchily known. Apparently, as the Indian Plate
collided with the Eurasian Plate to create the Himalaya Mountains, part of the stress was
transferred southeast to open up river valleys such as the Mekong (Tapponier et al., 1982).
The upper basin in China is only a valley cut into granitic and sedimentary rocks, several
kilometres wide. The river flows over Palaeozoic and Mesozoic rocks and local igneous
intrusions downstream, in places along faults. The Mekong flows in a narrow valley with
steep side slopes on rock; the water flow is steep and turbulent; and the channel is partly
filled with sediment, probably from slope failures on the steep valley slopes. Gorges, ter-
races and rapids are common. South of China in the mountainous terrain of Lao PDR
and north Thailand, the river flows through a landscape of narrow, sharp-crested ridges
separated by deep valleys, with a local relief between 500 and 1000 m. Near Vientiane, a
shallow layer of Quaternary alluvium overlies the rocks. Downstream of the alluvium, the
river continues on rock and then a mixed bed of rock and alluvium until it reaches northern
Cambodia, where it starts to flow over an alluvium of variable thickness, the thickness
increasing downriver (Gupta, 2007b).
The Annamite Mountains between Lao PDR and Viet Nam form the higher and steeper
divide to the east, with peaks reaching to 2000 m. The 200–500 m Korat Highlands of
Thailand form a lower but steep divide to the west. South of the Korat Plateau the western
divide is formed by the small but steep Cardamom and Elephant hills that reach the sea
abruptly at the northwestern corner of the Mekong Delta.
The basin is rugged, but the elevation of the valley floor drops below 200 m about
1500 km from the sea. The river, however, is still structure-controlled, and only becomes
free to move laterally 600 km from the sea in the Cambodian plain. This plain rapidly
widens to about 500 km between the two divides and has a western extension where the
regional drainage is collected in the large lake of Tonlé Sap. This lake is connected to
the Mekong by the Tonlé Sap River, whose direction of flow reverses seasonally. In the
wet season, water from the Mekong flows down the river and extends the lake. In the
dry season the lake drains via the link river to the Mekong and shrinks in size. The Tonlé
Sap River joins the Mekong near Phnom Penh, where about 330 km from the sea, the
first deltaic distributary, the Bassac, splits from the Mekong marking the apex of a very
large delta.
166 Large rivers in the tropics

9.6.2╇ Hydrology

The annual average basin precipitation is 1672 mm (MRC, 1997) but it varies geograph-
ically. The annual total drops below 1000 mm in China in the north and in the Korat
Highlands, Thailand in the west. To the east, over the northern and eastern basin, espe-
cially on the slopes of the Annamite Mountains, the annual rainfall is 2000–4000€mm.
The strongly seasonal rain comes from the southwestern monsoon, with 85–90 per cent
falling between June and October. This is reflected in the flow regime of the Mekong.
The river rises a little in May following the summer snowmelt over the Tibetan Plateau
and mountains, but 80 per cent of the discharge occurs between June and November
(Fig.€7.15); 20–30 per cent may arrive in a single month. Large floods tend to occur late
in the wet season, tailing off very slowly. Large floods, that inundate thousands of square
kilometres in the lower basin on the plains of Cambodia late in the wet season, are often
triggered by the arrival of tropical storms over the Annamite Mountains when the Mekong
is already high. The tributaries, especially the smaller hilly ones, are extremely seasonal
and rise sharply after rainfall. A disproportionately high amount of water arrives from the
steep northern hills of Lao PDR and the northern Annamite Mountains via the Nam Ngum
and Nam Theun systems and the southern Annamite slopes via the large tributaries of the
Kong, San and Srepok (Gupta, 2007b). These are the major source areas for both the nor-
mal discharge and floods of the Mekong.

9.6.3╇ Sediment

Sediment data on the Mekong are not of the same quality as the water discharge, but
certain generalisations can be made. Significant sediment contribution is from the nar-
row, steep part of the basin in China, the hills of northern Lao PDR and the Annamite
mountains. The high difference in river stage (30 m or more) between the wet and dry
seasons results in a wide difference in shear stress, unit stream power, channel ero-
sion and sediment transport. With the falling stage, sand and gravel accumulate around
cross-channel protrusions in the rock and form bars, fill depressions in the rock bed of
the river and form insets against vertical rock banks from massive episodic beds to plas-
ters over the rock in small quantity. Material ranging up to boulders is exposed on the
river bars during low flow, indicating the high competence of the Mekong in floods.
The Mekong River Commission (1997) mentions 75–85 × 106 t arriving from the part
of the basin in China and 150–170 × 106 t being discharged into the South China Sea
annually. This is comparable to Mead’s estimate of 160 × 106 t and Ta et al’s figure of
144€± 36 × 106 t derived from past sedimentation rates in the delta (Meade, 1996; Ta et al.,
2002). But the Mekong carries far less sediment than other large rivers of South and East
Asia, such as the Ganga or the Changjiang. Sediment transport is episodic in the Mekong,
mostly carried out during the high flows of the wet monsoon when the water depth may
rise by 10–30 m. Most of the sediment discharge in the South China Sea thus happens
later in the wet season, with the discharged sediment drifting west across the face of the
delta (Gupta et al., 2006).
167 The Mekong

9.6.4╇ The channel of the Mekong

The morphology of the channel of the Mekong changes several times along its course
(Fig. 9.12). As expected, the river on rock is entirely different from the river on alluvium.
The upper Mekong in China is a structure-guided river flowing over a very steep gradi-
ent, nearly 0.002. Straight sections are joined by short sharp bends in rugged gorges. The
channel of the Mekong here is steep, straight and laden with coarse sediment contributed
by short steep tributaries and mass failures on steep side slopes. Up to northern Cambodia,
the Mekong is essentially a rocky conduit, frequently fault-guided, and with very little
option to shift its course in the narrow deep valley. Accommodation space is limited in the
valley and the general direction of transport for both water and sediment is downstream.
The river, however, is not the same morphologically and Gupta and Liew (2007) identified
eight different sections in the 2000 km of the river south of China.
In cross-section, the river is either trapezoidal in shape or has a deep inner channel
bounded by rock benches. Scour pools and rock protrusions occur on top of the rock
benches and within the inner channel. The channel is marked in wider reaches by a profu-
sion of tens of metres high cross-channel rock ribs, isolated transverse rock piles and lines
of rapids. Local relief inside the channel of the Mekong controls sediment accumulation
and coarse sediment is visible during low flows, forming bars on rock benches and depos-
ited against rock ribs. The banks are in rock with insets in silt and sand. The channel pattern
is straight with sharp bends but a reach with six entrenched meanders in rock also occurs
(see Gupta, 2007b:€Fig. 20.10).
Near Vientiane, where the Mekong flows across an area of Quaternary alluvium, the
channel shallows and widens and a braided pattern develops during the dry season. This
is a near-straight part of the river with elongated rock-core islands located askew inside
the channel. Downstream of the Vientiane alluvium, the river is on rock with deep scour
pools and entrenched U-bends (Fig. 7.8). Rock-cored alluvial islands are the common in-
channel depositional features, where sediment starts to accumulate against a high feature
on the rocky bed. In southern Lao PDR, the Mekong travels over a 50 km reach where it is
anastomosed in rock before flowing over a series of rapids and waterfalls into the alluvial
lowlands of Cambodia. The river is still structure-controlled as reflected by its four 45–50
km long straight reaches joined by right-angled bends.
The Mekong flows freely in alluvium for the next 50 km, overflowing into its wide
floodplain during the wet season and more strikingly during times of high-magnitude
floods. Beyond is the delta of the Mekong, one of the largest in the world. The subaerial
delta can be divided into an upper and lower part dominated by fluvial and marine proc-
esses, respectively. The upper delta is an assemblage of levee-bounded tidal channels and
backswamps, whereas the lower delta is characterised by a series of parallel beach ridges
and inter-ridge swamps and flat plains (Nguyen et al., 2000).
The current river and its basin owe its present location and geometry to the collision
of the Indian Plate with the Eurasian Plate and the strengthening of the monsoon system
in the Early Holocene. This probably explains the narrow upper basin, the control on
river morphology by the regional pattern of folds and faults, the straight river segments
168 Large rivers in the tropics

(a)

(b)

Fig. 9.13 The Mekong River. (a) River channel in rock near Luang Prabang, Lao PDR. Note the alluvial bank in the background
and the sand deposited in the shelter of the rock to build a bar. (b) River channel in alluvium from the air near
Savannakhet, Lao PDR. Note the large midchannel bar. Such bars commonly have a rock core around which alluvium is
deposited. Photographs: A. Gupta

interrupted by sharp turns and U-bends, the inner channel and the rock ribs, and the
restriction on �channel movement for about 95 per cent of the river length. Other regional
big rivers such as the upper Changjiang, Sông Hóng (Red) and Salween exhibit similar
features. The shifting of the coastline across the South China Sea during the Quaternary
also affected the Mekong’s lower course. The current Mekong could possibly be an inte-
gration of several different river systems due to the formation of the Himalaya and the
sea-level change. The Mekong certainly is an unusual river, both in its morphology and
behaviour (Fig. 9.13).
169 Questions

9.7╇ The importance of major tropical rivers

Large rivers operate as huge systems for transferring water and sediment to the oceans.
Their nature and behaviour determine the morphology of their drainage basin to a large
extent as the entire drainage network is connected to these massive conduits. They are not
only important in the humid tropics with large rainfall but their passage, as in the case of
the Nile or the Niger, is also crucial for dry regions. It is therefore essential to study these
massive river systems.
The majority of large tropical rivers have been interrelated with human habitation for a
long time because of water, fertile fine-grained sediment, extensive floodplains of low relief
and ease of navigation. The floodplains and deltas have been anthropogenically modified
for a long time in human history and several of these rivers are now impounded and altered
from their natural form and function. An understanding of large rivers is therefore essential
for the sustainable development of their drainage basins, which cover a significant part of
the tropical land and ocean.

Questions

1. What is a large river?


2. Why does the Amazon have a very low stream power on average in spite of its huge
size?
3. Describe sediment transfer and storage in the valley of the Amazon. Is it all in the down-
stream direction?
4. The Amazon is a huge river and so are many of its tributaries. Do they still exhibit
�evidence of control by geological structure?
5. Does the channel area of the Zambezi increase uniformly with its length?
6. Describe the change in grain size of the bed material of the Ganga, using
Figure€9.7.
7. Explain the form and function of the Brahmaputra River in India and Bangladesh from
the MODIS image of Figure 9.10.
8. Does the Mekong receive water and sediment uniformly from all parts of its drainage
basin? If not, identify and explain the major sources.
9. In what way is the Mekong in rock different from the Mekong in alluvium?
10 The tropical coasts

Stately Spanish galleon coming from the Isthmus,


Dipping through the Tropics by the palm-green shores
John Masefield

10.1╇ Introduction

The coast is where geomorphic processes operate in combination. Waves and currents of
the sea play the dominant role but coastal forms are also shaped by fluvial processes, mass
movements and wind action. For example, the face of a coastal cliff is denuded by weath-
ering, rainwash, rill action and mass movements, while the sea erodes its base (Fig.€10.1).
Sediment derived from a coastal cliff is transported along the coast, as is sediment brought
down by rivers flowing into the sea and material moved landwards by waves from the
sea floor. The term 'coastal processes' includes all activities that take place at or near the
sea. Coastal processes, however, are restricted to a narrow vertical range primarily deter-
mined by the sea level, and confined within the extreme tidal limits. The reach of the sea,

Fig. 10.1 Coastal cliff and shore platform, Australia. Note the effect of sedimentary structure in cliff erosion and the operation of
multiple processes on the cliff shown. Photograph: A. Gupta
170
171 Introduction

however, extends in huge storms, e.g. tropical cyclones, or gigantic waves generated by
tsunamis, but these are low-frequency events.
Certain distinctive characteristics of the coastal zone are:
1.╇Coastal forms occur within a vertical range defined by the height of the highest cliff
or dune back of the coast and the depth of the lowest tidal limit.
2.╇The horizontal range of the coast is defined by the local slope. Steep offshore slopes
give rise to narrow coasts, gentle slopes to wide ones.
3.╇This vertical range and the absolute elevation of its boundaries vary over time as the
sea level fluctuates due to climate change or tectonic activities.
4.╇Coastal forms change swiftly, being different between the wet and dry seasons of the
year.
5.╇The sediment in the coastal zone may be derived from both land and water and con-
tributed to by a number of geomorphic processes.
6.╇A tropical coast often carries a rich biological diversity that creates a special envir-
onment:€mangroves, salt flats, coral reefs.
7.╇ Many coastal stretches are heavily populated.
The words shore and coast are used both interchangeably and definitively, the usage
depending on the context. Box 10.1 includes a list of definitions, including that of shore
and coast.

10.2╇ Types of coast

Inman and Nordstrom (1971) proposed a classification of the coasts of the world, based
on the premise that coasts are primarily controlled by plate tectonics. This classification
is based primarily on large coastal forms with linear dimensions at the scale of 103 km,
offshore–onshore (coastal plains to continental shelves) measurements at 102 km and ver-
tical dimensions of about 10 km from the ocean floor to the top of the coastal mountains.
According to their classification system, first-order features occur at this scale. They envis-
aged coasts primarily as (1) collision-edge coasts (occurring at the active plate margin
of a continent); (2) trailing-edge coasts (on the passive side of a continental plate); and
(3)€marginal sea coasts (where island arcs separate seas from the open ocean). Their coastal
classification is hierarchical. Second-order features are river deltas, sand dunes or glaciated
valleys. Smaller coastal forms such as beaches or bars are recognised as the third-order
features.
Apart from coastal tectonics, Inman and Nordstrom (1971) also divided the coasts
of the world as wave-eroded (45 per cent), glaciated (36 per cent), and wave-deposited
(10.7€ per€ cent). Rivers are primarily responsible for the remainder. Life forms such as
mangroves and corals thrive on the hot and humid coasts. Large salt flats tend to develop
along arid coasts. Large deltas form in protected physical environments, provided enough
sediment is available. Mangroves, salt flats and coral reefs are discussed in detail later in
this chapter and deltas in Chapter 11. In the tropics we should expect coastal processes to
172 The tropical coasts

Box 10.1 An introduction to coastal terminology

Shore: The part of the land–sea interface bounded by the extreme high and low tide levels.
Coast: Coast is shore plus the cliff or dune at the landward margin of the shore and, in some locations,
a small sector of land behind the cliff or the dune. A coast is wider than the shore, but both are usually
limited to several kilometres, generally much less.
The shore is further divided according to tidal levels.
Foreshore: It is the intertidal zone between the mean high and mean low tide levels. It is partially exposed
or covered with water, depending on the tidal conditions.
Backshore: Backshore is the zone landward of foreshore, inundated only during large storms; it includes
higher parts of the beach backed by cliffs or dunes.
Inshore: This part of the shore lies seaward of the foreshore and is commonly submerged.
Offshore: This is the furthest seaward zone, beyond breaking waves.
See Figure 10.2 for illustration.

Fig. 10.2â•… Coast and shore

be defined by tectonics, but modified by waves, currents, life forms or the presence of a
huge amount of river sediment in appropriate locations.

10.3╇ Moving water:€tides, waves and currents

Tides, waves and currents control the movement of water near coasts. It is there-
fore necessary to introduce these phenomena prior to discussing coastal forms and
processes. An in-depth discussion is found in a book on beach processes by Komar
(1998).

10.3.1╇ Tides

The tidal range controls the width of coast over which wave action operates. Strong tides
not only create a wide coast but also result in water flowing fast in and out of bays and
173 Moving water:€tides, waves and currents

lagoons, keeping such inlets open and carrying sediment out to the open sea and up inland
rivers. Tidal heights vary in different parts of the coasts of the world. Davies (1964) used
tidal heights to classify shores of the world into three classes:

• macrotidal (spring tidal range above 4 m)


• mesotidal (spring tidal range 2–4 m)
• microtidal (spring tidal range below 2 m).
Tides moving inland negotiate progressively narrower and shallower passages along rivers
and the inlets of coastal lakes. This leads to the front edge of the tide oversteepening and
resembling a wave front. This is the tidal bore. Off the Ganga–Brahmaputra Delta, where
the range is macrotidal, high tidal bores are formed as the rising tide moves into the delta
through funnel-shaped estuaries. The highest tidal bores probably occur on the Amazon
where they resemble upstream-moving waterfalls, 5 m in height above the level of the river
and advancing at a rate of about 10 ms-1 (Komar, 1998).
In large storms, the sea level is raised by wind pressure and huge waves approach
the coast. This phenomenon is known as storm surge. Storm surges are common on
many tropical coasts, especially those affected by tropical cyclones. Storm surges are
extremely effective in eroding and transporting coastal material and are destructive of
life and property. For example, the 29 October 1999 super cyclone from the Bay of
Bengal raised a 7–8 m storm surge that caused widespread destruction on the coast of
Orissa, India.

10.3.2╇ Waves

Waves are oscillatory movements that cause the water surface to rise and fall. A number
of terms are used to describe the physical properties of waves (Box 10.2). Waves are
generated by wind in the open ocean and travel for considerable distance to reach land
where they steepen and break against the coast, resulting in erosion, sediment transfer and
deposition.
As the energy of waves is derived from the wind blowing on the surface of the sea, the
following properties simultaneously determine the size of the waves and their function:

• wind velocity
• wind direction
• fetch (the distance over which a wave-generated wind blows).
The first step in the generation of waves usually happens in the open sea in an area of
strong winds. This area where waves start is known as the sea and its surface is complex
with waves of varying sizes and shapes. Some old waves disappear and are replaced by
new waves, and wave crests interfere with each other. Waves tend to move out of this
source area and travel downwind, and in the process the complexity is resolved into a
number of separate wave groups (also known as wave trains) called swells. The process
of transformation from sea to swell is known as wave dispersion. Swells tend to have
waves of similar heights and wavelengths and may travel huge distances across oceans
(Fig. 10.4).
174 The tropical coasts

Box 10.2 Properties of waves


wave crest: highest point of a wave
wave trough: lowest point of a wave
wave height (H): vertical distance between a wave crest and the following or preceding wave trough
wave length (L): horizontal distance between two similar points (e.g. crests) on successive waves
�measured perpendicular to wave height
wave period (T): time taken by two successive wave crests to pass a fixed point
wave set up: average water level raised by a wave above the stillwater level of the sea
wave steepness: ratio of wave height to wave length (H/L)
wave velocity (C): travel speed of wave motion in a specific direction (L/T)
See Figure 10.3 for illustration.

Fig. 10.3â•… Wave properties

Fig. 10.4 Wave formation

A number of relationships exist between wave properties


L = gT2/2π (10.1)
C = gT/2π, (10.2)
where g is acceleration due to gravity and the other symbols are as in Box 10.2.
175 Moving water:€tides, waves and currents

Fig. 10.5 Wave refraction

Such relationships indicate that waves with similar periods travel together and a confused
sea is transformed into a regular swell. Longer and faster waves outdistance themselves
from shorter and slower ones over time. In a swell, individual waves rise from the back of
the group, travel through the group and disappear towards the front. Thus individual waves
disappear but the group characteristic remains. Swells are wind-driven and therefore they
tend to travel in deterministic paths in long and narrow bands across the open water. Very
little energy is lost once the swell is formed and running. The period of these waves is usu-
ally shorter than 20 seconds (Komar, 1998). Waveforms and dynamics change as waves
approach a coast. The shallowing and irregularity of the coastline affects the wave trains in
three ways:€refraction, reflection and diffraction.
A line of waves refract as they approach the coastline. The water shallows and the base
of the waves slows down progressively. On a coast with alternate headlands and bays, the
section of the line of waves facing the headlands feels the effect of shallowness early and
slows down. The section of the wave facing the bay is still at a distance and therefore con-
tinues to move fast into the bay. This bends a straight wave crest until it becomes parallel
to the coastline. This is wave refraction (Fig. 10.5). The bending of the wave results in part
of the wave converging on the headlands and part of the wave diverging into the bays. This
results in the concentration of wave energy per unit of wavelength on the headlands and the
dissipation of energy in the bay. Thus high waves and erosion occur on headlands, whereas
low waves and deposition occur in bays.
When waves approach the coast through deeper water, such as in front of a cliff, a steeply
inclined beach or a seawall, waves are reflected back. If both the incoming waves and the
reflected waves are of similar dimensions, then standing waves or clapotis may develop off
the coast. If the reflected waves are of dissimilar dimensions, then partial development of
clapotis is seen. This kind of periodic oscillation is also known as seiches, a term originally
used for enclosed water bodies such as lakes, but it is now also used for this phenomenon
in harbours and open waters.
Wave diffraction occurs when waves approach a natural or engineered barrier, such as
an island or a breakwater, at an angle. This leads to a shadow zone in the lee of the barrier
where the waves are propagated with a change in direction of the crests (Fig. 10.6).
A wave approaching the coast moves into shallower water and at the depth where the
wave height is about half the wave depth, the base of the wave begins to be impeded and
176 The tropical coasts

Wave
crest

Fig. 10.6 Wave diffraction in the lee of an island

Fig. 10.7 Transformation of wave orbital motion near coast

a differential develops between the top and bottom of the wave. This difference in vel-
ocity increases as the water shallows and ultimately the waveform becomes so steep that
it breaks. The point where the wave breaks is determined by the ratio (γ) between wave
height (H) and water depth (h), γ = H/h.
Waves of different heights therefore break at different distances from the coast, giving
rise to several lines of breakers offshore. A wave of a given height breaks closer to the
shore on a steep coast than on a gentle one.
The nature of a wave changes as it moves from deep water to shallower depths near the
coast. Waves are oscillatory in deep water, where the waveform moves over the surface but
individual particles of water instead of travelling undergo an orbital motion (Fig.€10.7).
The orbit gets smaller with depth and stops beyond a threshold. The circular orbit becomes
elliptical as the wave moves into shallow water, as its base is impeded, and finally the
impediment causes the wave to break and the water at the crest moves forward in a linear
fashion. These are now translatory waves where the water particles move forward instead of
177 Moving water:€tides, waves and currents

Fig. 10.8 Types of breakers

oscillating, and impact the shore with the possibility of erosion. Instead of the low, rounded
swells of deep water, the waveform is now peaked with flat troughs between waves. At the
same time, the waves may undergo refraction, reflection or diffraction. The velocity€(V)
and wavelength (L) of a translator wave are given by V = √(gh) and L =€ T√(gh). The
�notations are the same as before.
The incoming wave finally breaks on the shore, giving rise to one of the three types of
breaker:€spilling, plunging or surging. The particular type formed depends on the local con-
ditions. Spilling breakers occur on beaches with low gradient. The top of the wave becomes
unstable and flows down the front of the waves with bubbles and foam, while the wave
travels for several wavelengths before breaking. Plunging breakers have near vertical wave
fronts that curl downwards and impact the base of the wave. These are therefore violent
breakers and, unlike the other two types, erode an impressive amount of sediment. Surging
breakers are seen on very steep beaches. The wave crest does not collapse; the wave top
curls over, but it is the base of the wave that travels up the beach causing the wave to col-
lapse (Fig.€10.8). Different breakers may operate at different locations on a given shore,
and the types may also change over time. Plunging breakers erode material out of the shore.
The other two types tend to deposit.
The area where waves break is termed the breaker zone. The shallow front part of the
breaker zone is the surf zone. Water from breaking waves runs up the surf zone as shal-
low sheets known as swash. The swash travels at high velocity and falls in the upper-flow
regime. The deposited sand on the beach therefore has a sub-horizontal lamination. As
swash runs out of energy, it flows backwards down the slope, known as the backwash,
leaving the beach temporarily exposed to the atmosphere. The zone over which swash and
backwash operates moves up and down the shore, depending on the height of the tide and
storm surges.

10.3.3╇ Currents

Waves move water at an angle to the shore. Nearshore water also flows in a direction paral-
lel or sub-parallel to the coast as a current. Two main types of current can be identified:
1. Longshore currents
2. A circulatory cell system.
178 The tropical coasts

Fig. 10.9 Rip currents

Longshore currents are created by waves coming at an angle to the shore. A wave com-
ing in at an angle to the shore may meet the retreating water from the previous wave which
returns downslope and orthogonal to the shore, and move it obliquely further down the
shoreline. In the process water and sediment may move longitudinally along the shore.
The velocity of longitudinal currents usually is 0.3–0.6 ms−1, although higher values have
been reported and a pulsating nature of velocity is common (Carter, 1988). A considerable
amount of sediment may move along the beach in this fashion, and longitudinal currents
tend to move sediment from headlands or river mouths to downcurrent bays to form the
depositional forms described later. These currents are visible at low tide when the surf zone
topography is visible and water can be seen flowing through troughs and channels close
to the shore.
A circulation system of water can be set up in a different way near the shore. As waves
move shoreward, the heights of waves change along the crest causing water to flow along
the shore from the highest towards the lowest breakers. This movement is often helped
by the presence of troughs running parallel to the shore in the surf zone. On reaching the
position of the lowest breakers, water flows seaward as a strong current in a narrow zone
through the surf. This is the rip current, which is very dangerous to swimmers. Velocities
approaching 2 ms−1 have been observed (Sonu, 1972), indicating that rip currents are cap-
able of moving coarse material. Water is returned to sea by rip currents and then moved
shoreward in waves, completing the circulation (Fig. 10.9). This circulatory system,
179 Rocky coasts

especially the location of rip currents, is enhanced by suitable bottom topography of the
surf zone, especially if rips are located over a trough. Rips, however, also occur off straight
beaches with little topographic variations in the surf zone.
Tides, waves and coastal currents erode part of the shore and deposit sediment �elsewhere.
A coast exhibits evidence of erosion and deposition simultaneously, the exact location
being dependent on the local environment, such as the position of headlands and bays,
location of river mouths, offshore gradient and presence or absence of rocks on the coast.
It is convenient to examine the rocky and non-rocky coasts separately, as their forms and
processes differ. Usually, however, a coastline includes both, and the forms and processes
transform from one type to the other.

10.4╇ Rocky coasts

Cliffs marked by rock-cut platforms at their base characterise rocky coasts (Fig. 10.10). A
coastal cliff is a steep slope whose base is eroded by wave action. The gradient of the cliff
and its appearance are determined by its geology and the nature of the waves that break at
its base. Certain properties can be highlighted.
1. Cliffs located at tectonically active coasts as along convergent or transform plate �margins
are usually steep and pounded by large waves at their base because of steep offshore
gradients. These cliffs are referred to as plunging cliffs.
2. Cliffs reflect their lithology, especially in tectonically passive areas. For example, cliffs
in granite or basalt are steep but those in soft, less indurated rocks tend to be low, gentle
and actively eroded.

Fig. 10.10 Base of a rocky cliff, shore platform and stacks exposed at low tide, east coast of western Malaysia, facing the South
China Sea. Photograph: A. Gupta
180 The tropical coasts

3. Structural variations are important. Waves tend to erode cliffs along lines of weak-
ness:€bedding planes, faults, joints. Cliffs in seaward-dipping sedimentary rocks tend to
break up in blocks which slide to the base. Cliffs in landward-dipping rocks are steeper
and more stable. If hard and soft rocks are bedded alternately, cliffs are often unstable.
4. If the strike of the rocks is at an angle to the coast, a headland-and-bay coast is formed,
with hard rocks at headlands and easily eroded softer varieties behind bays. If the strike
is parallel to the coast then hard rocks form a long line of cliffs, and bays are formed
only where such rocks are breached. The sea then can erode softer rocks at the back to
form near-circular bays with a narrow opening through the hard rock.

Waves erode the base of the cliffs, but most of the cliff is usually beyond the reach of
waves. The upper part of cliffs is therefore denuded by weathering, slope failures and rill
action. A coastal cliff is therefore formed by a combination of geomorphic processes whose
operations are influenced by the nature of regional and local geology. As a result, many
cliffs have an upper overhanging part as they are efficiently undercut by waves. This is fol-
lowed by collapse, and cliffs tend to retreat inland over time.
The erosion of cliffs by wave action takes place by any of the following sub-processes
or by a combination of multiple processes:

1. attrition:€disintegration of rock particles when waves bring them into forceful contact
with a rock or sand surface
2. corrasion:€erosion of coastal bedrock, commonly by the steady grinding action of sand
grains or rock fragments brought in by waves and coastal currents
3. corrosion:€dissolution of soluble coastal rock such as limestone by seawater
4. hydraulic action:€erosion by the waves themselves; caused by the water pressure of the
pounding waves or by the pressure exerted by collapsing air bubbles trapped between
sea and cliff (termed cavitation).

Sand and rock fragments are major tools for erosion and they could be derived from the cliff
itself. A coastal cliff therefore partially provides the tool for its own destruction. However,
the accumulation of too much material at the base may save the cliff from further erosion.
This may happen if material from slope failures on the upper part of the cliff falls to its
base or a large part of the lower cliff collapses because it has been undercut by waves.
Estimates of the annual retreat of coastal cliffs are usually very low and dependent on rock
type. Annual rates of cliff erosion have been estimated by laboratory experiment as 10 m in
volcanic ash, 10 cm–1 m in sedimentary rocks, 1 mm–1 cm for limestone, 1€mm in granite
and even less in basalt (Carter, 1988; referring to Sunamura, 1983). As expected, the rates
of erosion and steepness are indirectly related. Near-vertical cliffs are formed in basalt.
A retreating sea cliff is replaced by a near-horizontal or slightly dipping rock platform,
the level of which is a little below the high tide level and probably within a depth of 10 m
(Bradley, 1958). This is known as both a wave-cut platform and a shore platform. The gentle
slope and the limited vertical range of the wave action determine the width of shore platforms.
On an active plate margin coast, however, the platform may drop off steeply or the cliff may
plunge to depths without much of a shore platform. Waves move back and forth across the
platform in a tidal action, both trimming it and enlarging the irregularities in the rock.
181 Non-rocky coast

Fig. 10.11 Sea arch at low tide, Rawa Island, Malaysia. From Kale and Gupta (2001), © Orient Blackswan Pvt Ltd, India 2001

The retreating cliff leaves vertical erosion remnants in many places, which are
�
distinctly visible on the shore platform, especially at low tide. A narrow rock spur
may run down from the headland. If its nose is eroded at its base exposing a roofed
passage, it is known as a sea arch (Fig. 10.11). A sea arch frequently stands isolated
from the line of cliffs. When the roof of the arch collapses to form one or two separate
vertical bodies of rock, stacks are formed. Over time, stacks are eroded by the sea and
disappear. If soluble rocks like limestone are present, caves may develop into cliffs and
solutional features on the shore platform (Viles and Spencer, 1995). Even non-soluble
rocks like sandstone with well-developed zones of weakness may be eroded to form
tunnels and vertical holes through which seawater spurts in high tide. These are known
as blowholes.
About 80 per cent of the coasts of the world are lined with cliffs (Emery and Kuhn,
1982), although their height and appearance vary. A very long stretch of the subduction-
derived coast of South America is backed by high cliffs, running south from Peru. The west
coast of the Indian Peninsula, except the extreme north and south, is backed by coastal
cliffs. The northern part of this long line of cliffs is in flood basalt and the southern part in
gneissic material. Many mid-oceanic volcanic islands carry cliffs.

10.5╇ Non-rocky coast

A non-rocky coast is usually formed of material deposited either along a straight coastal
stretch or at the heads of bays in headland and bay type coasts. The material is sand, or in
less common instances, gravel. The depositional forms could be higher than the mean sea
level (beaches, spits, forelands, etc.) or submerged most of the time (bars).
182 The tropical coasts

10.5.1╇ Beach

A beach is probably the most common and striking form of coastal deposition (Fig. 10.12).
It is a cover of wave-deposited sand or gravel over rock or a pre-existing sediment back-
shore. The gentle surging breakers move sediment up the shore to build a beach. The top
of the leading wave approaches the beach at an angle, curls, breaks and its impact on the
shore carries a thin sheet of water upslope, depositing sediment. The water (generally with
very little sediment, if any) then runs down the steepest slope, orthogonal to the beach, as
backwash. Repeated swash and backwash thus cause water and sediment to drift parallel
to the coast; the process is known as beach drifting. The result is the same as the longshore
drifting mentioned earlier: water is transferred parallel to the coast and sediment is trans-
ferred with it, extending and building the beach.
Beaches are generally made of sand which comes from three sources:€rivers, headlands
and the sea floor. The available volume can be very large. The Ganga–Brahmaputra system,
for example, annually brings about 1000 million tonnes of sediment to the Bay of Bengal.
Large beaches, as expected, occur close to large sources of sand. The river sand can travel
long distances. Sediment carried to the South Atlantic by the Amazon is dispersed north-
westward along the northeast coast of South America. Much of this sand goes onshore
within the first thousand kilometres of coastal wetlands and estuaries, but about half the
amount, which Meade estimated as in the order of 100 million tonnes per year, reaches
much further to build the beaches of the Orinoco Delta, 1600 km away (Meade, 2007). This
of course is an extreme example; the journey of the sediment is usually much shorter.
The composition of sand varies. Commonly, it is silica sand, but where the mater-
ial is derived from a coastal limestone or a coral reef, the material is texturally sand but

Fig. 10.12 Beach and berm, east coast of Malay Peninsula. Note vegetation at the back of the berm. A spit extends to the right
from the beach enclosing a lagoon. Photograph: A. Gupta
183 Non-rocky coast

Fig. 10.13 Ridges and runnels and bars

compositionally calcareous. Such beaches are frequently dazzlingly white and found on
coral islands of the Caribbean Sea and Pacific and Indian Oceans. In contrast, sand derived
from volcanic rocks is dark. Pyroclastic material derived from the slopes of the volcano
Merapi has built black sandy beaches and dunes at Parangtritis on the south coast of Java
as discussed later in Chapter 13. If a beach receives sand from more than one source, then
different parts of the beach may have different colours. If coarser material is available, the
beach could be made of pebbles. The presence of calcium carbonate near the water table
may lead to induration of sand as the water table fluctuates with tides or seasonal changes.
The resulting consolidated material is known as beach rock. As expected, beach rock is
common in a lime-rich environment such as the coasts of the Caribbean Sea.
The alignment of the beach depends on the sediment supply and angle of the swash.
With a high supply of sediment, longshore transport extends the beach in the direction of
the drift. If sediment supply is low, the beach is extended following the angle of the swash.
Beaches can be long and straight, stretching for kilometres or they could be curved round
the head of a bay between headlands. Bay beaches may resemble the Greek letter zeta, and
are known as zetaform or log-spiral beaches. They acquire this shape because the direction
of swash is changed along the bay by wave refraction. Grains are sorted along beaches in
the direction of drift. Very small curved beaches are infrequently found attached to head-
lands. These are appropriately called headland beaches.
A rock cliff or a sand dune occurs at the back of the beach. A line of vegetation marks the
landward end of the beach in rare cases when cliffs or dunes are not present (Fig. 10.12).
Beaches slope towards the sea but not uniformly. The upper part of the beach is generally
flat, called a berm. This ends at a scarplet, below which the lower part of the beach slopes
down to the sea. The level of the berm is reached only by waves that are significantly
high, and berm levels are often markers of heights of storm waves. The berm is thus a
depositional form, modelled by sand left by the swash. Erosion by smaller waves gives it
a vertical step (Fig. 10.12). Waves from large storms, however, tend to flatten the beach
and recovery, with re-steepening of the beach surface, happens during inter-storm periods.
Beaches, however, do occur with no berms or multiple berms.
The lowest part of the beach, exposed only during low tide, exhibits lines of bars and
troughs between bars. A sequence of small parallel bars and troughs occur nearest to the
beach and are termed ridges and runnels (Fig. 10.13). Bigger bars occur in deeper waters,
except where the offshore profile is too steep for bars to form. Lines of breakers often indicate
the location of bars. Bars form as far into the sea as the limit of the wave base disturbs the
sediment on the sea floor, roughly up to 10 m depth. Even bigger bars are formed by storms in
deeper waters, which remain in position between storms. The geometry of bars varies. They
could be parallel or transverse to the shore; some bars are even crescent-shaped.
184 The tropical coasts

Certain beaches display special characteristics. A series of cuspate forms may develop
along the foreshore, consisting of uniformly spaced crescentic piles of sand or coarser
material that taper towards the water. These are known as beach cusps. Parallel ridges of
sand separated by linear depressions, called lagoons, may occur at the back of the beach.
These are known as beach ridges, formed either from old spits which have been welded
later to land, or as piled-up coarse sediment and shells by storm waves. Thus the presence
of beach ridges indicates either a former high sea level or a stormy coast. Many beach
ridges were formed in the last 5000 years, and are associated with the highest sea-stand
after the last glaciation. Excellent examples of this occur along the south Aceh coast in
Sumatra and the Kelantan and Trengganu coasts of the Malay Peninsula. Well-developed
beach ridges on a marshy coast are called cheniers.
A beach is an extremely ephemeral form and its size and shape change frequently in
storms or between wet and dry seasons. For example, beaches in the monsoon tropics
are eroded during the wet monsoon and rebuilt during the dry ones. Komar (1998) has
observed that beaches are eroded and bars are built in storms. Beaches with berms return
at other times when bars tend to cover less of the area of a beach.

10.5.2╇ Spit

A spit is an elongated depositional form like a very low embankment that extends from land
into the sea (Fig. 10.14). It is made of the same material as the beach and is built by long-
shore drifts carrying sediment. Longshore drifts tend to continue in the same direction even
if the coast bends following the outline of a bay, and spits are extended straight into the bay
as a continuation of the beach. A spit extending into open water is characterised by a well-
developed beach on the seaward side and a partially enclosed body of shallow water on the
other, known as a lagoon. When lagoons become filled with fine material and aquatic vege-
tation, they become tidal flats. Spits are common on the downcurrent side of river mouths,
and in places they divert river mouths or even block them, except for a small tidal inlet
between the sea and the river. Spits and their effect as barriers are common on the east coast
of the Malay Peninsula, the coasts of the Bay of Bengal and coastal Kerala (India).
Contact with a second set of longshore drift, running obliquely to the first set, curves the
distal end of a spit. This type is called a recurved spit or hook, a name borrowed from the
northeastern Atlantic coast of the United States where this feature is common. Recurved
spits are also formed from wave refraction round the distal end of the spit. Multipronged
recurved spits also occur, and the shallow water between prongs commonly changes to
tidal mudflats. Tombolos are spits that join two islands or an island with the mainland.
Spits are linear but triangular depositional features called forelands also are formed by the
accretion of material on both sides of a spit, or from deposition in a sheltered location such
as the lee of an island.

10.5.3╇ Barrier island

A barrier island is an elongated depositional feature, which appears above water as a long
line of sand separated from the mainland by a narrow and shallow body of water, called the
185 Non-rocky coast

Fig. 10.14 Spits, recurved spits and tombolos

lagoon or bay. These are much bigger features, up to 100 km in length and 5 km in width,
but they are low, commonly below 6 m in elevation (Ritter et al., 1995). As these are large
features, people have settled on them, but the low elevation makes them, and anyone who
lives on them, vulnerable to tropical storms and high waves. Their growth requires an
assured supply of sand, limited wave energy, a low offshore gradient and a limited tidal
range. Beaches tend to develop on the open side of a barrier island. Their tops carry low
sand dunes and flat areas under grass and shrubs. Higher barrier islands may even have
trees. Marshes and tidal flats occur between the island and the proper lagoon, and narrow
186 The tropical coasts

tidal inlets separate individual barrier islands, which may stretch one after another for a
long distance parallel to the mainland. The inlets, however, are temporary and often shift or
silt up following events such as a large storm. The origin of these islands is much debated,
but they are generally seen as relict forms from a lower sea level during the Quaternary
which have migrated landward with the rising sea level and been modified by current
coastal processes.

10.6╇ Coastal sand dunes

Wide sandy beaches are commonly backed by sand dunes. Two properties differentiate
coastal dunes from the dunes in the desert discussed in Chapter 12. First, although coastal
dunes are formed from wind-deposited sand, they are also found on humid coasts. In
Southeast Asia, for example, coastal dunes occur in areas that receive 2000 mm of rainfall
a year or more. Second, vegetation plays an important role by anchoring sand to develop
coastal dunes.
Wind blows sand inland from the beach. This implies that coastal dunes develop best in
areas with a wide beach, a large supply of fine sand, a strong wind blowing onshore and
the presence of low vegetation at the back of the beach. The width of the beach is usually
determined by the tidal range; a high range exposes a wide stretch of beach at low tide.
Coastal dunes, however, vary in their dimensions. In an earlier study of coastal dunes,
Smith (1954) stated that these dunes vary in width from 30 m to several kilometres and
in height from under 3 m to 600 m. Multiple lines of dunes may occur. The older dunes
may be semi-indurated and immobilised from the precipitation of CaCO3 from circulating
groundwater. The consolidated dune sand is called aeolinite. Smith (1954) provided a very
useful classification of coastal dunes (Table 10.1).

10.7╇ Coastal tropics

The general coastal characteristics and processes as discussed so far are the same in and
out of the tropics. The tropical coasts, however, have certain properties which are added
to the general ones. First, wetlands and offshore forms associated with climate and bio-
logical activities provide tropical coasts with a distinct appearance with the presence of
mangroves, dry salt flats (sabkhas) and coral reefs. Second, certain tropical coasts, because
of their location, are periodically modified by tropical storms. We deal first with the coastal
wetlands of the tropics.
Coastal wetlands appear as flat swampy areas behind bars or spits or as infilled deposi-
tional basins such as deltas and lagoons. These areas are characterised by very little relief,
standing water, brackish water and depositions of fine-grained sediment. Tropical coastal
wetlands are rich in life forms, and in many places prevent serious coastal erosion by
187 Coastal tropics

Table 10.1╇ Types of coastal dune


Types Description

Foredune Piles of sand not more than 3 m in height, running parallel to


the beach, immediately behind it. The form changes rapidly
with a shift in the wind unless anchored by vegetation.
Parabolic or U-shaped dune Curved sand ridges that open towards the beach. They are
formed when the middle part of a dune is removed by wind
or people. A former elliptical mound of sand thus is trans-
formed into two linear arms anchored by vegetation.
Barchan or crescentic dune Typical barchan (see Chapter 12) with the slip-off slope
towards the sea and with horns pointing inland.
Transverse dune ridge Sand ridge parallel or at a low angle to the shore and
�perpendicular to the direction of the prevailing wind. The
side towards the beach is gentle but the one towards land is
steep.
Longitudinal dune Highly oblique or perpendicular ridges to the shore but paral-
lel to the wind. They tend to be symmetrical in form.
Blow-out Hollows cut into a longitudinal dune with sand mounds
developing in the lee of the large hollows.
Attached dune Accumulation of sand around obstacles, mostly vegetation.

After Smith, 1954

�functioning as a barrier to waves. Mangrove swamps are associated with the humid trop-
ics, although some occur in arid areas. Arid coasts are characterised by salt flats. Although
deltas occur on coasts, they are formed by a combination of multiple processes (fluvial,
coastal, wind), and are discussed separately in Chapter 11. Deltas are the most important
type of coastal wetland.

10.7.1╇ Mangrove swamps

Mangroves form a vegetation community, which extends from about 32°N to 38°S, and thus
fringe a very large part of the tropical coast, especially the humid variety. The �vegetation
community commonly occurs in the zone between mean sea and mean high spring tide
levels (Fig. 10.15). Mangroves develop best in sheltered areas where a large supply of silt
from rivers as well as sand from the sea is available. However, this vegetation may grow
on various foundations:€rocky coast, sandy coast, sheltered mudflats. Spatially they extend
from saline conditions next to the sea to the brackish waters of inland river mouths. Plants
embedded in mud tend to slow down the flow of water and a basket-like pattern of man-
grove roots assist deposition. Mangroves therefore tend to accelerate sedimentation and
build up coastal wetlands, especially in shallow sheltered areas such as tide-dominated
deltas, river mouths and sheltered coasts behind a line of coral reefs. Fine-grained sedimen-
tation associated with mangroves occurs mostly by flocculation in tidal waters. Clay-size
particles (flocs) brought down by rivers as suspended load carry a positive charge on their
188 The tropical coasts

Fig. 10.15 Mangroves, South Andaman Island, India. Note the breathing root systems in the foreground. Photograph: A. Gupta

outer surface which results in flocs mutually repelling each other. However, where the
river water mixes with saline seawater, the dissolved salts react with the flocs neutralising
the charge. The flocs then coagulate and grow bigger (flocculation). They can no longer
remain suspended in water and are deposited at the bottom of the channel. In a tidal river,
flocculation spreads over a wider area, spreading the zone of sedimentation. Such sedi-
mentation raises the land, which is then colonised by mangroves. Once mangroves are in
place, sedimentation is accelerated. Mangroves also tend to protect coasts from the erosive
action of large waves.
The surface of a mangrove swamp is low and flat, dissected by small creeks and
�commonly tidal. The levees of such creeks and bigger streams tend to form linear high
grounds that may rise to several metres above the general surface and could be the only
location not inundated in high tide. Small low mounds are also formed by the swamp fauna,
such as mud lobsters. The soils tend to be fine-grained, rich in organics, poorly drained and
saline. Mangroves cope with high salinity and anaerobic soil conditions by physiological
adaptations such as the modification of leaves, sap and root systems. Species of mangroves,
however, vary from each other in physiological adaptation and seed propagation strategy.
189 Coral reefs

A mangrove community is very rich in biodiversity and mangroves have been compared in
this respect to tropical rain forests and coral reefs. Different species tend to adapt to vari-
ations in the micro-environment and different mangroves tend to exist in zones roughly
parallel to the shoreline at the genus level. Such zonations are characterised by belts of
Avicennia, Rhizophora, Brugeria, Ceriops, Sonneratia and Laguncularia. A dense com-
munity of mangroves is known as a mangrove forest or mangal. Individual plants could be
shrubs or trees that range up to tens of metres in height. The community may also include
non-mangrove plants such as nypa palms or ferns.
Mangroves are common along the tropical coasts but are best developed in tidal del-
tas. The Ganga–Brahmaputra delta is possibly the biggest mangrove community in the
world.

10.7.2╇ Sabkhas

Salt flats, resembling coastal deserts, are common in the arid tropics. They are known as
sabkhas, from Arabic. These barren salt flats can be extensive. In the Arabian Gulf, a set
of sabkhas on the United Arab Emirates coast is more than 300 km long and locally up to
24 km wide. These sabkhas are about 7000 years old, and are currently located landward
of coral reefs and beach ridges. Coastal salt flats are common in arid coastal areas, and
they have been studied in detail in other parts of the Gulf, the Red Sea coast of Egypt, Baja
California and Mexico. Generally, sabkhas are flat, salt-encrusted, geomorphologically
stable, sandy plains (Fryberger et al., 1984).
The origin of sabkhas is associated with a cycle of sedimentation related to a regress-
ing sea level displaying a simple upward change. Previously deposited subtidal carbonate
sediments are replaced over time, first with intertidal and then supratidal sediment. This
is reflected in the subsurface by algal mats, carbonate sands and gypsum being replaced
vertically first by gypsum and anhydrite and then nodular anhydrite. On an arid coast, sec-
ondary minerals grow by rapid evaporation leading to a concentration in the subsurface
fluids and the precipitation of gypsum, anhydrite and halite, as secondary minerals in the
interstitial space within the sabkha sediment. These evaporites fill the pore space within the
deposited grains and then expand to raise the surface level of the sabkha. Wind action and
rare floods scour the surface and keep it at a very low gradient.

10.8╇ Coral reefs

Coral reefs are biological structures built offshore by a number of organisms, the princi-
pal members being corals and algae. This implies the requirement for a favourable habi-
tat. Corals, for example, need light and a narrow temperature range, the optimal being
25–29°C. Coral reefs therefore grow only in tropical seas.
Three types of reef are built by the corals.
1. Fringing reefs, which grow next to land.
190 The tropical coasts

2. Barrier reefs, which occur within a short distance of the land and are separated from it
by a shallow body of water called the lagoon. The seaward side of a barrier reef may
drop abruptly to deep water. Small patches of reef may occur in the lagoon.
3. Atolls, which are near-circular or elliptical reefs, although some atolls have strange
shapes. These are found in the open sea, even in the middle of an ocean, and they
build upwards from a relatively shallow underwater basement such as a sunken vol-
cano or a seamount. The outer sides of an atoll drop off steeply to considerable depths,
but a shallow lagoon occurs inside. A large number of atolls occur off the north shore
of Java near Jakarta. The tropical Pacific Ocean and the Indonesian seas are dotted
with atolls.

All types of reef vary in size, usually from metres to tens of kilometres. The Great
Barrier Reef off the eastern coast of Australia stretches for 2000 km from the Torres
Strait (about 10°S) to a little beyond the tropic of Capricorn. It is separated from main-
land Australia by a lagoon which is 80 km wide towards the north and 300 km towards
the south. A number of small reefs and islands occur inside the lagoon. The Great Barrier
Reef itself consists of about 2500 individual reefs. Of course, other coral reefs are not
on this scale.
The basic building block of a coral reef is the coral polyp enclosed inside a calcium-
bearing skeleton. The coral grows by adding new material to the outside of the skeleton.
Parts of reefs are also built by algae. Reefs have a basal platform of calcium carbonate
with vertical ribs growing out of it. The final reef is flat-topped with vertical sides. Part of
the reef could be exposed, especially at low tide, but most of it would remain under water.
This submerged part is a wonderful architecture of walls, flat platforms and canyon-like
passages. Towards the top, corals build a hemispheric structure or a number of upward-
extending branches. Different species of coral inhabit different parts of the reef, as do
varieties of algae, fish, clams, sponges and foraminifera. Like mangroves and the tropical
rain forest, a coral reef is a species-rich ecosystem. Coral reefs may provide some defence
to the nearby coast from waves in tropical storms or small tsunamis.
Coral reefs have been divided into an Atlantic province centred on the Caribbean and
an Indo-Pacific province that includes most of the reefs of the two oceans. Species-
wise, the Indo-Pacific province is much richer. Morphologically, reefs on the wind-
ward side of an island may differ from those on the leeward side. The windward reefs
are straighter and better developed and may show a relationship between geomorpho-
logical zones and the ecological composition of reefs (Woodroffe, 2003). Figure 10.16
shows generalised sections across reefs from the two provinces. Reefs usually have
three zones:€reef front (towards the open sea), reef crest, back reef. The steep reef front
is a dynamic area with active corals. The reef flat is exposed at low tide, and at other
times waves break over it. The back reef is a sheltered environment, rebutting either on
land or, in the case of an atoll, sloping down to a central lagoon. Broken bits of reef form
ramparts of rubble on the reef flat in areas where storms are frequent. The locations
of coral reefs are �generally influenced by changes in sea level during the Pleistocene
(�section€10.9), and modern reefs often start on a base of Pleistocene limestone or sub-
merged volcanic material.
191 Tropical coasts and time

Fig. 10.16 Diagrammatic section through Indo-Pacific and Carribbean (West Indian) coral reefs. Figure reproduced, with
permission, from Woodroffe, 2003:€a) based on Emery et al. (1954), Stoddart (1969) and McLean and Woodroffe
(1994); b) based on Shinn et al. (1982) and Macintyre (1988)

10.9╇ Tropical coasts and time

Two time-based observations are crucial for understanding coastal geomorphology. First,
strictly speaking, all coasts are geologically young, not older than the last sea-level change
at the end of the Pleistocene about 10 000 years ago. The loss of ice has led to the isostatic
recovery of a number of extra-tropical coasts; the land is still being raised slowly against
the sea level. Tectonic movements have episodically raised coasts in active areas, disrupt-
ing the ongoing coastal forms and processes, so that a new coast has been formed several
times (Fig. 10.17). Second, the coastal forms, especially the forms of deposition, tend to
change rapidly.
A number of sea-level changes occurred during the Quaternary, as discussed in
Chapter€16. The peak of the last glaciation was about 18 000 years ago. At that time, the
sea level was about 120 m lower and a long distance away from the present coastline,
unless the offshore gradient was very steep. This led to coastal changes. For example, in
Southeast Asia a large part of the shallow South China Sea was exposed and major rivers
flowed on this land. As glaciation ended, the sea level started to rise, the old coastline was
192 The tropical coasts

Fig. 10.17 Raised coral reef, eastern Jamaica. Such reefs provide information about past sea levels and tectonic uplifts.
Photograph: A. Gupta

submerged, and the sea advanced over the then exposed part of the South China Sea. The
sea level continued to rise until about 6–7000 years ago when it stood several metres above
its present position, after which it came down to the present level (see Chapter 16). None
of the coastal forms in the tropics is older than several thousand years, even in areas not
affected by tectonics.
Even on an annual scale, coastal forms change rapidly. Beach profiles tend to change
between seasons and in storms. A shift in wind vector alters the shape and size of dunes at
the back of the beach, even if they are anchored by vegetation. Features in rock last longer,
but they change more rapidly than most landforms. Cliffs extending as spurs on the shore
platform change to arches and stacks and finally disappear. This may happen on a decadal
scale.
Large storms such as tropical cyclones (see Chapter 3) alter a coastline rapidly. This hap-
pens in the Caribbean, East Africa, South Asia, parts of Southeast Asia and Northeastern
Australia. Tropical cyclones not only erode coastal forms disastrously, but also remove
large quantities of sediment in order to deposit this material in lagoons, bays and other low
areas, at times completely filling them. Large storms also transfer sediment to form beach
ridges. The deltas on the Bay of Bengal coast are eroded by tropical cyclones at a span of
193 Tropical coasts and time

(a) (b) (c)

(d) (e) (f)

Fig. 10.18 Tsunami erosion and recovery, northern Aceh coast. (a) location of the image used for the Aceh coast; (b) coast on
January 2003; (c) December 2004 (tsunami); (d) February 2006; (e) January 2007; (f) April 2008. Note removal
of the beaches, destruction of the wetland and erosion along swales. Within 13 months, beaches have returned,
although the wetland in the centre is under water. Vegetation has returned to the swales. The effect of the tsunami
is not generally perceptible even in the 2006 image. The next two images show the continuation of the rebuilding
process. IKONOS satellite images © Centre for Remote Imaging, Sensing and Processing, National University of
Singapore, reproduced with permission. From Liew et al., 2010. See also colour plate section

several years. In between, more frequent, low-magnitude coastal processes tend to bring
back the depositional features. A coast or a shore may be reviewed as a collection of tran-
sitional features that change every wet season, high storm or tsunami, but recover quickly
to resemble their former appearance.
An excellent example of this destruction and recovery comes from the 26 December 2004
tsunami on the Aceh Coast, Sumatra (Liew et al., 2010). The same 175 km of coastline was
mapped from IKONOS satellite imagery at a 1 m resolution five times:€2003 (before the
tsunami), late December 2004 and early 2005 (immediately after the tsunami), 2006, 2007
and 2008. The Aceh coast was eroded back for about 500 m in the tsunami, except at the
rocky headlands, and almost the entire suite of depositional landforms (beaches, low sand
dunes, wetlands) was removed. A new coast started to appear within weeks, closely resem-
bling the pre-tsunami version (Fig. 10.18). Normal coastal processes returned to operate on
the post-tsunami coast, and as the environmental constraints remained the same, the new
coast is very much like the pre-tsunami one, although still several hundred metres back
from its previous location.
194 The tropical coasts

The response and recovery of a coast to a geomorphic process is fast. The only two
evidences of a powerful process such as tsunami or a hurricane storm that survive are (1)
masses of corals or boulders detached from the source, carried a short distance and left as
exotic blocks or (2) a buried layer of sand or coarser material in the subsurface, deposited
during the tsunami or the storm and subsequently buried under fine sediment.

Questions

1. Explain how multiple geomorphic processes simultaneously operate on a coast.


2. What happens geomorphologically when waves approach the coast?
3. Explain the movement of sediment along a coast.
4. What controls the nature of a coastal drift?
5. Describe the usual sources of sand for a beach. Is this sand transported for a short dis-
tance or brought from afar?
6. Can a coast or shore be seen as a collection of short-term features?
7. Describe a mangrove swamp. Does it have any control on the geomorphology of the
local coastal area?
8. In what way may coral reefs act as indicators of geological history?
9. How long-term are the effects of a particular tsunami or tropical cyclone on a coast?
11 Deltas in the tropics

This is a land half-submerged at high tide


Amitav Ghosh

11.1╇ Introduction

Deltas are partly subaerial and partly subaqueous accumulations of riverborne sediment
deposited at the mouth of the river, with the sediment reorganised by tides, waves and
currents. The characteristics of a delta reflect both the river basin and the dominant geo-
morphic process operating in the coastal region. Deltas therefore vary in geometry, morph-
ology, operating processes and sedimentary characteristics. They develop best when the
river contributes a large amount of sediment on a tectonically inactive, wide and shallow
continental shelf. Deltas can form in both seas and lakes.
The prime morphological division of a delta is between the subaerial and subaqueous
parts. The limit to which tidal processes operate divides the subaerial delta further into a
non-tidal upper part and a tidal lower part. In shape, deltas are approximately triangular or
lobate, the apex of the triangle starting where the main river first splits into more than one
channel. Channels then divide repeatedly and the surface of the delta is marked by a very
large number of such distributaries, their number generally increasing towards the depo-
sitional basin, usually the sea. The main flow of the river uses only some of these chan-
nels; other channels carry less water. The subaerial delta is thus also divided into an active
and an abandoned sector. Most of the discharge and sediment of the master stream flows
through a number of channels in the active part, and the abandoned part includes channels
that are old and moribund, and do not carry much water or sediment (Fig. 11.1).
Towards the end of its course, a river usually carries fine-grained sediment:€sand, silt
and clay. This is the material that generally builds deltas. Deltas tend to show a coarsening-
upward sedimentary sequence, with sand overlying clay and silt. River sediment flowing
into the sea is reorganised by tides, waves or currents, depending on local conditions, and
delta-building continues under water by sedimentation in the depositional basin.
As expected, the surface of a delta is nearly at sea level and has very little relief or gra-
dient. The highest areas are usually levees running parallel to the distributary channels,
with low alluvial basins separating them. In many deltas, almost the entire surface except
the levees becomes inundated during the rainy season or after a large storm. Under natural
conditions, freshwater swamp vegetation covers the upper part of tropical deltas and man-
groves flourish over the saline lower section.
195
196 Deltas in the tropics

Fig. 11.1 Landforms of a delta

Many deltas are highly populated and cultivated, because of their flat gradient, rich-
ness in aquatic resources, wide biodiversity, availability of water and navigability along
channels. On the other hand, deltas are extremely vulnerable to floods, storm damage and
large-scale channel avulsions. Tropical deltas are hazardous places during hurricanes and
typhoons, even during the average wet monsoon. As about a quarter of the world’s popula-
tion live at or near deltaic coastlines or wetlands (Giosan and Bhattacharya, 2005 referring
to Syvitski et€al., 2005), at times this translates into huge losses of life and property.

11.2╇ Distribution of deltas in the tropics

Rivers draining tectonic mountains with high relief and shattered rocks generally transfer a
large volume of sediment along their lengths to construct deltas at their mouths. Sediment
accumulation is faster if the coast is shallow and of low gradient. Large deltas are there-
fore associated with the passive edge of continents and shallow seas, e.g. the Niger Delta.
Figure 11.2 shows the location of large deltas in the tropics. In sum, deltas reflect the com-
plex interaction between sediment supply from rivers, sediment accommodation space on
the coast and coastal energy. Apart from rivers of significant size, smaller rivers also build
substantive deltas for their size, if they carry large volumes of sediment and discharge
onto a shallow sea. A number of rivers draining the islands of Southeast Asia, such as the
Solo in Java, have impressive deltas. The large sediment load is derived from young steep
197 Age and evolution of deltas

Fig. 11.2 Distribution of major tropical deltas

volcanoes and intensive cultivated land. The annual rainfall on these islands could be as
high as several thousand millimetres and the derived sediment accumulates on the broad
continental shelves of the South China Sea. Woodroffe (2005) stated that most Southeast
Asian deltas display a prominent anthropogenic impact. This effect is probably more dom-
inant for smaller deltas, as sedimentation rates have remained similar for 7000 years for
the Ganga–Brahmaputra (Goodbred and Kuehl, 2000a), 3000 years for the Mekong (Ta
et€al., 2002) and for the entire period of delta formation for the Chao Phraya (Tanabe et€al.,
2003).
Hilly islands and the collision side of continents do not have large deltas but are drained
by short steep streams. Such streams may carry a large amount of sediment, especially in
floods, which is deposited in small deltas with a steep gradient. For example, streams tens
of kilometres long drain the earthquake-prone Blue Mountains of eastern Jamaica, an area
which is also impacted by tropical storms, a number of which reach hurricane level. A
number of deltas are found on the northeastern and southeastern coasts of Jamaica:€steep,
built of coarse gravelly material and drained by braided streams. Such features are known
as fan-deltas (Wescott and Ethridge, 1980).

11.3╇ Age and evolution of deltas

Deltas are geologically young worldwide, as they expanded in area only in the Early
Holocene after the slowing down of the post-glacial sea-level rise, which provided accom-
modation space for rivers to deposit sediment and build deltas higher rather than forward
(Stanley and Warne, 1994). Stanley and Warne dated the beginning of modern deltas at
8500–6500 years before present (BP), based on radiocarbon dating of material collected
from boreholes sunk in the deltaic sediment. The dates, however, may vary regionally. For
example, Goodbred and Kuehl (2000a) dated the base of the present Ganga–Brahmaputra
Delta as 10–11 000 years BP. The delta of the Mekong started to prograde 6–7000 years
BP. The large deltas of South, Southeast and East Asia, Ganga–Brahmaputra, Irrawaddy,
Chao Phraya, Sông Hóng (Red River), Mekong, Chang Jiang and others, were formed in
the Holocene, following stable to slightly falling sea-level conditions, high sediment dis-
charge from the Himalaya and the Tibetan Plateau (Ta et€al., 2005), and strengthening of
198 Deltas in the tropics

the monsoon system. The highest Holocene sea level probably occurred about 6000 years
ago. It was several metres higher than the present sea, but varied in age and height in dif-
ferent locations. In some cases, deltas probably built up vertically during the sea-level
rise in the Early Holocene and, after the stabilisation of the sea level, prograded outwards
as observed for the delta of the Changjiang. In other instances, such as the Nile and the
Ganga–Brahmaputra, the present sea level is the highest, and fluvial sediments are actively
accumulating on the delta surface (Hori and Saito, 2007).
Sedimentation occurs where the river meets the sea due to deceleration of the river by
seawater and flocculation following the mixing of river and saline waters, as explained in
Chapter 10. As a result, when a river enters the sea as a jet of sediment-laden water, sand
and silt are deposited first followed further on by clay. The front of a subaqueous delta, its
lowest part, is therefore clay and as a delta is progressively built into the sea, sand is depos-
ited on top of the clay settled earlier. Delta sediments therefore show a vertical coarsening
pattern.
The river water floats above the saline seawater as it flows into the sea, unless it is very
heavily laden with sediment, and sinks. That, however, is a rare case requiring a very high
sediment concentration, which used to happen at the mouth of the Huang He before dams
and reservoirs blocked the flow of its sediment downstream. Mixing between the two sets
of water occurs at the edge of the intruding river jet and the deposition of sediment builds
underwater ridge-like forms. Over time and deposition of more sediment, these ridges are
exposed above water as levees. The seaward face of the delta becomes a mosaic of river
channels bounded by high levees with low waterlogged depressions such as bays and back-
swamps in between.
Breaks in levees give rise to distributary channels that radiate from the main river across
backswamps and bays carrying fine-grained sediment, and sandy sediment splays out
through crevasses. The advance of the delta into the shallow sea is thus followed by the
filling of the low ground between levees, transforming the bays and swamps into the sub-
aerial deltaic plain. Avulsions and the switching of the main river to a lower part of the
delta front focuses delta-building activities to a different area, leaving behind an aban-
doned sector. River-dominated deltas expand aerially in this fashion. The Mississippi is an
excellent example.
Tides, waves and currents, if strong, reorganise the sediment across the face of the delta
and determine the final form and sedimentary sequence (Fig. 11.3). In a strongly tidal zone,
the riverborne sediment moves back and forth over a length of the channel where floccula-
tion takes place. Tidal-dominated deltas are therefore distinguished by the following charac-
teristics:€funnel-shaped river mouths resembling estuaries; bars and islands in the channels,
elongated in the direction of tidal movement; and a number of small tidal creeks between and
through such islands. In the humid tropics, salt-resistant vegetation, mainly mangroves, tends
to grow on the islands and along the banks of coastal creeks, anchoring the fine-grained sedi-
ment (Fig. 11.4). The Ganga–Brahmaputra Delta exhibits all these characteristics.
Where waves are strong, riverborne sediment may be pushed back to give the delta
edge a straighter appearance; such deltas do not protrude into the sea. The deltas of the
Godavari, Senegal, Nile and Sâo Francisco are good examples. This often leads to the
development of wide beaches and, if onshore winds are strong, lines of beach ridges and
199 Age and evolution of deltas

Fig. 11.3 Different types of deltas. After Hori and Saito, 2007

Fig. 11.4 Tidal creek through mangrove forest, Sundarban, Ganga–Brahmaputra Delta. Note the steep bank in silt-clay.
Photograph: A. Gupta

dunes behind them. The lower Mekong Delta exhibits such beach ridges (Fig. 11.5). The
upper part of the Mekong Delta was primarily built by fluvial processes where multiple
channels are found bordered with levees and separated by backswamps. This part of the
delta was built in a sheltered bay. As the delta extended out beyond the headland, it was
exposed to waves and the lower delta became tide- and wave-dominated. Here divergent
200 Deltas in the tropics

Fig. 11.5 Generalised diagram of the Mekong Delta. The upper part is river-built; the lower one is wave and tide controlled.
Simplified from Ta et al., 2005

and bifurcating beach ridges rise 3–10 km above the sea level, separated by low inter-ridge
depressions (Ta et€al., 2005).
A strong longshore current tends to remove and transfer sediment away from the delta
face, limiting the size of the delta. The best example is probably provided by the Amazon
itself. The Amazon discharges a large volume of sediment into the South Atlantic which is
carried northwest parallel to the coast by the Guinea Current. For hundreds of kilometres
this sediment comes ashore to form the mudflats and beach ridges of north Brazil, French
Guiana and Surinam. Part of the Orinoco Delta is built with sediment from the Amazon,
201 Delta morphology

Fig. 11.6 Comparison of morphology and sedimentary environment of different types of delta. The top diagram shows a gener-
alised pattern of angle of sedimentary beds in a delta. From Hori and Saito, 2007. By permission of Wiley

that was originally eroded from the Andes but which has arrived by a very long and cir-
cuitous route. Sediment carried by the current has built mudcapes (linear, round-ended,
fine-grained promontories), several kilometres wide and up to 100 km long, a characteristic
coastal feature between the mouths of the Amazon and Orinoco (Warne et€al., 2002).
If the deposition of sediment in a receiving basin is carried out mostly by a river, it is a
river-influenced constructive delta. If the sediment is reworked and re-deposited by tides
or waves, the delta is recognised as a tide- or wave-influenced delta. Tide- and wave-
influenced deltas can be both constructive and destructive (Hori and Saito, 2007).

11.4╇ Delta morphology

Deltas have three components (Fig. 11.6):

• a low and flat delta plain forming the subaerial part of the delta
• the seaward-dipping part which extends offshore beyond the delta plain, called the delta
front
202 Deltas in the tropics

• the subaqueous low edge of the delta in front of and below the delta front, termed the
prodelta.
When deltas prograde, these components move forward, overriding each other.
A delta plain can be described as an extensive low flat area that includes a number of
active and abandoned distributary creeks that leave the main channels at a high angle. The
channels are bounded by levees, and an assemblage of bays, marshes, tidal flats and flood-
plains form the rest of the delta plain. These low areas are climate-dependent. Freshwater
swamps and mangroves provide a luxuriant surface cover in the humid tropics, as in the
Niger and Mekong deltas. Vegetation is scarce in the arid tropics and instead a saline crust
of gypsum and halite is found on the surface. If the sand supply is plentiful, dune fields
appear, as in the Sâo Francisco Delta. Channels in non-tidal, river-dominated deltas tend
to be sinuous but distributary channels may braid or anastomose in arid areas with discrete
high flows and coarse bedload (Elliott, 1986).
The interdistributary areas are flooded during the wet season but emerge as dry land at
other times. Overbank flooding and crevasse splays during high flows are common in the
humid tropics; so is channel avulsion.
The effect of the tidal passage of water is discernible in the lower parts of tide-dominated
deltas. The channels have a funnelled pattern resembling estuaries, and are of low sinuosity.
Bedforms and sand bars are common. For example, sand bars in low sinuosity channels have
been reported for the Mahakam Delta in eastern Kalimantan, Indonesia. The interdistribu-
tary areas in a tide-dominated delta are dominated by mangroves. For example, Sonneratia,
Avicennia and Rhizophora grow in the part of the Mahakam Delta next to the sea. An exten-
sive zone of nipa palm (Nypa fruticans) and Heritiera is found behind these mangroves,
which is replaced by a swamp forest landward (Woodroffe, 2005, and references therein).
The delta front is the subaqueous area in which sediment-laden river water meets the saline
seawater and sediment is actively deposited, as described earlier. Part of the sediment is also
delivered by traction. Bars are often deposited in front of a distributary mouth and the channel
bifurcates to flow past the bar. This is how the delta spreads into the receiving basin. Tide-
dominated deltas tend to have wide delta fronts (Nittrouer et€al., 1986). Waves redistribute
the sediment unless they are strong and persistent, when the shape of the bar is remodelled
by the direction of wave approach. Exposure to strong waves may straighten the delta face,
building beaches and beach ridges landward in sequence. The delta front advances, usually
with a gentle seaward slope, building a coarsening-upward pattern of vertical sequence.
The prodelta lies seaward of the delta front as a layer of clay and silt on the floor of
the receiving basin. It is the advanced edge of the delta. The sediment is intercalated with
silt stringers and thin shell beds and is highly bioturbated. The presence of shell beds and
coarse sediment in prodelta sediment is commonly attributed to storms.

11.5╇ Delta sediments and sedimentary structures

A structural pattern in delta sediment was recognised by G. K. Gilbert around lake mar-
gins in the western United States in 1885. This concept indicated that deltas start with
203 The Ganga–Brahmaputra Delta:€a case study

a �near-horizontal deposit of fine material in the receiving basin called the bottomset. A
clinoform called the foreset advances over the bottomset as a delta progrades. The foreset
is overlain by the topset, another layer of near-horizontal fine sediment. The angle of the
foreset or delta slopes depends on the coarseness of the deposited material. This concept
works for deltas in lakes and fan deltas with coarse material, where foreset beds have
an inclination of 10–25°. The majority of the deltas of the world are in fine sediment.
Milliman and Meade (1983) determined that the major deltas of the world receive 80–90
per cent fine-grained suspended load and 10–20 per cent coarse-grained bed load.
The subaerial delta plain, described earlier, carries a range of sediment that varies in
texture and structure. For example, levees, crevasse splays and beach ridges are usually
made of sand, whereas the interdistributary lowlands are of clay. In the humid tropics, the
mud is associated with decaying mangrove and peat. Interlaminated or thinly interbedded
alternate mud and sand layers occur on the delta front. The variation has been attributed to
a change in energy from processes such as changing tides. Where the beds are relatively
thick (about 2 cm) as in the delta of the Fly River, Papua, they have been associated with
seasonal changes in wind pattern.
The succession of prodelta, delta front and delta plain gives rise to a characteristic vertical
succession. A coarsening-upward sequence is seen from the prodelta sediment to that of the
delta front. This is overlain by a fining-upward succession from the top of the delta front to
the delta plain. The coarsest and best-sorted sediment occurs where the transporting energy
is the highest. This fining-upward sequence may be replaced by a coarsening-upward one
on the delta plain if dunes or beach ridges occur. This is a generalised succession which is
not necessarily found everywhere, as deltaic sedimentation is variable and complex (Hori
and Saito, 2007). The next section illustrates this complexity.

11.6╇ The Ganga–Brahmaputra Delta:€a case study

Huge quantities of sediment, about a billion tonnes, brought down annually by the Ganga
and Brahmaputra Rivers (see Chapter 9) have built one of the largest deltas of the world,
with a subaerial extent of 111 000 km2 in Bangladesh and India (Kuehl et€al., 2005). The
current delta also includes a prograding clinoform, more than 250 km across the contin-
ental shelf and extending about 125 km from land. The total area, summing the subaerial
and subaqueous parts, is about 140 000 km2. This large tidal delta in the humid tropics has
a monsoon setting, is highly populated and carries dense mangrove forests near the delta-
face (Fig. 11.7). A number of research programmes have recently expanded our knowledge
of this delta.

11.6.1╇ The background

Tectonics divided the Ganga–Brahmaputra Delta into subsiding basins and Holocene and
Pleistocene uplands (Fig. 11.7) which influence sedimentation pattern. The Bengal Basin
has been subsiding since the Eocene, providing space for a very thick accumulation of
deltaic sediment towards the south and the east beyond a hinge line marking the boundary
204 Deltas in the tropics

Fig. 11.7 Physiographic and tectonic map of the Ganga–Brahmaputra Delta. From Kuehl et al., 2005, © SEPM Online

between the Indian continent and the oceanic crust (Sengupta, 1966). The delta surface
slopes from 10–15 m elevation at the northwestern corner towards the southeast to the sea
level. It is always low in relief. The only exceptions are two uplands, known as the Barind
and Madhupur Terraces, that rise to about 15 m above the alluvial surface of the delta.
The surface dips gently offshore, except near Chittagong towards the east, where a
series of north–south-trending structural ridges and lows occurs offshore. A submarine
205 The Ganga–Brahmaputra Delta:€a case study

canyon, known as Swatch of No Ground, that reaches within 30 km of the coast towards
the west, opposite the India–Bangladesh border, is the other significant offshore relief fea-
ture. This canyon is believed to have been incised during the lowstand of the huge river
in the Pleistocene. The canyon is a link to the Bengal Deep Sea Fan. The offshore surface
otherwise is covered by the Holocene sediment of the Ganga–Brahmaputra (Kuehl et€al.,
2005).
The delta, in fact the Ganga and Brahmaputra drainage basins, is under the influence of
the Indian monsoon system. Rainfall is concentrated in a few months (late June to early
October) when the rivers are high and the joined flow of the Ganga–Brahmaputra to the
Bay of Bengal rises from 104 m3s−1 to 105 m3s−1 (Kuehl et€al., 2005). Most of the sediment is
also distributed during this period. The tropical storms, a proportion of which become trop-
ical cyclones, also inundate the lower delta and move a huge amount of offshore sediment
landward to raise the coast, especially the islands. This is a tidal area, with tides reaching
up to 5 m in the estuaries (Barua, 1990). The erosional damage from tropical cyclones is
accentuated if they arrive at high tide.
The two major rivers have been described in Chapter 9. They join in the centre of the
delta at a place called Aricha, and the combined waterway, known as the Padma, flows
southeast. The Padma joins another huge river, the Meghna, which drains the Sylhet Basin
(Fig. 11.7), and the combined mass of water flows into the Bay of Bengal past Noakhali.

11.6.2╇ Morphology

Kuehl et€ al. (2005) divided the entire Bengal Delta into three broad morphological
compartments:

• upper delta plain and flood basins


• lower delta plain and delta front
• subaqueous delta.
The upper delta plain covers a 200 km wide zone of the land part of the delta. Kuehl et€al.
(2005) extended the seaward boundary of this unit to the limit of the dry season inland
extension of salt water. This area is primarily controlled by fluvial processes, influenced by
tectonics and downstream coastal evolution and sea-level change. The river floodplains are
bounded by uplands. The upland areas are broad fluvially dissected surfaces of Barind and
Madhupur Terraces, which are probably Pleistocene in age, and less distinct, terrace-like
Chandina and Comilla surfaces, which are younger. These uplands partition the upper delta
in separate sub-basins, which are either narrow alluvial corridors or low and wide flood
basins such as the Sylhet. The floodbasins such as the Sylhet and Atrai remain inundated
during the wet monsoon under several metres of water that trap fine material (Fig. 11.7).
Both the Ganga and Brahmaputra flow through 40–80 km wide corridors with braid-
belts, overbank flooding and channel avulsions. Channel sand is the dominant subsurface
material (Fig. 9.11 and Fig. 11.8). These rivers transfer sediment through the corridors
downstream but a significant part of their sediment load is trapped there, leading to the
build-up of this very flat and low surface. Rivers have changed their course repeatedly on
the upper delta plain. The best example is that of the Brahmaputra, avulsing alternately
206 Deltas in the tropics

Fig. 11.8 The Brahmaputra (Locally known as the Jamuna) at Sirajganj, Bangladesh in low flow. Section of a sandy mid-channel
bar. For perspective, also look at Figure 9.11. Photograph: A. Gupta

into channels east or west of the Madhupur Terrace. Such major avulsions have occurred in
this enormous river at an interval of several thousands of years. The present course of the
Brahmaputra flows through the western channel.
The lower delta plain and delta front have been defined by Kuehl et€al. (2005) as the part
of the delta plain affected by salt water in the dry season, a zone about 100 km wide and
rising to hardly 3 m at its inland limit. This was originally under a mangrove forest, locally
known as the Sundarbans (Fig. 11.4); the forest is now much depleted. The coastline is a
series of peninsulas between major channels, the peninsulas are criss-crossed by minor tidal
creeks. The surface sediment is silt to clayey silt with a little sand. Clay-rich peat basins
occur, as internal drainage and standing water conditions are common. Interdistributary
islands (islands between distributary channels) extend seaward and in time transform into
peninsulas as the tidal channels separating them silt up. On the seaward side, the islands
extend for tens of kilometres under water as shoals that merge into an advancing wide
apron at a depth of several metres. Kuehl et€al. (2005) have referred to this zone as the
delta front.
The river system has built an accretionary subaqueous delta on the continental shelf. A
complete set of topset, foreset and bottomset forms has been identified on this subaqueous
delta. Apparently the Ganga–Brahmaputra delta, like a number of large river deltas, has
two sets of clinoforms both of which can be identified as delta fronts. Beyond the sub-
aqueous bottomset lies the outer shelf. Evidence of submerged former deltas dating back
to glacial lowstands occurs on the outer shelf, along with oolitic ridges that indicate a low
supply of silica-rich sediment due to climate change during the Pleistocene (Goodbred,
2003).
The submarine canyon (Swatch of No Ground) that dissects the shelf and starts from
about 30 km off the present coast is a prominent feature on the Bengal shelf. The canyon
207 Deltas in the tropics, a summary

is 20–30 km wide with a steeper eastern wall, and is believed to fill with deltaic sediment
which is then periodically emptied by turbidity currents that transfer the sediment to the
enormous Bengal Fan that extends for 2000 km.

11.6.3╇ Evolution

The evolution of the delta (Goodbred, 2003; Goodbred and Kuehl, 2000a) is related to
climate and sea level changes during the Pleistocene and Holocene, and is discussed in
detail in Chapter 16. Briefly, the low stand of the sea occurred around 18 000 years BP,
when both the Ganga and Brahmaputra flowed through incised valleys and the surface of
the Bengal Basin outside such valleys consisted of broad lateritic uplands at about 45–55
m below the present sea level (Goodbred and Kuehl, 2000b). Strengthening of the summer
monsoon started around 15 000 years BP and enhanced precipitation led to the supply of an
enormous amount of riverborne sediment to the delta. The first batch of sediment reached
the Bengal Fan but subsequently, as the sea level rose to transgress the low areas, it trapped
this sediment on top of the delta which then aggraded at an enormous rate between 11000
and 7000 years BP. After 7000 years, deltaic deposition was progradational, extending it
seaward. The surface of the delta was also selectively filled by the avulsions of large riv-
ers, principally that of the Brahmaputra. The main course of the Ganga migrated or avulsed
eastward about 5000 years ago.

11.7╇ Deltas in the tropics: a summary

Tectonics and climate determine the location and characteristics of deltas to a large extent.
The deltas in the humid tropics are further characterised by the availability of a large
amount of sediment, the occurrence of tropical storms and the presence of salt-resistant
vegetation, primarily mangroves. The deltas of the arid tropics display salt flats and, in
certain locations, dunes are formed on the delta plain.
Many tropical deltas are densely populated and farmed. This investment is getting big-
ger, not only because of the increase in population density but also due to changes in land
use. For example, a number of tropical deltas in the humid tropics were farmed for rice.
They are still farmed, but the intensive management of soil and water has given rise to
high-value crops such as vegetables for regional cities or aquaculture for the world. This
is happening especially near the coastline. At the same time, the flat deltas continue to be
extremely vulnerable to natural hazards such as river floods, channel avulsion and aban-
donment, tropical storms and wave surges, which not only destroy villages but also leave
the paddy fields inundated with salt or brackish water. The damage inflicted by the tropical
cyclone Nargis on the Irrawaddy Delta in the first week of May 2008 is a sobering example
(Fig. 11.9).
The spectre of such damages increasing due to climate change falls over the tropical del-
tas (see Chapter 19). We know from the history of adjustment of these deltas to climate and
sea-level changes in the Quaternary and Early Holocene that such hazards are possible.
208 Deltas in the tropics

(a) (b)

Fig. 11.9 The effect of a tropical cyclone on a major delta. (a) Part of the Irrawaddy Delta. (b) Same area after tropical cyc-
lone Nargis. IKONOS satellite image. © Centre for Remote Imaging, Sensing and Processing, National University of
Singapore (2009), reproduced with permission. See also colour plate section

Questions

1. List the major deltas in the tropics. How many of them carry mangroves? Group them
into river-dominated deltas, tide-dominated deltas and wave-dominated deltas.
2. Why do tropical deltas carry a large population?
3. How old are the current deltas?
4. List the factors that determine the geometry, morphology and sediment of the deltas.
5. Construct a hypothetical vertical section for a prograding delta.
6. Describe the morphological characteristics of the Ganga–Brahmaputra Delta.
7. Figure 11.9 includes two IKONOS images showing the destruction of a village in the
Irrawaddy Delta of Myanmar by tropical cyclone Nargis. Interpret the images to show
the type and extent of the damage.
12 The arid tropics

The low and level sands stretch far away.


P. B. Shelley

12.1╇ Arid areas

The tropics are not green and humid everywhere. Semi-arid and arid climates prevail over
about half of the tropics. Apart from a significant moisture deficiency, the arid tropics are
also characterised by high temperature, a high diurnal range of temperature, a theoretically
very high rate of evaporation (potential evapotranspiration > actual evapotranspiration)
and extreme variability in rainfall. Ground cover is limited to low and scattered vegetation,
and even this is absent from the extremely arid areas.
A number of attempts have been made to define the boundaries of the arid region using
various statistics on moisture availability. Grove (1977) defined semi-arid areas as receiv-
ing between 200 and 500 mm of rain annually. Even less rain was expected in the truly
arid areas. Annual precipitation (P) and potential evapotranspiration (PET) are often
used to construct an aridity index. For example, the aridity index of the United Nations
Environmental Programme is the ratio P/PET. Semi-arid, arid and hyper-arid areas have
aridity indices 0.50–0.20, 0.20–0.05 and <0.05, respectively (UNEP, 1992).
A number of more complicated formulae have been used in the search for a better delin-
eation of arid lands, but for our purpose we only need to identify regions where the tempera-
ture is high, the supply of moisture limited and the computed potential evapotranspiration
unachievable. Aridity is caused by several environmental factors and the relative importance
of these factors differs geographically. In general, aridity occurs at the latitude of subtropical
high-pressure belts, where the descending air is heated, clouds are rare and rainfall is limited
(Fig.€1.1). A long distance from oceans may also cause aridity, so continental interiors are
usually dry. Orographic barriers may cause aridity on their leeward side where the wind
descends and tends to pick up moisture. Lastly, cold ocean currents on the western side of
southern continents contribute to aridity by causing atmospheric condensation before the
winds reach the coast. There may be more than one factor behind the aridity of a place.
Although arid areas thus reflect current climate belts and topography, some of these
areas have been dry for a very long time. Certain areas in Africa and Australia have been
deserts since the Tertiary, and the Namib in Southwest Africa has been dated back 80 mil-
lion years to the Cretaceous (Thomas, 1997). Their sizes and boundaries have shifted over
time, especially during the Quaternary (see Chapter 16).
209
210 The arid tropics

Fig. 12.1 Arid lands in the tropics

Aeolian (wind-driven) processes and sand dunes are commonly perceived as associated
with an arid landscape, but at least half of the arid tropics is not covered by sand. These are
rock deserts, usually under low and scattered vegetation. After a rare occurrence of intense
rainfall, they become subjected to brief durations of surface runoff and even flooding. Arid
geomorphology therefore includes the examination of both fluvial and aeolian processes
and related landforms.

12.2╇ Geological characteristics of arid lands

Arid lands in the tropics tend to be on stable landmasses, mostly cratons (Fig. 12.1). Arid
landscapes commonly consist of highlands bounding an intermontane basin or a number of
mountain ranges or uplands rising out of low dry plains. Tectonics has given rise to a series
of block-faulted, steep-sided mountain ranges and basins in the tropical and subtropical
North and Central America over which the Sonoran and Chihuahuan Deserts have devel-
oped. The north–south-trending Chihuahuan Desert of Mexico is hemmed in between the
Sierra Madre Oriental and Sierra Madre Occidental Ranges. In the Sahara, lowlands and
dunes cover less area than rocky and stony mountains (Fig. 12.2). The Eastern Desert of
Egypt is a rugged highland that runs south to the Ethiopian Plateau, where extremely arid
areas occur towards the northern end of the Eastern Rift Valley of Africa at low elevations
that locally drop to depressions below sea level. The Arabian Desert also has hilly and
rocky surfaces. In Southwest Africa, the Namib Desert, located between the Atlantic coast
and the Great Escarpment of Southern Africa, is a narrow strip of rugged upland along with
dunes, plains and dissected hills.
In contrast, the Kalahari Desert in the interior of Southern Africa, east of the Great
Escarpment, is characterised by extensive plains of low relief and the huge swamps of the
Okavango Delta, Makgadikgadi Pan and Etosha Pan. Areas of low relief are also common
in the Thar Desert of India and Pakistan, but the land is also crossed by narrow ridges of
ancient metamorphic rocks. The huge, arid interior of Australia exhibits extensive areas of
riverine plains and sandy tracts, but ridges and rocky uplands are not uncommon (Fig. 1.2b;
note the presence of vegetation). The Atacama Desert of South America, located on the
lee side of the Andes, is somewhat unusual. It comprises a series of north–south-trending
ranges and valleys that rise in elevation from the coast to the Andes.
211 Arid hydrology

Fig. 12.2 Topography in Sahara indicating the extent of rock desert and major sand seas. From Goudie, 2002. By permission of
Oxford University Press

Deserts on old cratons, occurring in the Indian subcontinent, Africa, Arabia and
Australia, have less relief than the Atacama, but are not necessarily a dune-covered land
with low relief. The landform on such a geological framework is produced by processes of
erosion and deposition carried out by both water and wind. The surface material includes
bare rock, gravel, sand and riverine silt. A detailed account of the landforms of the arid
lands of the world is in Goudie (2002), but even this brief description illustrates the vari-
ability of desert landforms. The mountains and uplands may get more precipitation and
the relief may be shaped by rare instances of flowing water. In contrast, wind may take
over as the dominant geomorphic process in the lowlands.

12.3╇ Arid hydrology

12.3.1╇ Desert rainfall

Generalisations on rainfall in arid areas are difficult because rain gauges are few and far
between; the records are seldom long-term; and the rainfall is sporadic by nature. Schick
(1988) attempted a summary account of rainfall in the arid areas, based mainly on data
from the instrumented small drainage basin of Nahal Yael near Eilat, and certain records
from the Sahara. Averages may not mean much in such areas. For example, the mean
annual rainfall in the Mount Sodom area near the Dead Sea is 50 mm but the maximum
measured intensity is 50 mm in 30 minutes (Schick, 1988). The extreme 24-hour rainfall
212 The arid tropics

can exceed the annual mean by a factor of 3–4 once in 20–30 years in the Sahara. Most
of the rain that falls is highly concentrated in time and Schick refers to one case in Eilat
where the 24-hour rainfall was nearly twice that of the annual mean. The intensity of the
desert rain is usually high. At Nahal Yael, intensities exceeding 14 mmhr−1 were associ-
ated with almost half the total rainfall over a period of 17 years. Desert cloudbursts over
the Ahaggar Mountains in Sahara recorded about 1 mm per minute, which for partial
periods of rain rose to 2 mm per minute. Grodek et al. (2000) described the 18 October
1997 rainstorm over Eilat in southern Israel. Three spells of rainfall occurred within 16
hours, showing near vertical rises in the cumulative rainfall diagram. The first rain spell
reached an intensity exceeding 80 mmhr−1. The other two spells arrived with lower inten-
sity, but even then, as Figure 12.3 shows, the figures were impressive. Grodek et al. also
listed other storms, some of which had a very high short-term intensity. For example, the
storm on 12 November 1973 recorded 5- and 10-minute intensities of 120 mmhr−1. The
measures are usually not this high, but clearly rain can arrive with high intensity in the
desert.
The rain is usually from a travelling rain cell, whose velocities vary from nearly zero to
tens of kilometres per hour. The cell boundaries are very sharply delineated. On the ground,
the transformations from completely dry periods to blasts of intense rainfall are generally
near-instantaneous. Rainfall records reported by Schick from Nahal Yael show an insignifi-
cant pre-‘rain’ amount, then a period of sharp and intense rainfall, followed by a tail of low
intensity post-‘rain’ for individual rainfall events. Telemetric measurement of the frontal
advance of a storm cell indicated that it took only 18 minutes to cross the 2 km long Nahal
Yael basin (Schick, 1988).
In general, rainfall is localised, intense and erratic over both time and space. Widespread
rain occurs only when unusual meteorological incidents occur. A flood in 1976 in the Sinai
and the southern Negev from a 72-hour rainfall was caused by the unusual combination of
a Red Sea low-pressure trough and a Mediterranean frontal system that took an extreme
southward deviation.

12.3.2╇ Flood generation

Given this kind of intensity, storm rainfall from desert cloudbursts may cause brief periods
of flooding, especially in places where infiltration to the subsurface is low. Thus, streams
draining out of the mountains, where bare rock and thin soil and vegetation cover prevail,
may rise briefly. For flooding in arid lands, antecedent moisture is usually not a consid-
eration. Rainfall simulations in experimental basins show that the lag time from the initi-
ation of rainfall to the onset of runoff is highly dependent on the rock and soils of the area.
Schick (1988) provided useful data from infiltration experiments carried out in the Nahal
Yael Basin. On a 34° hillslope on widely jointed granite, runoff started 72 seconds after
the onset of rainfall. A total input of 1.05 mm of rain was required to stabilise the water
loss, which was primarily due to infiltration at a rate of 2.8 mmhr−1. In comparison, on a
colluvium underlain by schists and dykes and with a 95 per cent cover of angular 10–40
mm gravel, runoff started after 340 seconds. A total input of 6.4 mm of rain was required
to stabilise infiltration, which was as high as 10.5 mmhr−1.
213 Arid hydrology

Fig. 12.3 Example of type rainfall in arid areas from Nahal Yael and Eilat. From Grodek et al., 2000. By permission of IAHS

Following rain, floodwater may be trapped in the alluvial fills of the several kilometre
long small headwater channels of arid drainage systems. Such fills tend to become satu-
rated prior to the passage of the flood down-channel. Rapid infiltration occurs in these
channels as the rain continues and, after the entire fill resting on the rocky floor of the
channel is saturated, a flood wave proceeds down-channel on top of this saturated allu-
vium. Schick has suggested that the amount of moisture held in the alluvial fill of the
headwater channel acts in the same way as the antecedent moisture that saturates humid
214 The arid tropics

Fig. 12.4 Example of high-intensity rainfall in arid areas and the corresponding discharge hydrographs with steep rise and fall.
Data from the instrumented small drainage basin of Nahal Yael, near Eilat, Israel. Date of storm 18 October 1997. From
Grodek et al., 2000. By permission of IAHS

areas for flood generation. The nature of flood hydrographs in arid areas depends on the
subsurface material and storm characteristics. Usually a runoff-producing event is typi-
fied by a hydrograph with a near-vertical rise and a sharp fall (Fig. 12.4). This indicates
that flash floods in small and medium-sized basins have the ability to significantly modify
the alluvial channel.
At their peak, flash floods may flow with high velocity; even supercritical flow is pos-
sible. A metre-high wall of water may lead the flood wave down previously dry channels
(Leopold and Miller, 1956; Schick, 1988). It is, however, difficult to compute the recur-
rence interval of floods, as the rarity of floods requires a very long record of stream flow
for computing flood statistics. Usually such records are not available. If an alluvial fill is
present in the bigger main channel that the tributaries converge to, the downstream flood
size may be reduced. This may also happen when the floodwater from the channel reaches
the sandy floor of a wide basin. In sum, the protracted flows tend to modify the channel but
such flows occur rarely.
215 The rock desert

Fig. 12.5 River channel in a dry landscape near Alice Springs, Australia. Photograph: A. Gupta

12.4╇ Arid landforms

Landforms in the arid lands can be broadly divided into two classes depending on whether
the ground surface is (1) in rock or underlain by a thin veneer of coarse sediment on rock
or (2) in sand. In the first case, in spite of the aridity, running water is the chief geomorphic
agent; in the second case it is the wind.

12.5╇ The rock desert

Rocky slopes cover a substantive part of arid lands. Such slopes may be part of a steep
mountain front, from the foot of which other slopes run down to a low mountain-girt
basin of deposition. The mountain front and upper slopes are usually bare or under a thin
cover of material being transported. This sedimentary cover becomes deeper over the
lower slopes, and the low central basin is filled with fine sand, silt, clay and salt of dif-
ferent compositions. This tectonics-related basin and range topography is common in the
arid North and Central America, although similar landforms also occur in other places.
The contact of the mountain front with long slopes running down to the basin could be
sharp, or partly masked by a set of alluvial fans. The nearly bare upper slopes are called
pediment, the lower slopes under thicker sedimentary cover, bajada, and the depositional
basin, which may or may not carry a lake, is the playa. These are terms from the south-
western United States and Mexico, but the usage has become general. Bajadas are not
always present and a thinly veneered rock surface may directly reach the valley or basin
floor (Fig. 12.5).
216 The arid tropics

The erosional surface of the pediment may cut across different types of rock and struc-
ture. Its surface could be in bare rock, rocks under a thin veneer of coarse clasts, or par-
tially under alluvial fans. These surfaces are dissected by channels of ephemeral streams;
and small, scattered residual bedrock hills, known as inselbergs or bornhardts, rise from
the general surface. Both of these features are common near the mountains. Bryan (1922)
has described these pediments as transportation slopes, over which sediment travels either
under sheetflow or along small channels to major rivers or playas. Weathering profiles
have also been noticed on rocks of the pediment, indicating their stability.
Elsewhere, the country could be a mosaic of rocky ridges from which slopes run down
to a depression; but the country remains open and not surrounded by mountains (Fig. 1.2b).
Rock slopes from the ridges may end at a riverine plain instead of a playa, but either the
river or the playa would operate as the local base level.
Alluvial fans, common in the arid landscape at the contact between the steep mountain
slope and the plains, are associated with valleys emerging from the mountains. These tri-
angular sedimentary bodies emerge from an apex in the valley and then spread out over
the lower ground. They are fan-shaped in plan view, but their longitudinal profiles could
be concave or consist of a set of straight segments. The latter style is associated with post-
fan tectonic movements. The area of the fan is directly related to the area of the basin
supplying the sediment. Arid alluvial fans tend to be smaller and steeper than their humid
counterparts.
The deposited sediment could be from debris flow, hyperconcentrated flow or channel
flow. Often a fan displays evidence of all three processes. Fans are classified as debris
flow (also called mudflow) fans and fluvial fans. Debris flows are common in arid areas.
They result from brief periods of intense rain when accumulated lag sediment on channel
floors is periodically moved. Coarser grain sizes are associated with higher fan slopes and
variations in slope. However, Bull (1964) has shown that the abundance of sediment may
also steepen fan slopes, even for sediment derived from basins underlain by mudstone or
shale.
As the stream responsible for a fluvial fan emerges from the confined narrow and deep
valley in the mountains, it tends to change its form to flow through a wide, shallow, braided
channel. It also changes its course frequently. The surface of the fan is therefore marked
by deposits of variable processes and age that are related to the channel shifts. The older
part of the fan, if left undisturbed for some time, tends to attain a surface patina known as
desert varnish.

12.6╇ Running water in arid lands

Running water, both as sheet and channelised flow, is the prominent geomorphic process in
rock deserts. The water is derived from rain on the mountains or cloudbursts, as described
earlier. Given the nature of rainfall in the arid tropics, these ephemeral channels mostly
remain dry, being in flood for an extremely brief period of time. A number of floods on
rivers draining semi-arid areas have been studied in detail. Examples include floods on the
217 Running water in arid lands

Wadi Watir and Wadi El Arish in Israel (Schick, 1988; Schick and Lekach, 1987) and those
on the Finke or Todd River in Central Australia (Bourke, 1994; Bourke and Pickup, 1999).
One common observation in all papers dealing with stream channels in arid areas is the
episodic widening and filling of desert channels (Wolman and Gerson, 1978; Schick, 1988;
Bourke and Pickup, 1999).
It follows from earlier discussions that arid and semi-arid rivers need to adjust to varia-
tions in rainfall, rapid rises in runoff, flashy flood hydrographs, losses in the transmission
of water downchannel, high suspended sediment loads and asynchronous tributary activ-
ities (Bourke and Pickup, 1999). All these affect the channel characteristics of the arid
rivers as discussed for the Todd by Bourke and Pickup. They conclude that the concept of
dynamic equilibrium between discharge, sediment and channel form proposed for rivers in
temperate countries does not work for arid rivers which are subject to very high variabil-
ities in discharge. Many arid-zone rivers are non-equilibrium rivers (Graf, 1988).
The Todd River flows out of the rocky MacDonnell Ranges in Central Australia through
a piedmont zone with Pleistocene fans over rock to an alluvial plain downstream. The
river then disappears in the longitudinal dunefield of the northern Simpson Desert. It is
part of the large ephemeral stream system that drains into the Lake Eyre Basin (Fig. 12.6).
It passes the town of Alice Springs, which has a long-term rainfall record. The rainfall is
infrequent and variable. Rain tends to fall between November and March (the southern
hemispheric summer) as discrete storms of limited dimensions leading to asynchronous
flow in the basin tributaries in this large basin. The riverbed is dry 98 per cent of the time
(Fig. 12.7) and when the water arrives, both the rise and fall of the river are swift. The lar-
gest discharge measured at Alice Springs was 1290 m3s−1 in 1988, draining a catchment of
450 km2, which gives it a unit discharge of about 2.6 m3s−1km−2. This has been identified
as a 50-year flood. In general, all discharges in an arid river like the Todd are episodic and
flood-generated. The channel remains dry between floods (Fig. 12.8).
Immediately out of the MacDonnell Ranges, the river has a single straight or winding
channel, which is wide and shallow, and incised into Pleistocene alluvium, palaeoflood
deposits and modern fluvial sediments. The dry channel displays a coarse sandy load that
has formed large-scale ripples and tabular bars. Silt and clay occur in localised depressions
in the channel. Bourke (1994) and Bourke and Pickup (1999) provide detailed accounts
of the morphology and behaviour of the Todd, from which five major observations can be
highlighted, summarising the general characteristics of fluvial forms in arid and semi-arid
climates.
1. Unlike most streams of the humid tropics and temperate regions, the downstream vari-
ations in channel forms do not occur gradually, but as a set of step changes at tributary
junctions where abrupt morphological changes are found.
2. The tributary junctions could be discordant (Fig. 12.9), with one channel temporarily
barred with the sediment brought down by the other, due to localised rainfall and asyn-
chronous fluvial activities in different parts of the channel net as described by Bourke
and Pickup (1999). Schick (1988) has also referred to a study by Lekach on the Wadi
Mikeimin, a tributary of the Wadi Watir in southeastern Sinai. Following a localised
rainstorm, the Wadi Mikeimin built a fan at their confluence big enough to completely
218 The arid tropics

Fig. 12.6 The Todd river with reference to the Lake Eyre Basin. From Bourke and Pickup, 1999. By permission of Wiley

obstruct the channel of the master stream. The obstruction remained for 22 floodless
months, until a large flood removed the fan completely.
3. Transmission loss occurs in the river longitudinally, so the river ends in the middle of
dunefields. This phenomenon is locally known as the floodout. A floodout is the site
where channelised flow ends and floodwaters spill across the adjacent alluvial surface.
The river divides into a number of small channels prior to the complete loss of its water.
Floodouts in arid Australia have been investigated by Tooth (1999), who identified the
marked reduction in channel capacity in the floodout zone and the increased diversion
of floods overbank. The main causes of floodouts are either downstream reductions in
discharge or the presence of barriers to channelised flow. Downstream reductions of
discharge may happen due to the distance from the primary source of moisture, which
is the mountain range at the head of the basin, combined with the loss of transmission in
rivers flowing on coarse sandy alluvium. The barriers could be aeolian, alluvial, hydro-
logic or structural in origin (Fig. 12.10). A new alluvium burying older terraces marks
the location of a floodout zone. Floodouts can be intermediate, with channels reforming
downstream, or terminal where floodwaters dissipate completely.
219 Running water in arid lands

Fig. 12.7 The dry Todd River at Alice Springs. Note the width of the river in flood (as indicated by the length of the bridge) and
the size of the cross-section required to transport floodwaters. Photograph: A. Gupta

Fig. 12.8 Annual floods at Alice Springs, 1953–1993. From Bourke, 1994. By permission of IAHS

4. The sediment of the channel displays a chaotic cut-and-fill structure, related to floods
of varying sizes. As the channel effectively stays dry between floods coming down the
mountains, such structures survive until the arrival of the next significant flood that is
at least comparable in size to the one which last redistributed the sediment. Such floods
tend to (1) widen the river channel, (2) vertically strip the channel and floodplain sedi-
ment, (3) erode new flood channels on the floodplain, (4) form small back channels
on the floodplain and (5) mark the floodplain with macroturbulent scours, especially
220 The arid tropics

(a)

Ross

Todd

(b)

Ross

Todd

Fig. 12.9 The confluence of the Todd with a major tributary, the Ross. Note the blockage at the confluence created by an
asynchronous arrival of floods. From Bourke and Pickup, 1999. By permission of Wiley

around trees and bushes. Flood aggradation then builds a channel and floodplain as
insets within the flood-widened channel and deposits veneers of sediment overbank.
Floodplain formation thus happens primarily in large floods. A section in the alluvium
shows a sedimentation sequence repeatedly interrupted by erosional unconformities
because of the numerous cut-and-fill activities. Bourke (1994) has described it as cha-
otic. A cross-section of the floodplain and channel commonly shows a series of steps at
different levels on the floodplain.
221 Running water in arid lands

(a) (b)

(c) (d)

Fig. 12.10 Example of floodouts. From Tooth, 1999. By permission of Wiley

5. The vegetation, including large trees, tends to occur in channels and on the floodplain
close to the banks due to the restricted availability of water. In flood, macroturbulent
scours occur round the base of such trees. In the case of the Todd River, it most com-
monly happens round the large River Red Gums (Eucalyptus camaldulensis), where
asymmetric elliptical scours become visible after the flood. The different levels on
the floodplain also give rise to irregular flow boundaries (Bourke, 1994; Bourke and
Pickup, 1999).
222 The arid tropics

Fig. 12.11 Schematic block diagram of the Todd River channel and floodplain. From Bourke and Pickup, 1999. By permission of Wiley

Thus, arid rivers, as demonstrated by Patton et al. (1993), Bourke (1994) and Bourke and
Pickup (1999), cumulatively display the effects of floods at varying scale. Low-frequency
catastrophic floods are too large for the river channel to accommodate the discharge. The
high-magnitude flood therefore erodes the entire channel and floodplain to make room
for the water to pass. Subsequently, the channel and floodplain are built up over time by
smaller floods depositing within the greatly enlarged channel. Bourke and Pickup pro-
posed channel forms occurring at three scales, which they termed small (within-channel
flows), medium (flows which extend across floodplains) and large (catastrophic and rare
floods). Figure 12.11 illustrates the channel forms of the Todd River and the comparative
size of three valleys excavated by floods at different levels. Bourke and Pickup (1999)
indicated that dimensions across the three classes vary by an order of magnitude. In the
Todd River, they expect longitudinal bars in these categories to measure 10â•›100 and up to
2000 m, respectively. The estimated return periods for the three categories of these floods
are 10, 100 and 1000 years for the Todd. The effects of the smaller floods are more vis-
ible in the headwaters and the piedmont reaches because of the transmission loss further
downstream.
223 Aeolian geomorphology of sandy areas

The description provided for the Todd can be extended in a generalised fashion to arid
regions on rocks or a thin surface veneer on rocks. Rainfall and floods are rare, but run-
ning water is still the important geomorphic agent, at least in the arid mountains and the
pediment zone. The form of the headwaters in the mountains is structure-controlled and
can vary, but the streams widen as they emerge from the mountains to the lower plains.
The channel and the floodplain reflect only the work of floods and may demonstrate a
nested appearance related to floods of different dimensions and recurrence intervals. The
channels remain dry between such events, exhibiting forms on their bed that were created
by the last flood. The rivers, however, are likely to disappear downstream in some form
of floodout.

12.7╇ Aeolian geomorphology of sandy areas

The extent of sandy areas varies between the arid regions of the world. For example, sands
cover nearly half of the arid areas of the Sahara, Arabia and Australia. In contrast, sand
dunes are localised in the Americas. It is mostly quartz sand, with a small proportion of
feldspars and iron or aluminium silicates that have survived weathering in the dry climate.
The sources of the sand are riverine plains, beaches, coastal lagoons and playas. Sand from
playas also contains evaporites such as gypsum. Large sandy extents carrying dunes of
various sizes and forms are known as sand seas or ergs. Small areas with relatively smaller
dunes are called dune fields.

12.7.1╇ Aeolian process

Sand grains are picked up, carried and deposited by wind. Bagnold (1941) is recognised as
the pioneer researcher in aeolian processes and dune formation. He studied sand transport
and dune forms both in the field (mostly in North Africa and West Asia) and also in the
laboratory. Satellite imagery has recently been added to these techniques as a major tool
in dune research.
There is a limit to the size of grains that the wind may pick up from the ground and the
nature of aeolian transport determines the grain size (Fig. 12.12). Particles smaller than
about 60–70 μm are transported by suspension in turbulent wind eddies. Of these, the very
small particles (<20 μm) can be carried in the wind for a long time and transported a long
distance. Particles, limited to 60–70 μm in size, however, are carried for shorter periods
and distances before being deposited as dust grains. Bigger grains, measuring 60–500 μm,
move by saltation, the most effective means of transport that is primarily responsible for
building dunes. Saltation is a series of downwind jumps. A sand grain is picked up in wind
eddies, rises steeply and then comes down in a flatter trajectory, displaying a parabolic
path. As it impacts the ground surface the force of the impact causes other grains to jump
forward, but they move a shorter distance and at less speed. This process is known as rep-
tation. The original grain may also recoil from the impact and finally come to rest after a
series of decreasing hops. The impact of saltating particles also causes grains bigger than
224 The arid tropics

Fig. 12.12 Diagrammatic sketch showing sand transport by the wind. Generalised from Lancaster, 1995 and Kale
and Gupta, 2001

500 μm to move or roll along the surface. This is creep. Most of the sand grains in deserts
move by saltation.
In sandy areas, a cloud of fine sand moves downwind by saltation. The efficiency of
this process depends on several factors. Sand grains move when the force of the wind (lift,
drag) exceeds the weight of the grain and the grain can be airborne. As the velocity of the
wind increases over a surface of sand grains, the grains start to vibrate and at a certain
velocity determined by the grain size, they start to leave the surface. The velocity of the
wind at which this happens has been termed the fluid threshold. However, the grains which
become airborne following an impact with a saltating grain start moving at a velocity lower
than the fluid threshold. This second threshold is known as the dynamic or impact thresh-
old (Bagnold, 1941). One saltating grain therefore causes an exponentially increasing num-
ber of grains to move. However, as the velocity of the succeeding grains are less, about
a tenth of the velocity of the impacting grain, a steady state is reached in seconds with a
given number of grains moving in unit time. Wind entrainment of grains is a complex pro-
cess that is influenced by local conditions, such as the size, shape, sorting and packing of
surface grains. The presence of moisture, surface crusts and surface roughness including
vegetation all act as local impediments. As a moving grain is commonly deposited on a
dune, grains usually travel or come to rest on a sloping surface. A detailed review has been
given by Lancaster (1995). Effective wind action only occurs close to the surface, usually
within about 1 m of it.

12.7.2╇ Aeolian forms of erosion

Wind erodes either by blowing at strength or by bringing fine grains to impinge on stand-
ing objects, especially if the surface material is not indurated. Wind erosion is the sum
of three sub-processes:€ deflation (removal of material), abrasion (scratching with fine
grains) and attrition (breaking down of the grains being carried in the wind by mutual
impact). The aeolian erosional forms are almost always smaller in scale than the depo-
sitional forms.
225 Aeolian geomorphology of sandy areas

Deflation, which implies carrying away of material, happens where the wind blows
across a surface of unconsolidated fines, especially sand and silt. Abrasion causes ero-
sion of individual rock clasts or upright objects. The impact of silt and sand grains being
carried by the wind may provide a rock clast with smooth facets or the opposite, eroded
pits or lines on its surface. Such rock clasts are known as ventifacts. Ventifacts are gen-
erally pebble-sized but may range up to boulders. The impact of sand grains also erodes
soft bedrock rising above the land surface, giving it a characteristic ridge-like appearance.
Such features are known as yardangs. Yardangs are usually not more than 100 m in length.
Bigger yardangs have been reported from the central Sahara and Egyptian Deserts. These
are up to 1 km long and locally cut into sandstones. Not much is known about these mega-
yardangs.

12.7.3╇ Aeolian forms of deposition

Aeolian forms are usually built by the deposition of sand grains forming small ripples or
big dunes. The small-scale depositional form of wind ripples is common in sandy areas,
occurring both on flatter surfaces and on dune slopes. They are created by saltation and
reptation. The ripples are elongated orthogonal to the wind direction, although local sur-
face slopes may partially change their orientation. Their size depends on the sand grains,
with most ripples having a wavelength of 50–200 mm and an amplitude of 5–10 mm. Both
measures increase with texture. For ripples in coarse sand or granules, the wavelength and
amplitude increase to 0.5–2 m and 10 cm or more, respectively. Wind ripples tend to be
asymmetrical with a steeper lee slope that ranges between 20° and 30° (Bagnold, 1941;
Sharp, 1963).
Sand dunes are the expected depositional forms in arid areas. These dunes are formed
from the interaction between sand grains on the surface and the shearing flow of the wind
close to the surface. Desert dunes have a dynamic similarity with the subaqueous dunes
discussed in Chapter 8. Dunes vary in size and form, which has led to a number of classi-
fications. Several of these also use the relationship between dune form and wind direction
as a classifying criterion. It is, however, easier to classify dunes solely on morphology,
although the relationship with wind direction does exist. Table 12.1 summarises the morph-
ology and distribution of different types of dune. Four major types (Fig. 12.13) have been
recognised. Minor forms also occur due to vegetative or topographic control on wind
deposition.
A specific dune may exist as a discrete feature, in which case it is a simple dune. If dunes
of the same type are juxtaposed, for example, a small crescentic dune climbs up the back
of a large crescentic dune, the entire feature is a compound dune. If dunes of more than
one type occur together, for example, a crescentic dune is found next to a linear dune, the
entire feature is a complex dune. A star dune, which has a peaked form like a pyramid, may
have flanks that are linear dunes. Compound and complex dunes are common in large sand
seas.
According to Lancaster (1995), dunes of sand seas have a hierarchical structure regard-
ing their size and spacing. Wind ripples with a spacing of 0.1–1 m cover about 80 per
cent of the land area in all sand seas. All sand seas also exhibit discrete simple dunes or
226 The arid tropics

Table 12.1╇ Morphology of dune types


Type and subtype Common location Wind regime

Crescentic dunes Margin of sand seas and small Wind directions have a
Barchans sand corridors:€Namib Sand Sea, narrow range.
Skeleton Coast dune field; total
area coverage small.
Crescentic ridges Covers about 40% of all sand seas. Wind directions have a
Also called transverse Dominant dune forms in the narrow range.
dunes or barchanoid Thar Desert; Takla Makan and
ridges. Simple, Teneggar Sand Sea; Jafurah;
complex and com- Nafud; eastern and northern Rub’
pound forms occur. al Khali; northern Saharan Sand
seas; Namib Sand Sea; and North
American dune fields.
Linear dunes About half of all sand dunes in Wind regimes unidirectional
Simple, compound and certain areas. Examples include or bi-directional, rarely
complex forms Simpson-Strzelecki Sand Sea, complex.
occur. Rub’ al Khali, Sinai, parts of
the eastern and northern Sahara,
Akchar Erg (Mauritania),
Kalahari Sand Sea and Namib
Sand Sea.
Star dunes About 8.5% of all dunes. Grand Multi-directional or complex
Erg Oriental, Rub’ al Khali, Gran wind regimes, especially in
Desierto, Erg Fachi-Bilma (Niger) the months when the winds
and Namib Sand Sea. do most work.
Parabolic dunes Found in coastal and semi-arid dune Not well known, probably
fields, not in large dune areas, the unidirectional wind regime
only exception being the Thar associated with some
Desert. vegetation.
Other dunes
Nebkhas Namib coast, semi-arid areas Originates from sand trapped
(also called coppice Australia, Kalahari. by clumps of vegetation.
dunes)
Lunettes Occurs downwind from small
playas.
Topography-controlled Occurs upslope or downslope
dunes of a topographic barrier to
wind.

Note:€Not all sand deposits are in the form of dunes. Zibers or sand sheets are low-relief sand
surfaces common in the Namib Sand Sea and Gran Desierto. Sand sheets occur in areas where
dunes cannot develop due to factors such as a high water table, coarse sand and vegetation.
227 Aeolian geomorphology of sandy areas

Fig. 12.13 Diagrammatic sketches showing different types of dune

compound and complex dunes at a spacing that falls between 50 and 500 m. Wider-spaced
(> 500m) complex and compound dunes occur in some sand seas (Lancaster, 1995). In
general, the dunes of a region display a common regular pattern, comparable spacing and
a size affinity.
Dunes are aeolian depositional forms that are built by sand grains. The mineralogical
composition of the grains depends on the source. Quartz and feldspar are most common,
as the usual sources are weathered igneous rocks and sandstones, but fine rock fragments
and volcaniclastic material also occur, especially if a short distance separates the source
area from the dunes. The colour of the sand varies. Dunes in certain locations, such as
the Simpson-Strzelecki Desert of Australia and the Kalahari in Southern Africa, display a
reddish tint, which commonly comes from iron oxides embedded in the minute pits that
occur on the surface of the grains. Dunes in other areas lack this reddening effect and, as
Lancaster (1995) has described them, they are paler.
Sand grains are deposited mainly in three ways. Dune grains avalanche down the lee
face of the dunes, leading to dune movement. Some grains remain in suspension for a short
period on the lee side in the shadow of the crest where there is less wind movement before
falling on the dune. Wind ripples tend to migrate up the windward side of the dune add-
ing sand grains. Each process exhibits an identifiable stratification when a section is cut
through the dune. The processes are also reflected in the texture of the sand grains.
228 The arid tropics

Wind ripples are in relatively coarse grains because the material moves on or near the
surface by saltation and reptation. In contrast, grains that are deposited from temporary
suspension in the lee of the dune crest are finer and better sorted. In general, most dunes
are made of very well to moderately sorted fine to medium sands. Finer sand, as expected,
occurs on the dune crests and coarser material forms the bases or plinths. Case studies
indicate that avalanching deposits form most of the body of crescentic and parabolic
dunes, whereas wind ripples are the important deposition process for linear and star dunes
(Lancaster, 1995). In some sand seas, areas between dunes cover a significant proportion
of the region, and such interdune areas can be bare rock, alluvial material or wind ripple
deposits.
In sum, the origin and development of different types of dune are related to variations
in wind speed and direction and the transport of sand grains. The airflow accelerates on
the windward side of the dune, eroding and transporting sand to the crest. The airflow can,
however, be varied and complex in lee of the crest, leading to a complex pattern of erosion
and deposition of sand grains. In spite of the complexity, dunes tend to be similar in morph-
ology over wide regions, indicating a relationship between dune morphology and regional
wind direction and velocity, and the availability of sand, so that a condition in dynamic
equilibrium is achieved. Five factors determine dune morphology:

• wind regime, especially wind variability and strength


• amount of sand that is available
• texture and sorting of dune sands; these are expected to have control over the size and
spacing of dunes
• vegetation, if present
• time.
Simple dunes have faster transport rates and are controlled by the annual or seasonal pat-
tern of wind speed and direction. In contrast, much larger complex and compound dunes
move extremely slowly and are not influenced much by local wind conditions. It is the
plentiful supply of sand that affects their morphology and existence. The life spans of sim-
ple (10–102 years) and complex or compound dunes (103–105 years) are strikingly different
(Lancaster, 1995).
The basic dune morphology (Table 12.1) is related to wind direction. Crescentic dunes
are formed where wind blows from a single direction; linear dunes are found in regions of
bimodal winds; and star dunes originate in complex wind regimes, where the wind may
blow from different directions. The minor varieties are related to local environmental
parameters, such as vegetation for nebkhas or parabolic dunes, and topographic barrier
for dunes that are described as climbing or falling dunes. Sand seas, which cover large
areas, thus carry different types of dune (Fig. 12.14) due to spatial variations in the wind
regime, wind direction and sand supply (Kar, 1987; Lancaster, 1995). The simplest pat-
terns are those of sand sheets, zibars and low crescentic dunes close to the source area of
sand. Zibars are low rolling dunes without slip faces, usually in coarse sand. The dune
forms become bigger and more complex with increasing distance, which allows for the
accumulation of thicker sand. The dunes thin out and become simpler further downwind
from the source area.
229 Aeolian geomorphology of sandy areas

Fig. 12.14 Different dune types in the Namib Sand Sea, Southwest Africa. From Lancaster, 1995. By permission of Taylor and Francis

The huge sand seas like the Namib or Australian ones that occur over stable cratons
have been formed over a very long period of time, and their extents have fluctuated fol-
lowing climate changes during the Quaternary (Chapter 16). Wind ripples are formed over
small areas in a very short time, almost instantaneously. Dunes may form within a range
of months to a hundred years and are at least in partial equilibrium with the rate and direc-
tion of sand movement governed by surface wind characteristics. Large sand seas evolve
in appropriate tectonic settings, over a much longer period, millennia to millions of years
(Lancaster, 1995).
230 The arid tropics

Fig. 12.15 Semi-arid landscape near Yulara, Northern Territory, Australia. Photograph A. Gupta

12.8╇ Conclusion

Arid and semi-arid tropics extend over a large proportion of the Earth’s surface. More
than half of this area is in rock or with a thin sedimentary cover over rock; the rest is in
sand. Physical weathering tends to mechanically disintegrate the surface material, except
towards the arid coasts where the presence of moisture, commonly as fog, and salt leads
to chemical weathering. In spite of the general aridity and the sporadic nature of precipita-
tion, running water as sheetflows or streams is the dominant geomorphic process on rock.
The channels and depositional forms of the streams of the arid regions, however, reflect a
nested arrangement associated with past flood history. They have been described as non-
equilibrium rivers. The sandy deserts are commonly in geologically stable areas, and sev-
eral of them have existed for a very long time. In spite of stable cores, the distribution and
extent of arid areas have fluctuated in time, especially with climate changes during the
Quaternary, and relict forms do occur. This is discussed in Chapter 16. Rocky outcrops
and ridges occur in sand deserts, but the sand landscape is essentially formed by wind
deposition:€a mosaic of wind ripples, dunes of varying morphology and interdune deposits.
Figure 12.15 from Australia is a visual summary of a semi-arid landscape. Low hills are
flanked by a rocky piedmont replaced by lowland where sand has been deposited to form
low dunes under desert vegetation.

Questions

1. Where are the arid tropics?


2. Why is it difficult to generalise statistically on the rainfall in the arid tropics?
231 Questions

3. Describe an arid alluvial fan.


4. Is the role of running water important in the arid tropics?
5. Do the rivers of the arid tropics differ in morphology and behaviour from those in the
humid tropics?
6. How do the smaller rivers of the arid tropics end?
7. Explain how the origin and development of different types of sand dune are related to
variations in wind speed and direction, and the transport of sand grains.
13 Tropical highlands

Towards the north is the Himalaya, the lord of the mountains, the abode of the gods.
Kalidasa, translated from Sanskrit

13.1╇ Importance of highlands

Highlands are geomorphologically important regions. A large part of a river’s discharge


and nearly all of its sediment may come from mountainous areas. These elevated areas also
impact on the regional climate, especially precipitation. Precipitation arrives as snow at
high altitudes, even in the tropics, and as rain on the lower windward slopes. Such precipi-
tation can be very high in a tropical storm or when winds converge near the Inter Tropical
Convergence Zone (ITCZ). Highlands may therefore supply a large proportion of a river’s
discharge.
In a seminal paper on big rivers, Potter (1978a) discussed sediment transfer in South
America from the active to the passive margin by large east-flowing rivers. From the tec-
tonically active margin to the passive one is a common direction for many large rivers
crossing continents. Tectonic activities in the mountains not only create high relief but also
increase the supply of sediment by providing fractured and brecciated rocks, steep slopes,
earthquakes and volcanic activities (Milliman and Syvitski, 1992). Where high stratovol-
canoes are present, erupted material could be the major source of sediment for regional
rivers, as in the Andes and the Indonesian islands. On Java, frequent eruptions from the
Merapi and other volcanoes provide the material for floodplains and the channel bars of
streams that flow south to the Indian Ocean. Volcanic sand even builds the coastal features
including large sand dunes (Fig. 13.1).
Sediment from highland slopes enters the valleys of small headwater streams (Fig. 13.2)
and is ultimately transported out of the mountains by larger rivers. A high proportion of the
sediment is deposited at the highland–lowland contact at the foot of the mountains where
it builds alluvial fans, which may have huge dimensions. The fans south of the Himalaya
have been described as megafans, the name indicative of the volume of sediment coming
out of these mountains. A large part of the Andean sediment is similarly deposited in the
foothills region east of the tropical Andes. The rest of the sediment is stored in downstream
floodplains or conveyed to the coast. It is the Andean sediment that builds the northeastern
coastal plains and deltas of South America (Meade, 2007). Smaller fans are found where
less sediment is eroded from the highlands, as at the foot of the mostly forested Main Range
of the Malay Peninsula. A large river system may collect discharges from all tributaries
232
233 Importance of highlands

Fig. 13.1 Sketch map of part of central Java showing the volcanic complex of Merapi, the south-flowing drainage and the south
Java coast. Pyroclastic material erupted from the volcanoes is the principal source of sediment for these streams.
Sand of volcanic origin has built the coastal landforms including the well-known black sand dunes of Perangtritis. The
height of Merapi changes between maps, as it depends on the latest eruption!

draining the entire basin, but it is only the tributaries originating in the mountains that are
important contributors of sediment (Fig. 9.1).
It is therefore necessary to evaluate the importance of tropical highlands and the proc-
esses that operate on their slopes. Three major dynamic processes operate in the moun-
tains:€ glaciation, slope processes and fluvial action. These are reviewed in this chapter.
234 Tropical highlands

Fig. 13.2 Small stream high in meltwater carrying a high sediment load, Lidder Valley, Kashmir Himalaya, very late in April.
Photograph: A. Gupta

Effective glaciation and related slope activities, however, are currently confined in the
tropics to the two massive mountain chains:€the Himalaya and the Andes.

13.2╇ Glaciation in tropical mountains

In spite of their location in the low latitudes, the higher slopes of several tropical moun-
tains are perpetually under ice. Examples of such mountains include not only the Himalaya
and Andes, but also mountains that are much smaller, such as the high volcanic peaks in
East Africa:€ Mount Kenya, Kilimanjaro and Ruwenzori. The present glacier covers on
the Kilimanjaro and Ruwenzori are about 5 and 4.5 km2 respectively (Osmaston, 1989a,
1989b). Glaciers occur above 4620 m on Mount Jaya in Papua (Indonesia), where three
small permanent ice caps with a total area of about 7.5 km2 still survive although in retreat
(Allison et al., 1989). Glaciation was much more extensive on all these mountains dur-
ing the Pleistocene (Chapter 16). For example, geomorphological and sedimentological
evidences indicate glaciation on the Aberdares and Mount Elgon in East Africa, two peaks
that are no longer under ice (Mahaney, 1989). Landforms indicative of past glaciations are
found even on the top of Mount Kinabalu, a 4131 m high granitic pluton in the Malaysian
state of Sabah on the island of Borneo at a latitude just above 6° N.

13.2.1╇ Glaciation in the Himalaya

The Himalaya Mountains are a series of parallel and asymmetrical ranges progressively
rising in elevation northwards as described in Chapter 2. The high elevation of the northern
line of ranges supports glaciation, and the asymmetry in slope and the east–west decrease
235 Glaciation in tropical mountains

Fig. 13.3 Glacier in the Western Himalaya. Photograph: A. Gupta

in precipitation control the size of the glaciers. In general, glaciers are smaller on the steep
southern slopes and in drier areas such as Ladakh. The glaciated area on the south-facing
slopes of the Baspa River valley in Himachal Pradesh is only a quarter of the size of the gla-
ciated north slopes (Vohra, 1981). Given favourable conditions, some of the glaciers reach
considerable lengths, e.g. Siachen (72 km), Biafo (62 km), Hispar (61 km), Batura (59 km)
and Baltoro (58 km), but most glaciers in the Himalaya are 1–5 km long. The Himalayan
glaciers were longer in the Pleistocene, and descended to a lower elevation (Chapter 16).
At present, they descend to about 4500 m in eastern Nepal and Sikkim, 4000 m in the cen-
tral part of the range and about 3700 m in eastern Kashmir further to the west (Fig. 13.3).
Glaciers develop best in the northernmost range€– the high Greater Himalaya€– and tend to
be concentrated in the vicinity of the great Himalayan peaks. These areas of concentration
include, from east to west, Namche Barwa; the area overlooked by Everest, Makalu and
Kanchenjunga; the Dhaulagiri–Annapurna–Manasalu region; Nanda Devi; and finally the
Nanga Parbat–Kolohai–Nankun stretch (Vohra, 1981).

13.2.2╇ Glaciation in the Andes

An asymmetry prevails regarding the presence of permanent ice between the two sides
of the Andes. The coastal mountains to the west are arid or semi-arid, and a lack of mois-
ture prevents glaciation there even at high altitudes. On the eastern side, glaciers thrive
236 Tropical highlands

on abundant moisture from the Atlantic and are found as far north as the Caribbean. The
Andes, as expected, are heavily glaciated towards the south beyond the tropics, but even
the tropical ranges are glaciated because of their high elevation and the moisture availabil-
ity. The semi-arid ranges that run along the Argentina–Chile and Bolivia–Chile borders
currently support glaciers above 6000 m on massive volcanoes, but further to the north,
beyond about 20°S, glaciers occur at a lower elevation. Extensive glacial systems are found
on the high massive peaks of Cordillera Real, Cordillera Blanca, Cordillera Huayhuash,
Cordillera Apolobamba, Sierra Nevada del Cocuy and Sierra Nevada de Santa Marta, but
small ice masses may occur on small peaks and narrow ridges rising to 5000 m and above
(Clapperton, 1983).

13.2.3╇ Glacial geomorphology of high tropical mountains

The high ranges of the Himalaya and Andes have been magnificently eroded by glaciation
into a landscape of huge mountain peaks, steep ridges, high vertical rock walls, and wide
and deep valleys. Obviously it is the valley glaciation that needs to be studied in these
ranges, as ice caps did not develop in recent geological history over any part of the cur-
rent tropical landmass. There are, however, complications to the general pattern, as local
slopes determine the character of glaciers. The Raikot Glacier near Nanga Parbat, western
Himalaya descends from about 7000 to 5000 m very steeply in a series of ice cliffs. It then
moves on a gentler slope of about 7° to its terminus at 4250 m (Shroder, 1993).

13.3╇ Mechanics of mountain glaciation

A glacier is a well-defined body of ice that moves either forwards or backwards, and rests on
either bedrock or soft sediment. Some snow exists all year above a certain elevation in the
high mountains. This elevation is known as the snow line or firn line. Both low temperature
and precipitation are required to accumulate enough snow at this altitude to form glaciers.

13.3.1╇ Formation of glacial ice

Freshly fallen snow has two properties. It has a low density of about 0.05 gcm−3 and a hex-
agonal crystal structure. The density of glacial ice in comparison is 0.85–0.9 gcm−3. The
transformation from snow to glacial ice occurs by three changes:€compaction, recrystallisa-
tion and the refreezing of the water derived from part melting of the snow. Compaction by
overlying layers and partial melting packs snow tighter and changes the hexagonal crystals
to near spherical forms. This produces granular ice with a density of at least 0.4€gcm−3,
which is known as firn. The transformation from snow to firn may take hours or days in
high tropical mountains. The transformation from firn to denser glacial ice takes much
longer, decades or more. Pore spaces inside firm are eliminated over time by recrystal-
lisation and the freezing of meltwater. Only a few air bubbles remain as the density of ice
approaches 0.9 gcm−3. There may be further changes to the ice, but such changes are minor.
237 Mechanics of mountain glaciation

Fig. 13.4 Mass budget and mechanics of motion of a diagrammatic valley glacier

The melting point of ice is pressure-dependent and varies according to depth. Ice therefore
melts at slightly different temperatures at varying depths through the glacier and not neces-
sarily at exactly 0°C. The term pressure-melting point is used to indicate this property.
Glaciers occur in different parts of the world and they have been divided into morpho-
logical and behavioural groups by several glaciologists. As our review is limited to glaci-
ation in tropical mountains we need to be concerned with only two types:€cirque and valley
glaciers. A cirque glacier is smaller; it occurs when moving ice in the mountains is confined
to an amphitheatre-shaped depression, known as a cirque. A valley glacier extends beyond
the cirque, occupying at least part of a valley. Detailed discussions on glaciers of all types
are available in various textbooks (Sugden and John, 1976; Bennett and Glasser, 1996;
Benn and Evans, 2010).

13.3.2╇ The glacial mass balance

A glacier is an open system. Ice is added to it by precipitation and lost from it primarily by
melting. A budget for the mass of ice in a glacier explains its behaviour and movements. Ice
is added by snowfall and rain that freezes on the glacial surface. Ice is lost in mountains by
melting, sublimation and being blown off the glacial surface in strong winds. The addition
of ice is known as accumulation and the loss as ablation. Accumulation mostly happens
at the upper part of a valley glacier and ablation at its lower section. These two zones are
separated by a line (or zone) of equilibrium where the gain and loss of ice is balanced. Thus
a valley glacier is thickest at or around the line of equilibrium in the middle, and tapers off
at both ends (Fig. 13.4). The glacial mass balance (also known as glacial mass budget) is
computed by determining the difference between total annual accumulation and ablation.
As expected, accumulation increases during both the winter and the season of precipitation,
and ablation rises in summer.
If the mass budget is balanced and accumulation equals ablation, the front of the glacier
remains almost stationary, although ice moves through its body compensating ablation at
its terminus. The advance or retreat of a glacier is directly related to the positive or nega-
tive mass balance. The rate of movement, both up and down the valley, is dependent of
the relative proportion of budget surplus or deficit with reference to the dimensions of
the glacier. A large glacier may not move much if the budget is slightly unbalanced. The
238 Tropical highlands

glacier, however, needs to overcome the friction at its base and the valley sidewalls in order
to move.

13.3.3╇ The mechanics of glacial movement

We can observe or measure the advance or retreat of a glacier, but our inability to see the
ongoing, processes at the base of ice is a major problem in understanding glacial move-
ment. Such processes have been inferred by deductive reasoning from the landforms and
sediment left by past glaciers which are no longer in existence, or by inspecting the sedi-
mentary evidence left as a glacier retreats upvalley. Activities at the base of glaciers have
also been studied by sending instruments down boreholes sunk into the ice and by observa-
tions round the edge of glaciers.
Ice, in order to move, needs to overcome (1) basal friction between the ice and rock or
sediment at the valley bottom on which it rests and (2) internal friction between the ice
crystals and layers within the glacier. Ice resembles a plastic substance in that it can resist
a force up to a certain limit, beyond which it deforms. Ice deformation is caused when
the ice is stressed (acted on by a force per unit area). There are two types of stress in a
glacier:€hydrostatic stress and shear stress. Hydrostatic stress at a point inside the glacier
is derived from the weight of the ice on top and is equal in all directions. Shear stress is a
combination of the weight of the overlying ice and acceleration due to gravity determined
by the slope of the glacier at its surface. A glacier moves according to the resultant between
the shear stress and resistance to movement.
Movement of a glacier carries two components:€internal deformation (also referred to as
creep) and basal sliding. In most glaciers, basal sliding is responsible for about 80 per cent
of its movement. Shear stress which causes internal deformation can be expressed as
τ = ρgh sinα, (13.1)
where
τ = shear stress
ρ = density of ice
g = acceleration due to gravity
h = glacier thickness
α = slope of the upper surface of the glacier.
Shear stress and internal deformation therefore increase with surface slope and ice thick-
ness, and also directly with depth inside the ice mass. Ice deformation is the mutual dis-
placement of ice crystals relative to each other. Similar orientation of most of the crystals in
a body of ice helps. The deformation can be summarised by Glen’s Flow Law of 1955:
ε = kτn (13.2)
where
ε = the strain rate
k = a constant related to ice temperature, the value of k increasing with temperature
τ = shear stress
n = an exponent. (Nye, 1957).
239 Mechanics of mountain glaciation

Fig. 13.5 Ice movement

Ice is deformed more at depth and at a relatively higher temperature. There are other
models for internal deformation but Glen’s Flow Law is a robust and succinct summary
of a part of glacial movement. In sum, ice in a glacier can be deformed internally by both
shear stress and hydrostatic pressure.
Problems arise when ice has to overcome an upvalley slope over an irregular bottom top-
ography or the thickness of ice varies between parts of the glacier (Fig. 13.5). Nye (1952,
1957) suggested compressive and extending flows for overcoming such obstacles. When
ice is compressed against an upvalley slope or ice thins as it nears the ablation zone, com-
pressive flows occur with a slowing down of velocity. In contrast, extending flows occur
where the bedrock slope steepens or the volume of ice increases. This increases the vel-
ocity of the ice. Both these cases of departure from the ambient velocity create a series of
slip planes within the ice so that layers of ice can slide downvalley. In difficult cases where
deformation cannot be adjusted within the body of ice, blocks of ice may fracture off and
travel downslope, especially near the end of a glacier. Fractures produced by compression
frequently open up cracks at the surface of the ice, known as crevasses.
Glacial movement is also helped by ice slipping over bedrock on a thin layer of water.
The water could be from meltwater or precipitation that has made its way to the bottom
via fractures and openings in the ice. Water is also derived from ice melting near the base
where the pressure-melting point has been reached. Temperature is higher towards the bot-
tom of the glacier, as heat derived from radioactive disintegration of minerals coming up
through the crust is trapped by ice. Furthermore, the water acts as a lubricant against basal
friction. It also partially offsets the weight of the overlying ice. Uncertainty exists regard-
ing the location of the basal water€– whether it occurs as a continuous film or is found only
in cavities in the rock.
The problem with irregularities on the rock floor of a glacier could be surmounted by
basal water. If the irregularities are small in size, basal ice pressed against them melts and
the meltwater flows round the irregularities to freeze on the downstream side, a process
called regelation, thus transferring the mass of ice downslope. If the irregularities con-
stitute a large barrier, the ice at the upper end of the obstacle comes under high pressure,
deforms and passes round the obstacle. This is called enhanced creep.
If glaciers are on soft sediment instead of hard rock, they can also move downslope by
pushing and deforming the sediment. A mountain glacier therefore moves by a combination
of internal deformation, basal sliding and sediment deformation. Higher up the peaks, in a
much colder area, the movement could be entirely dependent on internal deformation.
240 Tropical highlands

(a) (b)

Fig. 13.6 Idealised velocity distribution in a valley glacier

The average annual surface velocity of glaciers on the lower slopes of the mountains
could be 13–200 m (Ritter et al., 1995). The velocity may be much slower on the colder
upper slopes. A glacier flows fastest near the line of equilibrium where the ice is thickest. It
is slower towards both the colder upper part and the lower section where the ice is thin and
melting. Cross-glacier the fastest velocity is achieved in the middle, from where it drops
off to the sides where the ice thins and friction becomes significant against the valley walls.
Vertically, the fastest velocity is at the surface, but the vertical distribution of velocity is
non-uniform (Fig. 13.6).
Figure 13.6 is a diagrammatic and simplified presentation. Local variations in velocity
always occur as departures from this general model. Variations in velocity also happen over
time, for example, the velocity at the ablation zone increases in the summer, presumably
because of a higher production of meltwater which assists basal sliding. The most striking
short-term change is a glacier surge, which is a sudden, large-scale and quick transfer of ice
down a glacier. The velocity in a surge could be as high as 1300 times the average figure.
The surging glacier may persist for some time, up to several years, before returning to its
average flow rate. During this period, a huge amount of ice and debris will be transferred
downslope. This may result in river floods beyond the terminus of a glacier. Surging hap-
pens in several ways. It can be caused by:
(1)╇the build-up of a large volume of ice in the upper part of the ablation zone and its sub-
sequent collapse
(2)╇position changes and interlocking of subglacial channels carrying meltwater enhan-
cing basal sliding
(3)╇ the accumulation of volcanic heat underneath glacial ice on volcanoes.

13.4╇ Glacial forms and processes

Cirque basins and deep U-shaped valleys are glaciated erosional forms in the mountains.
The depositional forms are usually limited towards the end or snout of a glacier. We will
first look at the process of glacial erosion and the erosional forms, followed by depositional
processes and the forms found lower in the mountains.
241 Glacial forms and processes

13.4.1╇ Glacial erosion

Glaciers are active agents of erosion, capable of sculpting out huge basins and deep and
wide valleys set in rock. Glaciers erode by abrasion, using rock fragments which become
embedded at the base of the ice and scratch the underlying bedrock as the glacier moves.
This requires a large supply of rock fragments that may range from sand to jagged boulders.
The source of the rock fragments varies. They may be derived from frost-shattered rocks
that fall on the ice surface from the valley walls above the glacier, material that drops down
crevasses and pieces of rock that originate from the scratching of bedrock by fragments
already embedded in the ice. These fragments may be transferred to the glacier surface,
move inside the glacier or remain at its base. Glacial meltwater is also a tool, as it enters
cavities and cracks in the rock and refreezes due to variations in pressure. This loosens and
removes chunks of bedrock, especially if the original rock is highly jointed, fractured, had
other lines of weakness or had been weathered before being covered by ice. This is called
plucking. A glacier erodes primarily by abrasion and plucking at the rate of several milli-
metres a year. Faster rates of erosion are expected from glaciers that are thick and flow fast.
Cirques and U-shaped valleys in the mountains are created by glacial erosion.

13.4.2╇ Cirques

Glaciers start in cirques, which are deep and semi-circular erosional basins with a high
headwall, a scooped-out floor and a raised sill at the lower end (Fig. 13.7). Cirques have
been compared to an armchair, an armchair with a high back. In size, cirques range from
small depressions cut into the hillsides to very large amphitheatres which could be sev-
eral kilometres long, over a thousand metres wide and with a high headwall of a thousand
metres. The slope of the headwall is hardly more than 30° for the lower part, but the upper
section can be extremely steep, 70–80°. The sidewalls are commonly curved, but rectangu-
lar and triangular cirques have been reported in the literature. When a cirque contains ice, it
usually covers the entire floor and stretches to the lower part of the headwall where a deep
crevasse, called a bergschrund, occurs near the wall. At the other end, ice spills over the
lip of the basin to flow down a valley. A number of cirques are now bare of ice, displaying
the full dimensions of the rock basin. Water may collect on the floor of an ice-free cirque
to form a lake called a tarn.
The beginning of a cirque is still not fully understood but, once started, it expands by
erosion. At the beginning, snow may accumulate in a small pre-existing depression to form
firn and later ice. Rock fragments are released by plucking, and they move down to the
bottom end of the depression to form a ridge called the protalus rampart. All the processes
that operate at this stage are collectively known as nivation. A curved base identifies the
formation of a nivation cirque. As more ice accumulates and the true glacial ice is formed,
its erosive effect transforms a nivation cirque to a cirque glacier with the proper armchair
shape. Erosion is carried out by headwall recession due to frost shattering and rock failures,
and basal sliding of the ice already inside the cirque.
As two neighbouring cirques increase their areas, the intervening ridge is eroded from
both sides until it becomes a steep narrow wall known as an arête. Several arêtes converge
242 Tropical highlands

Fig. 13.7 Shape and origin of a cirque

to a peak whose sides act as headwalls to several cirques. Such peaks are also known as
horns. Nearly all high Himalayan peaks are horns.

13.4.3╇ Glaciated valleys

A valley which carries a glacier or used to carry one is a glaciated valley or a glaciated
trough. Such valleys display characteristic features. As ice flows out of a cirque it travels
down a valley eroded into the mountains by a river. River valleys are steep, deep and nar-
row and circumvent noses of alternating spurs (called interlocking spurs).
Glacial erosion transforms such valleys into wide U-shaped ones (Fig. 13.8). It has been
suggested that the transformation takes place in two stages. In the first stage, the V-shaped
valley is given a parabolic shape because of differential erosion at different parts of the val-
ley. Valley side slopes, where the velocity of ice (hence shear stress) is higher, are eroded
more effectively than the narrow bottom. In the second stage, zones of high and low ero-
sion shift as the valley becomes parabolic in shape, and ultimately the entire valley side-
wall comes under erosion giving rise to a U-shaped cross-section (Harbor et al., 1988;
Harbor, 1992). Thus the characteristic form of a glaciated valley emerges:€ steep-sided,
flat-floored and with all spur-ends truncated. The floors of the major valleys are lowered
faster than those of tributaries, the effectiveness of erosion being dependent on the volume
of ice. When the ice disappears, the rock floors of the valleys are exposed and the valleys
are occupied by streams. The tributaries tend to hang, giving rise to waterfalls. Such val-
leys are known as hanging valleys.
The longitudinal profile of a glaciated valley is not smooth but descends in a series of
steps, with long treads at the back of the steps. Such treads may slope upvalley. Plucking
continues to erode the bottom of the steps. As glaciers retreat upvalley and the floor is
243 Glacial forms and processes

(a)

(b)

Fig. 13.8 Valley glacier

exposed, a string of lakes may occur in the middle of the valley where water is ponded
between the steps and the reverse slope of the lower treads. Such lakes are known as pater-
noster lakes, as they resemble the beads of a rosary to an imaginative observer.
Cirques, arêtes, horns and glaciated troughs form the glaciated topography in the
Himalaya and Andes. An enormous volume of sediment removed from the mountains dur-
ing the formation of such features is carried to the lower slopes where ablation is in oper-
ation. The processes of glacial erosion, transport and deposition continue as long as the
glacier is active. Even after the disappearance of the glacier, a glaciated valley may con-
tinue to supply the mountain streams with a high sediment load from the stored material.

13.4.4╇ Glacial deposition

The erosive power of glaciers in the mountains generates a large volume of sediment which
is carried through the ice and deposited at various suitable locations:€the end of the glacier,
at its sides and underneath the ice on top of the bedrock. It is necessary to recognise the
difference between the sediment and the forms produced by its deposition. The general
term for the glacial sediment (deposit) is drift. Glacial deposits commonly carry boulders.
In the early days of glaciology, such boulders were mistakenly identified as deposited not
by glaciers but drifting icebergs. The idea has been corrected but the term remains. Glacial
drift varies in thickness, being best developed in areas beyond the tropics where ice sheets
once prevailed during the Pleistocene. We are dealing specifically with glaciations in the
tropical mountains where the drift is thinner and restricted to valley floors and near the ter-
minus or snout of the glacier.
The character of the deposit varies depending on whether it has been deposited only by
ice, by a combination of ice and meltwater or reorganised completely by meltwater after
deposition. As the glacier advances and retreats along a valley, the sediment is considerably
244 Tropical highlands

modified. Two broad classes of drift are recognised:€non-stratified and stratified drift. Non-
stratified drift is primarily deposited by ice, whereas in stratified drift the sediment is first
deposited by ice and then modified by water. Hence stratified drift exhibits sorting and
stratification, differentiating it from non-stratified drift. Non-stratified drift is also known
as till or by an old term, boulder clay, which seems to have fallen out of favour but which
describes the wide textural range of till.
Till is extremely variable in character, but shares a few common properties:€absence of
stratification, absence of sorting, striation marks on clasts, grain angularity and a generally
bimodal texture. Some rounding of clasts, however, happens if they are transported over a
distance embedded in basal ice. In contrast, clasts that have fallen from the valley walls and
been carried on top of glacial ice are angular.
Stratified drift involves transportation by both glacial and fluvial processes and is rec-
ognised as a fluvioglacial deposit. It displays limited sorting, stratification and rounding of
clast edges and corners. Stratified drift is deposited by meltwater channels inside the gla-
cial ice, inside localised depressions and openings as between the rocky valley walls and
glacial ice, and in a washed-out form in front of the glacier.
Both types of drift produce a variety of depositional forms of varying size and internal
structure. We only review the forms commonly seen in the mountains:€a full coverage of
glacial deposition is available in textbooks on glaciology, which indicate many of the other
depositional forms associated with ice sheets at higher latitudes. The depositional forms
could be ice-contact features, i.e. formed in contact with the sides and front of the glacier,
or proglacial features, which are found in front of the glacier deposited by meltwater issu-
ing from the snout of the glacier. Glaciers tend to move forwards and backwards at the
lowest end in response to variations in accumulation and ablation. As a result, ice pushes
and thrusts sediment into ridges and mounds, while the zone of proglacial forms also shifts.
The zone near the glacial snout is characterised by a complicated assemblage of a range of
depositional forms.
Perhaps the commonest depositional form in mountain glaciations is the moraine, a
ridged glacial deposit that does not follow the local topography (Fig. 13.9). However,
ground moraine, which is areally extensive, is not a ridge. As ice disappears, glacial deb-
ris settles on the bedrock floor of valleys to create a depositional sheet, which is ground
moraine. Ground moraine is pushed into ridges several metres high near the edges of a
glacier. If it happens at the side of a glacier, it is called lateral moraine. Where two glaciers
join, two adjacent lateral moraines come together to create a ridge in the middle of the
combined ice to create a medial moraine. Moraine ridges, however, are best developed at
the end of a glacier where several ridges may be found; these are collectively known as
the end moraines. The end moraine that marks the furthest advance of the glacier is the
terminal moraine, which curves round the edge of the glacier to link up with the lateral
moraines (Fig. 13.9). All other end moraines are grouped together as end or recessional
moraines, marking locations where the glacier has stopped temporarily while retreating
and a moraine ridge has been formed. A subsequent advance may destroy moraine ridges,
at least partially. End moraines are also modified by meltwater issuing from the glacier.
Ice-contact features are formed at the side or snout of a glacier ending in a terminal zone
of thin and slowly disappearing ice. These features will be in stratified drift and formed by
245 Glacial forms and processes

(a)

(b)

Fig. 13.9 Moraines

meltwater streams. Deposition by meltwater may happen in channels on top of the glacier,
inside or under the ice, in a linear gap between the rocky sidewall and glacial ice, and near
a mass of stagnating and melting ice. Once the ice disappears, the sediment resettles and
develops structural contortions. The final forms of these ice-contact features are mounds of
stratified sand and gravel, collectively known as kames. Kames could be 102 m long and
10€m high. The form of a kame follows the depression in which meltwater had collected and
sediment deposited. If deposition occurs in a linear cavity, kames resemble ridges. A delta-
like feature is formed when a meltwater stream flows out of the ice depositing sediment.
This is the kame delta. When deposition occurs between the rocky sidewall and the ice, the
sedimentary form is left with a steep side where the ice used to be. This is the kame terrace.
Kames frequently occur on the valley floor with depressions called kettles, constructing the
kame and kettle topography. Kettles are formed when a mass of ice is trapped in the ground
sediment and melts. Kettles may form small lakes. Long meltwater channels through the
body of ice may deposit a linear winding body of sediment which is revealed when the ice
melts. This is the esker. Eskers, however, are best developed from ice sheets.
Proglacial features are found beyond the terminal moraine, where large quantities of
meltwater issue from a glacier carrying a high sediment load. The sediment is deposited
both inside the meltwater channels and overbank, building up a plain of washed-out gla-
cial sediment. This plain built by glacial outwash is known as a sandur, an Icelandic term.
The channels tend to continue beyond the sandur as the headwaters of a river system. A
sandur deposited within a valley bounded by rock walls is known as a valley sandur (val-
ley train in North America). If it terminates in a plain, as happens in the higher latitudes, it
is known as a plain sandur (outwash plain in North America). Several levels may be seen
in a sandur, related to several advances and retreats of the glacier upvalley. As expected,
sandurs display sorting, with coarse grains being deposited near the glacier and finer
material downvalley. Morphological changes also occur longitudinally. A sandur carries
one or a limited number of channels and a pitted appearance from kettles at its upper end.
Downslope, channels become multiple and braided with a wide and shallow cross-section.
246 Tropical highlands

These channels are mobile and a number of abandoned channels mark the surface of the
sandur. At the very distal end, the channels become very shallow and, when in flood, the
entire sandur could be under a sheet of water. Large meltwater floods modify the morph-
ology of sandurs periodically.
The morphology of high tropical mountains is thus modified by glaciations giving rise
to a craggy appearance. But the effect of the glaciations extends further in (1) controlling
the hydrology of streams originating in the mountains and (2) being an important source
of sediment transported by these streams. The valley floors in tropical high mountains
are characteristically covered with coarse clastic sediment that not only forms bars inside
mountain channels, but also floodplains, terraces and fans overbank. Sediment is also
derived from mountain slopes in mass movements. A large volume of sediment arrives in
rivers from the glaciated higher slopes.

13.5╇ Slopes and valley floors in high mountains

Below the high glaciated mountains, slope failures are a major supplier of sediment to
streams. The large mountains are tectonically unstable due to their association with conver-
gent or strike-slip plate boundaries. Slopes fail periodically due to such tectonic movements
and also following high rainfall events. If these are volcanic mountains, then pyroclastic
eruptions and lahars bring down vast amounts of material along river valleys.
Brunsden et al. (1981) described the geomorphological character of the lower slopes
of the Himalaya in Nepal. The terrain is a mosaic of deep trunk valleys, long and steep
(20–50°) side slopes and a dense network of steep V-shaped tributary valleys. The major
river valleys include channels laden with sediment, fragments of successive river terraces,
tectonically elevated and variably tilted fans, and boulder fields backed by major scars on
the hillslopes. The slopes also display a network of gulleys, especially if deforestation and
agricultural practices are common. Although mass movement is the major denudational
process, considerable sediment is brought down to the major rivers by surface and subsur-
face water erosion. High-magnitude floods in the trunk valleys tend to move most of the
sediment at intervals. This general description holds true for most of the high mountains
in the tropics.

13.5.1╇ Mass movements in the high mountains

Shroder and Bishop (1998) attributed the Himalaya with being the ideal laboratory for the
study of high-magnitude slope failures at both low and high frequencies. They indicated a
general relationship between altitude and types of failure in the Himalaya. Avalanches tend
to be common between 9000 and 4000 m, rockfalls and slides between 7000 and 2000€m,
debris flows between 4000 and 2000 m, and mudflows and earthflows between 3000 and
1300 m. In general, the failed material is finer and wetter downslope. Such failures contrib-
ute an enormous amount of sediment. Earthquakes in mountains also contribute episodic
instances of large quantities of sediment to streams. Goswami (1985) recorded repeated
247 Rivers in the tropical mountains

additions of extra sediment into the Brahmaputra River from tectonic slope failures in the
eastern Himalaya.
The massive slides and flows may block rivers in the mountains, giving rise later to dam-
burst floods. A commonly used example of such an event is the damming of the Indus in
the winter of 1840–41 by the earthquake-caused collapse of a section of the Nanga Parbat
in the river. The lake formed upstream of the barrier was 64 km long and over 300 m deep.
The dam was breached the next June by the Indus and a tremendous flood swept down the
river. Rising to 30 m in height, this flood washed away part of a Sikh army camping on the
floodplain at Attock, 400 km downstream of the dam.
Apart from episodic seismic disturbances and the long-term general degradation of steep
slopes leading to collapse, mass movements that arise out of high-intensity rainfall in large
storms erode mountain slopes and bring an enormous amount of sediment to the valleys.
Starkel (1972) described one incident of high-magnitude rainfall from a Bay of Bengal trop-
ical cyclone that, during 2 to 5 October 1968, deposited 700–1130 mm of water in the eastern
Himalaya, near Darjeeling. The valley flats of the Tista and Great Rangit rivers were blanketed
with very coarse material including a high proportion of boulders. Such sediment is deposited
not only inside the channel or on top of the valleyflat surface, but also accumulates at breaks
in slopes to create alluvial fans that may grow to huge dimensions. The material contrib-
uted by lahars can be very large, e.g. following the 1985 eruption of Mount Huascaran in the
Colombian Andes or periodically from Mount Merapi in central Java (see Chapter 14).
The huge amount of material brought down the mountain slopes is stored in the valleys
as fans, terraces and floodplains and also within channels as bars of various sizes (Fig.
8.6). As the rivers emerge from the mountains, they deposit a considerable amount of this
sediment at the break in slopes at the highland–lowland contact to build fans (Bruijnzeel
and Bremmer, 1989). Sediment is carried further downstream and stored temporarily in
channels and on floodplains in mixed bedrock and alluvial valleys.

13.6╇ Rivers in the tropical mountains

From the first-order headwaters to the major rivers that emerge from the mountains, the
general function of mountain streams is to convey water and coarse sediment over a steep
gradient to the plains. According to Douglas and Guyot (2005), tectonically active regions
provide very high sediment yield, of the order of 104 tkm−2yr−1. The nature of the sediment
depends on the regional geology, which includes active subduction zones, collision belts,
volcanic arcs, granitic plutons, rift valleys and uplifted sedimentary rocks (Scatena and
Gupta, in press). Many of the rivers carry runoff derived not only from glacier meltwater
in late spring and early summer, but also from a very large amount of annual precipitation
that may reach thousands of millimetres. The effect of this large rainfall is even more pro-
nounced in the seasonal tropics, where about 80 per cent of the annual rainfall may arrive in
4–5 months (Gupta, 1988). The steepness of the channel gradient and the coarseness of the
bed and suspended loads may leave their imprints on the channel forms in the mountains
by enhancing the seasonality in form and function of river channels.
248 Tropical highlands

Fig. 13.10 The Swift River in the Blue Mountains, northeastern Jamaica displaying the characteristic features (see text)
of a montane river. Photograph: A. Gupta

Episodic rainfall, as from tropical storms, may be very high. As expected, such rainfall
gives rise to large floods capable of transporting huge volumes of sediment downvalley.
Such sediment is derived from previously stored material in the valley and new material
that has moved en masse down the steep mountain slopes as a result of the storm. The
effect of these rainfall events could be catastrophic as the big storms may shed rain over
an entire river basin within a steep mountainous area. The steep slopes, saturated soils
and intense rainfall common in the mountains lead to high-discharge events, large floods,
sediment transportation and channel modification. The floods are powerful enough not to
allow long-term log jams and accumulation of coarse wood debris (CWD) in the chan-
nels, as happens in the temperate montane streams. In tropical streams, such wood debris
that arrives following landslides and large floods is usually removed in months or within
a couple of years. Bare solitary tree trunks deposited on a channel bar in a flood are more
common. Scatena has observed that although several CWD dams produced by hurricane
defoliation and uprooting may last for five years in the Luquillo Mountains of eastern
Puerto Rico, the majority of these barriers are broken up and redistributed within less
than six months. CWD dams have not been observed in streams higher than second-order
(Scatena and Gupta, in press).Wohl et al. (2009) concluded that in the Upper Rio Chagres
Basin of Panama, large wooden debris produced by floods and landslides lasts for two
years or less. The fluvial system seems to alternate between short periods of stored wood
in the valleys and long periods without.
Wohl and Merritt (2005) suggested that montane streams have at least a gradient thres�
hold of 0.002 m m−1. The gradients of tropical montane streams are commonly well above
this threshold. Their longitudinal profiles tend to be punctuated by waterfalls and alternate
reaches of step and gentle gradients determined by bedrock (Hancock et al., 1998). The riv-
ers carry the usual features of montane streams:€bedrock channels, boulder bars, boulder-
lined channels, step-pools and pool-riffle sequences (Fig. 13.10). A veneer of alluvium
and boulders cover the bedrock, and channels are in many places best described as mixed
249 Sediment from tropical mountains

bedrock-alluvial channels (Whipple, 2004). A thick cover of alluvium is rare. In places,


channels may be full of boulders. Such boulders could be either exhumed corestones or they
could be transported to the channel by mass movements, especially debris flows, in large
storms (Ahmad et al., 1993). Terry (1999) described the passage of sediment in an oceanic
island with high volcanic relief that is subjected to high rainfall events. This account can be
taken as representative of sediment derived from similar island environments.
Burbank et al. (1996) commented on the poorly understood competition between bed-
rock uplift and erosion in tectonically active mountains. They refer to the extreme high
annual incision rate of the Indus River (2–12 mm) in its middle gorge near Nanga Parbat,
Pakistan. An equilibrium is maintained in this tectonically active region between bedrock
uplift and river erosion, with slopes adjusting with failures to the fast downcutting of riv-
ers. Gabet et al. (2008) studied the relationship between high rainfall and rapid erosion in
the much wetter Nepal Himalaya using a network of ten river monitoring stations. They
conclude that the average erosion rates increase with discharge and precipitation. They
also suggest that long-term rates in parts of the Himalaya tend to reflect the effect of past
glaciations and not the present precipitation pattern. Interestingly, working in Andes catch-
ment, Aalto et al. (2006) concluded that erosion rates are directly related to the average
basin slope and relief rather than runoff. Gabet et al (2008) suggested that short-term rates
peak on local conditions, such as high precipitation during the present or intense glaciation
of the past, but such variations average out over time to provide a regional uniform rate for
the Nepal Himalaya. These studies probably indicate that a large volume of sediment may
emerge from high tropical mountains with elevated rates localised over time and space.
The mountains need to be studied in greater detail.

13.7╇ Sediment from tropical mountains

As Figure 13.10 illustrates, a huge amount of material comes off the slopes in high moun-
tains and is stored as fans, terraces, floodplains and channel bars within the river valleys.
Glaciers and slope failures produce most of the sediment which is redistributed on the
valley floor by mountain streams. Volcanic material is added to the assemblage where
volcanoes are present, as in the Andes and the islands of Indonesia and the Philippines
(Chapter 14).
A very large proportion of material coming off tropical mountains is stored on the lower
slopes to build alluvial fans, especially if the contact of the mountains with the plains is
sharp. This tends to happen in tectonically active areas and case studies on alluvial fans
have been reported from tropical South America, Central America, northern India, Papua
New Guinea, northeastern Australia and possibly other locations. A series of large allu-
vial fans, including several especially dynamic ones, extend along the entire length of the
Himalaya (Fig. 9.6) and the Andes. These are known as megafans and are differentiated
from the common alluvial fans (Leier et al., 2005) by their large size (103 km2 or greater),
low gradient (commonly < 0.1°), typical sedimentary texture (boulders at the apex and pre-
dominantly silt and mud at toes) and depositional process (mostly by wandering channels
250 Tropical highlands

of a single river). Horton and DeCelles (2001) expect megafans to drain a catchment area
of about 104–105 km2 and to have a depositional area of 103–105 km2. Leier et al. (2005)
stated that although alluvial fans may occur anywhere at breaks in the slope, fluvial mega-
fans are found within 15–35° of latitude in both hemispheres, indicating an association
with seasonal discharge, large sediment and a mountain front facing a large depositional
basin that allows space for the megafans to form.
The modern fluvial megafans, of the Rio Grande, Rio Parapeti and Rio Pilcomayo,
emerge from the central Andes in Bolivia to debouch onto the low-relief Chaco Plain.
These measure 5800–22 600 km2 (Horton and DeCelles, 2001). Alluvial fans of limited
size and catchment areas occur between two neighbouring megafans. The present active
channel for these fans has a straight pattern. Sedimentation in a channel occurs as mid-
channel and side bars of sand, whereas overbank deposition is characterised by crevasse-
splays and fine sediment. Abandoned divergent channels, suggesting frequent channel
avulsions, are found overbank and the main channel of a megafan remains stable only for
a brief period. Other Andes megafans occur towards the south in Argentina. The mobility
of the main channels may be related to large-scale river capture or lateral shifting in large
floods (Baker, 1978; Horton and DeCelles, 2001).
Megafans are important stores of sediment off tectonically active areas in the seasonal
tropics. Their construction is probably due to short-term deposition in one location fol-
lowed by avulsion, which shifts the centre of deposition to a new location. Smaller and
commoner alluvial fans develop in the tropics where the necessary attributes for megafan
formation are absent. Fine-textured sediment, low frequency or the absence of tectonic
movements, and a non-seasonal river discharge lead to the formation of a number of small
steep fans, for example, below the Main range of the Malay Peninsula, at the base of sand-
stone escarpments in tropical Australia and at the foot of Indonesian volcanoes.
The best-known of the Himalayan megafans (Fig. 9.6) is that of the Kosi River. A num-
ber of studies have described the properties of the Kosi Fan (Gole and Chitale, 1966;
Singh et€al., 1993; Wells and Dorr, 1987). The fan measures 154 × 147 km and longitu-
dinally slopes 0.89–0.06 mkm−1. The cross-fan slopes range from 0.1 to 0.25 mkm−1. The
area is affected by the Indian monsoon system and the regional annual rainfall averages
1300–1800 mm. Over 80 per cent of this arrives between June and October. Individual
large storms and considerable variability in the annual total are both superimposed on this
general pattern of rainfall. Large volumes of clastic sediment are delivered by the Kosi
from the Himalaya.
Singh et al. (1993) have recognised four distinct zonal distributions of facies in the depos-
its of the Kosi megafan:€zone 1 (gravelly, sandy, braided channels), zone 2 (sand, braided
channels), zone 3 (fine sand and mud, straight channel) and zone 4 (fine sand and mud, mean-
dering channel). Zone 4 has by far the longest reach. The deposits consist of multistoried sand
sheets interbedded with overbank muds. Individual sand sheets are commonly 8–10 m thick
and multistoried bodies measure 16–20 m. Sand is replaced upstream by gravel.
The Kosi has shifted 113 km westward in 228 years in episodic jumps of up to 19 km
a year. The reason for such shifting is not clear, although a number of explanations have
been put forward. Tectonism is a common cause of channel shifting in the Himalaya. Large
floods due to various factors have also caused river channels to shift. Shifting in floods is
251 Conclusion

feasible, as during the wet monsoon the Kosi turns into a 15–30 km wide river. The con-
cept of a 15 km wide river avulsing is difficult to comprehend. Unlike a number of rivers,
however, the Kosi shows a low statistical correlation between channel shifts and dates of
earthquakes or high-magnitude floods. Wells and Dorr (1987) suggested that the channel
of the Kosi shifts in annual floods in a random fashion, enabled by the flat gradient of the
fan. In 2008, following a major flood, the channel of the Kosi surprisingly shifted towards
the east.
The Tista River that flows from the Himalaya to the Brahmaputra has built an even
bigger megafan of 18 000 km2 that changes its slope from 0.19° at the apex to 0.01° at its
distal end. The slope across the fan is gentle as expected, about 0.01°. The river, like other
Himalayan rivers, demonstrates a pronounced seasonal pattern of discharge with an aver-
age annual discharge of 609 m3s−1 and the highest average monthly discharge above 2000
m3s−1. The fan displays several lobes, a radial framework of current channels and a network
of palaeochannels (Chakraborty and Ghosh, 2009).
The main rivers of the Himalayan megafans flow into the Ganga as a well-integrated flu-
vial system. Flowing into a major river, however, is not a global characteristic. The Andean
megafans tend to end in swamps and depressions. Megafans and large-scale swampy
depressions are also related in the Okovango, Southern Africa (Stanistreet and McCarthy,
1993). In such conditions, huge amounts of sediment remain stored in continental interiors
and do not reach the sea. Horton and DeCelles (2001) mentioned the deposition of a huge
amount of clastic sediment that is derived from the Central Andes between 18–22° S lati-
tudes on the Chaco Plain, primarily forming megafans.
Sediment that bypasses megafans finally reaches large rivers that collect the drainage
from high mountains. Sediment grains ultimately reach the sea, but in large rivers it may
take any length of time up to thousands of years to do so. Meanwhile, sediment is stored in
channel bars, floodplains and terraces along the valley. Several cycles of storage and trans-
fer could be required before an individual grain reaches the coast, where it contributes to
delta building or is deposited on a coastal shelf (Meade, 2007).
Very little sediment is derived from cratonic highlands in hard metamorphic rocks,
although tributaries draining such highlands may contribute a major part of a trunk river’s
discharge (Chapter 9). Isolated volcanoes may produce a considerable amount of sediment
for local rivers, mainly from pyroclastic flows and lahars (Chapter 14). Given the near-sea
location of subduction-zone volcanoes, a large part of volcanic sediment reaches the coast
quickly and tends to build up coastal features such as beaches and dunes. Volcanic material
draining east from the Andes is an exception. Such sediment travels a long distance along
the Amazon, like sediment of any origin.

13.8╇ Conclusion

Tropical mountains are the source of a large volume of water and almost the entire bed load
and suspended load of rivers that emerge from such high areas. The high sediment load
from the mountains is due to a combination of factors:
252 Tropical highlands

• glaciations
• extensive and frequent slope failures
• high and intense rainfall
• tectonics
• regional lithology and structure
• volcanic eruptions.
The sediment travels episodically. A large part of it probably stays stored for decades and
centuries in the foothills, forming fans and, further down the valley, forming floodplains.
In the case of large rivers, e.g. the Amazon or Ganga, sediment could remain stored for
thousands of years. Meade (2007) described the passage of the post-Pleistocene sediment
of the Orinoco and the Amazon from the eastern slopes of the northern Andes Mountains
to the Atlantic Ocean. Much of the Andes-derived sediment is stored for 102–103 years
or longer in the wide alluvial plains of the Andean foreland basin or the huge floodplains
along the mainstem of the Amazon and the Orinoco and a limited number of its large
tributaries. Meade attributes about 90 per cent of the sediment of the Amazon to the
Andes. He is of the opinion that the rest comes from the Andean foreland basins. In that
sense, all of the Amazon’s sediment is derived from a high tectonically active tropical
mountain range.
The source of the sediment transported by rivers also depends on the geology of the
highlands. The suspended load of the Nile is relatively modest, given its size. Most of the
load (72 per cent) arrives via the Blue Nile during the wet monsoon from the steep, vol-
canic Ethiopian Highlands. Relatively little is derived from the Central African Highlands
where the White Nile rises, and much of this load is then lost as the river traverses the flat
swamp of the Sudd (Woodward et al., 2007). Relative contributions of sediment from dif-
ferent parts of the drainage basin can be identified from chemical analysis of the isotopes
of Sr, Nb and Os in the sediment. It has been shown that almost all the sediment of the
Brahmaputra River comes from the Himalaya and 45 per cent from the part of the eastern
Himalaya known as the Eastern Syntaxis, where the upper Brahmaputra makes a U-turn in
a deep gorge around the Namche Barwa Peak in a region of heavy monsoon rainfall and
tectonic instability (Singh, 2007b).

Questions

1. Using Figure 13.1, describe the transfer of sediment from the volcanic slope of Merapi
to the sand dunes of Parangtritis in Java.
2. What are the depositional features associated with mountain glaciation in the tropics?
Do these act as significant sources of sediment to the regional rivers?
3. Read the description of the lower Himalayan slopes in Nepal by Brunsden et al. (1981).
Explain the origin of the morphological features described.
253 Questions

4. Shroder and Bishop (1998) described the Himalaya as the ideal laboratory for the study
of high-magnitude slope failures at both low and high frequencies. Why?
5. Describe the general character of tropical montane streams.
6. What are the form and function of megafans?
7. Tropical mountains act as the source for a large volume of water and almost the entire
bed and suspended loads of large rivers that start in the mountains. Why?
14 Volcanic landforms

The houses, people, traffic seemed


Thin fading dreams by day,
Chimborazo, Cotopaxi
They had stolen my soul away!
W. J. Turner

14.1╇ Introduction

The material discussed in Chapter 2 leads us to conclude that volcanoes occur in specific
locations. Such locations include convergent plate margins (for example, volcanoes of the
central Andes), islands that form higher parts of mid-oceanic ridges (Ascension Island in the
Atlantic) and volcanic islands formed over hot spots in the middle of the oceans (Mauritius
and Réunion) or continents (Tibesti, Chad). Volcanoes also occur in rift valleys. Mounts
Kenya and Kilimanjaro of the East African Rift are probably the best-known of the rift
valley volcanoes in the tropics. A number of less well-known ones, such as Nyamuragira
and Nyiragongo of the Virunga Mountains, Central Africa also form prominent features
on the regional landscape. Flood basalt erupted on a continental scale in the geologic past
over parts of the tropics such as the Deccan Plateau of India and the Drakensberg of South
Africa, giving these areas their characteristic landscape (Fig. 14.1).
In all these areas, the landscape is primarily determined by the lithology and relief created
by volcanism. The pattern persists even where geomorphic processes have considerably
altered the original rocks (Fig. 5.2B). Volcanic landforms are therefore an integral part of
the tropical environment. Almost 70 years ago, Cotton (1944) wrote a textbook on geo-
morphology indicating the importance of volcanoes in the title of his book. To an extent, the
book focused on New Zealand, where volcanoes are common in certain locations such as
the TaupoVolcanic Zone. An account of tropical landforms also requires a discussion of the
volcanic landscape for the same reason. We discuss the volcanic landforms of convergent
plate margins and flood basalts.

14.2╇ Types of volcano and the related landscape

A volcanic landscape is one of high relief but with a variety of forms. The type of vol-
canic form indicates its association with plate tectonics and lithology. Four broad types of
254
255 Types of volcano and the related landscape

Fig. 14.1 Distribution of volcanic landforms in the tropics

volcanic landform can be identified following the nature of the eruption involved:€mono-
genetic volcanoes, polygenetic volcanoes, shield volcanoes and calderas. A monogenetic
volcano is built by a single eruption; a polygenetic volcano is created by more than one
eruptive event; and a shield volcano is a large but gently sloping volcano, almost exclu-
sively basaltic, and commonly originating from a hot spot (Francis, 1993). Calderas are
formed by catastrophic events as described below. The following account is to a large
extent, based on Francis (1993) and Verstappen (2005).

14.2.1╇ Monogenetic volcanoes

Monogenetic volcanoes tend to build relatively small single cones of scoria or tuff and
small surface craters. Scoria refers to relatively light volcanic rocks with a fragmented,
cindery vesicular texture. Scoria cones are usually made of basalts, and are commonly
about 800 m in diameter and 200–300 m in height. They have relatively large craters for the
diameter of their edifice. Scoria cones are not always symmetrical, as they extend along fis-
sures or the ambient wind direction during eruption. If the magma comes into contact with
water during the eruption, shallow explosions blow out surface craters instead of building
cones. Such depressions later become lakes called maars. These circular depressions are
surrounded by low rims of ejected material. Tuff rings are similar features but they are built
by the accumulation of eruptive material in a circle above ground; they are not craters. Hot,
water-saturated material builds tuff cones, which resemble scoria cones morphologically
but are steeper and smaller. In contrast, scoria cones result from dry explosions. Once
formed, monogenetic volcanoes are subjected only to erosion, whereas polygenetic vol-
canoes may also experience several new explosions over time, rebuilding the volcano and
changing its appearance. Erosion is not the sole remodelling process.

14.2.2╇ Polygenetic volcanoes

Polygenetic volcanoes are much bigger and vary in their morphology and structure. Francis
(1993) listed simple cones, composite cones, compound cones and volcanic complexes.
Simple cones are scoria cones but with a history of multiple eruptions. They appear sym-
metrical with a single summit conduit and a small crater. Several well-known volcanoes
256 Volcanic landforms

fall into this category:€Mount Mayon, Philippines (2400 m) and El Misti, Peru (5822 m).
Volcanoes with a central vent have simple structures but eruptions may take place from
both the central vent that reaches the summit and lateral vents that feed eruptions on vol-
canic slopes; the latter phenomenon is known as the flank, parasitic or satellite eruption.
This happens in large volcanoes. Feeder eruptions may take place anywhere, but they are
very common in areas of crustal extensions or rifting and locations over hot spots.
Composite cones generally reflect volcanic activity at the same site, and hence the
edifices have a radial symmetry, but multiple eruptions give them a complex history of
volcanism. Compound volcanoes do not have individual cones that dominate, but their
massifs cover a large area which will include several cones, domes and craters. Volcanic
complexes are even more complicated. They incorporate an assemblage of several lava
flows, volcanic domes and pyroclastic rocks that are connected over space, time and origin,
showing vent migration over time. The Nevado Chuchuan Volcanic Complex near El Misti
in Peru is a good example.
The form of stratovolcanoes (volcanoes built by stratum over stratum of volcanic mater-
ial) may have complex forms due to lava outflows in mid-slope, flank eruption craters and
different slope processes operating over the middle or lower slopes. Faulting also results
in the alignment of eruption openings and thereby the shape of the volcano. Calderas are
conspicuous features in a volcanic landscape (Fig. 14.2). They result from violent erup-
tions that are accompanied by the collapse of the roof of the magma chamber or blowout
of a huge amount of material from the upper part of the volcano. Calderas can be 2–20 km
in diameter, with a circular or oblong appearance. The Toba caldera in north Sumatra is an
oblong caldera, elongated along a strike-slip fault.

14.2.3╇ Shield volcanoes

These are huge convex volcanoes (40 000–80 000 km3 in volume) with gentle (<10°) upper
slopes near the summit. Shield volcanoes are made of a large number of individual basaltic
lava flows and owe their origin to a hot spot underneath. The Hawaiian volcanoes are good
examples, especially Mauna Loa and Mauna Kea that rise more than 4000 m above the sur-
face and a total of 9000 m from the floor of the Pacific. Built on a thin oceanic crust, such
massive edifices tend to sag into the aesthenosphere (Francis, 1993). Huge landslides may
modify the shield volcanoes morphologically, as is happening on the southeastern flank of
Kilauea, Hawaii. The Galapagos Islands off Ecuador also provide good examples of shield
volcanoes.

14.2.4╇ A comparison

The dimensions and slopes of volcanoes depend on the composition of the construction
material and the eruptive stage. Volcanoes constructed of oceanic basalt tend to be large
and with gentle slopes. Andesitic and dacitic volcanoes of the subduction zones and island
arcs tend to be steeper and shaped by explosions due to the viscosity of their silicic and
alkaline magma.
The appearance of volcanic cones is also dependent on the erosion that follows the
building of the cone, especially from rain falling on fresh volcanic material (Fig. 14.3).
257 Lava and pyroclastic deposits

Fig. 14.2 The crater of the stratovolcano of Raung in east Java is about 2 km across. The volcano is high (3332 m) and erupts
regularly; it has already been active several times in the twenty-first century. Flow and explosive deposits and
screes of volcanic material mark the crater rim and the caldera floor. A second crater with a deep volcanic vent rises
asymmetrically from the floor of the caldera. IKONOS satellite image© Centre for Remote Imaging, Sensing and
Processing, University of Singapore (2009), reproduced with permission. See also colour plate section

V-shaped gullies, separated by ridges, cut into slopes displaying a radial pattern away from
the summit, give a pattern termed parasol-ribbing. A radiating drainage network with chan-
nels tens of metres deep develops rapidly, especially in tropical conditions with frequent
high-intensity rainfall as discussed in Chapter 3. This network is often very effective in
transferring volcanic material from the edifice to the surrounding area.
A set of volcanic landforms therefore not only incorporates individual volcanoes providing
high relief but also material erupted, eroded and transferred away from them, modifying the
surrounding landscape. The material is transferred by a number of processes:€volcanic flows,
mass movements and running water in channels down the slopes. If the coast is near, as in
Indonesia, the Philippines and the volcanic islands of the Caribbean, volcanic sand becomes
an important source of material for coastal landforms (Fig. 13.1). We need to examine the
processes that transfer volcanic material downslope and the nature of such material.

14.3╇ Lava and pyroclastic deposits

The morphological features of a volcanic surface depend on the nature of the lava and
pyroclastic deposits involved. Basaltic lava flows are usually grouped into two classes:€aa
and pahoehoe, both names derived from Hawaii. Aa lavas are more viscous, with a surface
258 Volcanic landforms

Fig. 14.3 Erosion on fresh volcanic deposits covering the slopes of Anak Krakatau (the child of Krakatau) growing at the same
location where the blown-apart volcano of Krakatau existed until 1883, between Sumatra and Java, Indonesia.
Photograph: A. Gupta

of irregular, sharp-edged blocks. Individual lava flows have two parts:€an upper part which
cools fast into rubbles and a lower part with a massive structure that cools slowly over a
scoria-rich base. Vesicles are seen in the lower part due to the escape of gases from the lava.
Pahoehoes are low in viscosity and flow fast to build a smooth, shiny, rope-like surface.
Andesitic lavas are more viscous than basaltic lavas. They travel a short distance from the
erupting vent, giving rise to large smooth-sided blocks on the surface. Dacites and rhyolites
form stickier lavas and are relatively rare, especially rhyolites. The surface expressions of
such lavas are often plugs and domes that rise vertically above the ground surface.
Pyroclastic deposits are products of volcanic explosions. The explosive act in a volcano
includes three stages. First, the magma is fragmented by coalescence and the bursting of
bubbles arising through it; second, the erupting mass is blasted through the vent to the
crater; and third, the erupted material rises as a column on top of the crater or spills over its
sides. The eruptions could be sustained jets (Plinian eruptions), short blasts separated by
near-instantaneous or longer intervals (Strombolian eruptions) or discrete blasts at inter-
vals (Vulcanian eruptions). The erupted material rises in a column high above the crater
that could reach tens of kilometres and enters the stratosphere before spreading radially to
an umbrella-shape that could be symmetrical or extended asymmetrically downwind. The
rate of eruption, and hence the eruption column itself, is controlled by the volatile content
of the magma, its exit velocity and the radius of the vent. The cloud consists of tephra.
Tephra is a general term for pyroclastic material that falls to the surface from an eruption
cloud. A rude reminder of the solid composition of this ash cloud is occasionally provided
when a jet aircraft flies through it and loses engine power. Volcanic eruptions also disrupt
air traffic schedule from time to time as after such experiences aircrafts normally try to
avoid eruption clouds. Most of the deposit falls to the ground as a grey blanket of fine-
grained material that varies in thickness, decaying exponentially away from the vent, and
259 Lava and pyroclastic deposits

Fig. 14.4 Eruption column over Pinatubo, 12 June 1991. The convective rise and the umbrella region at the top are visible.
Photograph by Dave Harlow, US Geological Survey

in many instances covering a wide area, but part of it may transform into flows by contact
with water along channels.
Tephra is divided into three size terms. If it is less than 2 mm it is described as ashes,
between 2 and 64 mm it is known as lapilli (little stones) and material coarser than 64 mm
is called blocks. Compositionally, tephra includes material from both the rising magma and
rocks from the walls of the vent, which are not necessarily volcanic. There are several types
of pyroclastic deposit.

14.3.1╇ Ash fall deposits

Major air-fall deposits result from Plinian eruptions. Plinian eruptions are huge eruptions
where high sustained exit velocities from the vent lead to strong convective columns that
rise into the stratosphere. Plinian tephra usually consists of pumice clasts of dacitic to rhy-
olitic composition, and Plinian eruptions generally occur at converging plate margins where
silicic material is available from the continental crust. Exceptions, however, do occur.
Pinatubo Volcano in Luzon, Philippines (Fig. 14.4) erupted in June 1991 with ash col-
umns that rose for about 40 km (Newhall and Punongbayan, 1996) with the paroxysm
peaking on 15 and 16 June. The eruption, however, lowered its summit elevation from
1745 m to near 1400 m, and left its top with a 2 km wide steep-sided caldera. The ash fall
buried more than 4000 km2 of area with tephra. The volume of the tephra has been esti-
mated to be between 8.4 and 10.4 km3, most of it loose sand-textured andesite–dacite frag-
ments. The tephra fall and pyroclastic flows filled a number of radial valleys, overtopped
the divides and created a new topography. The loose material is mobilised after rainfall
to create dangerous mudflows or lahars. The volcano has not erupted like this for several
hundred years (Francis, 1993; Newhall and Punongbayan, 1996; Nossin, 2005).
260 Volcanic landforms

14.3.2╇ Pyroclastic flow deposits

Pyroclastic flows are mixtures of rock fragments and gases that descend volcanic slopes hug-
ging the ground during explosive eruptions. They may originate from several causes:€par-
tial or complete collapse of an eruption column, collapse of an active lava dome or flow
and spilling of ash from an active crater. Pyroclastic flows travel at the high �velocity of
10–65€ms−1 and reach a temperature between 100 and 800°C (Blong, 1984; Cas and Wright,
1987; Francis, 1993).
Pyroclastic material ranges from vesiculated low-density pumice to dense clasts of lava
or dome fragments. The deposits left by pumice flows are known as ignimbrite and that
from the denser non-pumice type is called block and ash deposits, a term which describes
their textural characteristics. Correct terminology is important here to avoid confusion
between flows (process) and deposits (sediment). Pyroclastic flows of very low density
are called surges.
A reduction in the volatile content leads to instability in the erupting column. So does an
increase of the radius of the vent beyond a certain size. In such cases, pyroclastic material
begins to fall from the column (Francis, 1993). This may imply that the eruption column is
overloaded with too much pyroclastic material and convection cannot be sustained. Francis
(1993) attributed the mobility of a pyroclastic flow to two causes:€(1) transformation of
potential to kinetic energy due to fall from a considerable height in the eruption column
and (2) fluidisation of the material. A huge quantity of very fine material in the flow is
fluidised to act as a lubricant to transfer the large clasts. As fluidisation takes place, large
pumice clasts, which are of low density, float towards the top, whereas the fine-grained
material concentrates towards the bottom. This segregation structure is seen in deposits
of ignimbrite. Ignimbrite can therefore travel over a large distance in a process similar to
that of rock avalanches (see Chapter 6). The material descends to the surface of the ground
and spreads round as flows of pyroclastic material because of its high speed and large dis-
charge. This is pumiceous ignimbrite which is being deposited. It consists of bits of pum-
ice, dust and gaseous matter.

Ignimbrite
After deposition, ignimbrites can undergo welding and be anything from loose sandy ash to
a solid rock with the appearance of glass, termed vitraphyre. They can cover a wide range
of areas and a large-scale ignimbrite deposit may give rise to a distinctive topography.
They often give rise to near flat-topped plateaus dissected by channels. Ignimbrites are
gravity-driven flows and therefore tend to concentrate as flat-topped deposits in valleys
and lowland. Its distribution, unlike that of ash fall, is guided by the topography. Only in
exceptional cases does ignimbrite climb over topographic barriers. Even then the thickness
of the deposit varies considerably between the topographic highs and the valleys. Sumatra
has striking ignimbrite deposits from past pyroclastic eruptions. Canyon-like river valleys,
for example, have developed near Bukittingi in West Sumatra, between it and the neigh-
bouring village of Kota Gadang, which is well known for silver artefacts. Rivers have
eroded ignimbrite to create valleys with flat floors and cliff-like walls. The river itself is
261 Lava and pyroclastic deposits

entrenched within the flat valley floor. Landforms in ignimbrite are also prominent in north
Sumatra, where there have been several huge explosions from the Toba Volcano. The last
eruption about 75 000 years ago released 2000 km3 of ash which reached India. This cre-
ated Lake Toba, which is the subsiding caldera of the old volcano with the subsequent rise
of the island of Samosir in its middle. This is an example of a resurgent caldera. Ignimbrites
from the catastrophic 1815 eruption of Tambora formed vertical coastal cliffs on Sumbawa,
Indonesia.

Pyroclastic surge deposits


Pyroclastic surges have been described by Francis (1993). They consist of low-density
dilute material. As a result, pyroclastic surges travel with high velocity, are turbulent and
less constrained by topography. Fast surges that emerge radially from the base of the
erupting column are known as base or ground surges and are probably the best-studied.
In general, they form a wedge-shaped body that thins from the volcano with cross-bedded
dune structures in sand that change with distance into structureless material. This is then
replaced by planar deposits of fine grains at a further distance. Surges are extremely vio-
lent and turbulent, and erosive features are common in the deposited sand. The destructive
effects of the pyroclastic surges have been studied in detail for the 1965 one from Taal in
Luzon, the Philippines.
Taal is a volcano south of Manila on an island within a lake. There is another small
lake inside its caldera. It has an explosive history, but the region is highly populated with
paddy fields and fishing in the large lake. The eruption took place on 28 September 1965
when for 48 hours powerful surges swept out from the base of the erupting column. It
killed about 200 people, damaged trees on the mainland and left thick mud around the
trunks.

Block and ash flow deposits


Denser block and ash flows are deposited by pyroclastic flows from a hot avalanche, often
described as nuées ardentes. Pyroclastic flows on the Merapi in 2006 are described later in
the chapter. Block and ash deposits, as the term describes, include magmatic clasts in an
ash matrix. They flow under gravity and are therefore usually confined in valleys radiating
out from volcanoes. The material originates from viscous andesitic and dacitic lavas of
convergent-plate volcanoes. The clasts can be as big as boulders, metres in diameter. Such
boulders are often seen on top of the block and ash deposits. The front of a pyroclastic flow
is turbulent and highly fluidised. This leads to finer particles being elutriated up into the
ash cloud. The cloud then descends down the valley with a proclastic flow emerging from
its base. This gives rise to thin and fine-grained ash-fall deposits which are not preserved
as well as the denser flows.
Pyroclastic deposits therefore cover the slopes of a subduction-related volcano, with
thicker accumulation in the radiating valleys. This material with the addition of water gives
rise to volcanic debris flows called lahars, a term from Bahasa Indonesia. The material in
a dry state may also avalanche down the slopes of a volcano.
262 Volcanic landforms

14.4╇ Volcaniclastic flows:€debris avalanches and flows

A continuum exists between dry debris avalanches and wet mudflows called lahars. Debris
avalanches can occur in a range of sizes; almost an entire side of a volcano failed at Mount
St. Helens in the May 1980 eruption. Francis (1993) listed three types of collapse event
that lead to debris avalanches:€magmatic eruptions that blow out the side of a volcano, non-
magmatic explosions due to phreatic steam escapes, and earthquakes and any other event
which may cause the volcanic slope to fail in a cold state. The areal size of the large debris
avalanche deposits range from 1 to 20 km3 (Francis, 1993).
Past failures are indicated by distinctive crescent-shaped scars on the mountainside,
although later eruptions of lava may repair the damage. The debris avalanche deposits,
however, give rise to a characteristic hummocky terrain at the base of the volcano. The
oft-quoted example of such terrain is the 250 km2 on the lower slopes of the Galunggung
Volcano in west Java, popularly known as the ten thousand hills of Tasikmalaya. The 1982
eruption of Galunggung is notorious for blocking the engines of aircrafts that flew through
the eruption column. About 500 km2 of the surface of the Atacama Desert is covered by the
deposit of an avalanche that occurred on the Socompa Volcano, north Chile. Kilometre-size
blocks, known as torevas, were detached and slid downslope. In general the blocks slid
undisturbed. Part of the mass has been estimated to travel at velocities of 100–200 kmhr−1
and reach a distance of 35 km, but the original stratigraphic features were maintained
within the deposit (Francis, 1993). In general, the finest grains in a debris avalanche occur
at its base with boulders towards the top. Most of the material tends to travel as a plug with-
out much disturbance. Their long travel tracks have been explained by fluidisation and the
conversion of potential energy to kinetic energy, as discussed earlier. The presence of any
hydrothermally spread surges, ignimbrites or water acts as a lubricant.
Lahars are volcanic mudflows that can be described as slurries that carry boulder-sized
material and mud (Fig. 6.6). They originate from water coming into contact with tephra
that has accumulated on volcanic slopes and especially with the material in the channels
cut into the slopes of a volcano. Hence, material from pyroclastic flows deposited on upper
slopes may build lahars on lower ones. The arrival of water could be from rainstorms, the
spilling of water from crater lakes, snowmelt on the high volcanoes and pyroclastic flows
reaching a stream.
Lahars are common on tropical volcanoes during the wet season. The heavy monsoon
rainfall over the volcanic slopes of Indonesia and the Philippines and many other places
in the humid tropics periodically give rise to destructive lahars which may travel at a very
fast pace. The speed of lahars has been listed up to 90 kmph, but is usually much less at
4–8€ms−1. A tropical cyclone in 1991 that caused heavy rain to fall on the newly erupted
ashes on the slopes of Pinatubo gave rise to catastrophic lahars over a wide area that
destroyed life and property. Unlike Pinatubo, several volcanoes remain active in Southeast
Asia and on these volcanoes, such as Merapi, lahars are a perpetual hazard.
Mount Semuru, at 3676 m the highest volcano in east Java, is a persistently active vol-
cano (Thouret et al., 2007). Eruptive activities are common with short-lived eruption col-
umns appearing several times a day, frequent explosive eruptions and an unstable summit
263 Volcaniclastic flows: debris avalanches and flows

Fig. 14.5 Paths of destructive lahars originating from Navado del Ruiz eruption of 1985. After Francis, 1993 and Voight, 1996

cone. Small lahars travelling down channels are regular events, due to the daily supply
of pyroclastic debris over the summit cone and the high runoff during the rainy season.
Large pyroclastic flows tend to occur once in every five years. At least five large lahars,
each exceeding 5 × 106 m3, have occurred since 1884. This is a permanent threat to a
number of settlements at the foot of the volcano, including Lumajang, a town of 85 000
people. The threat of lahars is especially high towards the southern and southeastern val-
leys, which annually experience tens of rain-triggered lahars for a distance of 20 km from
Mount Semuru. The annual rainfall is 3700 mm, and 500 mm of rain has fallen in 48-hour
storms. The rain obviously falls on a plentiful supply of loose volcanic material, including
fine-grained ash which leads to hyperconcentrated flows. Thouret et al (2007) determined
the annual sediment yield in the local Curah Lengkong river valley to be 2.7 × 105 m3km-2
with an annual denudation rate of 4 × 105t km−2. This is a very high rate and so is any rate
approaching it. We may conclude that the slopes of pyroclastic volcanoes that are subjected
to high rainfall in the tropics contribute a huge amount of material to the lowlands and
ultimately to the coast (Fig. 13.1).
Volcanic mudflows are also created, very destructive ones at times, when water and
mud escape from crater lakes due to volcanic eruptions. Perhaps the best-studied example
comes from the Kelud Volcano on Java which has a deep crater lake. Lahars originate from
the lake periodically, unless the level of the lake is lowered. A system of tunnels was engi-
neered to do so, but later volcanic eruptions damaged the arrangement so the lower slopes
of the Kelud remain vulnerable to a mixed flow of mud and hot, acidic, sulphurous water.
Eruptions on high volcanoes with a snow cover on their upper slopes also give rise
to lahars. A catastrophic one occurred in November 1985 following a small eruption on
Nevado del Ruiz in Colombia (Fig. 14.5). The incident illustrates not only the origin of
this type of volcanic mudflow but also the need to have an effective hazard prevention sys-
tem which would have prevented the loss of life and property in the valleys. As described
below, the disaster killed 23 000 people and destroyed settlements, including the town of
264 Volcanic landforms

Armero, communication links, and crops and livestock. In terms of people killed from a
volcanic hazard it ranks 4th (Voight, 1996) surpassed only by Tambora (1815), Krakatau
(1883) and Mount Pelée (1902).
A small eruption on the 5200 m high Nevado del Ruiz generated a number of pyroclas-
tic flows and surges that scoured and melted ice and snow on the volcano. The meltwater
and pyroclastic debris descended the slopes and accumulated in narrow channels to form
lahars that reached the settlements within two hours of the eruption. West of the volcano,
lahars originated in the headwaters of the rivers Molinos and Nereidas and swept down
destructively along the River Chinchiná (Fig. 14.5). Lahars coming down both headwa-
ters reached the River Chinchiná with a flow rate of 13 000 m3s−1 and travelled for 70 km
more to reach the River Cauca. Towards the east, lahars came down the River Lagunillas
to destroy Armero and the River Guali to pass Mariquita. The eastern lahars were big-
ger, about 20 000 m3s−1 on the River Guali and an average of 25 000–30 000 along the
River Azufrado, which was the main supplier of lahars to the River Lagunillas. The peak
value attained was 47 000 m3s−1, one-fifth of Amazon’s discharge (Francis, 1993; Voight,
1996).
The peak flow velocity in most of the lahar-scoured channels ranged between 5 and
15 ms−1. Flow depths, except in the River Nereidas, exceeded 10 m. The depth and vel-
ocity of the lahars produced a high boundary shear stress and eroded the channel walls
to transfer sediment and interstitial water. This in turn increased their flow volumes and
velocities.
The event showed that a small eruption can produce lahars of huge dimensions and
destructive capability. It also illustrated the failure of disaster management in spite of
the expectedness of the event. As Voight described it, the ‘disaster happened because of
cumulative human error€– by misjudgment, indecision and bureaucratic shortsightedness’
(Voight, 1996:€764). The inhabitants of Armero could have been saved if an alarm had been
raised in time.
Lahars are thus extremely dangerous, and effectively modify the lower slopes of a vol-
cano, especially the channels. They can travel long distances; lengths of over 100 km have
been reported. In sum, material from inside the Earth’s crust that erupts through craters,
shapes the volcano and the lower grounds at its foot and, in suitable locations, a high sedi-
ment load reaches the sea via the regional streams. The chain of eruption, ash falls, pyro-
clastic flows and lahars can be seen clearly in high-resolution satellite images. Box 14.1 is
an example on the Merapi.
Volcanoes have smooth slopes on ashes and fragmentary material. Verstappen (2005)
mentioned that although slopes on the stratovolcanoes of Southeast Asia appear to be con-
cave, overall they can be divided into three sections. The top part is straight and essentially
formed by the transfer downslope of ashes and coarser material under gravity. These slopes
may reach 34°. The middle section is straight with about 8–12° of slope. Most of the vol-
canic slope may belong to this section which is modelled by deposits from lahars. Valleys
are filled and overflows may happen, but this phase of deposition is followed by rapid
erosion after an eruption. The lower slope of a volcano is modified by sediment deposited
in small radiating ravines in rainstorms. The approximately 2° slope of this section merges
with an alluvial plain below.
265 Volcaniclastic flows: debris avalanches and flows

Box 14.1 The 2006 pyroclastic deposits of the Merapi volcano


Merapi is a particularly active volcano of southcentral Java from which pyroclastic flows and lahars threaten
a populated region that includes the city of Yogyakarta, located close to the southern flank of the Merapi. A
number of eruptions and pyroclastic flows continued on the Merapi from May 2006. Four pyroclastic flows
due to dome collapse, occurred on 14 June, the largest among them reached 7.78 km. Two of these block and
ash flows were especially destructive, burying the village of Kaliadem under 7 m of deposit and killing two
people.
On 16 June, two IKONOS satellite images acquired by the Centre for Remote Imaging, Sensing and
Processing (CRISP), National University of Singapore captured an active block and ash flow. This was first sat-
ellite imagery of a transient pyroclastic flow (Thouret et al., 2010).The images also provided a synoptic view
of the fresh block and ash deposits of 14 June. It was possible to reconstruct the events from the summit of the
volcano to its lower slopes. Figure 14.6, dated 16 June 2006, shows a billowing ash cloud hugging the ground
at the front of the flow and rising further towards its back. It was supported by elutriated fine ash and hot
gases from the block and ash flow which is seen emerging from below the cloud.
One of the IKONOS images demonstrates the passage of the block and ash flow and surge deposits on the
lower slopes (Fig. 14.7). Several flow lobes and tongues were seen confined within the box-shaped gorge
of the Gendol River. These ash cloud surge deposits and knocked-down trees appear on both slopes where
the gorge narrows. A fan-shaped overbank deposit crossed an interfluve to invade a minor valley. Massive
overbank deposits reached the right bank of the Gendol River to destroy Kaliadem. Part of the flow avulsed
into a tributary valley.
The image showed evidences of avulsion into neighbouring valleys due to obstacles formed by previ-
ous flow deposits and channel banks plus anthropogenic barriers constructed in the channel to control the

Fig. 14.6 DEM of Merapi draped by the IKONOS image of 16 June 2006, showing the southern slope of the
volcano and the Gendol River Basin. A pyroclastic flow is emerging from below the billowing ash
cloud. The river channel is marked by deposition of earlier pyroclastic flows. From Thouret et al.,
2010. By permission of Elsevier. See also colour plate section
266 Volcanic landforms

Fig. 14.7 Lower slopes of the DEM of the Gendol Valley showing details of a volcanic flow. Segment 1:
upper section of the valley; Segment 2:€widening of the valley and a fan of volcaniclastic
material; Segment 3:€a wider valley entrenched in a volcaniclastic fan. A:€the moving pyroclastic
flow from upstream; B:€a ridge built by a previous flow; C:€an area stripped by an ash-cloud surge;
D:€spilling and re-routing of flow at a bend; E:€avulsion of re-routed flow to a neighbouring
channel; F:€lobes of flow reaching overbank which buried the village of Kaliadem (KA). From
Thouret et al., 2010. With permission from Elsevier. See also colour plate section
passage of flows. The volume of deposited material in the channel suggests the possibility of another lahar
after the next rainfall event.

14.5╇ Landscape on flood basalts

At certain periods of the Earth’s history, continental flood basalts (CFBs) emerged through
innumerable fissures fed by dyke swarms to cover extensive areas. The basaltic lava flowed
out in sheets and piled up to form highlands. Weathering and deep dissection give such
areas a characteristic stepped topography over time (Fig. 14.8). The upper part of each
lava flow is scoracious and softer than the lower parts. The lower part of a flow therefore
gives rise to a vertical step, whereas the upper flow is eroded to create a flatter tread-like
surface. Cumulatively, this creates a staircase-like appearance, described as a ‘trap’ in old
Swedish (Francis, 1993). The term is in common use. For example, the landscape of CFBs
in Deccan, India is known as the Deccan traps. The Deccan traps are 65 million years old
and they are estimated to have had an original extent of 1.5 × 106 km2, of which 0.5 × 106
km2 survives. The average thickness of the Deccan traps is more than 1 km. Other areas
in the tropics where large stepped plateau-like features on CFBs occur are Paraná, south-
eastern Brazil; an area centred on the Walvis Bay in Southwestern Africa; the Ethiopian
and East African Highlands; and the Drakensberg of South Africa (Fig. 14.1).
267 Landscape on flood basalts

Fig. 14.8 Plateau on flood basalt, Deccan, India

Fig. 14.9 Schematic representation of trap rivers, showing an inset river channel within a wide open valley and side slopes in
flood basalt. From Deodhar and Kale, 1999. By permission of Wiley

Figure 14.9 is a schematic representation of trap rivers from India. Most of the Deccan
trap is in a low-rainfall area. The rain shadow of the escarpment marks the edge of the CFB
plateau, which rises steeply from the narrow coastal plain of the Arabian Sea. At the top of
the escarpment, known as the Western Ghats, the rainfall is from the southwestern mon-
soon and exceeds 5000 mm annually. Towards the east in the rain shadow of the Western
Ghats, the figure drops to 500 mm within a distance of 200 km. As expected, the seasonal
pattern persists. The trap rivers tend to have low-gradient wide valleys. Within such val-
leys, the major rivers have channels incised into bedrock or alluvium. The tributaries, in
contrast, have steep upper courses with a change of gradient as they reach the valley flat of
the trunk streams. The channels are stable and, apart from seasonal high discharges, they
may, as in the Deccan, carry high flows of low-frequency floods. The landforms of stepped
plateaus in the CFB areas, as displayed in Figure 14.9, are associated with deep weathering
268 Volcanic landforms

and soil formation. Wide trunk valleys with the river channels meandering in and out of
bedrock canyons and incised upper courses of tributary streams are common. The channels
carry a fine suspended load of silt and clay and a coarse bed load, mainly of pebbles and
cobbles, derived from the jointed bedrock. The coarse material is generally contributed
by the tributaries draining the high side slopes of the valley and also by the trunk streams
flowing over jointed basalt in locally steep sections. The bar material is therefore coarse
and bed armouring is common. Floodplain building does not occur in many rivers and
an accumulation of old sediment may form terraces along with gravel fans at tributary
mouths, both of which are eroded in low-frequency, high-magnitude floods.
The main channels are box-shaped (see Chapter 7). Gorges in bedrock and deep chan-
nels in alluvium deposited over basalt are typical, but are not seen everywhere. In places,
rivers may be very wide with multiple channels and rocky islands where it flows over the
exposed bedrock. Such features have been described for the Mekong River where it crosses
an exposure of Mesozoic basalt at 4000 Islands near the Lao PDR–Cambodia border. The
maximum river width measured for the Mekong across islands and sub-channels is 15 km
(Gupta and Liew, 2007). Deodhar and Kale (1999) provide a list of erosional features such
as potholes, grooves, inner channels, scablands and boulder berms. There is almost no sedi-
ment on the channel floor at bedrock reaches because of secondary eddies, flow separation
and the formation of vortices during large floods (Kale and Hire, 2004).

14.6╇ Conclusion

Volcanic activities do not only form a distinctive landscape and provide a huge and wide-
spread volume of sediment; they also influence the local population. Surrounding areas may
be devastated periodically and the deposited pyroclastic ash fall and flow material is eroded
by rills and gulleys across the slopes. Volcanism, however, has its beneficial aspect. It provides
fertile black soils with a high water-holding capacity which are extensively cultivated. This
may lead to areas close to volcanoes being densely populated, thus exposing a large number
of people to volcanic hazards from time to time. Farming and high-density rural settlements
are also encouraged by the huge amount of groundwater that is stored in many places in the
large stratovolcanoes. The groundwater appears in a series of springs around the edifice where
the slope changes. Faults and old lava flows may complicate the picture (Verstappen, 2005).
Volcanoes may dominate a single island such as Montserrat or Lombok, where coralline lime-
stone surrounds the volcano to provide the island with its final shape.

Questions

1. Where in the tropics does volcanism primarily determine the landscape?


2. Explain how the dimensions and slopes of volcanoes depend on the composition of the
construction material. What else affects their appearance?
269 Questions

3. Compare the modifications brought to an earlier landscape by ash fall and pyroclastic
flow deposits.
4. Why are lahars important as a regional geomorphic process?
5. How hazardous are lahars?
6. Describe the passage of volcanic material down the slope of the Merapi in June 2006,
using Figures 14.6 and 14.7.
7. Describe the landscape formed on continental flood basalts. Where in the tropics do
such landscapes occur?
15 Tropical karst

Where Alph, the sacred river, ran


Through caverns measureless to man
Samuel Taylor Coleridge

15.1╇ Introduction to karst

The word karst denotes a set of special landforms that develop mostly on carbonate rocks
due to their high solubility. The rocks, however, do not dissolve uniformly, but mainly
along lines of structural weakness such as joints or bedding planes. This happens both
at the surface and inside the body of the rock. Water enters the subsurface through open-
ings on the exposed face of the rock and flows through subterranean passages and caves,
probably emerging some distance away in springs. As the drainage progressively passes
underground, a bare dry stony landscape evolves; this is karst. The word comes from
a high barren plateau in western Slovenia where such features are well displayed. The
original word kras was later modified to its present germanicised form, karst, which is
used to denote similar landscapes anywhere. The publication of Jovan Cvijíc’s book Das
Karstphaenomen in 1893 accelerated the study of karst geomorphology.
A terminological complexity exists in karst geomorphology. Almost all karstic features
are known under several names, each derived from a regional usage. This account carries
only the common international terms in order to avoid this difficulty. A complete list is
usually found in books specialising in karst geomorphology, such as Ford and Williams
(2007). Such volumes should also be consulted for greater details.
Complications also arise because of the use of the word ‘karst’ as a suffix to describe
landforms that are not really karst but strongly resemble it. Pseudokarsts are landforms
which resemble karsts but are not produced by solution or other usual processes by which a
karst landscape is formed. Thermokarsts are areas with topographic depressions caused by
the thawing of ground ice. Palaeokarsts are old karsts that are now covered with younger
sediment and are no longer affected by solution activities. Some geomorphologists do not
restrict the word to limestone features but extend it to solution forms such as furrows on
granite or basalt (Ford and Williams, 1989). This chapter covers only the true karst on
carbonate rocks:€ its forms and operating processes. Karst topography is best developed
on limestone which consists of calcite (CaCO3). Dolomitic limestone, Ca,Mg(CO3)2, and
limestone metamorphosed to marble may also display karst features but less distinctly.
Such rocks are not easily soluble.
270
271 karst in the tropics: the geographical distribution

Hollows, often with open shafts at their lowest point, pockmark the surface of a karst
area and lead to subterranean passages and subsurface caves in limestone. These features
are essentially formed by solution or structural collapse aided by solution. In certain cases,
small steep hills rise above the hollowed surface. Special features such as erosional notches
develop where the sea reaches the limestone. A sufficient amount of water is required to
dissolve part of the rock. Karsts are not associated with an arid climate.
It has been suggested that certain karstic features are more common in the humid tropics
or a modified version of the general form occurs in the humid tropics. Such forms have been
well described but their connection with the humid tropics has not been clearly explained.
It is generally suggested that the process of karstification operates at a high rate in the hot
humid climate where organic acids also are easily available at ground level (Jennings,
1985). In this chapter, the processes associated with the formation of a karst landscape are
discussed, with special reference to the features common in the humid tropics.

15.2╇ Karst in the tropics:€the geographical distribution

According to Ford and Williams (1989), surface exposures of carbonate rocks cover about
12 per cent of the dry ice-free land. Their estimation of karst areas, however, is a little less,
between 7 and 10 per cent. A number of environmental conditions are essential for the
development of a good karst landscape, but these do not always coincide with carbonate
areas (Thornbury, 1960). These are:
1. The presence of a soluble rock, preferably limestone, at or very near the surface.
2. Such a rock should be dense, highly jointed and preferably thinly bedded, encouraging
the circulation of water along marked paths.
3. The area should have at least a modest amount of rainfall. Sweeting (1972) suggested
a minimum of 250–300 mm annually for limestone karst development. Karstic features
found in arid areas are therefore relict and indicate a wetter climate at the time of their
formation.
4. The presence of vegetation which increases the acidity of the circulating water by add-
ing biogenic CO2 and organic acids.
5. The presence of entrenched trunk valleys below the limestone with karst, allowing
groundwater to descend through the body of limestone and emerge as surface streams
at a lower elevation. A continuous passage of water through the rock is extremely
helpful.
The requirements for the good development of a karstic landscape therefore explain why
(1) karst is not everywhere associated with surface exposures of carbonates and (2) karst
is expected to be well developed in the humid tropics. Jennings (1985) has described the
tropical variety as the botanical hothouse karst.
In the tropics, good karst landforms occur in south China and countries of Southeast
Asia such as north Viet Nam, Lao PDR, Thailand, east Myanmar, north Sumatra, north and
central Java, central and south Sulawesi, islands of eastern Indonesia, Papua, New Guinea,
272 Tropical karst

Sabah and Sarawak. Karst landscapes have also been reported in India and Southern Africa
and Madagascar. They are well developed in Central and South America and the Caribbean
islands, including northern Yucatan, Jamaica, northern Puerto Rico, western Cuba and
Barbados. Karst probably occurs in more areas in the tropics, but some locations have not
always been well reported. Karstic features are also seen in the evaporites (gypsum and
salt) in the arid tropics.

15.3╇ Karst hydrology

Water in karst can be divided into three types, based on location.


1. Atmospheric water from direct precipitation over a karstic area enters the subsurface.
This is known as autochthonous or autogenic water. Water also arrives from outside the
karstic area by either surface flow or runoff from rivers, subsequently passing into the
subsurface only through specific openings in the rock surface. Such flows from external
noncarbonate rocks to karstic areas are allochthonous or allogenic water. Generally both
types of water are present in the bedrock (Fig. 15.1).
2. Water that enters karst accumulates in the vadose zone, the zone of aeration above the
groundwater table. This water passes through the body of the rock either by percolation
or, more effectively, along openings provided by joints, fractures, etc. In the vadose
zone above the water table, usually only a fraction of the holes are filled with water;
the rest are filled with air. The proportion of holes filled with water depends indirectly
on the length of time since the arrival of water originating from precipitation or runoff.
The entry of water through the pores can be complicated. The passage of water may
be impeded by air bubbles or locally modified by perched aquifers above a localised
impermeable layer of shale or chert.
3. Water that has reached the groundwater table and augmented it. It is controversial
whether the pores in the rock below the groundwater table are all filled with water
or water is stored mostly along various passages and openings and thereby occurs at
different depths in the rock. According to the second view, water is stored in an inter-
connected network of passages. These passages do not occur at the same level, and no
typical water table as such exists. Presumably, the pattern of water storage in the sub-
surface depends on the physical property of the local rock. Pores in some rocks may be
too small and not interconnected enough to store water.
Williams (1983) recognised a special zone in the weathered upper part of the bedrock
below the soil:€the subcutaneous or epikarstic zone. This zone is periodically saturated,
especially after storms when the water table rises. Cavities and passages are efficiently
enlarged at this level but not at the level immediately underneath. A perched water table
therefore occurs episodically. This then slopes towards larger passages that originated
from joints, fractures, etc. This lateral component of flow results in enlarged openings at
the surface through which rain or surface water subsequently enters the bedrock.
The nature of the passage depends on the size of the openings. If the openings are small
then the flow is assumed to be laminar. Flow velocity rises in large openings, leading to
273 Dissolution of karst rocks

Fig. 15.1 Both types of karstic water in bedrock

fluctuating eddies and turbulent mixing. Flow through large openings is therefore more
capable of both reacting with and mechanically eroding bedrock. Ford and Williams (1989)
have discussed the hydrology of a karst landscape in detail.

15.4╇ Dissolution of karst rocks

The general reaction between a limestone and water can be written as


H2O ⇔ H++ OH− (15.1)
CaCO3 ⇔ Ca2++ CO32− (15.2)
Ca + CO32− + H+ ⇔ Ca2+HCO3−
2+
(15.3)
The equations are unbalanced and the reaction moves forward at rates proportional to the
concentration of reactants (Ford and Williams, 1989). The reaction is accelerated by the
presence of CO2 in the soil, which may occur in the pore spaces that are not filled with
water. Green plants and soil fauna and flora respire quantities of CO2. The microfauna
and flora in soil are densest in and near the rooting zone, from which CO2 diffuses out
to the soil base and ground surface. The production of CO2 increases with temperature
and it is likely that this is a factor behind well-developed karst in the tropics. The amount
of CO2 in the soil, however, varies tremendously. Ford and Williams (1989) listed a
number of factors that influence the reaction between the rock and the passing solution.
These factors are various properties of the soil (types, texture, horizon, depth, drainage
and exposure), vegetation cover, soil flora and fauna, and seasonality in warming and
wetting.
274 Tropical karst

Fig. 15.2 The result of mixing of two saturated solutions (A and B) in a CO2 or H2S system. It produces an undersaturated
solution (C) resulting in subsequent dissolution along CD. Shown for both calcite and dolomite. From Palmer, 1991

Soil CO2 is a heavy gas and drains into fissures and caves below. The gas is further
increased in volume by the decomposition of vegetal matter and animal droppings that
wash into caves. Subsurface drainage, especially streams in flood flowing through under-
ground caves, picks up the increased amount of CO2 and becomes more corrosion-prone.
Mechanical mixing of karst waters from two different sources increases the aggressiveness
even if both types of water are saturated. The mixing leads to water being unsaturated and
more carbonate is dissolved from the bedrock, especially if the concentration of CaCO3 in
any of the waters is below 250 mgl−1 (Fig. 15.2).
The solution of limestone thus depends on various factors. It has been suggested
that among these factors precipitation acts as the principal control. Ford and Williams
(1989) referred to a study by Maire in very wet Papua New Guinea (annual rainfall
5700–12 000€mm) where the calculated annual solution denudation rate is very high,
270–760 mm in 1000 years. Karst is also denuded by mechanical erosion as turbulent
water carrying sediment rushes through subsurface openings and passages. The amount
of available water, its CaCO3 content, the percentage of limestone in the basin and the
time that is available all control the solution denudation rate.

15.5╇ Karst landforms

The development of karst landforms in soluble rocks depends on the ambient hydrological
and geochemical conditions. Such landforms are developed best in a humid climate where
enough water is available to circulate through the bedrock. Temperature is also an import-
ant factor, as it influences the water budget, rate of chemical reactions and biochemical
processes. The lithological and structural properties of the bedrock determine the solute
pathways, the rock strength of the support roofs of underground passages and the suscep-
tibility of the ambient bedrock to solution and mechanical erosion (Ford and Williams,
275 Karst landforms

1989). Although similar landforms are found in karst worldwide, landforms in the coun-
tries of the humid tropics tend to display certain characteristics. We first introduce the
general case of karst landforms and, in the next section, review the special case of tropical
karst. The dimensions of karst features range from millimetre-scale dissolution pits and
groves on limestone surfaces (also called limestone pavements) to hills several hundred
metres high and underground caves that stretch for kilometres.

15.5.1╇ Limestone pavements


Small-scale solution features, millimetres to centimetres in scale, with a circular or linear
geometry, evolve on flat or nearly horizontal exposed limestone. It is thought that such
small features are also associated with algal or bacterial activities, and certainly bacteria,
fungi, lichens and mosses, once established in the indentations, can grow. Larger circular
features are known as solution pits and linear features as karrens. Karrens elongate along
joints and veins in rock, or as rills along lines of dissolution. Such channels may widen and
deepen downslope. These surface indentations vary in appearance and size, and large areas
with such features (karrenfeld) are found either on bare rock or under soil and vegetation.
Large fissures of this type are known as lapiés and, if exposed, they are difficult to cross,
as they may measure up to metres in scale. Lapiés and other indentations, however, may
occur under a cover of soil and weathered material, which is coloured red from the release
of iron impurities from the carbonate bedrock as it dissolves. The red regolith is known as
terra rossa. The term pinnacle karst is used to denote an area where these depressions are
separated by large carbonate spikes that are sharpened towards the top. The term ‘stone
forest’ is aptly used to describe such karst in China. Pinnacles more than 45 m in height
have been reported in Papua New Guinea (Williams, 1971) and Sarawak (Osmaston and
Sweeting, 1982). Karren forms may also develop on inclined surfaces that are bare.

15.5.2╇ Dolines or sinkholes

Dolines, small openings through which surface drainage passes underground, are probably
the most common feature of the karst landscape. The term, in common use, is derived from
Slovenian. These depressions are known as sinkholes in North America; this is another
term widely in use. The openings are circular or near-circular in shape, with surface diam-
eters ranging from several metres to a kilometre and depths from several to hundreds of
metres. Their side slopes vary between gently sloping to near vertical, and they can resem-
ble a saucer, a funnel with a shaft at its bottom, or even a vertical hole. Dolines may occur
as single individuals or as a densely packed group giving the landscape an egg-box appear-
ance. Their distribution pattern may also be skewed by lines of structural weakness.
Dolines are classified according to their origin. The common types are solution doline,
collapse doline, subsidence doline and suffusion doline. Other less common types have
been listed (Ford and Williams, 2007). In general, a single process dominates, although
more than one is usually involved in the formation of dolines (Fig. 15.3).
Solution dolines are formed by the mass transfer of carbonate material in solution from
the surface to the subsurface, creating a surficial depression. This requires the establishment
276 Tropical karst

Fig. 15.3 Types of doline

of a drainage system below the doline to connect the surface recharge to an underground
passage. Usually while limestone is either exposed or under a thin cover of alluvial mater-
ial at the centre of the doline, its flanks could still be under an overlying impermeable rock.
Once the doline is formed, a positive feedback comes into operation because of the centri-
petal concentration of flowing water and associated corrosion inside the depression. The
biogenic production of CO2 in soils overlying dolines helps. The connection to the subsur-
face passage usually improves with time, thereby accelerating drawdown and the passage
of water underground, speeding up the formation of dolines.
Collapse dolines have smaller openings and steeper sides, although both characteristics
may change over time. They are formed by the collapse of the surface layer exposing a
shaft that connects with underground passages. The collapse may happen in three ways:
1.╇ final collapse of part of a cavern roof earlier weakened by solution
2.╇ undermining of the roof from below as the cavern expands
3.╇collapse of a rock layer due to removal of buoyant support by the lowering of the water
table. This makes part of the rock unstable.
In all cases, the collapse of a rock layer provides a connection pathway for the surface
drainage towards the flow in the underground passage. In some cases, the shaft is quite
deep and, after rainfall and recharge, a temporary underground waterfall to the cavern may
operate.
Subsidence dolines are formed when the ground surface sags down forming a cone-
shaped depression without a significant faulting of the rocks. No rupture in the rock hap-
pens as in the case of collapse dolines, but the sagged depression concentrates surface
water and forms and behaves like a doline.
Water in suffusion dolines seeps through thick unconsolidated regolith or detritus over-
lying a karstic rock. This creates a depression by the removal of fines in solution or down-
washing, a process known as suffusion. Water then percolates through the sagged surficial
material to develop a karren vertically, and finds a connection with the cave drainage.
Other terms have been used by researchers to identify dolines of different origin, but
these four are the ones normally used. Where dolines are closely spaced and all the space
is taken up by depressions with shared divides, the landscape is known as polygonal karst.
Thousands of dolines may occur inside one square kilometre, although their frequency is
often in single figures within this area. It has been suggested that dolines vary in morph-
ology in different climatic zones, but the concept is still in a hypothetical stage. White
277 Karst landforms

et€ al. (1984) showed that the average depth of depressions is high in the tropical karst
regions of Puerto Rico and the Dominican Republic when compared to temperate karst in
various regions of the United States (White et al., 1984). It could be due to higher rainfall
in the humid tropics and a positive water balance. The density of dolines, however, varies
widely.

15.5.3╇ Uvalas and poljes


Dolines are closed depressions of limited size, but closed depressions bigger than dolines
may also occur as part of a karst landscape. Uvalas are formed by merging dolines and, as
expected, are characterised by an uneven floor with multiple topographic lows. They can
be of irregular shape, and with sizes ranging between 5 and 1000 m in diameter and 1 and
200 m in depth.
Poljes are bigger depressions with steep sides and usually a flat floor, several kilometres
in length. A number of them are likely to be structurally controlled, and some are elongated
along lines of structural weakness. They are surrounded by impermeable rocks that overlie
the soluble carbonates, and poljes have been described as the inliers of a fluvial landscape
(Ford and Williams, 1989). External streams flow across poljes, frequently disappearing at
a doline on the polje floor. The dolines are usually not efficient enough to cope with flood
conditions and a shallow short-term lake may episodically cover the floor of the polje. This
enables the depression to expand by lateral corrosion. The floor of poljes often acts as a
window on the local water table, which fluctuates leaving the polje with either a dry flat
floor or under an aggressive sheet of water.

15.5.4╇ Karst valleys

River valleys that cross the karst have distinctive features. An allochthonous river with
a large discharge may succeed in crossing a karst landscape without losing all its water
underground. Such rivers usually flow in gorges that are created due to the combination
of the solution of carbonates by the chemically aggressive river discharge flowing across
a carbonate floor and abrasion by the clastic sediment it carries. Perennial allogenic rivers
function as local base levels for the section of the karst they flow through. Karst gorges
also originate from the collapse of roofs of underground caves, making part of the subter-
ranean river visible.
Smaller rivers may not succeed in crossing the karst. Such rivers tend to lose water
through the dolines along its valley and ultimately no water is left, so the valley below the
last active doline becomes dry. This lower valley only carries water in large floods which
exceed the capacity of the upstream dolines to capture water. Such valleys are known as
semiblind valleys. Over time, a difference in elevation builds up in the form of a rise that
separates the upper valley, which carries water and is thus capable of eroding and lowering
its level, from the lower valley, which, in the absence of erosion, remains near the original
elevation. The lower part of the valley may remain dry even when floodwaters arrive. It is
then a blind valley.
278 Tropical karst

Pocket valleys are the opposite of blind valleys. Pocket valleys occur where a stream
descends through the soluble rock and emerges at its contact with an impermeable layer
underneath. This results in a resurging spring and a U-shaped valley with steep walls and
a receding headwall at the spring head. The size of the valley is determined by the volume
of water discharged. Pocket valleys are also known as steepheads.

15.5.5╇ Cone and tower karsts

The sharply edged standard dolines, the commonest feature of the karst landscape of the
temperate latitudes, are often replaced in the humid tropics by star-shaped depressions sur-
rounded by cone-shaped hills. This feature is known as cockpit karst, a name derived from
Jamaica where such depressions were used in the past for holding battles between fighting
cockerels with an uninterrupted view for the spectators. Like dolines, they usually occur
in concentrations. Cockpit karst has been reported from outside the tropics, but it is more
frequent in tropical countries such as Jamaica, Indonesia (especially Java), the Philippines,
Viet Nam and southern China.
Several explanations have been forwarded to explain the morphological difference
between cockpit karst and standard dolines. A rapid enlargement of dolines to create cock-
pit karst has been attributed to the increased acidity in surface and subsurface water from
rapidly decomposing vegetal matter in the humid tropics. It has also been suggested that
cockpit karst develops where dolines expand faster than they are deepened. Dolines can
only deepen to the water table, which in the humid tropics is usually shallow, but they can
continue to enlarge laterally.
Several neighbouring cockpits may ultimately overlap and merge. The low hills marking
the boundary of a cockpit are converted over time into cone-shaped rises by the expansion
of cockpits from various sides. This kind of landscape is known as the conekarst. Cones
also display remarkable symmetrical shapes. Kangning (1992) measured the morphometry
of about 750 cones in Guizhou, southern China. The cones were generally symmetrical,
and their slopes measured between 45 and 47°, regardless of their structure.
A tower karst is a steeper variety of karst risers (Fig. 15.4). In general, towers are
steeper and higher than cones, and have a bigger height–diameter ratio. They are consid-
ered to be common tropical karst forms although, like cones, temperate examples have
been reported. These residual hills come in a range of shapes, from towers to almost
conical or hemispheric. The common dominant feature is their sharp rise from the sur-
rounding plains, although some may have pedestals. Towers are commonly grouped,
although they do occur in isolation. It is better not to expect a typical appearance as
tower karst exists across a range of forms, any of which may carry the suggestion of a
tower.
Wilford and Wall (1965) grouped different types of tower karst in Sarawak, where all the
towers are residuals on a planed surface:

• residual hills that emerge from a planed carbonate surface covered with alluvium
• residual hills emerging from limestone inliers in a surface which is a plane mostly across
non-carbonate rocks
279 Karst landforms

Fig. 15.4 Diagrammatic section showing tower karst, caves and surface drainage. After Jennings, 1985

• residual hills emerging through a surface cover of thick alluvium


• isolated towers that rise from steeply sloping bases of varying lithology.
Karst towers have been recognised as time-transgressive forms, developing slowly, and
being older near their tops and younger at their bases (Ford and Williams, 1989).
They have been studied in Puerto Rico, Cuba, Malaysia, Viet Nam, south China and
other places inside the humid tropics. Towers as high as several hundreds of metres above
the local plains have been reported (Wilford and Wall, 1965). There is a general inclin-
ation to attribute their shape to structural control, the rock disintegrating along joints and
fracture zones. Tower karsts are known under different regional names:€pepinos, haystacks
and mogotes. The term mogote is common in Spanish-speaking countries. Asymmetrical
mogotes in Puerto Rico have evoked considerable research interest. Such an asymmetry
with steeper windward eastern slopes has been attributed to case-hardening of limestone
by repeated cycles of solution and reprecipitation of CaCO3. Water runs down the eastern
slopes due to frequent rainfalls in the prevalent trade winds, followed by drying of the same
slopes by the same winds between rain events. This leads to precipitation on the slopes of
amorphous calcium carbonate, which is harder and difficult to dissolve. Hence the steeper
slopes on the windward side (Monroe, 1976). This explanation, however, has been ques-
tioned and the asymmetry attributed to other factors, such as basal erosion.
Only a fraction of towers are asymmetrical in Puerto Rico and they have not been reported
to be asymmetrical elsewhere. A number of hypotheses have been offered to explain the
origin of towers and why they are found in certain areas instead of cones. These sugges-
tions involve the depth of regional water table, surface fluvial processes, intense corrosion
along zones of structural weakness, river erosion, etc. It is possible that towers may owe
their origin to a variety of processes (Williams, 1987).

15.5.6╇ Caves

Morphology in karst is studied both on the surface and underground. Karst rocks, espe-
cially limestone, carry a number of underground openings, called caves, through which
water in the subsurface circulates. The standard definition of a cave is a solutional opening
underground, large enough for a human being to enter. According to Ford and Williams
280 Tropical karst

Fig. 15.5 Diagrammatic sketch of a cave in limestone. The upper part (left side) carries vadose water via dolines and a higher
entrance. This part stays dry periodically. The lower part (right side) carries phreatic water, mainly from groundwater
rising through a fissure. The water from the cave emerges at a spring to feed a surface stream

(1989), a karst cave is a solutional opening, bigger than 5–15 mm in diameter or width.
This is effectively the smallest aperture that allows turbulent flow to form. Larger caves are
also called caverns. They are common in wet climates, and allogenic bigger rivers tend to
flow through larger caves. However, large caves do occur in arid regions, possibly as relicts
from an earlier wetter climate.
Caves may occur as isolated openings in rock but they usually have input and output
points connected by conduits, at least 5–15 mm in diameter. The entire set-up behaves as an
integrated system. Caves occur at different depths through the bedrock. A long integrated
cave system may continue at various depths (Fig. 15.5). The upper part of a cave may be
in the vadose zone and dry, but its lower part may be in the phreatic zone and filled with
water. As the water table inside the rock fluctuates, the proportion between dry and wet
parts of a cave changes. The form of the passage also tends to be different in the two zones.
Vadose caves tend to have passages resembling canyons, with steep walls and vertical
shafts. Phreatic caves are tube-shaped and more equidimensional. The form of a cave pas-
sage depends primarily on the solution in vadose (free fall) conditions, phreatic (pressure
flow) environment, or on conditions that alternate between the two during floods.
Caves therefore have entrances via dolines and vertical shafts (input points); passages;
larger openings called rooms or chambers; and spring outlets (output points). They tend
to form along paths of greatest discharge of groundwater and solutional aggressiveness
(Palmer, 1991). Caves may end abruptly with rockslides or collapses, the filling of the pas-
sage by silt and clay, and the narrowing of the passage due to a change in rock characteris-
tics which allows water to permeate but does not allow people to pass. Such instances are
known as termination. Cave chambers can be huge. The biggest known chamber (Sarawak
Chamber, Good Luck Cave) occurs in the Mulu National Park of Sarawak, East Malaysia.
It measures 20 million m3. The Belize Chamber in the Tun Kul Cave in Belize is about a
million m3. Many large chambers, 100 000–500 000 m3 in volume, have been reported
(Ford and Williams, 1989). Palmer (1991) stated that caves usually need a time-span of
104–105 years to reach a traversable size. The pattern of linked underground passages in
a cave depends on the mode of groundwater recharge. For example, discharges that pass
underground via sinkholes tend to form what is called a branching cave, where subterranean
tributaries join to create a higher-order system. Other cave forms, often more complicated
281 Karst in the tropics

and with passages at different levels, are created by factors such as steep gradients, greater
undersaturation and lines of geological weakness (Palmer, 1991). Palmer is of the opinion
that climate affects caves only indirectly via changing patterns of rate and character of
recharge. Tropical caves formed at a higher presence of CO2 may enlarge quickly but their
morphology remains comparable to caves formed in other climatic regions.
The underground passages and chambers are formed along lines of structural weak-
ness, and the geological structure of the bedrock may influence the shape of the cave. The
enlargement of the openings is carried out both by corrosion by circulating water and abra-
sion by sediment that arrives in high flow from the surface via input points such as dolines
with vertical shafts. Small sculptured markings, formed by corrosion and abrasion, also
occur on the walls, floors and ceilings of such openings. Caves display forms made with
chemically precipitated carbonate deposits; the general term for such forms is speleothem.
The deposited material is called travertine. Such precipitation comes from circulating
waters because of the mixing of water and air in caves, each with a different CO2 concen-
tration. CO2 is then diffused to the atmosphere of the cave because the partial pressure of
the gas is much higher in water. This leads to precipitation of CaCO3. Water in the cave
may also evaporate to some extent, causing precipitation of CaCO3. In either case, CaCO3
is precipitated from the ceiling to form elongated near-vertical bodies of lime stretching
from the roof, and known as stalactites. Alternatively, lime-rich water may be evaporated
from the floor leading to the precipitation of CaCO3, which slowly builds upwards to form
incipient pillars called stalagmites. Stalagmites are also formed on the floor by water drip-
ping from the roof. In some cases stalactites and stalagmites coming from opposite direc-
tions meet to build pillars. The passages of floodwater entering from input points and
flowing through caves are geomorphologically very efficient. Floodwaters tend to enlarge
caves and openings both by abrasion from clastic load carried in suspension and along the
bed of the underground streams, and by corrosion. Corrosion may occur high up the walls
and even at roof levels.
The hydrologic effect of caves varies. A number simply operate as an underground drain-
ing system. Others may transfer water across topographic divides as underground water
flows follow hydraulic gradients, not surface slopes. Such underground transfers may lead
to stream piracy or cutoffs across rock meanders.

15.6╇ Karst in the tropics

Research on karst landscapes located in the tropics really started in the twentieth century
(Lehmann, 1936; Sweeting, 1958). The morphological difference between the tropical forms
and the type of landscape common in the standard karst regions, such as the high plateau
on the eastern Adriatic coast, northern England and the southern Appalachian Plateau, was
noted. Cockpits, cone karsts and tower karsts were accepted as characteristic features of
the tropical karst. Such forms have been described in the literature but a complete account
of their origin is still elusive. The morphological characteristics of the tropical karst have
been attributed in a general way to high temperature and precipitation, dense vegetation
282 Tropical karst

and the high production of organic acids. The karst of Southeast Asia, for example, was
described by Jennings (1985) as the ‘botanical hothouse extreme’. Viles (1984) discussed
karstic landforms produced largely by direct biological processes. Boring plant filaments
from colonies of blue-green algae, red algae and diatoms have been observed as etching
agents below the colonies. These activities are found on bare rocks in warmer climates and
particularly on seacoasts, as described from the small Caribbean island of Hell (Folk et al.,
1973).
Gillieson (2005) recognised the karst in Southeast Asia as most diverse, and listed a
number of factors€– varied geology, uplift history, eustatic change, past and present cli-
mate€– to account for this diversity. In Southeast Asia, karst occurs from the sea level to
nearly 4000 m, comprising large plateaus with dolines, cone karsts, tower karsts and low-
lying swamps. All karst in the humid tropics probably displays a diversity of similar forms,
some of which are less common in the temperate regions.
A standardised terminology for large-scale karst morphology of the type that is common
in the tropics has been borrowed from karst in southern China. The type occurs near Guilin,
Guanxi Province, China, where karst is formed on pure, thick limestone. This limestone
has been uplifted and subsequently subjected to a warm and humid monsoon climate over
a long period. Three principal karst styles have been recognised, the description of which
has been taken from Gillieson (2005).
1. Fenglin (tower karst):€a number of limestone towers that rise above a plain, commonly
under alluvium.
2. Fengcong (cone karst):€ a limestone massif with a large number of closed polygonal
depressions, sharing a common rock base and usually at a higher elevation than the sur-
rounding area. The shapes of the conical hills separating the depressions are determined
by local structure and rock type.
3. Gufeng (isolated towers):€single towers rising out of an alluvial plain formed by extreme
dissection of fenglin by meandering rivers. The rivers isolate blocks of limestone and
they are subsequently eroded by a combination of undercutting, cave development at
the foot of the blocks/hills, and material spalling off cliffs.
Such karst areas of high relief occur throughout the humid tropics, modified by local envir-
onmental factors. For example, the karst near Ipoh in the Kinta Valley of Malaysia shows
types of both the fenglin and fengcong style of karst. On the eastern side of the valley,
the rugged towers resemble fenglin; on the westen side they are more like fengcong with
gentler relief. Gillieson (2005) has ascribed the difference to greater metamorphism of the
rocks on the eastern side, as they are next to a granitic upland.
Some tropical examples of the high relief karst of towers and cones are better known
than others. The Chocolate Hills of Bohol Island of the Philippines are a good example.
This is a landscape of smooth, cone-shaped, isolated residuals of limestone. The cones have
been locally metamorphosed and reflect the gentle folding on a regional scale. Several of
these present the classic cone karst (fengcong) or modified forms of high-relief karst. The
Gunung Sewu area of southern Java is also a good example of both. Here karst and vol-
canic deposits have been juxtaposed, with thick volcanic ash filling the depressions and
cones and cockpits being aligned with faults and major lineaments. Highly concentrated
283 Karst in the tropics

organic acids passing through the bedrock and across the surface of the limestone mark the
tropical towers and limestone walls with solution rills and ripples. Swamp slots up to sev-
eral metres deep, filled with fine-grained alluvium, also occur at or near the base of towers
and limestone walls, indicating the location of past or present water levels. These are at the
elevations where the alluvial plains met the limestone towers.
Another characteristic landform of the humid tropics karst is pinnacles, a spectacular
landform covering a small area. The pinnacles are sharp-edged vertical blades of lime-
stone, up to 50 m tall and closely packed. They may also penetrate deep into the subsurface.
Thick organic soils, root mats and deep fissures separate individual pinnacles from each
other. Their origin has been attributed to structural control and intense rainfall. Pinnacles
are best developed in massive limestone with well-developed horizontal and vertical joints
(Osmaston, 1980; Waltham, 1995). Several hectares may contain hundreds of pinnacles.
A classic location occurs on Gunnung Api, Sarawak, at an elevation of 1100 m, with pin-
nacles crowding a shallow bowl-shaped depression about 2 ha in area.
Certain landforms such as poljes and dolines may occur in the tropics without mor-
phological modification. Poljes are common on tropical limestone plateaus in Malaysia,
Viet Nam and New Guinea. They tend to be underlain by alluvium or an impermeable
sedimentary rock such as shale or sandstone. Spring-fed streams may traverse such poljes.
Dolines may occur with the standard morphology or modified as star-shaped depressions.
The star-shaped depressions are enclosed by convex hillslopes, and steep gullies flow to a
sink in the centre of the depression.
The most spectacular example of a tropical karst arguably comes from the Gunung Mulu
National Park in the Malaysian state of Sarawak, on the island of Borneo. This involves a
1500 m thick Tertiary limestone that has given rise to a spectacular line of hills deeply dis-
sected by river gorges into four separate units, as shown in Figure 15.6. The limestone is
pure, massively bedded, and dips steeply to the northwest while faults and joints run north-
east. The upper surfaces of the hills are very rugged with structure-guided pinnacle karst,
deep gullies and cliffs. Locally, limestone hills display the standard forms:€dolines, valleys
and caves. Several of these features are impressive, e.g. the 1 km diameter collapsed doline
towards the south known as the Garden of Eden. The rivers shown in Figure 15.6 have cut
deeply incised gorges overlooked by 300 m cliffs and high-level caves with alluvial fans at
the contact between the limestone hills and the gorge floors. The presence of past alluvial
fans and the wide vertical range of caves in the hills indicate a history of regional uplift.
The caves of Gunung Mulu (Mulu Hills) are famous, with about a hundred known so far
and a total mapped length of 295 km (Brook et al., 1982). The caves include large passages
that follow strikes, excellent examples of speleothems, and evidence of past floods. The
assemblage includes Deer Cave, which with a 120–150 m diameter is the largest natural
cave passage in the world; and the fault-guided Sarawak Chamber in Gua Nasib Bagus
(Good Luck Cave), the largest chamber reported inside a karst cave. It is 600 m long,
415€m wide and 80 m high. A number of other very large caves and passages occur in the
Mulu Hills. Flow through the caves is still active. Seasonal rivers with outside sources rise
in floods and continue to aggressively develop the openings to these caves.
Karst features near sea level have been repeatedly submerged and exposed due to sea-
level changes during the Quaternary. This has given rise to magnificent tropical limestone
284 Tropical karst

Fig. 15.6 Limestone massifs in the Gunung Mulu National Park, Sarawak, Malaysia. From Gillieson, 2005. By permission of
Oxford University Press

coasts as in the Ha Long Bay of north Viet Nam, where a large number of towers emerge
above water as spectacular steep-sided islands separated by narrow channels.

15.7╇ Tropical karst as an environment

Karst geomorphology in the humid tropics is fascinating because of the rich development
of the standard karst landforms, the prevalence of high-relief karst and the scale of the
underground openings. Tropical karst regions are also associated with resource develop-
ment and ecological niches. As a number of tropical countries with karstic areas are also
285 Tropical karst as an environment

Fig. 15.7 IKONOS image of coast near Phang Nga, southern Thailand. Limestone towers are partially inundated by a Holocene
rise in sea level and surrounded by mangroves. Boats provide scale. Seawater frequently invades sinkholes in karst
giving rise to a spectacular landscape. IKONOS satellite image © Centre for Remote Imaging, Sensing and Processing,
National University of Singapore (2007), reproduced with permission. See also colour plate section

developing countries, such associations are important and require in-depth studies of trop-
ical karst geomorphology.
For example, limestone in the Malay Peninsula, as in the Kinta Valley discussed earlier in the
chapter, occurs next to granitic bodies. Hydrothermal mineralisation has given rise to deposits
of tin, tungsten and other minor ores, part of which occurs inside the limestone, like the vein
deposits of tin-bearing cassiterite in the walls and ceilings of caves of the Kinta Valley.
The change in sea level associated with glacial advances and retreats during the
Pleistocene affected karst topography near sea level. During the last advance of ice, lime-
stone next to the sea was exposed subaerially, leading to the development of karst topog-
raphy. The subsequent Holocene rise of sea level inundated part of the topography, creating
a coast with cliffs, caves and fascinating forms such as blowholes which developed from
limestone shafts. This has resulted in the development of tourism in many areas where
the coastal features are essentially partially submerged karst (Fig. 15.7). Good examples
are the innumerable islands and narrow passages of Ha Long Bay of north Viet Nam that
developed from a rising sea level inundating the lower part of a tower karst landscape.
Similar features are also seen in the Malaysian coastal resort of Langkawi and part of the
coast of southern Thailand. This connection between a partially submerged karst landscape
and tourism is common in many parts of the world.
286 Tropical karst

In the tropics, karst areas are often recognised as maintaining a range of biodiversity not
found elsewhere. The range of hydrology, temperature and light all lead to the develop-
ment of specialised flora and fauna on the exposed surface of the karst and in the dark pas-
sages underground. Research in tropical karst still has the lure of the dark unknown. The
karst environment in the tropics also provides the backdrop for agriculture (polje floors
and limestone pavements), specialised forestry (in high-relief karst areas) and water man-
agement. A basic knowledge of the geomorphology of karsts in the tropics is therefore
extremely useful.

Questions

1. What are the essential environmental conditions for the development of a good karst
landscape? Are such conditions common in the humid tropics?
2. Where would you expect good karst topography in the tropics?
3. What characteristics of karst are traditionally attributed to the tropical climate?
4. Describe cockpits, cones and tower karsts.
5. Examine Figure 15.7. What can you say about the origin of the landscape from this
IKONOS satellite image? Why are such areas becoming economically important?
16 Quaternary in the tropics

This is like déjà vu all over again.


attributed to Yogi Berra, New York Yankees

16.1╇ Introduction

The present landforms, geomorphic processes and sedimentary deposits in the tropics have
been influenced by events which took place in the Neogene. Events in the Tertiary included
the building of two huge mountain chains, the Himalaya and Andes, which in turn modi-
fied the tropical climate. For example, the origin of the Himalayan Ranges and the uplift
of the Tibetan Plateau either created the South Asian Monsoon system, or strengthened it
immensely from its previous state. This gave rise to a strong seasonality and storminess in
regional rainfall that translated into seasonal river discharges and high-magnitude floods in
the rivers of South, Southeast and East Asia. Prell and Kutzbach (1992) associated the ris-
ing of the eastern Himalaya with an east–west decrease in precipitation along the mountain
chain, enhanced aridity in Rajasthan and western Sahara, and cooler conditions over north-
ern Asia and Europe. The very high mean elevation of the Tibetan Plateau (4000–5000€m)
prevented the influx of the monsoon from the south, leading to aridity and a low level of
glaciation north of the Himalayan chain (Gasse and Derbyshire, 1996; An et al., 2001).
The world subsequently went through a repeated pattern of glacial and interglacial times
in the Quaternary.
Although explanations of Quaternary ice ages are primarily based on variations of the
Earth’s orbital geometry that changed the seasonal distribution of solar radiation at inter-
vals of tens of thousands of years (Imbrie and Imbrie, 1979), events that occurred in the
Tertiary€– such as the closure of the Panama Isthmus, the expansion of ice over Antarctica,
the building up of ice sheets in temperate latitudes, the expansion of tropical grasslands
replacing forests and the development of high plateaus€– also helped to create favourable
conditions for a cold period (Ruddiman and Kutzbach, 1991; Raymo and Ruddiman, 1992;
Quade et al., 1995; Derbyshire, 1996; Williams et al., 1998). For example, the closing of
the Panama Isthmus prevented warm water from the Pacific entering the Atlantic to keep
the North Atlantic warm.
In the Quaternary, the temperate and polar areas of the world underwent tremendous
changes due to alternating glacial and interglacial periods. In the tropics, the Quaternary
was also a time of repeated change associated with cold and warm periods, although the
effect of direct glaciation, unlike in the higher latitudes, was limited to high mountains.
287
288 Quaternary in the tropics

Such mountains included the Himalaya and Andes, the high volcanoes as in East Africa
and isolated elevated peaks like Mount Kinabalu in Sabah, East Malaysia. On the other
hand, repeated climate and sea-level changes have left strong impressions in the tropics.
The Quaternary in the tropics is not as well studied as in the higher latitudes. Certain
summary observations, however, can be made. The landforms and geomorphic processes
in the tropics were mainly affected by three factors:

• enhanced glaciation in the mountains


• repeated climate change
• repeated changes in sea level.
These three changes operated simultaneously and often in an interrelated fashion. Large
rivers tend to originate in high mountains, flow through a long valley and finally reach
the sea beyond a delta or a coastal plain. The upper basin of such a river is affected by
glaciations and climate change, the middle part in the valley by climate change and the
lower part mostly by changes in sea level. Thus, different but simultaneous changes
affected different parts of the rivers, and the river as a system had to adjust to all the
changes. Such events may have happened multiple times. The evidences of the three
major changes are incorporated in current tropical landforms, geomorphic processes,
soils and subsurface alluvium. Events in the Quaternary explain the presence of relict
alluvium in tropical river valleys, palaeochannels of major rivers, and the origins of
deltas and coastal plains.
These changes are discussed separately in the following pages, followed by two case
studies which illustrate the simultaneous modifications of different parts in a geomorphic
system. The two examples come from the Ganga River system and coastal Southeast Asia.

16.2╇ History and structure of the Quaternary

The Pleistocene epoch of the Quaternary period covers approximately the last 2 million
years and is marked by a number of glacial advances and retreats. The cooling started in
the late Pliocene, but the Pleistocene is the period of repeated advances of glaciers sepa-
rated by warmer times. From the middle of the nineteenth century, land-based glacial sedi-
ment and relict landforms contributed immensely to our knowledge about glaciations in the
Pleistocene. But the sequence and dates were only worked out much later in the twentieth
century, when cores from the ocean floor could be examined.
Scientists examine cores from the deep oceans for various attributes. D. B. Ericson
plotted the fluctuations in a planktonic foram (Globorotalia menardii) in sea-floor cores
from the Caribbean. Its abundance varied at different depths; the deeper the core sample,
the older was the material. As the foram is temperature-dependent, the fluctuation indi-
cated a time-based pattern of alternating cold and warm intervals. When C. Emiliani plot-
ted oxygen isotopes from the same cores, he noted a similar pattern of variation. Skeletons
of marine organisms carry two isotopes of oxygen, O16 and O18, the same as seawater.
O16 evaporates at a higher rate from the oceans than the other isotope, and when glaciers
289 History and structure of the Quaternary

Fig. 16.1 Estimated sea-surface temperature from the core V23–82, Northwest Atlantic. After Bradley, 1985

expand on land a progressively greater amount of O16 is locked up in the ice, and the
oceans become richer in O18. As the climate becomes warm, the ice melts and the ratio of
the isotopes in the oceans changes back in favour of O16. This type of change in the ratio of
these isotopes in the core samples indicated the repeated advance and retreat of glaciers,
of cold glacial and warm interglacial times. Further investigations standardised the series
of changes, which is known as the marine oxygen isotope series (MIS). Dates for specific
changes, numbered sequentially from the top, have also been established (Shackleton and
Opdyke, 1973; Imbrie and Imbrie, 1979).
Twenty-six isotope stages have been identified for the last one million years, and the
number rises to 65 when these are traced to the beginning of the Pleistocene. It has also
been established that the change from a warm to a cold climate was slow and gradual,
but the other type of change, from cold to warm, was very swift, a characteristic that has
been described as a sawtooth pattern. Figure 16.1 illustrates the last five MIS changes.
The Holocene (the present interglacial) is MIS 1, the previous full-interglacial is stage
5e (128€000–111 000 years ago). In between, there are several glacial stages and warmer
periods, which are not real interglacials.
In the tropics, the repeated advance and retreat of glaciers only happened in the higher
parts of tropical mountains. In comparison, the effect of the other two changes, climate and
sea level, affected the entire tropics, although not with the same intensity everywhere. The
landforms and sediments in the tropics are partly relicts from the Pleistocene. The modi-
fications were not as direct as that of ice sheets moving across land surfaces in the higher
latitudes, but more subtle, such as changing the discharge pattern and amount of sediment
load in rivers. This in turn affected alluvial valleys and deltas. Our knowledge about such
modifications is derived mostly from the last glacial advance which lasted approximately
290 Quaternary in the tropics

35 000–10 000 years BP, peaking roughly around 18 000. The previous climatic changes
appear to have followed similar patterns.

16.3╇ Quaternary glaciation in the tropics

Glaciers tend to advance during the cold stages and retreat during the warm ones. The
snow line therefore moved up and down tropical mountains repeatedly, with the same
happening for the upper limit of vegetation. Besides the low temperature, advancing gla-
ciers require a considerable supply of moisture. This is illustrated by the limited move-
ment of glaciers in many parts of the tropics around 18 000 years BP, identified as the
Last Glacial Maximum (LGM). It was a very cold period, but the rather dry climate pre-
vented glaciers from expanding in places such as the Himalaya. A current illustration is
the location of the present glaciers on the high volcanic massifs of the Andes:€Cayambe,
Antisana, Cotopaxi and Chimborazo. They are all on the wet windward eastern side
(Clapperton et al., 1997).
Van der Hammen et al. (1981) mentioned glaciers in the Colombian Andes reaching
their greater extent earlier than the rest of the world, as precipitation was high in this loca-
tion before the LGM but not later when the climate remained cold but very dry. In contrast,
Himalayan glaciers advanced briefly during the Early Holocene, when the climate was not
the coldest but considerable moisture was present. Since the Middle and Late Holocene,
glaciers on tropical mountains have been in retreat.
The Andes in South America went through periods of accelerated glaciation during the
Pleistocene (Clapperton, 1983). The largest development of ice was in the south beyond the
tropics, in Chile and Argentina. Enlarged cirques and valley glaciers, however, also devel-
oped in the tropical north, on the higher peaks. A number of the high peaks of the Andes
rise to more than 5000 m and currently carry glaciers. The Andean snowline of the glacial
period has been identified as 400–1000 m lower than the present. A much larger area was
then under ice. A number of plateaus, higher than 4000 m, occur in the Andean Cordillera,
plateaus which are ice-free at present. These supported small ice caps and glacier systems
during the colder times in the Pleistocene. Batholithic massifs and folded ranges that rose
to higher elevations developed large mountain ice caps. Huge troughs, up to 16 km long
and 500–800 m deep were eroded into such massifs, in places exploiting zones of structural
weakness. It has been suggested that during this period, lakes occupied a number of inter-
montane basins or tectonic troughs of the Andes. Such pluvial lakes have been attributed
to meltwater from neighbouring glacial systems and higher precipitation. Some of these
basins no longer carry a lake, whereas others, like Lake Titicaca, became smaller. Erosional
benches and terraces round Titicaca indicate a past bigger area. The warmer Holocene
probably started around 11 000 years BP in the Andes, taking a couple of thousand years to
peak. This period of hypsithermal conditions was followed by a worsening of the climate,
and the last 5000 years have been colder and marked by minor glacial advances. This pat-
tern has changed from the beginning of the eighteenth century, resulting in a marked retreat
291 Quaternary glaciation in the tropics

of Andean ice (Clapperton, 1983). Clapperton has referred to the occurrence of massive
arcs of both lateral and terminal moraines extending downvalley beyond the moraines of
the Holocene epoch. Stadials commonly occur behind the older terminal moraines.
During glacial times, ice accumulated not only on the Andes or Himalaya, which are
very high mountains, but also on highlands as in Australia and Papua New Guinea. In the
latter case, snow lines were lowered by about 1000 m. This led to the development of small
cirques and valley glaciers, and extended periglacial activities, which now are identified as
relict features (Williams et al., 1998). In Southeast Asia, the current snowline is at 4650€m.
Glaciers occur above 4620 m on Mount Jaya in Papua (Indonesia) which rises to 4884 m.
Currently, three very small glaciers, covering a total of about 7.5 km2 are found in these
mountains (Allison et al., 1989). Even at this near-equatorial location, glaciation occurred
during the last cold period, and at the LGM 18 000 years ago, ice covered 1600 km2 in
Papua New Guinea and a small area of 5 km2 on the summit of the 4101 m high granitic
pluton of Mount Kinabalu in Sabah (Hope, 2005). Hope (2005) suggested that the snowline
was lowered by about 1000 m. Snow could have been common on volcanic peaks above
3000 m in the islands of Sumatra, Java, Lombok and Sulawesi, but these peaks do not seem
to be high enough to support glaciers.
Similarly, enhanced glaciations happened only on high mountains in Africa, which
usually are volcanic peaks and not a continuous range. Knowledge of glaciations in
Africa is fragmentary and the best account available is for the mountains of East Africa.
Some of the high peaks, Mounts Kenya, Kilimanjaro and Ruwenzori, currently carry
glaciers, but the geomorphological and sedimentological evidence indicate a more
extensive glaciation that also occurred on the Aberdares and Mount Elgon, which are no
longer under ice (Mahaney, 1989). Hastenrath (1984) estimated that a maximum possible
total of 800 km2 was glaciated on these mountains. The present extent of glacial ice is
about 10 km2. It is likely that multiple phases of glaciations occurred on the East African
Mountains, separated by warm interglacial periods. Evidence of past glaciation has also
been reported from the mountains of Ethiopia:€ over 600 km2 in the Bale Mountains,
160€km2 on Mount Badda and 10 km2 in the Simen (Messerli et al., 1980). These high
areas no longer carry glaciers. Glaciation has been suggested on mountains in the Sahara,
in the Drakensberg and the mountains of the Western Cape, but such suggestions are still
at a hypothetical stage.
Extensions of glaciers to lower elevations left their mark on the tropical highlands by
eroding large U-shaped valleys that are currently drained by river systems. A number of
valleys also carry a large amount of glaciogenic sediment, which now provides the rivers
with considerable amounts of clastic load. It is also possible that the retreat of glaciers has
produced pressure-release joints on the valley walls, leading to enhanced slope failures. A
cold climate has left its mark on the mountain landscape and the sediment of headwater
streams in parts of the tropics. The glacial–interglacial change happened a number of times,
but in most cases, only the last glacial advance near the LGM and the sequence of events
after that can be clearly inferred.
292 Quaternary in the tropics

16.4╇ Climate change

Information regarding climate change in the Late Pleistocene and Holocene is derived
indirectly. For example, it comes from ice and sea-floor cores, the reconstruction of vege-
tation history using pollens and the palaeohydrology of arid areas such as old lakes and
river sediment. Collectively these are known as climate proxies. Computer simulations of
palaeoclimates also provide valuable insights. In general, fluctuations in the climate of the
tropics match the isotope stages. It has been suggested (though it is still not a definite con-
clusion and probably did not happen everywhere in the tropics) that glacial periods were
associated with colder, drier, more seasonal climates, whereas interglacials were warmer
and wetter times.
A number of case studies from different parts of the tropics indicate a cooling of several
degrees C on land, and even a cooling of the ocean waters. Based on terrestrial pollens, Van
der Kaars and Dam (1995) inferred a 4–7°C drop in temperature in Java. Similar figures
have been reported for the Amazon Basin, e.g. Colinvaux et al. (1996) indicated a glacial
cooling of 5–6°C from pollen information. The tropical oceans were found to be cooler by
approximately the same amount near Barbados (Guilderson et al., 1994), in the tropical
Atlantic (Thompson et al., 1995) and the Southwest Pacific (Beck et al., 1997).
Evidences for climate change are common in rain forests (Kellman and Tackaberry,
1997) and deserts (Williams et al., 1998), two completely different ecosystems. Of the
three large blocks of rain forest, Indo-Malesian, African and South American, the Indo-
Malesian forest of South and Southeast Asia was the least affected, presumably because of
its occurrence on peninsulas and islands where a wet maritime influence persisted to some
extent. Controversy exists about the exact extent of the Amazonian rain forest during the
glacial periods. It is likely that it retreated to a smaller area during glacial times and the rela-
tively drier edges were probably replaced by grassland. Rain forests over West and Central
African lowlands were probably the most affected. Vegetation change at such a scale would
undoubtedly impact the regional river discharge, erosion and sediment transfer.
Williams et al. (1998) described deserts as remarkable repositories of palaeoclimatic
information. Such information comes from relict landforms that are the result of processes
that have stopped operating or have changed. Evidence such as sediment from past lakes,
the location of old and defunct river channels, fossil dunes and fossil fauna provide infor-
mation about past climates in arid areas. As discussed in Chapter 12, landforms in deserts
may be very old but that age does not necessarily reflect the surficial sediment (Williams
et al., 1998). Climate change is reflected in such sediments and depositional forms created
by them, e.g. sand dunes, palaeosols, and river and lake deposits. Collectively, these young
forms and sediment indicate periods of wetter and drier phases in the Pleistocene.
An onset of aridity from about 2.5 million years has been identified in several parts of the
tropics. Large inland lakes started to dry out in the Sahara, Ethiopia, the Arabian Peninsula,
India, Australia and southern Africa. Glacial periods in tropical deserts were arid times.
Examination of equatorial Atlantic deep-sea cores indicates an association between dust
from the Sahara, times of glacial maxima and low sea surface temperature (SST). Data
from the Gulf of Aden and the Red Sea match this pattern. Williams, et al. (1998) included
293 Climate change

a summary discussion of this characteristic, using data from several sources, and recognis-
ing dry periods during glacial maxima and wet ones during interglacials, although a caution
against oversimplification is included. It has been established, however, that at the LGM,
18 000 years ago, tropical areas were more arid and cold, trade winds blew more strongly,
active desert dunes were much more widespread than their present distribution, and desert
dust and loess were carried out and deposited on both land and sea.
Rivers in the tropics, especially those that originate in high mountains and flow into
the sea or an inland lake, also went through a series of adjustments during the Quaternary.
Their base level changed repeatedly with the rise and fall of the sea level (as discussed in
section 16.5); their discharge and flood patterns fluctuated due to climate change; and the
amount and nature of their sediment also altered, matching a changing precipitation pat-
tern and vegetation cover over their drainage basins. Evidence for this, at various scales, is
derived from the subsurface alluvium in the river valleys, preserved palaeochannels, evi-
dence of former large floods in slackwater deposits (Box. 16.1) and deltas. Deltas include
both the present onshore one and, in many cases, an older submerged extension. The story
varies for each river, and a case study on the Ganga is discussed in section 16.7.
The size and volume of lakes also altered between the glacials and interglacials. This
has been studied for many desert lakes, and evidences for lakes caused by meltwater and
natural dams in glaciated mountain valleys are also available. In Australia, such former
lakes are referred to as megalakes. Megalakes covered larger areas of a structural basin and
the extra water was available due to a wetter climate (higher rainfall, reduced evaporation,
increased cloud cover) or changes in runoff following a different pattern of vegetation
cover in catchment basins (Bowler, 1981).
Climate change in the tropics, however, is a complex story, further complicated by the
feedback from changes in sea level. A higher sea level during a warm climate increased
maritime characteristics over land and probably increased precipitation. A lower level in a
glacial stage brought in continentality and a drier climate.
Changes in climate continued in the Holocene. The Early Holocene was wetter in a
number of locations including the Sahara and monsoon Asia. Changes in climate also influ-
enced regional palaeohydrology, as seen in the alluvium and morphology of a number of
rivers including the Amazon, Nile, Narmada and Murray Darling. Pollen studies from peat
swamps (commonly on tropical mountain slopes) also indicate similar climate fluctuations
in the tropics.
Our knowledge regarding climate change is best for the last pair (Fig. 16.1) of glacial and
interglacial times. Duplessy (1982) showed a weak southwest monsoon and a strong north-
east one using oxygen isotopes of the planktonic formaminifera deposited on the floors of
the Arabian Sea, northern Indian Ocean and the Bay of Bengal near the LGM, about 18 000
years ago. A number of rivers that drain to the Bay of Bengal€– Ganga, Brahmaputra and
Salween€– carried reduced discharges during this period of weak southwest monsoon.
In contrast, the southwest monsoon at the onset of the Holocene and the Early Holocene
was a warmer and wetter time. Goodbred and Kuehl (2000a) estimated that between 11 000
and 7000 years BP, the sediment released by the Ganga–Brahmaputra system was double
that of the present amount and this increased rate was sustained for 4000 years. Given that
the present annual sediment discharges of these rivers total about a billion tonnes, the Early
294 Quaternary in the tropics

Box 16.1 Slackwater deposits


Slackwater deposits tend to be fine-grained sand and silt that accumulated rapidly from suspension during
large floods. These are deposited in areas where the velocity of a river is reduced, and are preserved in loca-
tions which are protected from erosion in subsequent floods (Baker et al., 1983). As slackwater deposits fre-
quently contain datable material, usually for radiocarbon dating, and as several slackwater deposits may occur
in a vertical sequence, the flood history of a river can be extended for a long period of time. It is thus possible
to estimate the frequencies of past catastrophic floods.
River gorges in both rock and indurated alluvium are potential environments for accumulating and preserving
slackwater deposits. The commonest location is the mouth of tributaries where fine sediment is deposited by the
backflooding of tributary channels from the main river. As the water backfloods, its velocity rapidly decreases
and suspended sediment is deposited (Fig. 16.2). High-angled river confluences are particularly good for accu-
mulating slackwater deposits. Other suitable areas include the lee of bedrock protrusions in the channel, shallow
scallops in bedrock wall (rock shelters), abrupt width changes in channels and the inside of tributary meanders.
Thin horizontal laminations are frequently seen in slackwater deposits because of their fine texture and
process of deposition. Slackwater deposits can also be made of structureless fine sand or silt. Cross-bedded
units with foresets indicating an upvalley movement of water from the main river are occasionally seen.
Individual sedimentation units could be capped by fine-grained organic material. Other evidences, such as
buried palaeosols, mudcracks and rapid changes in sedimentation pattern, can be used to separate the slack-
water deposits of individual floods in a vertical sequence.
Slackwater deposits have been used to reconstruct past large floods and compute their recurrence inter-
vals. The flood history of a river can thus be extended beyond the recorded period. They are extremely useful
for determining the flood climate of the Holocene, going back thousands of years.

Fig. 16.2 Diagrammatic sketch of accumulation of slackwater deposit (S) in the Katherine Gorge, Northern Territory, Australia.
From Baker et al., 1983. By permission of Wiley
295 Sea-level change

Holocene must have been a time of strong monsoons and accelerated erosion. A number
of studies on lake sediments from the Thar Desert of western India similarly indicate a
wetter Early Holocene when the lake levels were higher. This was followed by drier condi-
tions and less moisture availability and ultimately the present-day arid conditions (Wasson
et€al., 1984). Large-scale climate changes between the last glacial advance and the Early
Holocene have also been reported from Australia. Kershaw (1978) reviewed the case of
increased rainfall and a rapid spread of rain forest in northeastern Queensland.
Enzel et al. (1999) demonstrated changes in climate in northwestern India during the
Holocene from a careful analysis of sediments from the Lunkaransar dry lake in Rajasthan.
The shallow lake levels fluctuated frequently in the Early Holocene and then abruptly
rose to a higher level around 6300 C14 years BP. It dried out completely around 4800 C14
years BP along with a period of intense dune destabilisation. We should recognise that
climate changed repeatedly even in the Holocene in the tropics. Wasson (1995) concluded
that a peak in available moisture occurred in many regions between approximately 6000
and 8000 years BP. He listed monsoon Australia, India, Arabia, China, Japan, Taiwan and
Thailand. The departures from this widespread pattern occurred in southern India and east-
ern China.

16.5╇ Sea-level change

Repeated glacial advance and retreat during the Pleistocene were associated with the rise
and fall of the sea level. The fall in sea level during the last glaciation, MIS 2, and the rise
afterwards are the best-known. Similar instances have occurred with other isotope stages,
but the history of such changes is less well documented. The falling sea level tends to be
slow; a rate of 1–5 m per thousand years has been suggested by Williams et al. (1998). The
rise is faster; the average rate is 5–10 m in a thousand years (Williams et al., 1998).
It is generally accepted that the drop in sea level during MIS 2 was about 120 m, but
the exact amount has been different in places. Areally, such a drop is most effective where
shallow seas existed on cratons:€ exposing large areas of the sea floor, linking offshore
islands to the mainland and providing land bridges. This happened globally, and a good
tropical example comes from Southeast Asia where the retreat of the sea for hundreds of
kilometres across the shallow Sunda Shelf linked mainland Asia with islands of western
Indonesia and those in the South China Sea (Biswas, 1973; Batchelor, 1979). Similarly, the
exposed Sahul Shelf connected northern Australia to Papua New Guinea (Fig. 16.3).
Such changes not only affected coastal landforms, the distribution of mangroves and
coral reefs, but on a bigger scale, they also affected regional climate, ocean currents, river
systems, and the diffusion of flora and fauna. These changes can be reconstructed from
species distribution and relict sediment. The story is more complicated along active coasts
and coasts of volcanic islands, where corrections need to be made for post-change tec-
tonic movements. It is a curious fact that clearer evidences for former sea levels tend to
be recorded in such unstable regions, presumably because the regional uplift preserves
a former sea-level marker by raising its elevation. Such records come from Papua New
296 Quaternary in the tropics

Fig. 16.3 Approximate extent of emergence of land around Southeast Asia (exposing part of the Sunda Shelf) and Australia
(exposing part of the Sahul Shelf) during the last glacial maximum. The land connection permitted distant migration
of flora and fauna from the mainland of Southeast Asia. Later, a rising sea level separated part of the land mass into
islands, isolating such flora and fauna from the mainland. Note the major river systems on the Sunda Shelf. These
drainage networks were connected with headwaters which now form detached river systems on land after the rising
sea level inundated lower trunk valleys and alluvial sediments. The exposure of the Sahul Shelf connected Australia
with Papua New Guinea and interrupted ocean currents and climatic parameters. The present pattern of ocean
currents and climate started with the drowning of the exposed land by a rising sea level

Guinea, Timor Leste (Chappell and Veeh, 1978), Sumba (Bard et al., 1996), Barbados
(Bender et al., 1979), Haiti (Dodge et al., 1983) and other places.
The most striking example of exposed coral communities due to a falling sea level dur-
ing the Pleistocene comes from the Great Barrier Reef of Australia. The fall of the sea level
left this line of reef as a ridge, approaching 160 m above the LGM sea level. The exposed
limestone would have undergone active dissolution, giving rise to the development of caves
and pinnacles. A subsequent rise in sea level inundated the reef, but its present appearance
297 Sea-level change

is thus partly due to geomorphic processes in operation during the former fall in sea level
(Carter and Johnson, 1986).
The distance from the sea could be responsible for reduced precipitation and an increase
in continentality on the landmasses. Lower sea surface temperatures (SSTs), as mentioned
earlier, may lead to less storminess, as warmer waters are required for a tropical storm
to develop. Climate change also followed the diversion of ocean currents following the
emergence of the former sea floor as a land barrier. The linking of northern Australia with
New Guinea across the Sahul Shelf diverted the warm South Equatorial Current and iso-
lated North Australia from an important supply of moisture from incoming maritime air
(Fig. 16.3). As a result, aridity was enhanced in northern Australia (Williams et al., 1998).
Afterwards, in the Early Holocene, summer moisture returned with a rising sea level, and
rain forests replaced eucalyptus forests and woodlands in northeastern Australia (Kershaw,
1978).
In some cases, a falling sea level exposed large land areas. Rivers extended their channels
across this area to meet the distant sea. In the process they incised their lower courses due
to the fall in base level and also linked up with other rivers to form bigger systems. Land
bridges also allowed the migration of flora and fauna from the mainland to the present-day
islands. The subsequent Holocene rise in sea level drowned the exposed continental shelf,
disrupting the lower parts of the extended river systems, and isolating the migrated flora
and fauna on islands. This happened in many places but with local variations. The case for
Southeast Asia is discussed as an example in section 16.7.
Current deltas at the mouths of rivers flowing to the sea were formed when the sea level
started to rise at the beginning of the Holocene. Stanley and Warne (1994) constructed a
model of delta-building using data from boreholes sunk into delta material. Radiocarbon
dates from the base of deltaic deposits suggest that delta-building started between 8500
and 6500 years ago. The rise of sea level slowed down after an impressive start during
which the exposed sea floor was inundated and the lower parts of former rivers were sub-
merged. The slowing provided accommodation space at the mouth of the rivers, sediment
was deposited and deltas started to build out. Thus, although a particular river may be
old, its delta is Early Holocene, although the exact dates may differ between deltas (see
Chapter 11).
From LGM onwards (about 18 000 years BP), the geomorphology of the tropics needed
to adjust simultaneously to both climate and sea-level changes. Climate change determined
the supply of water and sediment to rivers either directly by controlling temperature and
precipitation patterns, or indirectly by influencing the frequency of slope failures and vege-
tation cover. At the same time, the rise in sea level affected the lower course of rivers, when
water and sediment, adjusted to the environment of the upper basin, arrived. Usually this
is a complicated story which needs to be sorted out regionally. Figure 16.4 summarises
the changing environment of South and Southeast Asia, using data from various sources
including climate proxies (Kale et al., 2003).
Two regional examples are summarised in sections 16.6 and 16.7. They illustrate the
multiplicity of factors changing simultaneously in different parts of a geomorphic system.
The first is a summary of the changing form and behaviour of the Ganga River from MIS 3
(58–24 000 years) to the present, based on a source-to-sink study of the river by Goodbred
298 Quaternary in the tropics

Fig. 16.4 Changes in sea level and climate over Southeast Asia and South Asia. The two changes operated simultaneously.
Generalised from Kale et al., 2003

(2003). The second is an account of the sea-level change over the shallow Sunda Shelf in
Southeast Asia (Gupta et al., 1987).

16.6╇ The Ganga River system:€Quaternary adjustments

River systems as large as the Ganga integrate signals from a series of changes over a huge
area. The headwaters of a large river system produce most of the sediment that the river
transports. This is due to high precipitation on the windward side of the mountains and
tectonic movements that provide uplift, steep slopes and fractured rocks. Low-gradient,
mixed bedrock–alluvial valleys below the mountains mainly function as conveyor belts
transporting sediment (Fig. 8.1). Such sediment is deposited near the sea in the form of a
prism that extends to the highstand level of the sea. Incised valleys may extend under water
from the high to low stands of the shoreline, marking the distance the sea retreated during
glacial times (Castelltort and Van Den Driessche, 2003; Blum, 2007). Large rivers tend
to be polyzonal, indicating that not all of their tributary basins respond in the same way
to climate and other changes. The rivers may also traverse rock-cut and alluvial sections
alternately, requiring an adjustment at each transition. During the Pleistocene, such rivers
299 Quaternary changes around the Sunda Shelf

had to adjust near-simultaneously to enhanced glaciation in the mountains, climate change


over the basin and a falling sea level in glacial times. The direction of change reversed
during the interglacials.
The Ganga–Brahmaputra system is described in Chapter 9. The Ganga rises in the high
Himalaya from the Gangotri glacier, flows eastward though the Himalayan foreland basin
to meet with the Brahmaputra and Meghna, and with them builds the biggest delta in the
world before emptying to the Bay of Bengal. The Himalayan tributaries of the Ganga flow
through steep valleys, many of which have been glaciated and a number of which still
carry glaciers. As it flows east, the Ganga collects tributaries from the north, the larger
ones of which originate in the Himalaya. These rivers and the Ganga itself have deposited
megafans at the contact between the Himalaya Mountains and the plains. The southern
tributaries of the Ganga drain the Indian Peninsular craton and have no history of glaci-
ation. The system is very seasonal as far as water and sediment discharges are concerned,
and almost all the work of these rivers is carried out during the 4–5 months of the Indian
southwest monsoon.
The river was affected almost simultaneously by glaciations, climate change and sea-
level change in different parts of the system; and the changes were different at each marine
isotope stage. Goodbred (2003) summarised such changes from the high mountains to the
floor of the Bay of Bengal (Table 16.1). The length of the river varied, and the extended
river of glacial times left its mark on a submarine canyon, the Swatch of No Ground, and a
very large submarine fan extending into the Bay of Bengal (Fig. 16.5). The summary table
illustrates that temperature and precipitation in combination affected glaciers in mountains
and erosion/deposition in valleys, the length of trunk streams fluctuated, and, over time,
the centre of deposition shifted between deep-sea fans and deltas. The growth of the Bengal
Delta is discussed in detail in Chapter 11.

16.7╇ Quaternary changes around the Sunda Shelf

Sea-level cores from the middle of the South China Sea have revealed plant debris, oxi-
dised claystone, mottled sediment and laterite wash, indicating a former marine regression
that exposed a large area of the current sea floor (Biswas, 1973). Some controversy exists
regarding how much of this shallow sea over the cratonic Sunda Shelf was lowered, but a
figure of 120 m is generally accepted. Figure 16.3 indicates the extent of the retreat of the
sea and the exposure of land. It is likely that the retreat of the sea to such an extent would
bring a less maritime climate to the present coastal areas of Southeast Asia, which would
thus have a drier and seasonal climate. Verstappen (1975) proposed a rainfall gradient with
elevation, with reduced and seasonal rainfall in the lower areas. He postulated a tropical
forest and grassland vegetation replacing part of the present rain forest (Verstappen, 1980).
The exposed sea floor would have functioned as a land bridge, allowing the migration of
flora and fauna. The existing river systems that now flow into the South China Sea would
have extended across the exposed area and merged to form larger systems. At least three
such systems have been identified (Fig. 16.3).
Table 16.1╇ Source-to-sink response of the Ganga River system since MIS 3 (from Goodbred, 2003)
Climate of the River Himalayan Megafans and Delta and
Time region discharge Sediment valleys alluvial plains coastal shelf Deep-sea fan

MIS 3 Slightly cooler Moderate High sediment Marked glacial Aggrading Incision and Possible
(58–24 ka) and drier than transfer advance and megafans, possibly aggradation
present valley upper Ganga local
aggradation plains and aggradation
southern
tributaries
MIS 2/ LGM Cold and dry Low Low Limited Largely Inactive. Probably
(24–18 ka) glaciation inactive Possible inactive
and valley fans, minor incision River linked
aggradation incision, with the
alluvium Bengal Fan
in southern via canyon
tributaries
Hypsithermal Warm and very Very high Very high Brief glacial Widespread Very rapid Aggradation
(18–7 ka) wet, stronger advance. rapid ero- aggradation shifting to
monsoon Erosion of sion, nearly delta
valley fill to bedrock in
valleys
MidHolocene More relaxed High High Retreating gla- Low Aggradation Less
(7 ka to now monsoon, less ciers, sediment aggradation but slower aggradation
rainfall removal and
aggradation
301 Quaternary changes around the Sunda Shelf

Fig. 16.5 Diagrammatic sketch showing the relative position of the catchment of the Ganga, the Ganga–Brahmaputra Delta,
the Swatch of No Ground and the submarine Bengal Fan. Generalised from Goodbred, 2003

Molengraff (1921) recognised a North Sunda River flowing north between the present-
day Malay Peninsula and the island of Borneo, collecting tributaries from both these regions.
Emmel and Curray (1982) suggested that a group of rivers came down the western side of
the Malay Peninsula, met rivers coming off east Sumatra and the joint system flowed north-
westwards. They traced morphological changes along this system, identifying a braided
pattern in the upstream area which was replaced downstream by a meandering channel that
ended in a delta. A third major river flowed east, collecting left-bank tributaries from south
Borneo and right-bank tributaries from northeastern Sumatra and northern Java.
Afterwards, the rise in sea level inundated the exposed sea floor and truncated the river
systems. The higher parts of the land appeared as peninsulas and islands leading to the
302 Quaternary in the tropics

(a)

(b)

Fig. 16.6 Geomorphological processes operating in the Blue Nile Basin during (a) the Last Glacial Maximum (global scale) and
(b) the early Holocene. From Woodward et al., 2007. By permission of Wiley

isolation of flora and fauna on various islands with no further communication between
them. The coarse alluvium deposited by the braided headwaters of the past river systems
can be found in the subsurface in many parts of Southeast Asia. For example, it is known as
the Old Alluvium in Singapore (Gupta et al., 1987) and as the Simpang Formation in West
Malaysia. Similar deposits have been found in Sarawak, Thailand and southern Viet Nam.
It has been traced under the waters of the Malacca Strait, as a sedimentary cover lying over
a granitic regolith (Aleva et al., 1973; Batchelor, 1979). A complete picture of what exactly
happened, however, requires more investigation.

16.8╇ Conclusion

The tropics, like the rest of the world, underwent repeated changes during the last two
million years of the Quaternary. These changes included enhanced glaciations in the high
303 Questions

mountains, and changes in climate and sea level. The geomorphological effects of such
changes were enormous (Fig. 16.6). The explanation of current landforms and sediments
requires a careful investigation of such changes. It is also likely that the present climatic
conditions are part of this sequence of changes, and geomorphology in the tropics has
altered repeatedly throughout the Quaternary. A knowledge of what happened in the tropics
during the Pleistocene is therefore essential for understanding the current landforms and
processes (Nott and Hayne, 2001; Thomas et al., 2007) and, as discussed in Chapter 19,
the knowledge of Pleistocene and early Holocene conditions in the tropics should help us
in forecasting the possible geomorphic effects of climate change.

Questions

1. What kind of changes happened repeatedly in the tropics during the Quaternary?
2. List the mountainous areas where glacial ice accumulated during the colder phases in
the Pleistocene.
3. How do we deduce the nature of climate change during Late Pleistocene and Early
Holocene?
4. What is the evidence for a warmer Early Holocene?
5. It has been said that repeated glacial advances and retreats during the Pleistocene were
associated with the fall and rise of the sea level. What effect did it have on the river
systems in the tropics, starting from MIS 2?
6. It is generally accepted that during MIS 2, the drop in sea level was about 120 m. What
kind of tropics did exist during this time?
7. Slackwater deposits are considered as important indicators of palaeohydrology. Why?
Part III

Anthropogenic changes
Anthropogenic alteration of geomorphic
17
processes in the tropics

Man has too long forgotten that the earth was given to him for usufruct alone, not for
consumption, still less for profligate waste.
George Perkins Marsh

17.1╇ The beginning

Anthropogenic transformation of the physical environment changes processes and land-


forms. Large-scale alteration of the landscape started about 10 000 years BP, with the
beginning of agriculture which led to the destruction and replacement of natural vege-
tation. Additional demands for construction, urbanisation and shipbuilding made further
inroads into the forests. As natural vegetation was depleted, many areas were subjected to
a changing hydrology and geomorphology.
Rapp et al. (1972) provide a striking example from Tanzania (Fig. 17.1). Their study
area was a pediment sloping at 3.5° from an inselberg to a floodplain. A grass-covered,
thin, red, sandy-loam soil rested over the bedrock. The grass cover effectively prevented
soil and water loss. A number of experimental plots with a variety of land use were estab-
lished in mid-slope, and water and sediment loss measured from such plots. As Figure
17.1 shows, no sediment was lost from a plot under ungrazed thicket and surface runoff
was equivalent to only 0.4 per cent of rainfall. When thickets were replaced with grass, the
water lost by runoff increased to only 1.9 per cent of rainfall, but there was still no percep-
tible erosion. Losses of both water and sediment increased dramatically when millet was
planted in rows or the land was left as bare fallow. This pattern of increased runoff and
sediment loss happens with removal of vegetation everywhere; the numbers depend on the
local environment.
In 1864, George Perkins Marsh wrote Man and Nature, the title of which was changed
subsequently to The Earth as Modified by Human Action:€ a New Edition of Man and
Nature. He felt compelled to write this book in order to point out how much erosion and
sedimentation have increased due to anthropogenic actions. Similar landscape changes
have occurred in the tropical countries for a very long time, but it became considerably
more large-scale and intensive in the twentieth century. Given the nature of tropical rain-
fall, once natural vegetation that protects the thick soil and weathered material in the humid
tropics is destroyed, accelerated erosion and sediment transfer happen quickly (Fig. 17.1).
The increased runoff and erosion marks the slopes with gullies and scars of mass move-
ment, and the streams need to adjust to a new regime of water and sediment supply. In
307
308 Anthropogenic alteration of geomorphic processes

Fig. 17.1 Soil erosion with different land use covers from mid-slope at Mpwapwa, Tanzania. From Rapp et al., 1972.
By permission Wiley/Blackwell

many places, such excessive sediment even reaches the coast. The transformation of the
landform and process is particularly effective where the rainfall is intense, slopes are steep
and the region is tectonically unstable.
Destruction of vegetation and alteration of the environment is mainly due to:

• deforestation
• changes in land use and rural migration
• river impoundments
• urbanisation.
309 The beginning

Table 17.1╇ A summary set of sediment measurements from Southeast Asia


Type of land use Sediment yield (tkm−2yr−1)

Forest 100–102
Large river basins (mixed land use) 102–104, mostly 103
Urbanisation 103
Shifting cultivation 103 from plot measurements, 102 for entire basins
Agriculture 102–103, depending on conservation measures used
Logging, early stage 103, maximum recorded 17 000

Note Range of sediment yield compiled from a large number of published measurements in
Southeast Asia. From Gupta, 2005a

The effect on the physical environment from the first three is reviewed in this chapter and
the effect of urbanisation is discussed in the next one. Observations on anthropogenic modifi-
cation of landforms and processes in the tropics have increased since the 1960s. When vege-
tation is removed, soil and regolith are de-structured, natural slopes are altered, and erosion
rates and sediment yields reach figures which are accelerated and no longer natural. A com-
bination of anthropogenic alteration and fragile landforms give rise to very high local sedi-
ment yields. The resulting changes may affect not only proximal slopes and rivers, but also
distal deltas, beaches, mangroves and coral reefs. The changes have become large enough to
be mapped from satellite images:€both the areas that are being eroded and also the resulting
sediment plumes on the coast (Gupta and Krishnan, 1994; Gupta, 1996; Gupta, 2005a).
The degree of change varies with location, and it is not always possible to make a quan-
titative measurement with reasonable accuracy and robustness. Erosion and sediment loss
have been estimated for various locations using diverse methods, such as the calculation
of ground lowering in small erosion plots (Box 17.1), sediment-sampling in streams and
the rate of extension of a delta into the sea. Variations in local conditions and measurement
techniques make it difficult to rely on a single figure to represent the effect of a particular
type of anthropogenic change, such as deforestation or urbanisation. It is safer to use a range
of figures than an average of all the measurements available. It can, however, be stated that
the present rates of erosion and sedimentation are distinctly higher than earlier ones.
This is illustrated in Table 17.1, which summarises sediment measurements in Southeast
Asia. In spite of the generalisation involved, it is clear that sediment yield increases con-
siderably once the vegetation is removed and the soil is exposed to erosion. The increased
loss happens cumulatively over time and also with episodic jumps, as illustrated by the
landslide devastation of logged slopes and lower valleys in southern Thailand in November
1988 after about 1000 mm of rain in five days. A similar incident of soil erosion, huge
landslides and flooded rivers happened in the same year following deforestation in the
Bengkulu Province, Sumatra.
Therefore, the ongoing geomorphological processes and evolving landforms in many
parts of the world should not be considered entirely natural. The current rates of erosion,
sediment yield and building of sedimentary bodies such as bars in the river or beaches on
the coast may differ from past ones. Destruction of vegetation and other anthropogenic
modifications of the landscapes tend to degrade the environment and, at times, lower the
310 Anthropogenic alteration of geomorphic processes

Box 17.1 Measuring erosion rates and sediment loss


Various techniques exist for measuring erosion rates from a geomorphic unit such as a slope. They range from
very simple to rather complicated, and practices vary in efficiency and robustness. Sediment loss by erosion
can be measured in two ways. It can be determined as a measure of ground lowering in the area undergoing
erosion or the rate at which sediment is deposited at the lower end of the system. The measure of erosion or
sediment accumulation, depending on which technique is used, is generally expressed in volume or weight of
sediment per unit area, such as m3km−2 or tkm−2. Low sediment loss can also be expressed in kilograms and
small areas are often calculated in hectares, e.g. kgha−1. As the sediment loss (also called sediment yield) is
usually expressed as a rate, the total weight of sediment eroded in a year is expressed as tkm−2yr−1.
Several common techniques are listed below, but there are others. For details of application in the field
please consult a field manual. It is always good to have rates from several techniques on the same area so that
an idea about reliability can be formed.

1. Erosion plots. Probably the easiest technique. An area on a sloping surface is marked off and long pins are
inserted into the ground with washers below the pinhead lying directly on top of the ground surface. Over
time, material is eroded from the surface and the washer drops down the pin. Thus the amount of erosion
at various points of the marked area can be measured from which an overall rate can be computed. The
problem is in extrapolating the data from such a small area to a region.
2. Experimental plots. A small area, usually at the scale of about 10–20 m2 is selected (Fig. 17.2). A perimeter
trench is cut and partly filled with clay on top of which a metal sheet is planted which should extrude well
above the ground surface. The plot is thus insulated from subsurface flow and rainsplash. The plot should
be of trapezoidal or similar shape, so that water and sediment move towards the narrow end where storage
tanks can catch sediment-laden water after a spell of natural or artificial rainfall. Both rates of runoff and

Fig. 17.2â•…Experimental plot for measuring runoff and sediment loss looking downslope. Sediment-bearing
water is collected in the two containers in the background. The surface is currently bare. It is possible
to control the amount of groundcover for experimental purpose. Experimental plot designed by
Ausafur Rahman. Photograph: A. Gupta
311 Deforestation, land use changes and rural migration

sediment loss thus can be determined. The small-scale of the plot allows the plot land use to be changed
(bare, grass, shrubs, etc.), which is an advantage, but it too has the problem of extrapolation to a large area.
3. Surveying. Periodic surveys can be carried out over an area and with reference to monumented points. This
indicates which area is being eroded and where the sediment is going. This has been done with a wide
range of instruments, from simple hand levels to total stations, presumably depending on the resolution
required.
4. Suspended sediment in rivers. Suspended load in a stream can be measured in various ways, by using a
vertically integrated sampler or a set of sample bottles fixed to a base and collecting at various depths of
water. If the river does not carry much solution or huge amounts of bed load, the figure can be used to
represent the load of the river, as bed load is less than 10 per cent in most rivers, below the precision of
suspended load instruments. The amount is representative of a large area, the entire upstream drainage
basin, but it provides a general figure and as such would not indicate the variation in erosion between the
different parts of the basin.
5. Boreholes. Sedimentation rates can be determined by sinking boreholes in floodplains and deltas. Often it
is possible to date sediment from carbon samples collected at various depths. This provides a rate of sedi-
mentation which reflects upstream erosion, but this technique is applicable only for a long time-period,
such as a delta built up since Early Holocene.

quality of life. A geomorphologist may use this knowledge to ameliorate environmental


degradation associated with anthropogenic alteration of the landscape. This is one of the
fascinations of tropical geomorphology.

17.2╇ Deforestation, land use changes and rural migration

17.2.1╇ Causes of forest destruction

By the mid-twentieth century, forests of the world had been reduced to between one-half
and one-third of the original extent (Eckholm, 1976). A substantial part of this was the
tropical rain forest. Although other types of tropical vegetation were also destroyed to meet
various demands, it is the accelerated destruction of the rain forest that attracted many
researchers, including geomorphologists. Tropical light hardwood trees of the rain forest
are in high demand for industrial purposes. These trees are tall, frequently festooned with
creepers, and when they are felled they tend to bring down smaller trees with them. A very
large area therefore is cleared when the tropical rain forest is logged. Mechanical extraction
of such trees requires logging roads into the forest, which leads to further forest clearing. A
limited amount of firewood is collected from the rain forest, but firewood collection is far
more destructive for drier forests in the tropics. The collection of firewood is particularly
disastrous near desert fringes or on mountain slopes. In both cases depletion of the vegeta-
tion cover significantly increases erosion.
312 Anthropogenic alteration of geomorphic processes

Rain forests are destroyed not only for their resources but also for their land. Large-scale
land clearance was carried out in the second half of the twentieth century to settle migrant
farmers. This has happened both under government planning and from the spontaneous
migration of poor and landless people. A spread of pastures has destroyed forest in Central
and South America. A variety of other demands and developmental policies such as mining
and highway construction are also destroying tropical rain forest in various countries.
The causes of forest destruction vary in importance from forest to forest. Rain forests
have been destroyed for logging and plantation agriculture in Southeast Asia, unsustain-
able shifting cultivation in Central Africa, settlement and ranching in South and Central
America, and firewood collection in drier areas and on the slopes of the tropical mountains.
In the ten years between 1980 and 1990, it has been estimated that tropical forests have
been lost at an annual rate of 174 000 km2 (FAO 1993, as reported in World Resources
Institute, 1994). Another estimation by Trexler and Haugen (1994) put this �figure at
170€000 km2. The estimates are very close. The highest annual deforestation rates occur
in two regions:€(1) continental Southeast Asia and (2) Central America and Mexico. It is
the easily accessible lowland forest, commonly covering a rolling landscape, which was
destroyed most between 1980 and 1990. The fastest rate of loss, however, was from hill
and montane forests.

17.2.2╇ Effects of forest destruction

Forest clearance on the slopes of the humid tropics gives rise to slope failures and gully
formation, which periodically contribute large slugs of sediment to a river at the bottom
of the slope. If deforestation continues over a wide area, such sediment would accumulate
in a number of river channels, changing their flow pattern and physical character. In time,
the eroded material from deforested slopes would reach the coast as plumes of sediment
affecting coastal beaches, mangroves and coral reefs.
A longitudinal research project, measuring water and sediment changes that followed
alterations in land use started at the Danum Valley, Sabah, Malaysia in October 1987
(Douglas et al., 1992). The data were collected from hilly catchments between 0.5 and
10€ km2 in area and generally on easily eroded lithology. An unlogged catchment was
compared with one in a logging concession, where road construction started in June 1988
and logging in January 1989. After storm rainfalls, both the discharge and sediment load
increased in the stream which drained the logged basin. Sediment in the logged catchment
increased to four times that of the natural catchment, after a logging road was built across
the upper part of the catchment; it increased to five times after logging within 37 m of the
road; and increased to a peak of 18 times in the five months after logging of the rest of the
catchment. The amount of sediment lost was reduced a year after logging stopped.
The combined pattern of accelerated supply of sediment that reaches a peak after log-
ging and then tails off and logging roads operating both as water conduits and as a second
sediment source due to gully formation can be taken to represent the general sequence
of events that follow active deforestation. The logging tracks, cleared of vegetation and
stamped hard by heavy machinery, often function as run-off channels following intense
tropical rainstorms. Thus both the flood potential and sediment load of local streams
313 Deforestation, land use changes and rural migration

Fig. 17.3 Land use as main control in sediment production. Kenya. From Dunne, 1979. With permission from Elsevier

increase. Sediment accumulates in the channels and floodplains of forest streams and slugs
of sediment later travel down-channel following episodic rainfalls. Sediment from small
coastal streams may reach the sea in a few years. Coastal plumes of sediment in Southeast
Asia are now clearly visible even in low-resolution satellite imagery (Gupta and Krishnan,
1994). It is the same for other tropical regions. The cover of the book presents an excellent
illustration.
Post-logging sediment yield can be very high in the rain forest (103tkm−2yr−1) for the first
one or two years and then it exponentially tails off as secondary vegetation returns to the
bare slopes. The highest figures on record from deforestation are more than 7000 tonnes
from Sabah, Malaysia (Greer et al., 1994) and 17 000 t from Java, Indonesia (Ruslan and
Menan in Lal, 1987). It is difficult to choose a representative figure, but a number of more
than a thousand tonnes is probably a reasonable estimate. Lai et al. (1996), working in the
forests of Malaysia, suggested that the sediment yield from a logged catchment could be as
much as 50 times of the yield from a natural basin.

17.2.3╇ Land use changes and sediment yield

The effect of changes in land use on sediment yield was clearly shown by analysing sedi-
ment derived from 61 drainage basins in southern Kenya (Dunne, 1979). Although climate
and topographic variables are important factors, it is the land use that primarily controls
erosion and sediment yield in these basins (Fig. 17.3). A very large portion of the sedi-
ment is eroded during the wet period in the ambient seasonal climate and then transferred
along rural roads. The roads behave as conduits for water and sediment transfer, and in
consequence they become deeply eroded. This resembles the water and sediment transfer
following deforestation. Sediment yield primarily depends on vegetation cover and basin
steepness. Within a single type of land use, however, sediment yield indicates a direct
314 Anthropogenic alteration of geomorphic processes

relationship with runoff. The annual erosion rates in forested basins were 20–30 tkm−2. For
farmed basins this figure varies enormously. The extreme values given by Dunne (1979)
are about 10 tkm−2yr−1 for a flat basin on permeable lavas and low rainfall, and more than
4000 in wet and steep cultivated catchments. In sum, lack of vegetation cover, steep slopes
and considerable rainfall is the worst combination for erosion and sediment transfer in this
semi-arid landscape, and it is the removal of vegetation or insensitive farming that primar-
ily increases sediment yield.
It is generally accepted that sediment loss increases under agriculture, especially on
hillslopes. Reliable measurements, such as carried out by Dunne and his colleagues in
southern Kenya, however, are limited. For example, Douglas et al (1993) stated that the
only valid soil erosion data for Southeast Asia are measurements by Hatch (1981) on 40
m2 plots in Sarawak. It is, however, difficult to extend such data to a regional scale. Weera
and Kittipong (1987) computed figures for large areas in Thailand. Their work suggests
that sediment yield of up to several thousand tkm−2yr−1 may be expected from river basins
when the fields are open and no proper conservation measures are adopted. According
to Murtedza and Ti (1993), such figures may be reduced to hundreds of tonnes when the
entire river basin is considered, part of which could still be under forest. Agriculture there-
fore increases erosion and soil loss, but the numbers may vary widely depending on local
conditions (Table 17.1). Moving and resettling a large number of people, as has happened
in Malaysia, Indonesia and Brazil, also increases erosion and sediment load.
Shifting agriculture is a farming system common among the local settlers of the rain forest
who usually live in small communities. In shifting agriculture, a small tract of land is cleared
for farming. Such areas are abandoned after several years as the fertility of the land falls, and
new tracts are cleared for the next batch of farming. A return to the originally cleared area
should not happen for years until the soil recovers its fertility. Such a rotation with a reason-
able number of years also allows the soil to repair the damage from erosion. Shifting agricul-
ture has thus been described as a rotation of fields rather than a rotation of crops. Used within
its constraints, it is a sustainable system. As the area of the rain forest decreases, it becomes
difficult for the forest population to properly rotate the fields, and a quick return to old areas
or longer occupations leads to accelerated erosion of the surface material.

17.3╇ Temporal and seasonal patterns of sediment transport

Erosion of the regolith and associated sediment yield from a disturbed site, however,
does not continue at a uniform rate but changes over time (Douglas et al., 1992). Data on
the impact of deforestation from their research sites in Sabah show that sediment yield
increased to 18 times the pre-forest disturbance rate, but started to tail off in 2–3 years.
Secondary vegetation grows rapidly in the humid tropics, and after several years the yield
figures dropped to numbers representative of undisturbed areas (Greer et al., 1996). This is
a common pattern and unless the disturbance to the natural vegetation continues, second-
ary vegetation covers up the bare ground and the erosion rate and sediment yield decrease.
315 Spatial transfer of sediment

The adjustment of slopes and rivers to changing geomorphic processes thus continues over
time.
Temporal and seasonal changes in erosion and transfer of sediment were described for a
225 km2 area of northern Lao PDR that is drained by the Mekong River (Gupta and Chen,
2002). This is an area of steep-sided ridges and valleys with the ridge crests rising to over
1000 m, where the Mekong flows at an elevation of 250 m within 5 km of the ridge crests.
The annual rainfall is around 1700 mm. The rain comes mostly from the southwestern
monsoon, with 85–90 per cent falling between May and October. The steep slopes are gen-
erally under forest (primary or degraded) or shifting cultivation. Plots are cleared for shift-
ing cultivation at the beginning of the dry season in December to January and the biomass
on the ground is burnt later (April to May). The rural population is low (tens per km2) but
the cultivated slopes are very steep. Areas under vegetation or cleared were mapped repeat-
edly from multispectral SPOT satellite images, as such measures changed every month.
Multiplying the values of these areas with a measure of average sediment yield from the
literature (Table 17.1), allowed estimation of the sediment removed from the slopes on a
monthly basis. This is only a semi-quantitative estimate, but it does indicate (Table 17.2)
that (1) a seasonal pattern of erosion and sediment transfer and storage occurs each year,
and (2) the amount of sediment loss and transfer is directly related to rainfall.
This exercise indicates that erosion and sediment yield vary between years. For
example, significantly less sediment would be released in a dry El Niño year. Working
in the Philippines, White (1995) attributed erosion, short-distance sediment transfer and
its storage to frequent thundershowers; and efficient long-distance transfer of sediment to
tropical cyclones, which occur at much longer intervals. Sediment production from farm-
ing on steep slopes would therefore produce a fluctuating, but not necessarily cumulative,
sediment production. This trend will change with continuous cultivation, without proper
conservation measures.
The rivers of many parts of the tropics therefore repeatedly adjust to such fluctuations in
hydrology and sediment yield. The changing characteristics of drainage basins indicate that
such streams may no longer be entirely dependent on natural conditions.

17.4╇ Spatial transfer of sediment

Latrubesse and Stevaux described the connection between accelerated sedimentation in


the upper basin and channel form and behaviour downstream for the Araguaia Basin of
central Brazil, where the ‘intensive and indiscriminate expansion of agricultural activities’
(Latrubesse and Stevaux, 2002:€109) have degraded the natural environment. Gullies have
rapidly increased in size in the upper basin (Fig. 17.4). Large gullies (locally known as
voçorocas), hundreds of metres long and with headcuts that are more than 20 m wide,
developed in 30 years. These are formed by downcutting, piping and failure along pres-
sure-release joints in alluvial fills. Latrubesse and Stevaux referred to a study by Silva in
which 97 large and recently formed gullies were identified in about 3 km2. The sediment
produced by the easy erosion of thick soils and weathered material of the upper basin leads
316 Anthropogenic alteration of geomorphic processes

Table 17.2╇ Middle Mekong Basin:€seasonal erosion and sediment transfer


Period Slopes Tributaries The Mekong

Dec–Apr Increase in clear- Low flow; past Sediment stored


Dry, a few showers ance, very little sediment stored on beds, as bank
erosion of sedi- mainly on bed, as insets, and at
ment transfer due bank insets and confluences
to lack of rainfall at confluences with major
tributaries. The
river stage falls
gradually and
sediment transfer
diminishes
May Slopewash, gullies, Transfer of sedi- Increased river
Rains start, rising small debris ment pulses fol- velocity and
discharge flows; sedi- lowing showers, transport of
ment transfer to sediment arriving channel sedi-
channels from slopes ment, increased
competence
Approximately Jun–Jul Considerable Transfer of tribu- Erosion of
Early wet season erosion and sedi- tary channel bed mater-
ment transfer to sediment to the ial, insets and
channels by proc- Mekong; more bars; increased
esses described sediment arrives competence;
for May from slopes; bedrock erosion;
accumulation at localised critical
confluences and supercritical
flows; water tens
of metres deep in
the inner channel.
Aug–Nov Growth of vege- Continuous trans- Continued erosion
Late wet season tation; reduced fer of sedi- of bed mater-
erosion; very ment already ial, insets and
little sediment in channels, bars; increased
transfer accumulation at competence;
confluences bedrock erosion;
localised critical
and supercritical
flows; deep water
in the inner chan-
nel; sediment
accumulation at
confluences

Source:€Gupta and Chen, 2002


317 Spatial transfer of sediment

Fig. 17.4 Gulley eroding alluvial fill in the basin of the Upper Araguaia River, Brazil. Photograph: A. Gupta

to dynamic changes in the middle river with localised aggradation and bank erosion. The
extensive deforestation and land use changes in the upper basin have changed the fluvial
dynamics of the Araguaia.
Anthropogenic disturbances that caused the destruction of the original vegetation cover
allow intense tropical rain to fall on bare slopes mantled with thick soil and weathered
material. The usual sequence is a large number of small failures on hillslopes along with an
increase in the number and size of gullies, the accumulation of clastic sediment in stream
channels and floodplains, and an increase in the flood potential of rivers (Fig. 17.5). The
destruction of forests over a large area and on steep slopes gives rise to extensive slope
failures. Large landslides and debris flows occur, especially after a period of heavy rain-
fall (Chapter 6). A huge volume of sediment reaches the rivers at the foot of the slopes.
These rivers tend to flood more frequently because of the alteration in the water balance in
their drainage basins following widespread deforestation. This is followed by accelerated
growth of bars and floodplains. Mapping of the slopes near the Mekong River in Lao PDR
from 20-m resolution SPOT satellite images (Gupta et al., 2002) indicates vegetation clear-
ance in structure-guided tributary valleys leading to the pulsatory passage of sediment to
the Mekong as described in Table 17.2. Fresh young sand is seen in the Mekong as insets
(Fig. 17.6), around rock protrusions in the channel and as river-mouth fans where tributar-
ies join the main river. A very large volume of sediment is making its way to the coastal
waters, although a large part of it may remain stored for a considerable time in the alluvial
lowlands of the Lower Mekong in Cambodia and the delta in Viet Nam. Sediment plumes
have been measured off the mouth of the Mekong from satellite images, the seasonal vari-
ation in the size of the plumes being related to the varying discharge of this monsoon river
(Gupta et al., 2006). Such plumes occur along coasts where inland basins are undergoing
considerable changes in land use and at the mouths of major rivers in South and Southeast
Asia (Gupta and Krishnan, 1994). They are common along large sections of the coastline
318 Anthropogenic alteration of geomorphic processes

Fig. 17.5 Diagrammatic sketch of a river system with vegetation destruction in its upper basin

Fig. 17.6 Fresh sand in the Mekong River as inset near Luang Prabang, Lao PDR. From Gupta, 2007. By permission of Wiley

in many parts of the humid tropics. The IKONOS image on the front cover of this book
illustrates the discharge of such sediment-laden river water to the sea.
Anthropogenic alterations of the drainage basins therefore modify the hydrology and
sedimentation on slopes and along stream channels. In most cases, such alterations lead to
an episodic and floodprone pattern of discharge and increased storage and transfer of sedi-
ment. Impoundments across rivers, such as dams and reservoirs, modify drainage directly,
leading to a changed hydrology, reduced sediment transfer and channel modification.
319 Impoundments along rivers and their effects

17.5╇ Impoundments along rivers and their effects

Dams act as barriers to the down-channel passage of water and sediment, forcing rivers
to adjust their morphology and behaviour (Kondolf, 1997). Such barriers have been used
for thousands of years to control rivers. About 5000 years ago, King Menes of Egypt used
clay bunds to hold back the floodwaters of the Nile for basin irrigation when the annual
flood receded. Since then, the number of dams has increased significantly. The International
Commission on Large Dams (ICOLD) estimates that around 45 000 large dams exist in the
world, about 35 000 of which have been built since 1950. According to ICOLD, a large dam
is at least 17 m high from its foundation to the crest. Dams 5–17 m high and with a reservoir
capacity of more than 3 million m3 of water are also listed as large dams (World Commission
of Dams, 2000). Such a large number of dams cumulatively store a very large volume of
water and sediment in their reservoirs. A 1987 World Bank study indicated that about 50 km3
of sediment was being trapped annually in the reservoirs of the world (Mahmood, 1987).
Following the trend in Western countries, a very large number of dams, some huge in
size, were constructed across tropical rivers in the second half of the twentieth century.
The practice continues, although big dams are now hardly built in the developed world. As
many countries in the tropics achieved independence and concentrated on development,
dams and reservoirs were seen as instrumental in providing water for irrigation and power,
and also for flood control. They were not only engineering marvels but were also perceived
as essential symbols of national economic development. As a result, the number of dams in
the tropics proliferated, their sizes also increased.
Vörösmarty et al. (2003) reviewed the global impact of sediment retention from river
impoundments, mostly using data from 633 large reservoirs. They identify a large reser-
voir as one with at least 0.5 km3 maximum storage capacity. These reservoirs impound
runoff and trap sediment. Half of these reservoirs have a local sediment trapping capacity
of 80€per cent or more, and the total volume is rising. The trapping tends to be more effi-
cient in drier areas. Arid Asia, Australia/Oceania and Africa show a tendency towards high
trapping. The scale of trapping is less for South and Southeast Asia and parts of South and
Central America (Vörösmarty et al., 2003:€181, Fig. 5). The global sediment retention in
regulated basins was a modest 5 per cent of the moving sediment in 1950. This figure had
risen to 15 per cent by 1968 and to 30 per cent by 1985. Apparently, there has been a sta-
bilisation since then. This amounts to an estimate of 4–5 Gt annually. Vörösmarty et€al.
(2003) estimated that this has reduced the annual global sediment flux that reaches the
oceans to 10–16 Gt. If all the reservoirs, including smaller ones, are included, this esti-
mated total would be reduced even further. Impoundments on rivers therefore significantly
modify fluxes of water, sediment, carbon and nutrients on a global scale.
This volume of impeded sediment increases cumulatively each year, as does the num-
ber of large dams built or under construction, which includes huge examples, such as the
Three Gorges Dam over the Changjiang and the Xiaowan Dam over the Lancang Jiang
(Upper Mekong). Under normal circumstances all this sediment would have built bars and
floodplains downstream and ultimately reached the coast. Rivers downstream from dams
therefore need to adapt to a decrease in sediment and a controlled water release. This alters
320 Anthropogenic alteration of geomorphic processes

Fig. 17.7 Examples of sediment change downstream of dams. From Williams and Wolman, 1984. Courtesy of USGS

their form and function. Rivers in alluvium and rock tend to adjust differently. We need to
determine the adjustment of rivers downstream of dams.

17.5.1╇ Downstream effects of dams on alluvial rivers

Williams and Wolman (1984) discussed changes downstream from 21 dams constructed
on alluvial rivers in the United States. The listed changes are applicable to impounded
alluvial rivers globally. The study used very good records on discharge, sediment, channel
and floodplain morphology, and valley-floor vegetation. The records covered a number of
years. It was possible to trace changes from the pre-dam conditions to the post-dam envir-
onment, over time. The major findings of the study are listed below.

• Flood peaks generally decrease after the construction of dams, but the effect on other
post-dam discharge characteristics varies between rivers.
• Sediment concentrations and suspended loads decrease markedly for hundreds of kilo-
metres downstream from dams (Fig. 17.7).
• This results in bed degradation and a lowering of the longitudinal profile, although the
level of degradation varies. Most of the degradation occurs during the first decade or two
after the dam closure (Fig. 17.8).
321 Impoundments along rivers and their effects

Fig. 17.8 Degradation and long profile change downstream form dams. From Williams and Wolman, 1984. Courtesy of USGS

• Channel width can increase, decrease or remain the same downstream from the dam.
• Bed material may coarsen in the initial stage after dam closure, but this trend may change
over time.
• Riparian vegetation commonly increases downstream from dams, probably because of
decreased peak flow.
A river downstream of a dam therefore carries a more uniform and reduced discharge and
less sediment. This results in the deepening of the channel and erosion of its bed. If the
riparian vegetation builds up next to the river, then it binds fine material to build bigger
bars and floodplains, which are often common features below a dam in an alluvial river.
The down-dam channel pattern also may need to change to adjust to variations in discharge
and sediment. Williams and Wolman showed that rivers may require a distance of about
500 km to ameliorate these changes. Such amelioration may also happen over a shorter dis-
tance, depending on the scale of the river and the alteration due to impoundment.
322 Anthropogenic alteration of geomorphic processes

Fig. 17.9 Changes in flood discharge and variability due to Glen Canyon Dam on the Colorado River. From Webb et al., 1999.
Courtesy of AGU

17.5.2╇ Downstream effect of dams on rivers in rock

Very few detailed case studies exist on the downstream effect of dams on rivers in rock.
The case study on the Glen Canyon Dam on the Colorado River upstream of the Grand
Canyon is an excellent example. The pre-dam Colorado was a river with large summer
floods which moved a considerable amount of sediment in the channel. The post-Glen
Canyon Dam discharge has become seasonally uniform and floods no longer enter the
Grand Canyon. Flow regulation has also reduced discharge variability from year to year
(Fig. 17.9). This reduced the competence of the river downstream from the Glen Canyon
Dam, resulting in the aggradation of debris fans at the mouth of tributary streams in the
canyon and a changing channel morphology in places, particularly at rapids. Very little
323 Application of geomorphology towards a better environment

sediment is now carried by the Colorado, as most of the former high sediment load is
deposited in Lake Powell, the reservoir behind the Glen Canyon Dam. The texture of
the sediment is also proportionally much finer in the canyon. As a result, bed scour has
occurred and the bed of the Colorado has been armoured with coarse sediment. The river
also erodes sand bars which are no longer replenished, as the former sediment-carrying,
bar-building floods are absent. This has resulted in a degradation of ecological habitats,
changes in flora and fish species in the river, and the destruction of tourist campsites
(Webb et al., 1999). An experimental flood was sent down the canyon on 26 March 1996
by releasing water from the dam to see whether it had any beneficial effect. The experi-
ment has been repeated since then. Obviously damming rocky streams, as commonly
done in the mountains, degrades rivers.
Both sets of studies are from outside the tropics, but there is no reason to expect
dammed tropical rivers to behave differently. However, good studies of the effect of
dams on geomorphology of tropical rivers are rare. This is unfortunate, as a very large
number of dams, several of which are huge, have been constructed in tropical countries
as part of development projects. The Zambezi Basin, for example, carries a series of
dams on both the mainstem and the tributaries:€Kariba, Cahora Bassa, Itezhi-tezhi and
Kafue Gorge. Multipurpose dams and reservoirs are necessary to meet growing demands
for power and irrigation, but it is also necessary to determine the effect of such impound-
ments on the river and cumulatively on the coastal waters of the tropics. Dams can
influence river morphology and behaviour over a long distance. For example, China is
constructing a cascade of eight dams on the upper parts of the Mekong River that origi-
nates in China. It is suspected that the impact of the dams will mostly be on the lower
river, thousands of kilometres downstream (Gupta et al., 2006; Kummu and Varis, 2007;
Kummu et al., 2010).

17.6╇ Application of geomorphology towards a better environment

Only a fraction of the landforms of the world can claim to be natural, whether inside
or outside the tropics (Fig. 4.8). Tropical landscapes have been settled for a long time
and, since the middle of the twentieth century, the pressure of high population and the
need for developing the economy have strikingly modified the landforms and operating
geomorphic processes. Understanding the processes across hillslopes, along rivers and on
beaches therefore requires understanding the effect of anthropogenic modifications. This is
also necessary to manage rivers or ameliorate slope failures.
This chapter is a general introduction to the applied aspect of tropical geomorphology. In
the next two chapters we focus on specific problems:€geomorphology in urban areas, where
the intensive use of land creates a special type of environment with its associated hazards;
and the effect of climate change on tropical rivers. The rapid anthropogenic changes of the
tropical environment provide geomorphologists with a problem-solving capability which
unfortunately is seldom activated.
324 Anthropogenic alteration of geomorphic processes

Questions

1. Explain how a combination of anthropogenic alterations and fragile landforms lead to


high sediment yield. Give several examples where this is prevalent.
2. How far does the effect of anthropogenic alteration in the upper valley extend down a
long river?
3. What are the effects of deforestation on the slopes of the humid tropics?
4. How did Dunne demonstrate that land use is the primary control of sediment loss in
small river basins in southern Kenya? What type of climate prevails over this area?
5. There is a widespread belief that shifting agriculture leads to accelerated erosion and
high sediment yield. Is it just a belief or the truth?
6. What effect would the time-based variation in sediment yield have on local rivers fol-
lowing deforestation?
7. What effects do dams have on
(a) the longitudinal profile of a stream downstream from the dam?
(b)╇ changes in suspended and bed load downstream from the dam?
(c)╇ sediment that used to reach the coast?
8. How would you measure the rate of erosion of a place?
9. Can geomorphologists play an important role in managing the environment?
18 Urban geomorphology in the tropics

The mountain held the town as in a shadow.


Robert Frost

18.1╇ Introduction to urban geomorphology

Urban geomorphology evaluates the impact of urbanisation on landforms and the ambient
geomorphological processes of an area. It focuses on alterations to the physical environ-
ment caused by the spread of towns and cities. Urban geomorphology developed from a
number of case studies in the developed countries (mainly the United States) in the 1950s
and 1960s. Wolman (1967) structured a three-stage sequence of changes in an area under-
going urbanisation:

• Stage 1. Pre-Urban:€land under natural vegetation or agriculture; river channels adjusted


to existing conditions.
• Stage 2. Brief period of construction:€ land stripped bare; soil and weathered material
disturbed; extensive erosion; channels receiving large quantities of sediment; channels
in disequilibrium.
• Stage 3. Post-construction urban landscape:€impervious surfaces of streets, parking lots,
rooftops, etc; concrete drains and sewers replacing natural drainage to streams; insignifi-
cant erosion and sediment supply to channels; river channels still in disequilibrium but
over time may adjust to new conditions or persist in a state of disequilibrium.
Patterns of water and sediment supply from drainage basins change after urbanisation
(Wolman, 1967; Leopold, 1968; Dunne and Leopold, 1978). For a brief period during con-
struction, an impressive amount of sediment is moved from the land, building bars, dunes
and tributary-mouth fans in channels and constricting their capacity, thus making such
channels floodprone. The subsequent establishment of an impervious cover significantly
reduces infiltration and increases runoff. The degree of imperviousness may range from
near 20 per cent in low-density residential areas to about 90 per cent in the central business
district. The transfer of water to drainage channels is further accelerated by the construc-
tion of efficient storm sewers that take the water off the street and convey it speedily to
drainage channels The percentage of area in a city served by storm sewers may range from
0 to nearly 100 depending on the degree of urban development. This reduces the lag time
between the peak in rainfall and peak in channel flood (Fig. 18.1). Dunne and Leopold
(1978) suggested that on average a partly sewered basin has a lag time of about 20–25
325
326 Urban geomorphology in the tropics

Fig. 18.1 Shortening of lag time between rainfall and runoff due to urbanisation

per cent of the rural value. The increasingly impervious nature of the drainage basin also
reduces base flow, often dramatically.
Urban channels therefore undergo a set of modifications as listed in Table 18.1. If
left in their natural state after urbanisation, they flood more frequently and increase
their cross-sectional area, mostly by bank erosion. Usually, however, they are modified
by engineering, which involves straightening, widening and lining the channels. They
become more efficient for transferring floods downstream. Leopold (1968) has shown
that urbanisation increases the frequency of floods of a given size. Based on the records
of the US Geological Survey he listed several estimates of increase in discharge (Fig.
18.2) from a drainage area of one square mile (2.59 km2). For example, the mean annual
flood (defined as the flood with a recurrence interval of 2.33 years) increases to between
two and three times its former value in a basin with 50 per cent storm sewers and 50
per cent impervious cover, a situation often found in medium-density residential areas.
Urbanisation, however, does not seem to change the size of large meteorological floods.
A large volume of rainfall, as from a tropical storm, is transferred rapidly into surface
runoff, irrespective of any land use. It is the smaller floods that illustrate the effect of
urbanisation.
Such hydrological changes cause unlined urban channels to erode their beds and banks,
and expand their cross-sectional areas. Such channel erosion may produce as much sedi-
ment annually as derived cumulatively over five years from an undisturbed basin (Leopold,
1968). Various measurements indicate that channel enlargement could range between 50
and 300 per cent. Most urban channels left in a natural state probably tend to look like
Wolman’s description of an urban channel that reflects an erosive regime:€flood debris,
eroding banks, scour holes and exposed bars.
The effects of urbanisation are heightened in the humid tropics for several reasons:€the
presence of thick weathered regolith that is no longer protected by vegetation; the inten-
sity and high amount of tropical rainfall; and the ongoing progressive urbanisation of the
327 Introduction to urban geomorphology

Table 18.1╇ Geomorphological changes in a drainage basin following urbanisation


Type of change Description

Hydrological changes Changes in flow duration


modification of flow–duration curve:€lower base flows and
larger and quicker floods
Changes in flood frequency
increase in frequency of floods of a given size
increase in size of a given flood, e.g. mean annual flood
Modification of runoff from individual storms
decrease in lag time between rainfall and streamflow
increase in size of flood peaks
increase in peak velocity
Changes in water quality
Changes in hydrologic amenities
Sedimentological changes Deposition of relatively coarse-textured sediment in channels
in the early stages of urbanisation, followed by a reduction
in the amount of sediment reaching the channels
Morphological changes Changes in size, width and depth of channels
Changes in depositional forms
Changes in drainage density

Fig. 18.2 Changing flood–frequency curves for a hypothetical 1 square mile basin at various stages of urbanisation.
Note figures are in FPS system. From Leopold, 1968. Courtesy of USGS
328 Urban geomorphology in the tropics

tropical countries. We are familiar with the first two factors. The next section reviews the
case of rapid urbanisation in the tropics.
In general, urbanisation increases slope instability, accelerates flooding, increases sedi-
ment production and, in certain cities, results in ground subsidence. All such impacts are
discussed in this chapter. The problems are made worse if the city is located in a naturally
hazardous landscape (see section 18.3).

18.2╇ Urbanisation in developing countries

Urban settlements in the tropics started at least 5000 years ago, but a number of major
modern cities date back only to colonial times. In developing countries, a number of cities
started as the trading or military posts of a foreign power a few hundred years ago, and sub-
sequently evolved into large cities. Quite a few industrial and port cities are even younger,
having grown up in recent years, largely after independence from colonial powers.
The location of a number of major tropical cities can be justified in economic terms with
respect to their hinterland but not their actual sites, which often are fortuitous. Kolkata
started on the eastern bank of the Hooghly River at the location where the boatmen of
the British trader Job Charnock managed to moor their boat at the end of a rainy day in
1690. Nairobi grew up in the middle of a swampy plain where a tented depot of railway
stores was established in 1899 as the rail line was being laid across East Africa. Kingston
(Jamaica) developed when the residents of the nearby Port Royal left their homes following
the destructive earthquake of 1692. Once established, a number of these centres prospered
and grew into large cities with millions of inhabitants. Some are still growing rapidly.
The ongoing rapid rate of urbanisation in the tropics is also a modern phenomenon.
The urban population of the less-developed world, roughly synonymous with the tropics,
increased from 310 000 million in 1950 to 2.38 billion in 2007, and is projected to rise to
5.33 billion by 2050 (United Nations, 2008). The urban inhabitants, however, are unevenly
distributed between cities of different size. Just over half of the urban population in the
less-developed region live in small urban centres with fewer than 500â•›000 people. At the
other end of the scale, megacities (defined as cities with more than 10 million people) carry
about 10 per cent (United Nations, 2008). The number of megacities in the world was 19
in 2007, a number which is expected to rise to 27 in 2025. The number of megacities is
increasing in the tropics (Table 18.2), and such cities are expanding to cover a very large
area. This phenomenon is not confined to very large urban agglomerations. Urbanisation is
happening across the range and almost all cities are expanding.
Table 18.2 indicates that tropical cities are growing faster than cities outside the tropics.
Of the 17 megacities in 2000, 10 were in the tropics. The pattern probably also holds for
smaller cities. This implies an impressive expansion of built-up areas replacing rural or
natural landscapes. For example, Sâo Paulo covered an area of 180 km2 in 1930. By 1988
its area covered more than 900 km2 and its metropolitan region had spread over 8000 km2.
Usually, but not always, the urban periphery records the fastest growth, an area that is not
always suitable for city expansion. The urban fringe of Jakarta, on the north coast of Java,
329 Urbanisation in developing countries

Table 18.2╇ Ranked distribution of twenty largest urban agglomerations, 1950–2025


(Approximate population in millions within parentheses)
Rank 1950 1975 2000 2025 (projected)

1 New York area (12.3) Tokyo (26.6) Tokyo (34.5) Tokyo (36.4)
2 Tokyo (11.3) New York area (15.9) Mexico City Mumbai (26.4)
(18.0)
3 London (8.4) Mexico City (10.7) New York area Delhi (22.5)
(17.8)
4 Paris (6.5) Osaka-Kobe (9.8) Sâo Paulo Dhaka (22.0)
(17.1)
5 Shanghai (6.1) Sâo Paulo (9.6) Mumbai (16.1) Sâo Paulo
(21.4)
6 Moscow (5.4) Los Angeles area (8.9) Shanghai (13.2) Mexico City
(21.0)
7 Buenos Aires (5.1) Buenos Aires (8.7) Kolkata (13.1) New York area
(20.6)
8 Chicago (5.0) Paris (8.6) Delhi (12.4) Kolkata (20.6)
9 Kolkata (4.5) Kolkata (7.9) Buenos Aires Shanghai (19.4)
(11.8)
10 Beijing (4.3) Moscow (7.6) Los Angeles Karachi (19.1)
area (11.8)
11 Osaka-Kobe (4.1) Rio de Janeiro (7.6) Osaka-Kobe Kinshasa (16.8)
(11.2)
12 Los Angeles area (4.0) London (7.5) Rio de Janeiro Lagos (15.8)
(10.8)
13 Berlin (3.4) Shanghai (7.3) Cairo (10.5) Cairo (15.6)
14 Philadelphia (3.1) Chicago (7.2) Dhaka (10.3) Manila (14.8)
15 Rio de Janeiro (3.0) Mumbai (7.1) Karachi (10.0) Beijing (14.5)
16 St. Petersburg (2.9) Seoul (6.8) Moscow (10.0) Buenos Aires
(13.8)
17 Mexico City (2.9) Cairo (6.5) Manila (10.0) Los Angeles
area (13.7)
18 Mumbai (2.9) Beijing (6.0) Seoul (9.9) Rio de Janeiro
(13.4)
19 Detroit (2.8) Manila (5.0) Beijing (9.8) Jakarta (12.4)
20 Boston (2.6) Tianjin (4.9) Paris (9.7) Istanbul (12.1))

Note:€(1) Increase of population and (2) the rising number of tropical cities among the top twenty
over time. Source:€United Nations, 2008

is spreading at a pace of up to 18 per cent annually; a part of this expansion is replacing


coastal swamps, the other part is moving inland upslope beyond the upper boundary of allu-
vial fans into the hills. Development also alters previously unwanted degraded land inside
the city limits. For Sâo Paulo, this has meant steep vegetated slopes and wetlands next to
rivers (World Resources Institute, 1996). Cities located in close proximity to volcanoes
330 Urban geomorphology in the tropics

tend to climb their lower slopes, as in the case of Mexico City or Bandung (Java). All such
areal spreads come with associated geomorphic problems (Gupta, 1984). The alteration
of the landforms and processes due to urbanisation, as described earlier, is common in the
tropics. The data, however, tend to come from large cities, as these are the ones usually
studied.
We can group the problems of landforms and processes due to urbanisation into three
hazard classes:
1. Standard problems of alterations of slopes and channels due to urbanisation, as men-
tioned in section 18.1.
2. The problems are heightened in certain cases due to the location of the cities in natur-
ally hazardous areas, such as a tectonically active zone or where tropical cyclones are
common.
3. Hazards created or accentuated by accelerated resource utilisation and city metabolism,
e.g. excessive use of groundwater.
A city frequently demonstrates multiple hazards. We review three major cities to illustrate
such hazards, and evaluate whether these hazards could be at least partially ameliorated
using knowledge of local geomorphology. Details of these three cities are in Gupta and
Ahmad (1999). We will follow up these case studies with a discussion on the application of
geomorphology for improving the quality of urban life. The examples that follow indicate
different types of problem in connection with urban geomorphology and the practicality of
solving such problems. The case studies include:

• Singapore, where the geomorphological problems are small-scale flooding and slope fail-
ures, that are almost entirely anthropogenic, and which have been reasonably managed.
• Kingston (Jamaica), sited in an extremely hazardous location which exacerbates slope
failures and flooding, and where the management of geomorphological hazards is
extremely difficult.
• Bangkok, where an existing local problem has been magnified to a nearly unmanageable
level by city development, which led to the unrestricted use of groundwater.

18.3╇ Three examples of geomorphic hazards and


their amelioration

18.3.1╇ Urban geomorphology of Singapore

Singapore was founded in 1819 by Sir Stamford Raffles, on authorisation from Lord
Hastings, the Governor-General of India, as a commercial station in the Johore-Riau area
for serving British mercantile interest. The location is extremely favourable. Tectonically,
it is a stable area, away from any plate boundary disturbance. The relief is low, although
short steep slopes do occur on the island. The average annual rainfall ranges between 1650
and 2250 mm across Singapore. Usually the tidal range is less than 2 m, and the city is
331 Geomorphic hazards and their amelioration

Fig. 18.3 Diagrammatic sketch of land use change and its effect in a granitic hill area in the humid tropics, central Singapore.
From Gupta and Ahmad, 1999. With permission from Elsevier

protected by a small group of islands to the south, the large Indonesian island of Sumatra
to the southwest and the southern tip of the Malay Peninsula towards the north and north-
east. It is safe from tropical cyclones, being almost on the equator. Most rainstorms are
small (less than 10 km across), intense and of short-duration (usually not lasting for more
than an hour). According to the Intensity–Duration–Frequency study of the Singapore
Meteorological Service, about 90 mm of rain may fall in a 30-minute, 2-year rainstorm. The
runoff from such storms has the capacity to erode land that is unprotected by vegetation.
The city started on a coastal plain and the slopes of a nearby low hill (the present Canning
Hill) and gradually spread inland along small river valleys and over the rolling low hills of
Triassic sedimentary and granitic rocks. The natural regional landform was that of small,
steep, forested hills in sharp contact with valley flats, which comprise the floodplain of
a river channel in the middle. These floodplains used to act as wetlands, attenuating the
downstream passage of rainstorm-created floods. An impervious cover spread over the
hillslopes and arterial valleybottom roads were built along the floodplains (Fig. 18.3).
There was rapid expansion after the 1960s, when planned urban development was vigor-
ously pursued by the government. As Singapore expanded, arterial roads followed valley
flats with residential expansion climbing the hills on either side. The natural propensity for
flooding the valleyflats was thus heightened by urbanisation. By the 1980s, about half the
island was built-up area, with the percentage rising steeply towards the centre of the city.
332 Urban geomorphology in the tropics

Fig. 18.4 Bukit Timah Basin, Singapore illustrating intensity of urban land use that necessitates a large lined canal and drainage
diversions (not shown in the photograph). The cars indicate the scale of the canal which at this point has only about
6 km2 of catchment. Photograph: A. Gupta

Local newspapers reported 100 floods in less than 30 years, although most of these
occurred before 1985. Floods in Singapore tend to relate to urbanisation:€they are localised
and last for a short time (Gupta, 1992). The urbanised nature of Singapore floods has been
studied in the 6.4 km2 Upper Bukit Timah Basin. Here the densely built-up area increased
from less than 10 per cent of the basin area to over 45 per cent in 36 years (1950–1986).
By 1960, the stream had been engineered into a straightened earth canal. Between 1967
and 1972, this had to be replaced by a larger trapezoidal lined canal with a flood diversion
arrangement to the neighbouring basin to the west, as there were six to seven floods each
year that affected the arterial roads that ran on the former swampy floodplain. The calcu-
lated size of the flood with 5-year recurrence interval increased from 73.68 m3s−1 in 1950 to
91.65 in 1986 m3s−1 (Chuah, 1987). Progressive housing development on the steep slopes
draining into the canal required further modifications, as floods were becoming common
after rainstorms. A second diversion was built to the neighbouring basin to the east and part
of the 16 m trapezoidal canal was redesigned to a U-shaped one, 20–26 m wide.
The need to ameliorate the disturbance created by frequent urban floods transferred
natural drainage to lined channels. This was an ongoing process, but rapid urbanisation
increased the total lined channel length from 320 km in 1979 to 650 km in 1984, more than
doubling in five years at a cost of $200 million for channel improvement and extension.
Singapore now has about a kilometre of lined channel for each square kilometre of area.
Flood alleviation in Singapore depends on the capacity of drains keeping up with urbanisa-
tion of the upper valleys, and urbanisation is continuing (Fig. 18.4).
Given their catchment areas, the new lined channels of Singapore are very large, a
reflection on the anthropogenic flood propensity in the humid tropics. Finding space for
this enlargement is difficult in the older parts of cities and requires some ingenuity. For
333 Geomorphic hazards and their amelioration

example, this has been achieved in parts of Singapore by hiding the drainage canals under
decorative pedestrian malls and walkways. In the peripheral areas of new townships, large
lined canals are built from the very outset, and are designed to take runoff when the con-
struction of the new township is complete. Huge concrete canals, however, are not very
attractive, but the physical alteration of the landscape also provides the opportunity to
flank them with trees, paths and grassy stretches. Such a practice provides recreational and
nature corridors through the city and creates surrogate rivers. Amelioration of urban floods
ideally requires good practices of geomorphology, engineering and ecology.
The other principal geomorphological hazard in Singapore is the failure of engineered
slopes, including both rock and fill slopes. Slopes tend to fail after regrading, quarrying,
road building and residential development. Such alterations activate weathering processes
and also change the groundwater regime. The circulation and accumulation of subsurface
water, particularly following a rainstorm, is controlled by relict structures in the weathered
material (Grade V) including iron-crusted layers. Even weathered granite, which occurs
over about a third of Singapore, tends to show stress relief sheeting structures or openings
of pre-existing discontinuities when excavated. The slopes then fail along definite surfaces,
depending on both antecedent and trigger rainfalls (Pitts, 1992).
Pitts (1992) discussed slope failures in Singapore. The slides tend to be large in number,
small in size and repetitive in occurrence. The positioning of many of these slides is due
to the location of bedding, joints and cleavages on backscars, sliding surfaces and release
surfaces. A number of past slides are primarily due to engineering activities, such as align-
ing new roads parallel to the strike of sedimentary rocks, oversteepening of cut slopes and
failure to provide adequate subsurface drainage.
Both floods and slope failures are mainly dealt with in Singapore by post-event engin-
eering construction of lined canals, flood diversions, porous pavements and slope protec-
tion devices. The city is supported by efficient technical and managerial capacities, and
financial resources are available to tackle such hazards. The scale of the hazards is also
limited, and they occur only when the local geology and geomorphology are ignored.
Singapore is an example where the hazards are small-scale, the resources are available
and it is possible to ameliorate the problems to a large extent. The next example is differ-
ent. Kingston (Jamaica) is prone to hazards of different types and large magnitudes and its
resources are limited.

18.3.2╇ Kingston:€the hazard-prone urban landscape

Jamaica is located within a 200 km wide left-lateral strike-slip deformation zone of


Neogene age, marking the boundary between the North American and Caribbean Plates.
Neotectonics and bedrock control the geomorphology of this island. This is also an area of
high rainfall from tropical storms, which may reach hurricane force. The city of Kingston
is located in such an environment, where there is also a high population and widespread
poverty. It is one of the most environmentally hazardous cities in the world.
Kingston is a 200 km2 mosaic of coastal plains, gravelly palaeofans and steep hills.
All areas are hazardous. The rocks in the hills are highly faulted and deeply weathered.
The fault zone continues to be active. The epicentre of the 13 January 1993 earthquake
334 Urban geomorphology in the tropics

Fig. 18.5 Sketch map of the city of Kingston with steep hills surrounding the Liguanea Fan

was north of Jacks Hill, a northern suburb of Kingston. The maximum thickness of the
Liguanea gravel fan, over which most of Kingston is situated, is at least 259 m. The top
8 m of the fan appears to be an old debris flow deposit associated with palaeodrainage
of the Hope River (Fig. 18.5). Landslides and liquefaction-related ground failures occur
in Kingston during each major earthquake. Shepherd and Aspinall (1980) determined the
recurrence interval of earthquakes of MMI VII, VIII, IX and X as 38, 97, 137 and 237
years, respectively. Slopes also fail frequently in tropical storms (Ahmad, 1994). About
200–300 mm of rain in 24 hours initiates shallow slides on more than 25° slopes. Such
slopes cover about 80 per cent of the hilly areas of Kingston and this amount of rain is
expected to arrive once in 5 years. Flooding is also associated with transfer of a very high
amount of sediment with a high coarse fraction. The damage from hurricanes is greater but
less frequent. The most common types of slope failure are shallow slides and debris flows
on steep, weathering-limited hillslopes.
The natural vulnerability of Kingston is further increased by three anthropogenic acts:

• deforestation on slopes
• occupation of hazard-prone lowlands and steep hillslopes by the less affluent part of the
society; they are joined on the upper slopes by the rich community occupying plots with
commanding views
• reoccupation of slopes partially stabilised after a failure.
335 Geomorphic hazards and their amelioration

Such removal of vegetation and soil erosion raises the probability of numerous shallow
slides and debris flows in heavy rain (Maharaj, 1993), resulting in the destruction of prop-
erty and infrastructure. Furthermore, urbanisation of the Liguanea Fan has led to increased
overland flow and flash floods.
Geomorphological problems are difficult to resolve at this scale. The shortage of funds
and resources and the difficulty in moving people from particularly hazardous areas do
not help. The Office of Disaster Preparedness and Emergency Management of Jamaica is
based in Kingston. Both the government and the University of the West Indies are involved
in active research on geomorphological hazards and a database is growing. The implemen-
tation of the findings, however, is far more difficult than raising awareness. Kingston is not
the only city with this kind of problem. A densely populated city in a particularly hazardous
location is quite common.

18.3.3╇ Bangkok:€flooding and subsidence

Bangkok is located in the flat delta of the Chao Phraya, about 25 km inland. The ground
elevation is extremely low, varying between 0.5 and 2 m above sea level. The average
annual rainfall is 1500 mm. Drainage has always been a problem in the low swampy area,
especially during the wet southwestern monsoon that operates between late May and early
October.
The old city was located next to the river, partly on its levee. Over time the city expanded
east and south into the flat swampy lands of the delta. The expansion has been phenomenal
in recent years. The 1950 population of Bangkok was less than 1.5 million. The current
population is over 10 million.
The drainage is directed to the Chao Phraya via a series of canals, originally constructed
for both drainage and transportation. Heavy rain, a high river level, the flatness of the land
and high tides combine to create an environment that is difficult to drain during the wet
monsoon. Widespread inundation occurs regularly in Bangkok. Inundation during the rainy
season is a natural phenomenon, but Bangkok’s flood problem has been disastrously exac-
erbated by anthropogenic extraction of its groundwater.
The urban expansion of Bangkok resulted in a rapid increase in demand for water for both
domestic consumption and industrial use. In the east and southeast of the city, this demand
was met by pumping water out of sand beds that function as aquifers in the Quaternary thick
marine clay underneath the city. The quality of water was acceptable and as the normal
water rates were not levied, the use of groundwater was a favourable economic proposition
for new housing and industries such as breweries and paper manufactures. By 1985, about
1.3 million m3 of groundwater was being pumped out daily. The piezometric level in the
wells, which was originally at ground level, as expected in a deltaic plain, fell to about −9 m
by 1959 and locally to nearly −50 m by 1983. Saline water started to leak out of the marine
clays in between the sandy aquifer beds in the groundwater (ESCAP Secretariat, 1988).
Withdrawal of such a large amount of groundwater from sandy subsurface aquifers led
to subsidence, in places up to nearly 10 cm each year. The cumulative effect of this subsid-
ence created a bowl-shaped depression over eastern Bangkok, which has been estimated to
be 4550 km2 in area (Nutalaya et al., 1996). This is a shallow depression, but easily flooded
336 Urban geomorphology in the tropics

Fig. 18.6 Cumulative subsidence in the Bangkok area. From Noppadol and Nutalaya, 2005. By permission of Oxford
University Press

and difficult to drain. Several days of waterlogging are now common in Bangkok. In the
exceptionally wet year of 1983, when 1029 mm of rain fell in August and September, water
covered an extensive part of Bangkok for months. The subsidence progressively damages
the drainage systems, as drains and canals tend to sink with the ground. Subsidence also
structurally damages roads and buildings in Bangkok (Noppadol and Nutalaya, 2005).
Figure 18.6 demonstrates the level of cumulative subsidence in Bangkok. Subsidence has
reached the coast at the head of the Gulf of Thailand south of Bangkok, resulting in the
sinking of a part of the seashore, marine inundation and ecological destruction, especially
mangroves. Further lowering of the ground surface at Bangkok, which is only a short dis-
tance inland, and simultaneous global-warming driven sea-level rise may increase the level
of hazard disastrously.
The scale of the problem has accelerated research on the local Quaternary geology that
forms the subsurface material in Bangkok, and the surface of the city has been carefully
surveyed. Nature, causes and rates of Bangkok’s subsidence are now well known and the
solution obviously lies in stopping the subsidence and, if possible, raising the level of the
land. This, however, is very difficult. Groundwater pumping is now controlled and water
is no longer free. Perhaps the best option for flood control is a combination of careful land
337 General nature of urban geomorphological problems

zoning, control of the growth of Bangkok and its suburbs, and increased restriction on
groundwater withdrawal. Subsidence, however, cannot be reversed once the marine clays
below Bangkok have been dewatered.

18.4╇ The general nature of urban geomorphological problems

These examples illustrate the degradation of slopes and streams as a result of urbanisation
in the tropics. If the scale of degradation is limited, the problems can be resolved by careful
planning and management. However, in certain cases, the hazardous nature of the city’s
location may make it almost impossible.
A number of cities are located in such hazardous areas, with the hazards arising from
tectonic instability, volcanic eruptions or the expected passage of large storms at short
intervals. Certain cities suffer from more than one of these factors. Kingston is vulnerable
to both earthquakes and hurricanes. Mexico City is very close to volcanoes and parts of it
suffer from subsidence due to groundwater withdrawal. The nature of local geology may
cause problems, if there are vulnerable sites such as buried river channels, old landslide
deposits or unmapped karst landforms in the subsurface. All these cause problems in con-
struction and later the collapse of buildings, at times with loss of life as has happened in
Kuala Lumpur (Douglas, 2005a).
Problems also arise from excessive resource utilisation, such as the withdrawal of
groundwater stored in the subsurface. This, as has happened in Bangkok, results in sub-
sidence followed by increased inundation of the city and structural damage. This happens
regularly in certain types of environment, such as cities in swampy basins, coastal plains
and deltas. More than 70 per cent of the water required by Mexico City is drawn out of its
aquifer. The groundwater is lowered about a metre annually and in the 50 years between
1935 and 1985, the surface of the central area of the city, the location with the highest
demand of water has subsided 7 m (World Resources Institute, 1996).
Subsidence and related hazards due to groundwater extraction are quite common and
are experienced by the inhabitants of many cities. About 70 per cent of the population of
metropolitan Jakarta depends on groundwater, most of which is extracted from tens of
thousands of shallow wells and more than 3000 deep wells. The rapid increase of ground-
water extraction after 1970 caused a significant drop in the level of groundwater, leading
to saline water intrusion which particularly affected the older parts of the city near the
coast. Saline water intrusion in the deeper aquifer system has advanced between 3 and
7€km inland. The usual problem of subsidence has accompanied this deterioration in water
quality (Douglas, 2005b and references therein).
Kolkata developed in the abandoned delta of the Ganga on levees, floodplains and
swamps. As the city expanded, it extended into less desirable swampy ground, exacerbat-
ing the flood effect of the wet monsoon. The flooding is partly anthropogenic, caused by
three factors:€extension of the impervious landscape, disruption of the natural drainage and
the disappearance of the old swamps where floodwater used to accumulate. Furthermore,
the increased demand for water is partially met by pumping out groundwater from layers of
338 Urban geomorphology in the tropics

sand and gravel, primarily from a depth of 60 and 180 m, an aquifer sandwiched between
two clay beds. Large-scale groundwater extraction has caused the piezometric surface to
drop by metres, leading to ground subsidence in south-central Kolkata (Biswas and Saha,
1985). The subsidence is still not at a critical stage but should not be allowed to continue.
The problem of increasing subsidence associated with groundwater withdrawal has been
reported so far from a limited number of cities, but as the generic causes (a Quaternary
alluvial subsurface with sandy aquifer beds between impervious clay strata and large-scale
demand for water) are common, subsidence-related problems may become far more wide-
spread as urbanisation of the tropical countries continues.

18.5╇ Geomorphology and urban management

The case studies illustrate the progressive degradation of the physical environment of cit-
ies with urbanisation and also the potential for the application of geomorphology for urban
management. Such management requires (1) the acquisition of selective geomorphological
data and (2) the ability to interface with specialists who plan and manage cities:€ urban
planners, engineers and city managers. This is required globally, but the rapid urbanisa-
tion in tropics and subtropics has made it crucially important for developing countries.
Geomorphological information required for managing tropical cities includes:

• hard rock geology with special reference to neotectonics and rock structure
• Quaternary geology, especially for river valleys, coastal plains and deltas
• a descriptive account of landforms with special reference to steep slopes, breaks in slopes
and lack of slope
• depth of weathering and relict and precipitated structures in the regolith
• ambient operation processes with special reference to drainage, landslides and
subsidence
• nature of rainfall, especially as caused by storms
• prevalent water balance with details of storage and flux
• distribution of vegetation
• characteristics of resource use and extraction.
The availability and quality of such information varies among cities. Douglas (2005a), for
example, has reviewed the urban geomorphology of Kuala Lumpur to illustrate the type of
information that is required. In a number of cities there need to be plans for special hazards,
for example, Mexico City, Yogyakarta and Bandung, which are all located in the vicinity
of volcanoes.
In the tropics, certain cities have been well studied and even mapped for possible geo-
morphological hazards, but their number is small. This includes cities with possible exten-
sive urban flooding such as Singapore and Kuala Lumpur, where the intensity and amount
of tropical rainfall accentuate such floods. A number of cities on steeplands have also been
observed and mapped. Kingston, Kathmandu and Hong Kong are good examples. Cities
339 Geomorphology and urban management

Fig. 18.7 Hazardous location of Arequipa City and its surroundings close to the Misti stratovolcano. From Thouret et al., 1999.
(www.borntraeger-cramer.de)

located near convergent plate margins, where seismicity, slope failures and, in certain cases,
pyroclastic flows and lahars are hazards have also been studied to some extent. Thouret and
Lavigne (2005), for example, mapped the hazards and risks in the city of Yogyakarta from
the nearby volcano of Merapi in central Java. Thouret et al. (1999) identified the hazards
for the Peruvian city of Arequipa, located below the Misti stratovolcano (Fig. 18.7). Only
17 km separates the city centre from the crater of the volcano; the suburbs are even closer
and extend up the slopes. Arequipa is vulnerable to ash fall, pyroclastic flows, lahars and
debris avalanches. Cities tend to expand into hazardous zones, Nossin (2005) referred to
the expansion of the city of Bandung northwards to Lembang, building luxury settlement
estates which brings it very close to the volcano of Tangkuban Perahu.
Urban geomorphology should be studied in far more cities in order to provide an exten-
sive, in-depth and reliable database. Information from such databases could then be trans-
mitted to urban planners and city engineers in a format that they can usefully use. A map
is a useful format. Information supplied prior to the expansion of the city is useful for
sensitive planning, so that areas with potential hazards which degrade a number of cities
at present can be avoided. The information database is useful even for existing cities, as
340 Urban geomorphology in the tropics

(a)

(b)

Fig. 18.8 (a) Hazardous urban location on an alluvial fan at the mouth of a mountain stream in northern Venezuela and (b)
destruction of a city in the debris flows and large floods of 1999. Photograph:€M. C. Larsen, US Geological Survey

efficient management and correct engineering techniques could ameliorate some of the
geomorphic hazards.
The idea has been attempted in a few places. For example, Hong Kong has a well-
established geotechnical organisation for building such a database. UN-ESCAP, in its
role as a promoter of urban and environmental geology, published a series of volumes
called the Atlas of Urban Geology from its headquarters in Bangkok. Detailed studies and
maps of geology, landforms and hazards are therefore already in print for certain cities.
Such an approach is required, especially for tropical cities located in hazardous environ-
ments. Given the location of a number of tropical cities, an escape from calamity may be
impossible. The destruction of townships such as Carmen de Uria, located on an alluvial
fan in northern Venezuela in the rainstorms of December 1999, is a relevant example.
Landslides (mainly debris flows) and floods from nearly 1200 mm of rain in the first 16
days of the month inundated coastal communities, killed a number of people (estimated
between 15€ 000 and 30 000), and severely damaged buildings and infrastructure (Fig.
18.8). The settlements in this area are commonly on alluvial fans, the only relatively flat
terrain available between the coast and the steep mountains. The fans, as expected, are
areas of high geomorphic activity (Larsen et al., 2001a, 2001b). Such vulnerability exists
for a number of cities, as discussed by Chardon (1999) for Manizales, which is at a haz-
ardous location in the Colombian Andes. Even if technical capabilities exist, unsuccessful
341 Questions

communication with urban managers and planners leaves such capabilities unutilised, as
evidenced in the 1985 Nevado del Ruiz catastrophe (Voight, 1996).

Questions

1. What is urban geomorphology? Why is urban geomorphology becoming progressively


more important in the tropics?
2. Are cities in the tropics optimally located from the viewpoint of a geomorphologist?
3. What conclusions can we reach from Table 18.2?
4. Floods in Singapore tend to be localised and short-term. Why?
5. It has been said that in Singapore flood alleviation is based on the capacity of drainage
canals to keep up with urbanisation. Is this a correct summary statement?
6. Describe the problems of a tropical city of your choice which is located in a hazard-
prone environment.
7. Describe the common scenario for cities using large volumes of groundwater.
8. What kind of geomorphological information is required for managing tropical cities?
How could such information be transmitted to city managers and planners?
19 The future with climate change

The future ain’t what it used to be.


attributed to Yogi Berra, New York Yankees

19.1╇ Climate change and the future

We did not experience any significant climate change over the last several thousand years.
This allows us to extrapolate the present processes and rates over a length of time and
assume that we can statistically stretch any time-based record, such as river discharge,
if need be. We therefore are able to compute the size of 50-year floods and other useful
measures.
It seems that we can no longer make such an assumption (Milly et al., 2008). The ris-
ing global temperature is changing climatic components and forcing a rise in sea level.
The nature and dimension of such changes are not likely to be uniform but vary across the
planet (IPCC, 2007). The general pattern of environmental changes due to global warm-
ing has started to emerge, but we are still not in a position to predict with confidence the
specific effect of global warming on small areas such as a 100 km2 drainage basin or the
southern slope of an isolated plateau.
A generalised list of global warming-related changes in the tropics is set out in the next
section. Our knowledge of this area is still rather limited and it is preferable to discuss only
the robust and large-scale changes. We still do not have enough information to model the
future appearance and behaviour of small-scale individual geomorphological features like
a hillslope due to climate change. It is also difficult to work out the changes for a large
geomorphic system. For example, a large river like the Ganga or the Amazon would have
to adjust to the shrinking of glaciers at the source, the changing pattern of rainfall and
storminess over most of the basin, and a rising sea level at the delta. All such changes
may occur near-simultaneously. Such an evaluation is a difficult task, but even a qualita-
tive understanding of the possible state of a river or a hillslope or a delta face would be
extremely useful. A basic knowledge of geomorphology is essential for this purpose, and
as mentioned in Chapter 1, the geomorphologist may have the opportunity to fulfil the role
of a prophet.

342
343 A robust prediction of the effects of climate change in the tropics

19.2╇ A robust prediction of the effects of climate


change in the tropics

We can anticipate certain geomorphological changes as a result of global warming. In this


section we describe the changes in climate which appear to be near-certain. Subsequently,
we examine the possible effect of such changes on landforms and geomorphic processes in
the tropics. These are the modifications that the tropics would go through in the near future.
We restrict the analysis to robust changes qualitatively and do not explore the finding of
various models which are currently at the frontier of research and still being finalised. The
effect of the changes in climatic characteristics discussed here are summarised from IPCC
(2007).

19.2.1╇ Nature of climate change

Concentrations of carbon dioxide, methane and nitrous oxides (collectively known as the
greenhouse gases) in the global atmosphere have increased significantly since 1750 as a
result of human activities. The concentration of CO2 has increased from 280 ppm dur-
ing pre-industrial days to 379 ppm in 2005. Concentration of all greenhouse gases has
increased, resulting in a warmer atmosphere. As a result a number of long-term changes in
climate have been observed at continental, regional, and ocean-basin scales. These include
changes in arctic temperature and ice, precipitation amounts, wind patterns, and aspects of
extreme weather (drought, high precipitation, and tropical cyclones). Such changes are a
departure from the normal trend. For example, atmospheric warmth recorded in the second
half of the twentieth century is higher than what prevailed for the last 1300 years at least.
The last time temperature rose so high for an extended period was about 125 000 years ago
in an interglacial time. A number of global projections expect a warming of about 0.2°C
per decade for the next twenty years, an estimate that may need to be revised upward if
greenhouse gas emissions continue at or above the current rate. It is therefore very likely
that high temperature and heavy precipitation events would become more frequent.
Model-based projections of future rainfall are more reliable for the higher latitudes than
the tropics, but certain characteristics may be noted. Summer monsoon rainfall over South
and Southeast Asia is generally expected to increase along with increased rainfall over
East Africa. Model projections indicate an increase in precipitation and variability in the
Asian monsoon and the southern part of the West African monsoon. The pattern over the
drier regions in Africa or the basin of the Amazon is not clear at this stage, although a
decrease in summer precipitation has been projected for the Sahel. A similar decrease in
summer precipitation is expected also for Mexico and Central America, and an increase for
monsoon Australia. Meteorological observations already show that droughts have become
more common over the tropics and subtropics since the 1970s.
Changes in the radiative flux at the surface of the Earth alter heat and moisture budgets,
and thus the hydrologic cycle. As the sea surface temperature (SST) of the tropical seas and
oceans increases, tropical cyclones of the future would become stronger, with larger peak
344 The future with climate change

wind speeds and heavier precipitation. For example, post-1970 observations in the North
Atlantic indicate an increase in intensity of tropical cyclonic activities. Since about 1970,
the number and proportion of hurricanes that reached categories 4 and 5 have increased
globally, although the total number of tropical cyclones has decreased. Such changes are
expected to continue even if we manage to stabilise greenhouse gas emissions, as once
started this type of global warming and rise in sea level may persist for centuries, given the
timescale at which climatic processes and feedbacks operate.

19.2.2╇ Warming of oceans

Since 1955, the oceans have warmed, accounting for about 80 per cent of the energy flux of
the Earth’s climate, and such a warming is widespread over the upper 700 m of water in the
oceans. The sea level is rising. For example, the global rate of sea level rise as measured for
the 1993–2003 period by sensors carried by TOPEX/Poseidon satellite is 3.1 ± 0.7 mm yr−1.
This figure is close to the estimated total of 2.8 ± 0.7 mm yr−1, which is reached by summing
the thermal expansion of ocean water and the contribution from the melting of land ice by
global warming. Satellite measurements, which tend to be precise, further indicate that the
amount of sea-level change tends to vary from region to region. For example, the rise in sea
level tends to be higher for the Southeast Asian and East Asian seas and, to a smaller extent,
for the tropical Atlantic. Sea level may continue to rise for several decades. Greater melt-
ing of glacial ice from Greenland and Antarctica can raise the sea level much higher. About
125 000 years ago, during the last interglacial period, the mean global sea level was 4–6 m
higher. The amount of ice on Earth is currently decreasing from most mountain glaciers and
the Greenland Ice Sheet. Ice thinning and breaking up also occur in Antarctica.

19.2.3╇ Identifying the changes

If we select only the robust changes we come up with the following list:

• retreat of mountain glaciers


• melting of snow and ice earlier in the year
• changes in annual rainfall and regional variations
• enhanced seasonality
• regional increase in droughts
• increase in strength of large storms (for example, hurricanes of category 4 and 5)
• higher rainfall from extreme events
• a rising sea level, not globally uniform but with regional variations.
These changes that rise out of global warming and are expected to continue at least for
decades will affect geomorphic processes in the tropics. We should expect different rates
of erosion and sedimentation, and the modification of landforms. We may explore such
alterations based on the basic concept of geomorphology and analogues arising out of nat-
ural changes during the Quaternary. Such Quaternary modifications have been explored
for large rivers in general by Blum (2007) and specifically for the Ganga by Goodbred
(2003).
345 Geomorphological adjustments in the tropics

19.3╇ Geomorphological adjustments in the tropics


from climate and sea-level changes

The changes listed in the previous section will affect landforms and geomorphic processes
in the tropics. It is difficult to determine such effects precisely, due to the lack of reliable
quantitative measurements, but we can work out the robust changes based on our general
knowledge of fluvial geomorphology and analogues from the past, especially from the
Quaternary. We can use the reconstruction of past river systems (Goodbred, 2003; Blum,
2007). We will construct the changes in a hypothetical large river basin, to illustrate pos-
sible future scenarios. Given the scale of our current knowledge, it is best to keep this ana-
lysis large-scale and qualitative.
Major river systems integrate signals from large areas that stretch from the mountains of
the upper basin through a long mixed bedrock–alluvial valley to sea-facing deltas. Climate
change forces alterations in the discharge and sediment supply to the river, causing it to
go through a series of morphological and sedimentological adjustments. As a result, more
sediment is either stored in the valley or removed downstream. Large river basins are
polyzonal, implying different types and rates of processes in operation in different tributary
basins. Thus, different parts of the basin of a large river will adjust differently to climate
change, complicating the picture.
Rivers are conduits that transfer water and sediment from the land to the ocean. The
nature of the river channels depends primarily on four variables:€ water, sediment, tex-
ture of sediment and channel slope (Chapter 7). One or all of these four prime variables
may be altered by climate change, causing rivers to adjust to new conditions. The type
of change would probably be similar for all rivers, but the degree of change will also be
determined by the specific physical environment of the basin. This limits the level of con-
fident prediction.
Let us assume we are dealing with a hypothetical long river that originates in a trop-
ical high mountain like the Himalaya and flows for thousands of kilometres in a mixed
bedrock–alluvial channel through a wide valley to the sea via a large delta. This region is
undergoing the appropriate robust climate changes as listed in section 19.2.3. We would
expect glaciers to be on the retreat in the mountains and an earlier melting of ice and snow
in the spring. Normally about 80–90 per cent of the sediment of a major river is derived
from such mountains. The supply of the sediment would be enhanced by increased slope
failures in the upper valleys due to the melting and retreat of glaciers. Episodic high rain-
fall events would be common if the mountains stretch across the tracks of major storms.
Thus, early into climate change huge volumes of sediment may be produced, and stored
temporarily in bedrock valleys in the mountains or as alluvial fans at the highland–lowland
contact. Such sediment would be gradually transferred and stored along the mixed bed-
rock–alluvial valley and floodplain of the river.
The sediment transfer is likely to be episodic, due to the enhanced seasonality in dis-
charge and high rainfall from an increased number of large storms. The seasonality and
large storms would also modify the annual discharge hydrograph of the river with a longer
346 The future with climate change

Fig. 19.1 Probable geomorphic changes in the basin of a hypothetical large river due to climate change

period of low flow and increased peakedness. We may also expect overbank flooding and
transfer and storage of sediment in both cross-valley and downstream directions. This type
of behaviour would make the river wider, shallower and floodprone. The river would have
a high rate of sediment transfer in pulses and a long-term storage of sediment in between,
if accommodation space for it is available in the valley.
The lower course of this hypothetical river and its delta may need to adjust to both the
changing pattern of water and sediment arriving from upstream and a rising sea level. A
rising sea level would allow more accommodation space for sediment; increase coastal ero-
sion and saline intrusion; and allow more effective inland penetration by enhanced tropical
cyclones. The expected morphological changes in the delta would include:€levee breaks
and crevasse splays; sedimentation in flood basins and filling of wetlands; and avulsions
and the abandonment of distributary channels in low flat deltas (Fig. 19.1).
The river is a source-to-sink system and will need to adjust to all these modifications
resulting from climate change. The future of this hypothetical river has been derived from
material discussed earlier in the book. This is a simplified illustration of the application
of geomorphology in an uncertain world. Similar scenarios could be constructed for other
geomorphological units such as alluvial fans, coastal plains or steep slopes.

19.4╇ The noise effect of anthropogenic changes

A large part of the Earth’s surface is no longer natural. Anthropogenic alterations have
modified the prevalent geomorphic processes and, in certain cases, even landforms,
as detailed in Chapters 17 and 18. Such alterations are extensive and their cumula-
tive effect could be enormous. It has been estimated that reservoirs behind dams are
capable of holding back 15 per cent of total annual river discharge (Nilsson et al.,
347 Tropical geomorphology in the near future

2005). In the world, there are 45 000 dams more than 15 m high or with a large reser-
voir (World Commission on Dams, 2000). The number of smaller reservoirs is much
bigger. Vörösmarty et al. (2003) have computed with some reservation that annually
4–5 gigatonnes of sediment is deposited in the reservoirs behind such impoundments.
This is between 20 and 33 per cent of global natural flux that would reach the sea in an
unmodified world. In Chapter 17, we discussed increased erosion and the accelerated
production of sediment due to various development projects. Even such an increased
amount of sediment is arrested locally by dams and does not travel to the coast. Syvitski
et al. (2005) concluded that due to such impoundments, sediment delivery to the seas
has fallen globally by 10 per cent from the natural rate, and by much more for selected
regions.
Geomorphic systems that are relatively untouched anthropogenically are expected to
respond significantly to climate change. They have done so during the Quaternary and are
expected to do so in the near future (Blum, 2007). However, given the scale of anthropo-
genic modifications, the effect of anthropogenic changes is probably larger in certain sys-
tems than the impact expected from climate change. Under such circumstances, the signal
of climate change could be masked, at least partially, by the noise of anthropogenic modi-
fications. Prediction of the health and efficiency of geomorphic systems in the future is
therefore going to be a complicated challenge. The tropics, and indeed the world, are likely
to be maintained by natural geomorphic systems that are attempting to respond to climate
change while also being affected by anthropogenic alterations. Any prediction under such
circumstances is difficult.

19.5╇ Tropical geomorphology in the near future

Geomorphology in the near future therefore will reflect three factors:€the normal pattern
of landforms and processes, their anthropogenic modification, and the effect of climate
change. All these factors carry their tropical components. We have discussed in detail the
natural tropical environment for most of the book (Chapters 1–16). Anthropogenic modifi-
cations are currently extensive and significant in the tropics (Chapters 17–18). As most of
these modifications are essential to development, their effects are increasing progressively
in tropical countries. Certain aspects of climate change are also likely to be significant in
the tropics. For example, rainfall from large tropical cyclones is expected to increase in
the higher humid tropics. Certain areas of the arid tropics are expected to become drier.
Rainfall zones may shift (section 19.3).
The combined effect of these factors for a particular location or for a large geomorphic
system will be difficult to determine. Given the state of our knowledge, numerical predic-
tions such as the level of increase in rainfall or decrease in river discharge at a location are
still not very reliable. We can, however, predict some of the robust changes on a qualitative
basis, based on our knowledge of geomorphology. Even that could be a very important
component of tropical geomorphology in the future.
348 The future with climate change

Questions

1. What aspects of climate change due to global warming are geomorphologically


crucial?
2. How would a hypothetical long river, rising in a high tectonic mountain and flowing in
a mixed alluvial-bedrock valley to the sea, adjust to climate change?
3. What is more important in tropical geomorphology:€climate change or anthropogenic
alterations?
References

Aalto, R., Dunne, T. and Guyot, J. L. (2006). Geomorphic controls on Andean denudation
rates, Journal of Geology, 114:€85–99.
Aalto, R., Maurice-Bourgoin, L., Dunne, T., et al. (2003). Episodic sediment accumula-
tion on Amazonian floodplains influenced by El Niño/Southern Oscillation, Nature,
425:€493–497.
Abrahams, A. (1984). Channel networks, a geomorphological perspective, Water Resources
Research, 20:€161–188.
Ahmad, R. (1994). Landslides in Jamaica:€extent, significance and geological zonation, in
Environment and Development in the Caribbean:€Geographical perspectives. D. Barker
and D. J. F. McGregor (eds.). Mona, Jamaica:€University of the West Indies Press.
Ahmad, R., Scatena, F. N. and Gupta, A. (1993). Morphology and sedimentation in
Caribbean montane streams:€ examples from Jamaica and Puerto Rico, Sedimentary
Geology, 85:€157–169.
Ahuja, P. R. and Majumdar, K. C. (1959). A sampling approach to the estimation of floods.
Indian Journal of Power and River Valley Development, 9:€23–49.
Aleva, G. J. J., Bon, E. H., Nossin, J. J. et al. (1973). A contribution to the geology of part
of the Indonesian Tinbelt:€the sea areas between Singkep and Bangka islands and around
the Karimata Islands, Bulletin of the Geological Society of Malaysia, 6:€257–272.
Allison, I., Peterson, J. A. and Chinn, T. J. H. (1989). Glaciers of Irian Jaya, Indonesia, and
New Zealand. U.S. Geological Survey Professional Paper 1386-H, Washington, DC.
An, Z., Kutzbach, J. E., Prell, W. L. et al. (2001). Evolution of Asian monsoons and phased
uplift of the Himalaya-Tibetan Plateau since Late Miocene times, Nature, 411:€62–66.
Anderson, J. M. and Spencer, T. (1991). Carbon, Nutrient and Water Balances of Tropical
Rain Forest Ecosystems Subject to Disturbance:€Management Implications and Research
Proposals. MAB Digest 7, Paris:€UNESCO.
Bagnold, R. A. (1941). The Physics of Blown Sand and Desert Dunes. London:€Chapman
and Hall.
Bagnold, R. A. (1966). An approach to the sediment transport problem from general phys-
ics. U.S. Geological Survey Professional Paper 722-I, Washington, DC.
Baker, V. R. (1978). Adjustments of fluvial systems to climate and source terrain in trop-
ical and subtropical environments, Memoir, Canadian Society of Petroleum Geology,
5:€211–230.
Baker, V. R. (1981). Catastrophic Flooding:€ The origin of the channeled scabland.
Stroudsburgh, PA, Hutchinson Ross.
Baker, V. R. (2007). Greatest floods and largest rivers, in Large Rivers:€Geomorphology
and management, A.Gupta (ed.). Chichester:€Wiley, 65–74.
349
350 References

Baker, V. R. and Costa, J. E. (1987). Flood power, in Catastrophic Flooding. L. Mayer and
D. Nash (eds.). London:€Allen and Unwin, 1–21.
Baker, V. R., Kochel, R. C., Patton, P. C. et al. (1983). Palaeohydrologic �analysis of
Holocene flood slack-water sediments, Special Publication of International Association
of Sedimentologists, 6:€229–239.
Bard, E., Jouannic, C., Hamelin, B., et al. (1996). Pleistocene sea levels and tectonic uplift
based on dating corals from Sumba Island, Indonesia, Geophysical Research Letters,
12:€1473–1476.
Barnes H. H., Jr. (1987). Roughness characteristics of natural channels. U.S. Geological
Survey Water-Supply Paper 1849, Washington, DC.
Barua, D. K. (1990). Suspended sediment movement in the estuary of the Ganges–
Brahmaputra–Meghna river system, Marine Geology, 91:€243–254.
Batchelor, B. C. (1979). Discontinuously rising late Cainozoic eustatic sea-levels, with
special reference to Sundaland, Southeast Asia, Geologie Mijnbouw, 58:€1–20.
Beck, J. W., Récy, J., Taylor, F., et al. (1997). Abrupt changes in early Holocene tropical sea
surface temperature derived from coral records, Nature, 385:€705–707.
Benbow, M. C. (1990). Tertiary coastal dunes of the Eucla Basin, Australia, Geomorphology,
3:€9–29.
Bender, M. L., Fairbanks, R. G., Taylor, F. W., et al. (1979). Uranium-series dating of
the Pleistocene reef tracts of Barbados, West Indies, Bulletin, Geological Society of
America, 90:€577–594.
Benn, D. I. and Evans, D. J. A. (2010). Glaciers and Glaciation. Oxford:€Oxford University
Press.
Bennett, J. D. (1984). Review of Superficial Deposits and Weathering in Hong Kong, GCO
Publication No. 4/84, Hong Kong, Geotechnical Control Office, Engineering Development
Department.
Bennett, M. R. and Glasser, N. F. (1996). Glacial Geology:€ Ice sheets and landforms.
Chichester:€Wiley.
Berner, E. K. and Berner, R. A. (1996). Global Environment:€Water, air and geochemical
cycles. Upper Saddle River, New Jersey:€Prentice Hall.
Best, J. L., Ashworth, P. L., Sarker, M. H. et al. (2007). The Brahmaputra–Jamuna River,
Bangladesh, in Large Rivers:€ Geomorphology and management, A. Gupta (ed.).
Chichester:€Wiley, 395–433.
Billi, P. and el Badri Ali, O. (2010) Sediment transport of the Blue Nile at Khartoum,
Quaternary International, doi:10.1016/j.quaint.2009.11.041.
Biswas, A. B. and Saha, A. K. (1985). Environmental hazards of the recession of piezo-
metric surface of groundwater under Calcutta, Proceedings, Indian National Science
Academy, 51A:€610–621.
Biswas, B. (1973). Quaternary changes in sea-level in the South China Sea, Bulletin,
Geological Society of Malaysia, 6:€229–256.
Blong, R. (1984). Volcanic Hazards:€ A sourcebook on the effects of eruptions. San
Diego:€Academic Press.
Blum, M. D. (2007). Large river systems and climate change, in Large Rivers:€Geomorphology
and management, A. Gupta (ed.). Chichester:€Wiley, 627–659.
351 References

Blum, M. D. and Törnqvist, T. E. (2000). Fluvial response to climate and sea-level change:€a
review and look forward, Sedimentology, 47 (supplement):€1–48.
Bourke, M. C. (1994). Cyclical construction and destruction of flood-dominated flood
plains in semiarid Australia, in Variability in Stream Erosion and Sediment Transport,
L.€J. Olive, R. J. Loughran and J. A. Kesby (eds.). Wallingford:€International Association
of Hydrological Sciences, Publication 224, 113–123.
Bourke, M. C. and Pickup, G. (1999). Fluvial form variability in arid central Australia, in
Varieties of Fluvial Form, A. J. Miller and A. Gupta (eds.). Chichester:€Wiley, 249–271.
Bowler, J. M. (1981). Australian salt lakes:€a palaeohydrological approach, Hydrobiologia,
82:€431–444.
Bradley, R. S. (1985). Quaternary Paleoclimatology. London:€Allen and Unwin.
Bradley, W. C. (1958). Submarine abrasions and wave-cut platforms, Bulletin, Geological
Society of America, 69:€967–974.
Brandt, J. (1988). The transformation of rainfall energy by a tropical forest canopy in rela-
tion to soil erosion, Journal of Biogeography, 15:€41–8.
Brice, J. C. (1983). Planform properties of meandering river, in River Meandering, C. M.
Elliot (ed.). New York:€American Society of Civil Engineers, pp. 1–15.
Bronger, A. and Bruhn, N. (1989). Relict and recent features in tropical alfisols from South
India, Catena, Supplement 16, 107–128.
Brook D. B., Eavis, A. J. and Lyon, M. K. (1982). Caves of the limestone:€Gunung Mulu
National Park, Sarawak, Sarawak Museum Journal, 51(1):€95–120.
Brookfield, M. E. (1998). The evolution of the great river systems of southern Asia, during the
Cenozoic India–Asia collision:€rivers draining southwards, Geomorphology, 22:€285–312.
Bruijnzeel, L. A. (1989). Nutrient cycling in moist tropical forests:€the hydrological frame-
work, in Mineral Nutrients in Tropical Forest and Savanna Ecosystems, J. Proctor (ed.).
Oxford:€Blackwell, 383–415.
Bruijnzheel. L. A. and Bremmer, C. N. (1989). Highland–Lowland Interactions
in the Â�Ganges–Brahmaputra River Basin:€ A review of the published literature.
Kathmandu:€ICIMOD Occasional paper 11.
Brunsden, D., Jones, D. K. C., Martin, R. P. et al. (1981). The geomorphological char-
acter of part of the Low Himalaya of Eastern Nepal, Zeitschrift für Geomorphologie,
37:€25–72.
Bryan. K. (1922). Erosion and sedimentation in the Papago County, Arizona with a sketch
of the geology, U.S. Geological Survey Bulletin, 730-B:€19–90.
Buchanan, F. (1807). A journey from Madras through the countries of Mysore, Canara and
Malabar. London:€East India Company.
Büdel, J. (1982). Climatic Geomorphology. Translated by L. Fischer and D. Busche,
Princeton: Princeton University Press.
Bull, W. B. (1964). Alluvial fan deposits in western Fresno County, California. U.S.
Geological Survey Professional Paper 437-A.
Burbank, D. W., Leland, J., Fielding, E., et al. (1996). Bedrock incision, rock uplift and
threshold hillslopes in the northwestern Himalaya, Nature, 379:€505–513.
Carson, M. A. and Kirkby, M. J. (1972). Hillslope Form and Process. London:€Cambridge
University Press.
352 References

Carter, R. M. and Johnson, D. P. (1986). Sea-level controls on the post-glacial development


of the Great Barrier Reef, Queensland, Marine Geology, 71:€137–164.
Carter, R. W. G. (1988). Coastal Environments:€An introduction to the physical, ecological,
and cultural systems of coastlines. London:€Academic Press.
Cas, R. A. F. and Wright, J. V. (1987). Volcanic Successions:€ Modern and ancient.
London:€Unwin Hyman.
Castelltort, S. and Van Den Driessche, J. (2003). How plausible are high-frequency sediment-
supply driven cycles in the stratigraphic record? Sedimentary Geology, 167:€3–13.
Chakraborty, T. and Ghosh P. (2009). The geomorphology and sedimentology of the
Tista Megafan, Darjeeling Himalaya:€ Implications for megafan building processes,
Geomorphology, 115:€252–266.
Chalcraft, D. and Pye, K. (1984). Humid tropical weathering of quartzite in southeastern
Venezuela, Zeitschrift für Geomorphologie, 28:€321–332.
Chappell, J. and Veeh, H. H. (1978). Late Quaternary tectonic movements and sea-level changes
in Timor and Atauro Island, Bulletin, Geological Society of America, 89:€356–368.
Chardon, A.-C. (1999). A geographic approach of the global vulnerability in urban area:€case
of Manizales, Colombian Andes, GeoJournal, 49:€197–212.
Chow, V. T. (1964). Handbook of Applied Hydrology. New York:€McGraw-Hill, 7–25.
Chuah, G. C. M. (1987). The hydrological impact of land use changes on runoff in the upper
Bukit Timah Basin. Unpublished academic exercise, National University of Singapore,
Singapore.
Clapperton, C. M. (1983). The glaciation of the Andes, Quaternary Science Reviews,
2:€83–155.
Clapperton, C. M., Hall, M., Mothes, P., et al. (1997). A younger Dryas icecap in the equa-
torial Andes, Quaternary Research, 47:€13–28.
Clark, M. K., Schoenbohm, L. M., Royden, L. H., et al. (2004). Surface uplift, tectonics,
and erosion of eastern Tibet from large-scale drainage patterns, Tectonics, 23, TC1006d
oi:€10.1029/2002TC001402.
Coleman, J. M. (1969). Brahmaputra River:€ channel processes and sedimentation,
Sedimentary Geology, 8:€129–239.
Colinvaux, P. A., de Oliviera, P. E., Moreno, J. E., et al. (1996). A long pollen record from
lowland Amazonia:€forest and cooling in glacial times, Science, 274:€85–88.
Corlett, R. T. (2009). The Ecology of Tropical East Asia. Oxford University Press.
Costa, J. E. (1974). Response and recovery of a piedmont watershed from tropical storm
Agnes, June1972, Water Resources Research, 10:€106–112.
Costa, J. E. (1984). Physical geomorphology of debris flows, in Developments and
Applications of Geomorphology, J. E. Costa and P. J. Fleisher (eds.). Berlin:€Springer-
Verlag, 268–317.
Costa, J. E. (1988). Rheologic, geomorphic, and sedimentologic differentiation of water
floods, hyperconcentrated flows, and debris flows, in Flood Geomorphology, V. R.
Baker, R. C. Kochel and P. C. Patton (eds). Wiley:€New York, 113–122.
Costa, J. E. and O’ Connor, J. E. (1995). Geomorphically effective floods, in Nature and
Anthropogenic Influences in Fluvial Geomorphology, J. E. Costa, A. J. Miller, K. W.
353 References

Potter and P. Wilcock (eds.). Washington, DC:€American Geophysical Union, Monograph


89, 45–56.
Cotton, C. A. (1944). Volcanoes as Landscape Forms. Christchurch, NZ:€ Whitcombe and
Tombs.
C. W. Thornthwaite Associates (1962–65) Climatic Water Balance Data of the Continents.
Laboratory of Climatology, Publications in Climatology, vols, 15–18.
Dana, J. D. (1850). On denudation in the Pacific, American Journal of Science, 9:€48–62.
Dardis, G. F. and Moon, B. P. (eds.) (1988). Geomorphological Studies in Southern Africa.
Rotterdam:€Balkema.
Das, P. K. (1968). The Monsoons. New Delhi:€National Book Trust, India.
Das Gupta, S. P. (1984). The Ganga Basin, Part I. New Delhi:€ Central Board for the
Prevention and Control of Water Pollution.
Davies, J. L. (1964). A morphogenic approach to world’s shorelines, Zeitschrift für
Geomorphologie, N.F. 8:€107–142.
Davies, J. L. and Williams, M. A. J. (eds.) (1978). Landform Evolution in Australasia.
Canberra: A.N.U. Press.
Davis, W. M. (1899). The geographical cycle, Geographical Journal, 14:€481–504.
De Martonne, E. (1951). Traité de géographie physique. Tome II. Le relief du sol. Paris:
Colin.
Dearman, W. R. (1974). Weathering classification in the characterisation of rock for engin-
eering purposes in British practice, Bulletin, International Association of Engineering
Geology, 9:€33–42.
Dearman, W. R. (1976). Weathering classification in the characterisation of rock:€a revi-
sion, Bulletin, International Association of Engineering Geology, 13:€123–127.
Dendy, F. E. and Bolton, G. C. (1976). Sediment-yield-runoff-drainage area relationships
in the United States, Journal of Soil and Water Conservation, 31:€264–266.
Deodhar, L. A. and Kale, V. S. (1999) Downstream adjustments in allochthonous riv-
ers:€western Deccan Trap upland region, India, in Varieties of Fluvial Form, A. J. Miller
and A. Gupta (ed.). Chichester:€Wiley, 294–315.
Derbyshire, E. (1996). Quaternary glacial sediments, glaciation style, climate and uplift
in the Karakoram and northwest Himalaya:€review and speculations, Palaeogeography,
Palaeoclimatology, Palaeoecology, 120:€147–157.
Dhar, O. N., Kulkarni, A. K. and Sangam, R. B. (1984). Some aspects of winter and mon-
soon rainfall distribution over the Garhwal-Kumaun Himalayas − a brief appraisal,
Himalayan Resource Development, 2:€10–19.
Dietrich, W. E., Day, G. and Parker, G. (1999).The Fly River, Papua New Guinea:€infer-
ences about river dynamics, floodplain sedimentation and fate of sediment, in Varieties
of Fluvial Form, A. J. Miller and A. Gupta (eds). Chichester:€Wiley, 346–376.
Dodge, R. E., Fairbanks, R. G., Benninger, L. K. et al. (1983). Pleistocene sea levels from
raised coral reefs of Haiti, Science, 210:€1623–1625.
Douglas, I. (1967). Erosion of granitic terrains under tropical rain forest in Australia,
Malaysia and Singapore, in International Association of Scientific Hydrology, Publication
75, 31–39.
354 References

Douglas, I. (1977). Humid Landforms. Cambridge:€MIT Press.


Douglas, I. (2005a). The urban geomorphology of Kuala Lumpur, in The Physical
Geography of Southeast Asia, A. Gupta (ed.). Oxford University Press, 344–357.
Douglas, I. (2005b). The urban environment in Southeast Asia, in The Physical Geography
of Southeast Asia, A. Gupta (ed.). Oxford University Press, 314–335.
Douglas, I. and Guyot, J. L. (2005). Erosion and sediment yield in the humid tropics, in
Forests, Water and People in the Humid Tropics, M. Bonell and L. A. Bruijnzeel (eds.).
Cambridge University Press, 407–421.
Douglas, I., Greer, T., Bidin, K. et al. (1993). Impacts of rainforest logging on river sys-
tems and communities in Malaysia and Kalimantan, Global Ecology and Biogeography
Letters, 3:€245–252.
Douglas, I., Spencer, T., Greer, T., et al. (1992). The impact of selective commercial logging
on stream hydrology, chemistry and sediment loads in the Ulu Segama rain forest, Sabah,
Malaysia, Philosophical Transactions, Royal Society of London, B, 335:€397–406.
Dudal, R. (2005). Soils of Southeast Asia, in The Physical Geography of Southeast Asia,
A. Gupta (ed.). Oxford University Press, 94–104.
Dunne, T. (1979). Sediment yield and land use in tropical catchments, Journal of Hydrology,
42:€281–300.
Dunne, T. and Leopold, L. B. (1978). Water in Environmental Planning. San Francisco:€W.H.
Freeman.
Dunne, T., Mertes, L. A. K., Meade, R. H., et al. (1998). Exchanges of sediment between
the floodplain and channel of the Amazon River in Brazil, Bulletin, Geological Society
of America, 110:€450–467.
Duplessy, J. C. (1982). Glacial to interglacial contrasts in the northern Indian Ocean,
Nature, 295:€494–498.
Eckholm, E. P. (1976). Losing Ground:€Environmental stress and world food prospects.
New York:€W.W. Norton.
Elliott, T. (1986). Deltas, in Sedimentary Environments and Facies, H. G. Reading (ed.).
Oxford:€Blackwell Scientific Publications, 113–154.
Emery, K. O. and Kuhn, G. G. (1982). Sea cliffs:€their processes, profiles and classifica-
tion, Bulletin, Geological Society of America, 93:€644–654.
Emery, K. O., Tracey, J. I. and Ladd, H. S. (1954). Geology of Bikini and nearby atolls,
United States Geological Society of America Bulletin, 93:€644–654.
Emmel, F. J. and Curray, J. R. (1982). A submerged late Pleistocene delta and other features
related to sea level changes in the Malacca Strait, Marine Geology, 47:€197–216.
Enzel, Y., Ely, L. L., Mishra, S., et al. (1999). High-resolution Holocene environmental
changes in the Thar Desert, northwestern India, Science, 284:€125–128.
ESCAP Secretariat (1988). Geological information for planning in Bangkok, Thailand, in
Geology and Urban Development; Atlas of Urban Geology 1. Bangkok:€UN-ESCAP,
24–60.
Faniran, A. and Jeje, L. K. (1983). Humid Tropical Geomorphology. London:€Longmans.
Flores, J. F. and Balagot, V. F. (1969). Climate of the Philippines, in Climates of
Northern and Eastern Asia:€ World Survey of Climatology, H. Arakawa (ed.) vol. 8.
Amsterdam:€Elsevier, 159–213.
355 References

Folk, R. L., Roberts, H. H. and Moore, C. N. (1973). Black phytokarst from Hell, Cayman
Islands, West Indies, Bulletin, Geological Society of America, 84:€2351–60.
Food and Agricultural Organization (1998). World Reference Base for Soil Resources.
World Soil Resources Report 84. Rome:€FAO.
Ford, D. C. and Williams, P. W. (1989). Karst Geomorphology and Hydrology.
London:€Unwin Hyman.
Ford, D. C. and Williams, P. W. (2007). Karst Hydrology and Geomorphology. Chichester:
Wiley.
Fournier, F. (1960). Climat et érosion. Paris:€Presses Universitaires de France.
Francis, P. (1993). Volcanoes:€A planetary perspective. Oxford University Press.
Franzinelli, E. and Igreja, H. (2002). Modern sedimentation in the Negro River, Amazonas
State, Brazil, Geomorphology, 44:€259–271.
Fryberger, S. G., Al-Sari, A. K., Clisham, T. J., et al. (1984). Wind sedimentation in the
Jafurah sand sea, Saudi Arabia, Sedimentology, 31:€413–431.
Gabet, E. J., Burbank, D. W., Putkonen, J. K., et al. (2004). Rainfall thresholds for landslid-
ing in the Himalayas of Nepal, Geomorphology, 63:€131–43.
Gabet, E. J., Burbank. D. W., Pratt-Sitaula, B., et al. (2008). Modern erosion rates in the
High Himalayas of Nepal, Earth and Planetary Science Letter, 267:€482–494.
Galy, A. and France-Lanord, C. (2001). Higher erosion rates in the Himalaya:€geochemical
constraints on riverine fluxes, Geology, 28:€23–26.
Gansser, A. (1973). Facts and theories of the Andes, Journal of the Geological Society,
London, 129:€93–137.
Gardner, T. W., Back, W., Bullard, T. F., et al. (1987) Central America and the Caribbean, in
Geomorphic Systems of North America W. L. Graf (ed.). Boulder:€Geological Society of
America, Centennial Special Volume 2, 343–399.
Gasse, F. and Derbyshire, E. (1996). Environmental changes in the Tibetan Plateau and
surrounding areas€ – preface, Palaeogeography, Palaeoclimatology, Palaeoecology,
120:€1–3.
Gibbs, R. J. (1967). The geochemistry of the Amazon River system:€I. The factors that
control the salinity and the composition and concentration of the suspended solids,
Bulletin, Geological Society of America, 78:€1203–1232.
Gibbs, R. J. (1972). Water chemistry of the Amazon River, Geochemica Cosmochemica
Acta, 36:€1061–1066.
Gillieson, D. (2005). Karst in Southeast Asia, in The Physical Geography of Southeast
Asia, A. Gupta (ed.). Oxford University Press, 157–176.
Giosan, L. and Bhattacharya, J. P. (2005). New directions in deltaic studies, in River Deltas:
Concepts, models and examples, L. Giosan and J. P. Bhattacharya (eds.). Tulsa:€SEPM
(Society for Sedimentary Geology), special publication 85, 3–10.
Gole, C. V. and Chitale, S. V. (1966). Inland delta-building activity of the Kosi River,
Journal, Hydraulic Division, American Society of Civil Engineers, HY2:€111–126.
Goodbred, S. L., Jr. (2003). Response of the Ganges dispersal system to climate
change:€ a source-to-sink view since the last interstade, Sedimentary Geology,
162:€83–104.
356 References

Goodbred, S. L. and Kuehl, S. A. (2000a). Enormous Ganges–Brahmaputra sediment dis-


charge during strengthened early Holocene monsoon, Geology, 28:€1083–1086.
Goodbred, S. L. and Kuehl, S. A. (2000b) The significance of large sediment supply, active
tectonism and eustasy on margin sequence development:€Late Quaternary stratigraphy
and evolution of the Ganges–Brahmaputra delta, Sedimentary Geology, 133:€227–248.
Goswami, D. C. (1985). Brahmaputra River, Assam, India:€physiography, basin degrad-
ation, and channel aggradation, Water Resources Research, 221:€858–878.
Goudie, A. (1973). Duricrusts in Tropical and Subtropical Landscapes. Oxford:€Clarendon
Press.
Goudie, A. S. (2002). Great Warm Deserts of the World:€Landscapes and evolution. Oxford
University Press.
Graf, W. L. (1988). Fluvial Processes in Dryland Rivers. Berlin:€Springer.
Graham, N. E., Barnett, T. P., Wilde, R., et al. (1994). On the roles of tropical and midlati-
tude SSTs in forcing interannual to interdecadal variability in the winter northern hemi-
sphere circulation, Journal of Climatology, 7:€1416–1441.
Greer, T., Bidin, K. and Douglas, I. (1994). Tropical rain forest disturbance and sus-
pended sediment discharging variation, in Variability in Stream Erosion and Sediment
Transport:€Poster contribution, L. J. Olive and J. A. Kesby (eds.). Canberra:€Australian
Defence Force Academy, 34–38.
Greer, T., Sinun, W., Douglas, I. et al. (1996). Long-term natural forest management and land-
use changes in a developing tropical catchment, in Erosion and Sediment Yield:€Global
and regional perspectives, D. E. Walling and B. W. Webb (eds.). Wallingford:€ IAHS
Publication 236, 453–461.
Griffiths, J. F. (1972) Climates of Africa, World Survey of Climatology, vol. 10.
Amsterdam:€Elsevier.
Griffiths, J. W. and Ranaivoson, R. (1972). Madagascar, in Climates of Africa:€ World
Survey of Climatology, J. W. Griffiths (ed.) vol. 10. Amsterdam:€Elsevier, 461–499.
Grodek, T., Lekach, J. and Schick, A. P. (2000). Urbanizing alluvial fans as flood-
�conveying and flood-reducing systems:€lessons from the October 1997 Eilat flood, in
The Hydrology–Geomorphology Interface:€ Rainfall, floods, sedimentation, land use,
M. A. Hassan, O. Slaymaker and S. M. Berkowicz (eds.). Wallingford:€ International
Association of Hydrological Sciences Publication 261, 229–250.
Grove, A. T. (1977). The geography of semi-arid lands, Philosophical Transactions of the
Royal Society of London, Series B, 278:€457–475.
Guilderson, T. P., Fairbanks, R. G. and Rubenstone, J. L. (1994). Tropical temperature
variations since 20,000 years ago:€modulating interhemispheric climate change, Science,
263:€663–664.
Gupta, A. (1975). Stream characteristic in eastern Jamaica, an environment of seasonal
flow and large floods, American Journal of Science, 275:€825–847.
Gupta, A. (1984). Urban hydrology and sedimentation in the humid tropics, in Developments
and Applications of Geomorphology, J. E. Costa and P. J. Fleisher (eds.). Berlin:€Springer-
Verlag, 240–267.
Gupta, A. (1988). Large floods as geomorphic agents in the humid tropics, in Flood
Geomorphology, V. R. Baker, R. C. Kochel and P. C. Patton (eds.). New York:€Wiley,
301–315.
357 References

Gupta, A. (1992). Floods and sediment production in Singapore, in Physical Adjustments


in a Changing Landscape:€ The Singapore story, A. Gupta and J. Pitts (eds.).
Singapore:€Singapore University Press, 301–326.
Gupta, A. (1993). The changing geomorphology of the humid tropics, Geomorphology,
7:€165–186.
Gupta, A. (1996).Erosion and sediment yield in Southeast Asia, in Erosion and Sediment
Yield:€ Global and regional perspectives, D. E. Walling and B. W. Webb (eds.).
Wallingford:€IAHS Publication 236, 217–222.
Gupta, A. (2000). Hurricane floods as extreme geomorphic events, in The Hydrology–
Geomorphology Interface:€ Rainfall, floods, sedimentation, land use, M. A. Hassan,
O. Slaymaker and S. M. Berkowicz (eds.). Wallingford:€ International Association of
Hydrological Sciences, Publication 261, 215–228.
Gupta, A. (2005a). Accelerated erosion and sedimentation in Southeast Asia, in The Physical
Geography of Southeast Asia, A. Gupta (ed.). Oxford University Press, 239–249.
Gupta, A. (2005b) Rivers of Southeast Asia, in The Physical Geography of Southeast Asia,
A. Gupta (ed.). Oxford University Press, 65–79.
Gupta, A. (2007a). Introduction, in Large Rivers: Geomorphology and Management, A.
Gupta (ed.). Chichester:€Wiley, 1–5.
Gupta, A. (2007b). The Mekong River:€ morphology, evolution, management, in Large
Rivers:€ Geomorphology and Management, A. Gupta (ed.). Chichester:€ Wiley,
435–455.
Gupta, A. (2009). Geology and landforms of the Mekong Basin, in The Mekong:€Biophysical
environment of an international river basin, I. C. Campbell (ed.). Amsterdam: Elsevier,
29–51.
Gupta, A. (2011). The effect of global warming on large rivers and deltas, in Developing
Countries Facing Global Warming, M. De Dapper, D. Swinne and P. Ozer (eds).
Brussels:€Royal academy for Overseas Sciences, 123–136.
Gupta, A. and Ahmad, R. (1999). Geomorphology and the urban tropics:€building an inter-
face between research and usage, Geomorphology, 31:€133–149.
Gupta, A. and Asher, M. G. (1998). Environment and the Developing World:€Principles,
policies and management. Chichester:€Wiley.
Gupta, A. and Chen, P. (2002). Sediment movement on steep slopes to the Mekong
River:€an application of remote sensing, in The Structure, Function and Management
Implications of Fluvial Sedimentary Systems. F. J. Dyer, M. C. Thoms and J. M. Olley
(eds.). Wallingford:€IAHS Publications 276, 399–406.
Gupta, A. and Dutt, A. (1989). The Auranga:€ description of a tropical monsoon river.
Zeitschrift für Geomorphologie, 33:€73–92.
Gupta, A. and Fox, H. (1974). Effects of high-magnitude floods on channel form:€a case
study in Maryland Piedmont, Water Resources Research, 10:€499–509.
Gupta, A. and Krishnan, P. (1994). Spatial distribution of sediment discharge to the coastal
waters of South and Southeast Asia, in Variability in Stream Erosion and Sediment
Transport, L. J. Olive, R. J. Loughran and J. A. Kesby (eds.). Wallingford:€ IAHS
Publication 224, 457–463.
Gupta, A. and Liew, S. C. (2007). The Mekong from satellite imagery:€a quick look at a
large river, Geomorphology, 85:€259–274.
358 References

Gupta, A., Kale, V. S. and Rajaguru, S. N. (1999). The Narmada River, India, through space
and time, in Varieties of Fluvial Form, A. J. Miller and A. Gupta (eds.) Chichester:€Wiley,
pp. 113–143.
Gupta, A., Liew, S. C. and Heng, A. W. C. (2006). Sediment storage and transfer in the
Mekong:€generalisations on a large river, in Sediment Dynamics and Hydromorphology
of Fluvial Systems, J. S. Rowan, R. W. Duck and A.Werritty (eds.). Wallingford:€IAHS
Publication 306, 450–459.
Gupta, A., Lim, H., Huang, X. et al. (2002). Evaluation of part of the Mekong River using
satellite imagery, Geomorphology, 44:€221–239.
Gupta, A., Rahman, A., Wong, P. P. et al. (1987). The Old Alluvium of Singapore and the
extinct drainage system to the South China Sea, Earth Surface Processes and Landforms,
12:€259–275.
Habib, P. (1975). Production of gaseous pore pressure during rock slides, Rock Mechanics,
7:€193–197.
Hancock, G. S., Anderson R. S. and Whipple, K. X. (1998). Beyond power:€bedrock river
incision process and form, in Rivers Over Rock:€Fluvial processes in bedrock channels,
K. J. Tinkler and E. E. Wohl (eds.). Washington, DC:€American Geophysical Union,
35–60.
Harbor, J. M. (1992). Numerical modelling of the development of U-shaped valleys by gla-
cial erosion, Bulletin, Geological Society of America, 134:€1364–1375.
Harbor, J. M., Hallet, B. and Raymond, C. F. (1988). A numerical model of landform devel-
opment by glacial erosion, Nature, 333:€347–349.
Harvey, A. M. (2002) Effective timescales for coupling within fluvial systems,
Geomorphology, 44:€175–201.
Hastenrath, S. (1984). The Glaciers of Equatorial East Africa. Dordrecht:€D. Reidel.
Hatch, T. (1981). Preliminary results of soil erosion and conservation trials under pepper
(Piper nigrum) in Sarawak, Malaysia, in Soil Conservation:€Problems and prospects,
R. P. C. Morgan (ed.). Chichester:€Wiley, 255–262.
Henry, W. K. (1974) The tropical rainstorm, Monthly Weather Review, 102:€717–725.
Hjulström, F. (1939). Transportation of detritus by moving water, in Recent Marine
Sediments:€A Symposium, P. Trask (ed.). Tulsa, OK:€American Association of Petroleum
Geologists.
Hope, G. (2005). The Quaternary in Southeast Asia, in The Physical Geography of Southeast
Asia, A. Gupta (ed.). Oxford University Press, 24–37.
Hori, K. and Saito, Y. (2007). Classification, architecture, and evolution of large-river deltas,
in Large Rivers:€Geomorphology and management, A. Gupta (ed.). Chichester:€Wiley,
75–96.
Horton, B. K. and Decelles, P. G. (2001) Modern and ancient fluvial megafans in the fore-
land basin systems of the central Andes, southern Bolivia:€Implication for drainage net-
work evolution in fold-thrust belts, Basin Research, 13:€43–61.
Horton, R. E. (1945). Erosional development of streams and their drainage basins, Bulletin,
Geological Society of America, 56:€275–370.
Hovius, N. (1998). Controls of sediment supply by large river, in Relative Role of Eustasy,
Climate and Tectonism in Continental Rocks, K. W. Shanley and P. J. McCabe (eds.).
Tulsa:€Society for Sedimentary Petrology, Special Publication 59, 2–16.
359 References

Hsü, K.J. (1975). Catastrophic debris streams (sturzstroms) generated by rockfalls, Bulletin,
Geological Society of America, 86:€129–40.
Hudson, N. (1971) Soil Conservation. Ithaca:€Cornell University Press.
Hudson, N. (1981). Soil Conservation. London:€ English Language Book Society and
Batsford Academic and Educational.
Hutchinson, J. N.(1968). Mass Movement, in The Encyclopaedia of Geomorphology, R.€W.
Fairbridge (ed.). New York:€Reinhold, 688–696.
Imbrie, J. and Imbrie, K. P. (1979). Ice Ages:€ Solving the mystery. Short Hills:€ Enslow
Publishers.
Inman, D. L. and Nordstrom, C. E. (1971). On the tectonic and morphological classifica-
tion of coasts, Journal of Geology, 79:€1–21.
Intergovernmental Panel on Climate Change (2007). Climate Change 2007:€The Physical
Science Basis. Contribution of the Working Group I to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change, S. Solomon, D. Qin, M. Manning,
Z.€Chen, M. Marquis, K. B. Averyt, M. Tignor and H. L Miller (eds.). Cambridge and
New York:€Cambridge University Press.
Jamaica Weather Reports 1909. No. 372, Kingston, Jamaica.
Jennings, A. H. (1950). World’s greatest observed point rainfalls, Monthly Weather Review,
78:€4–5.
Jennings, J. N. (1985). Karst Geomorphology. Oxford:€Blackwell.
Jennings, J. N. and Mabbutt, J. A. (eds.) (1967). Landform Studies from Australia and New
Guinea. Canberra:€A.N.U. Press.
Johnsson, M. J., Stallard, R. F. and Meade, R. H. (1988). First-class quartz arenites in the
Orinoco River Basin, Venezuela and Colombia, Journal of Geology, 96:€263–277.
Junk, W. J. (ed.) (1997). The Central Amazon Floodplain:€Ecology of a pulsating system.
Berlin:€Springer.
Kale, V. S. and Gupta, A. (2001). Introduction to Geomorphology. Calcutta:€Orient Longman/
Orient Blackswan Pvt Ltd.
Kale, V. S. and Hire, P. (2004). Effectiveness of monsoon floods on the Tapi River,
India:€ role of channel geometry and hydrologic regime, Geomorphology, 57:
275–291.
Kale, V. S. and Rajaguru, S. N. (1987). Late Quaternary alluvial history of the northwest
Deccan Trap region, Nature, 325:€612–614.
Kale, V. S., Gupta, A. and Singhvi, A. K. (2003). Late Pleistocene-Holocene palaeohydrol-
ogy of Monsoon Asia, in Palaeohydrology:€Understanding global change, K. J. Gregory
and G. Benito (eds.). Chichester:€Wiley, 213–232.
Kale, V. S., Ely, L. L., Enzel, Y. et al. (1994). Geomorphic and hydrologic aspects of
monsoon floods on the Narmada and Tapi Rivers in central India, Geomorphology,
10:€157–168.
Kangning, X. (1992). Morphometry and evolution of fenglin karst in Suicheng area, west-
ern Guizhou, China, Zeitschrift für Geomorphologie, NF 36, 227–248.
Kar, A. (1987) Origin and transformation of longitudinal sand dunes in the Indian desert,
Zeitschrift für Geomorphologie, 31:€311–337.
Kellman, M. and Tackaberry, R. (1997). Tropical Environments:€The functioning and man-
agement of tropical ecosystems. London and New York:€Routledge.
360 References

Kershaw, A. P. (1978). Record of the last interglacial cycle from northeastern Queensland,
Nature, 272:€159–161.
Khan, A. A., Dubey, U. S., Sehgal, M. N. et al. (1982). Terraces in the Himalayan tributar-
ies of Ganges in Uttar Pradesh, Journal Geological Society of India, 23:€392–401.
King, L. C. (1951). South African Scenery. Edinburgh:€Oliver and Boyd.
King, L. C. (1962). Morphology of the Earth. Edinburgh:€Oliver and Boyd.
Komar, P. D. (1998). Beach Processes and Sedimentation. Upper Saddle River, NJ:€Prentice-
Hall.
Kondolf. G. M. (1997). Hungry water:€effects of dams and gravel mixing on river channels,
Environmental Management, 21:€533–551.
Krishnan, M. S. (1982). Geology of India and Burma. Delhi:€ CBS Publishers and
Distributors.
Krynine, P. D. (1936). Geomorphology and sedimentation in the humid tropics, American
Journal of Science, 232:€297–306.
Kuehl, S. A., DeMaster, D. J. and Nittrouer, C. A. (1986). Nature of sediment accumulation
on the Amazon continental shelf, Continental Shelf Research, 6:€208–225.
Kuehl, S. A., Allison, M. A., Goodbred, S. L. et al. (2005). The Ganges–Brahmaputra delta,
in River Deltas: Concepts, models and examples, L. Giosan and J. P. Bhattacharya (eds.).
Tulsa:€SEPM (Society for Sedimentary Geology), special publication 85, 413–434.
Kummu, M. and Varis, O. (2007). Sediment-related impacts due to upstream reservoir trap-
ping, the Lower Mekong River, Geomorphology, 85:€275–293.
Kummu, M., Lu, X. X., Wang, J. J. et al. (2010) Basin-wide sediment trapping efficiency
of emerging reservoirs along the Mekong, Geomorphology, 119:€181–197.
Lacey, G. (1930). Stable channels in alluvium, Proceedings, Institute of Civil Engineers,
London, 229:€259–384.
Lai, F. S., Ahmad, J. S. and Zaki, A. M. (1996). Sediment yields from selected catchments in
Peninsular Malaysia, in Erosion and Sediment Yield:€Global and regional perspectives,
D. E. Walling and B. W. Webb (eds.). Wallingford:€IAHS Publications 236, 223–231.
Lal, R. (1976). Soil erosion on alfisols in western Nigeria, III: effects of rainfall character-
istics, Geoderma, 16:€389–401.
Lal, R. (1987). Tropical Ecology and Physical Edaphology. Chichester:€Wiley.
Lamb, P. (1975). Slope failures in Hong Kong, Quarterly Journal of Engineering Geology,
8:€31–65.
Lambiase, J. J. (1989).The framework of African Rifting during the Phanerozoic, Journal
of African Earth Sciences, 8:€183–190.
Lancaster, N. (1995). Geomorphology of Desert Dunes. London:€Routledge.
Lane, E. W. (1955). The importance of fluvial morphology in river hydraulic engineering,
Proceedings of the American Society of Civil Engineers, 81:€1–17.
Langbein, W. B. and Leopold, L. B. (1964) Quasi-equilibrium states in channel morph-
ology, American Journal of Science, 262:€782–794.
Langbein, W. B. and Schumm, S. A. (1958). Yield of sediment in relation to mean annual
precipitation, Transactions, American Geophysical Union, 39:€1076–84.
Larsen, M. C. and Simon, A. (1993) Rainfall-threshold conditions for landslides in a
humid-tropical system, Puerto Rico, Geografiska Annaler, 75A:€13–23.
361 References

Larsen, M. C. and Torres Sánchez, A. J. (1992) Landslide triggered by Hurricane Hugo in


eastern Puerto Rico, September 1989, Caribbean Journal of Science, 28:€113–125.
Larsen, M. C., Vásquez Conde, M. T. and Clark, R. A. (2001a). Flash-flood related haz-
ards:€landslides, with examples from the December 1999 disaster in Venezuela, in Coping
with Flash Floods, E. Gruntfest and J. Handmer (eds.). Dordrecht:€Kluwer, 259–275.
Larsen, M. C., Wieczorek, G. F., Eaton, L. S. et al. (2001b). The rainfall-triggered landslide
and flash flood disaster in northern Venezuela, December 1999, Proceedings, Seventh
Federal Interagency Sedimentation Conference, 2001, Reno, IV-9€– IV-16.
Latrubesse, E. and Franzinelli, E. (2002). The Holocene alluvial plain of the middle Amazon
River, Brazil, Geomorphology, 44:€241–257.
Latrubesse, E. M. and Stevaux, J. C. (2002). Geomorphology and environmental aspects of
the Araguaia fluvial basin, Brazil, Zeitschrift für Geomorphologie, 129:€109–127.
Lehmann, H. (1936). Morphologische Studien auf Java, Geographische Abhandlungen,
9:€15–67.
Leier, A. L., DeCelles, P. G. and Pelletier, J. D. (2005). Mountains, monsoons and mega-
fans, Geology, 33:€289–292.
Leopold, L. B. (1968). Hydrology for urban land planning. U.S. Geological Survey Circular
554, Washington, DC.
Leopold, L. B. (1973). River channel changes with time:€an example, Bulletin, Geological
Society of America, 84:€1845–1860.
Leopold, L. B. and Maddock, T. (1953). The hydraulic geometry of stream channels and
some physiographic implications, U.S. Geological Survey Professional Paper 282-A,
Washington DC.
Leopold, L. B., and Miller, J. P. (1956) Ephemeral streams€ – hydraulic factors and
their �relation to the drainage net, U.S. Geological Survey Professional Paper 282-A,
Washington DC.
Leopold, L. B. and Wolman, M. G. (1957). River channel patterns:€braided, meandering
and straight, U.S. Geological Survey Professional Paper 282-B, Washington D.C.
Leopold, L. B., Wolman, M. G. and Miller, J. P. (1964). Fluvial Processes in Geomorphology.
San Francisco:€Freeman.
Liew, S. C., Gupta, A., Wong, P. P. et al. (2008a). Coastal recovery following the destruc-
tive tsunami of 2004:€Aceh, Sumatra, Indonesia, The Sedimentary Record, 6(3):€4–9.
Liew, S. C., Gupta, A., Wong, P. P. et al. (2010). Recovery from a large tsunami mapped
over time:€The Aceh coast, Sumatra, Geomorphology, 114:€520–529.
Liew, S. C., Thouret, J.-C., Gupta, A. et al. (2008b). First satellite image of a moving pyro-
clastic flow, Eos, Translations, American Geophysical Union, 88(22):€202.
Lirios, J. F. (1969). Rainfall-Intensity-Duration-Frequency Maps for Kingston and
St.€Andrews, Jamaica, Mimeographed. Jamaica:€Caribbean Meteorological Institute.
Loughnan, F. C. (1969). Chemical Weathering of the Silicate Minerals. New York:
Elsevier.
McCarthy, T. S., Cooper, G. J. R., Tyson, P. D. et al. (2000). Seasonal flooding in the
Okavango Delta, Botswana€– recent history and future prospects, South African Journal
of Science, 86:€25–33.
McGregor, G. R. and Nieuwolt, S. (1998). Tropical Climatology. Chichester:€Wiley.
362 References

Macintyre, I. G. (1988). Modern coral reefs of Western Atlantic:€new geological perspec-


tive, Bulletin, American Association of Petroleum Geologists, 72:€1360–1369.
McLean, R. F. and Woodroffe, C. D. (1994). Coral atolls, in Coastal Evolution:€ Late
Quaternary shoreline morphodynamics, R. W. G. Carter and C. D. Woodroffe (eds.).
Cambridge University Press, 267–302.
Mackin, J. H. (1948). Concept of the graded stream, Bulletin, Geological Society of
America, 59:€463–512.
Madej, M. A. and Ozaki, V. (1996). Channel response to sediment wave propagation and
movement, Redwood Creek, California, USA, Earth Surface Processes and Landforms,
21:€91–927.
Mahaney, W. C. (ed.) (1989). Quaternary and Environmental Research on East African
Mountains. Rotterdam:€A.A. Balkema.
Maharaj, R. J. (1993). Landslide processes and landslide susceptibility analysis from an
upland watershed:€ a case study from St. Andrew, Jamaica, West Indies, Engineering
Geology, 34:€53–79.
Mahmood, K. (1987). Reservoir Sedimentation:€Impact, extent and mitigation. Washington,
DC:€World Bank, Technical paper 71.
Marsh, G. P. (1965). Man and Nature, D. Lowenthal (ed.). Cambridge, Mass.:€The Belknap
Press of Harvard University Press.
Meade, R. H. (1996). River-sediment input to major deltas, in Sea-level Rise and Coastal
Subsidence:€Causes, consequences and strategies, J. D. Milliman and B. U. Haq (eds.).
Dordrecht:€Kluwer, 63–85.
Meade, R. H. (2007). Transcontinental moving and storage:€ The Orinoco and Amazon
rivers transfer the Andes to the Atlantic, in Large Rivers:€Geomorphology and manage-
ment, A. Gupta (ed.). Chichester:€Wiley, 45–63.
Mekong River Commission (MRC) (1997). Mekong River Basin Diagnostic Study:€Final
Report. Bangkok:€Mekong River Commission, MKG/R.88/013.
Mertes, L. A. K. and Dunne, T. (2007). Effects of tectonism, climate change and sea-level
change on the form and behaviour of the modern Amazon River and its floodplain, in
Large Rivers:€ Geomorphology and management, A. Gupta (ed.). Chichester:€ Wiley,
115–144.
Mertes, L. A. K., Dunne, T. and Martinelli, L. A. (1996). Channel-floodplain geomorph-
ology along the Solimões-Amazon River. Brazil, Bulletin, Geological Society of America,
108:€1088–1107.
Messerli, B., Winiger, M. and Rognon, P. (1980). The Saharan and East African Uplands
during the Quaternary, in The Sahara and the Nile, M. A. J. Williams and H. Faure
(eds.). Rotterdam:€Balkema, 87–132.
Meybeck, M. (1987). Global chemical weathering of surficial rocks estimated from river
dissolved loads, American Journal of Science, 287:€401–428.
Milliman, J. D. and Meade, R. H. (1983). World-wide delivery of river sediment to the
oceans, Journal of Geology, 91:€1–21.
Milliman, J. D. and Syvitski, J. P. M. (1992). Geomorphic/tectonic control of sediment dis-
charge to the ocean:€ the importance of small mountainous rivers, Journal of Geology,
100:€525–544.
363 References

Milly, P. C. D., Betancourt, J., Falkenmark, M., et al. (2008). Stationarity is dead:€whither
water management?, Science, 319:€573–574.
Mohr, E. C. J. and Van Baren, F. A. (1954). Tropical Soils. The Hague:€Van Hoeve.
Molengraff, G. A. F. (1921). Modern deep-sea research in the East Indian Archipelago,
Geographical Journal, 58:€95–121.
Monroe, W. H. (1976). The karst landforms of Puerto Rico, U.S. Geological Survey
Professional paper 899, Washington DC.
Moore, A. E. and Blenkinsop, T. (2002). The role of mantle plumes in the development of
continental-scale drainage patterns:€the South African example revisited, South African
Journal of Geology, 105:€353–360.
Moore, A. E. and Larkin, P. A. (2001). Drainage evolution in south-central Africa since the
break-up of Gondwana, South African Journal of Geology, 104:€47–68.
Moore, A. E., Cotterill, F. P. D., Main, M. P. L. et al. (2007). The Zambezi River, in Large
Rivers:€Geomorphology and management, A. Gupta (ed.). Chichester:€Wiley, 311–332.
Murtedza, M. and Ti, T. C. (1993). Managing ASEAN’s forests:€deforestation in Sabah, in
Environmental Management in ASEAN, M. Seda (ed.). Singapore:€Institute of Southeast
Asian Studies, 111–140.
Nemcok, A., Pasek, J. and Rybar, J. (1972). Classification of landslides and other mass
movements, Rock Mechanics, 4:€71–8.
Newhall, C. and Punongbayan, R. S. (eds.) (1996). Fire and Mud:€Eruptions and lahars
of Mount Pinatubo, Philippines. Quezon City:€Phivolc Press and Seattle:€University of
Washington Press.
Nguyen, V. L., Ta, T. K. O. and Tatheishi, M. (2000). Late Holocene depositional environ-
ments and coastal evolution of the Mekong River Delta, Southern Vietnam, Journal of
Asian Earth Science, 18:€427–448.
Niekerk, A. W. van, Heritage, G. L., Broadhurst, L. J. et al. (1999). Bedrock anasto-
mosing channel systems:€morphology and dynamics in the Sabie River, Mpumalanga
Province, South Africa, in Varieties of Fluvial Form, A. J. Miller and A. Gupta (ed.).
Chichester:€Wiley, pp. 33–51.
Nieuwolt, S. (1977). Tropical Climatology. Chichester:€Wiley.
Nieuwolt, S. (1981). The climates of continental Southeast Asia, in Climates of Southern
and Western Asia:€World Survey of Climatology, K. Takahashi and H. Arakawa (eds.)
vol. 9. Amsterdam:€Elsevier, 1–66.
Nilsson, C., Reidy, C. A., Dynesius, M. et al. (2005). Fragmentation and flow regulation of
the world’s large river systems, Science, 308:€405–408.
Nittrouer, C. A., Kuehl, S. A., DeMaster, D. J. et al. (1986). The deltaic nature of Amazon
shelf sedimentation, Bulletin, Geological Society of America, 97, 444–458.
Noppadol, Phienwej and Prinya Nutalaya (2005). Subsidence and flooding in Bangkok,
in The Physical Geography of Southeast Asia, A. Gupta (ed.). Oxford University Press,
358–378.
Nossin, J. J. (2005). Volcanic hazards in Southeast Asia, in The Physical Geography of
Southeast Asia, A. Gupta (ed.). Oxford University Press, 250–274.
Nott, J. and Hayne, M. (2001). High frequency of ‘super-cyclones’ along the Great Barrier
Reef over the past 5,000 years, Nature, 413:€508–511.
364 References

Nunn, J. A. and Aires, J. R. (1988). Gravity anomalies and flexure of the lithosphere at the
middle Amazon basin, Brazil, Journal of Geophysical Research, 83:€415–428.
Nutalaya, P., Yong, R. N., Chumnankit, T. et al. (1996). Land subsidence in Bangkok dur-
ing 1978–1988, in Sea-Level Rise and Coastal Subsidence, J. D. Milliman and B. U. Haq
(eds.). Dordrecht:€Kluwer, 105–130.
Nye, J. F. (1952). The mechanics of glacial flow, Journal of Glaciology, 2:€82–93.
Nye, J. F. (1957). The distribution of stress and velocity in glaciers and ice sheets,
Proceedings, Royal Society, Series A, 239:€113–133.
Osmaston, H. (1980). Patterns in trees, rivers and rocks in the Mulu Park, Sarawak,
Geographical Journal, 146(1), 33–50.
Osmaston, H. (1989a). Glaciers, glaciations and equilibrium line altitudes on Kilimanjaro,
in Quaternary and Environmental Research on East African Mountains, W. C. Mahaney
(ed.). Rotterdam:€A.A. Balkema, 7–30.
Osmaston, H. (1989b). Glaciers, glaciations and equilibrium liem altitudes on Ruwenzori,
in Quaternary and Environmental Research on East African Mountains, W. C. Mahaney
(ed.). Rotterdam:€A.A. Balkema, 31–104.
Osmaston, H. A. and Sweeting, M. M. (1982). Geomorphology (of the Guning Mulu
National Park), Sarawak Museum Journal, 51:€75–94.
Palmer, A. N. (1991). Origin and morphology of limestone caves, Bulletin, Geological
Society of America, 103:€1–21.
Patton. P. C., Pickup, G. and Price, D. M. (1993). Holocene paleofloods of the Ross River,
central Australia, Quaternary Research, 40:€201–212.
Patzelt, A., Li, H., Wang, J. et al. (1996). Palaeomagnetism of Cretaceous to Tertiary sedi-
ments from southern Tibet:€evidence for the extent of the northern margin of India prior
to the collision with Eurasia, Tectonophysics, 259:€259–284.
Peltier, L. (1950). The geographic cycle in periglacial regions as it is related to climatic
geomorphology, Annals Association of American Geographers, 40:€214–236.
Pickup, G. and Rieger, W. A. (1979). A conceptual model of the relationship between chan-
nel characteristics and discharge, Earth Surface Processes, 4:€37–42.
Pisahorty, P. R. (1991). Indian rainfall and water conservation, Journal of Geological
Society of India, 37:€608–612.
Pitts, J. (1992). Slope stability in Singapore, in Physical Adjustments in a Changing
Landscape:€ The Singapore story, A. Gupta and J. Pitts (eds.). Singapore University
Press, 259–300.
Portig, W. H. (1976). The climate of Central America, in Climates of Central and
South America:€ World Survey of Climatology, W. Schwerdtfeger (ed.) vol. 12.
Amsterdam:€Elsevier, 405–478.
Potter, P. E. (1978a). Petrology and chemistry of modern big river sands, Journal of
Geology, 86:€423–449.
Potter, P.E. (1978b). Significance and origin of big rivers, Journal of Geology, 86:€13–33.
Prell, W. L. and Kutzbach, J. E. (1992) Sensitivity of the Indian monsoon to forcing param-
eters and implications for its evolution, Nature, 360:€647–652.
Quade, J., Cater, J. M. L., Ojha, T. P., et al. (1995). Late Miocene environmental change
in Nepal and the northern Indian subcontinent:€stable isotopic evidence from paleosols,
Bulletin, Geological Society of America, 107:€1381–1397.
365 References

Queen, E. (1934). The Chinese Orange Mystery, A problem in deduction. Stokes Publishing
Company.
Rao, Y.P. (1981) The climate of the Indian subcontinent, in Climates of Southern and
Western Asia:€World Survey of Climatology, K. Takahashi and H. Arakawa (eds.) vol. 9.
Amsterdam:€Elsevier, 67–182.
Rapp, A., Murray-Rust, D. H., Christiansson, C. et al. (1972). Soil erosion and sedi-
mentation in four catchments near Dodoma, Tanzania, Geografiska Annaler,
54A:€255–318.
Raymo, M. E. and Ruddiman, W. F. (1992). Tectonic forcing of late Cenozoic climate,
Nature, 359:€117–122.
Reynolds, R. (1985). Tropical meteorology, Progress in Physical Geography, 9:€157–186.
Richards, K. S. (1982). Rivers:€Form and process in alluvial channels. London:€Methuen.
Ritter, D. F. Kochel, P. C. and Miller, J. R. (1995). Process Geomorphology. Dubuque,
Iowa:€Wm C. Brown Publishers.
Ruddiman, W. F and Kutzbach, J. E. (1991). Plateau uplift and climate change, Scientific
American, 264:€42–50.
Ruhe, R. V. (1975). Geomorphology:€ Geomorphic Processes and Surface Geology.
Boston:€Houghton Mifflin.
Ruxton B. P. and Berry, L. (1957). Weathering of granite and associated erosional features
in Hong Kong, Bulletin, Geological Society of America, 68:€1263–1293.
Sah, M. P. and Virdi, N. S. (1997). Geomorphic signatures of neotectonic activity along
the Sumdo Fault, Spiti Valley, district Kinnaur, Himachal Pradesh, Himalayan Geology,
18:€81–92.
Said, R. (1994). Origin and evolution of the Nile, in The Nile:€Sharing a scarce resource,
P. P. Howell and J. A. Allan (eds.). Cambridge University Press, 17–26.
Saunders, I. and Young, A. (1983). Rates of surface processes on slopes, slope retreat and
denudation, Earth Surface Processes and Landforms, 8:€473–501.
Scatena, F. N. and Gupta, A. (in press). Streams of the montane humid tropics, in Treatise
on Geomorphology, E. E. Wohl (ed.). Amsterdam: Elsevier.
Schick, A. P. (1988). Hydrological aspects of floods in extreme arid environments, in Flood
Geomorphology, V. R. Baker, R. C. Kochel and P. C. Patton (eds.). New York:€Wiley,
189–203.
Schick, A. P. and Lekach, J. (1987). A high magnitude flood in the Sinai desert, in Catastrophic
Flooding, L. Mayer and D. Nash (eds.). Winchester, MA:€Allen and Unwin, 381–410.
Schlesinger, W. H. (1991). Biogeochemistry:€ An analysis of global change. New
York:€Academic Press.
Schumm, S. A. (1960). The effect of sediment type on the shape and stratification of some
modern fluvial deposits, American Journal of Science, 258:€177–184.
Schumm, S. A. (1968). River adjustment to altered hydrologic regimen€– Murrumbidgee
River and paleo channels, Australia, U.S. Geological Survey Professional Paper 598,
Washington DC.
Schumm, S. A. (1977). The Fluvial System. New York:€Wiley.
Schumm, S. A. (1985). Patterns of alluvial rivers, Annual Review of Earth Sciences,
13:€5–27.
Schumm, S. A. (2005). River Variability and Complexity. Cambridge University Press.
366 References

Schumm, S. A. and Khan, H. R. (1972). Experimental study of channel patterns, Bulletin,


Geological Society of America, 83:€1755–1770.
Schumm, S. A. and Lichty, R. W. (1963). Channel widening and floodplain construction
along Cimarron River in southwestern Kansas, U.S. Geological Survey Professional
Paper 352-D, 71–88.
Selby, M. J. (1993). Hillslope Materials and Processes. Oxford University Press.
Sengupta, S. (1966). Geological and geophysical studies in the western part of the Bengal
Basin, India. Bulletin, American Association of Petroleum Geologists, 50:€1001–1017.
Shackleton, N. J. and Opdyke, N. D. (1973). Oxygen isotope and paleomagnetic stratig-
raphy of equatorial Pacific core V28–238:€oxygen isotope temperatures and ice volumes
on a 105 and 106 year scale, Quaternary Research, 3:€39–55.
Sharp, R. P. (1963). Wind ripples, Journal of Geology, 71:€617–636.
Sharpe, C. F. S. (1938). Landslides and Related Phenomena. New York:€ Columbia
University Press.
Shepherd, J. B. and Aspinall, W. P. (1980). Seismicity and seismic intensities in Jamaica,
West Indies:€ a problem in risk assessment, Earthquake Engineering and Structural
Dynamics, 8:€315–335.
Sherman, G. D. (1952). The genesis and morphology of the alumina-rich laterite clays, in
Problems of Clay and Laterite Genesis, American Institute of Mining and Metallurgical
Engineers, 154–161.
Shinn, E. A., Hudson, J. H., Halley, R. B., et al. (1982). Geology and sediment accumula-
tion rates at Carrie Bow Cay, Belize, in The Atlantic Barrier Reef Ecosystem at Carrie
Bow Cay, Belize. I:€Structure and Communities, K. Rützler and I. G. Macintyre (eds.).
Smithsonian Contributions to Marine Science, Washington DC, 63–75.
Showstack, R. (2010). Haiti earthquake underscores need for better use of seismic informa-
tion, Eos, Transactions, American Geophysical Union, 91(4):€30–31.
Shreve, R. L. (1965). The Blackhawk landslide, Geological Society of America Special
Paper 108.
Shreve, R. L. (1967). Infinite topologically random channel networks, Journal of Geology,
75:€175–186.
Shroder, J. F., Jr. (1993). Himalaya to the sea:€ Geomorphology and the Quaternary of
Pakistan in the regional context, in Himalaya to the Sea:€Geology, geomorphology and
the Quaternary, J. F. Shroder (ed.). London:€Routledge, 1–42.
Shroder, J. F. and Bishop, M. P. (1998). Mass movement in the Himalaya:€new insights and
research directions, Geomorphology, 26:€13–35.
Shukla, U. K., Singh, I. B., Srivastava, P. et al. (1999). Palaeocurrent patterns in braid-bar
and point-bar deposits:€examples from the Ganga River, India, Journal of Sedimentary
Research, 68:€882–1002.
Sidle, R. C., Tsuboyama, Y., Noguchi, S., et al. (2000). Stormflow generation in steep
forested headwaters:€ a linked hydrogeomorphic paradigm, Hydrological Processes,
14:€369–385.
Sikka, D. R. (1977). Some aspects of the life history, structure and movement of monsoon
depressions, Pure and Applied Geophysics, 115:€1501–1529.
367 References

Simons, A., Larsen, M. C. and Hupp, C. R. (1990). The role of soil processes in determin-
ing mechanisms of slope failure and hillslope development in a humid-tropical forest,
eastern Puerto Rico, Geomorphology, 3:€263–286.
Simon, D. B. and Richardson, E. V. (1971). Flow in alluvial sand channels, in River
Mechanics, H. W. Shen (ed.). Fort Collins, 9.1–9.89.
Simonett, D. S. (1967). Landslide distribution and earthquakes in the Betwani and Toricelli
Mountains, New Guinea, in Landform Studies from Australia and New Guinea, J. N.
Jennings and J. A. Mabbutt (eds.). Canberra:€ Australian National University Press,
64–86.
Simpson, R. H. and Riehl, H. (1981). The Hurricane and its Impact. Oxford:€Blackwell.
Singh, H., Parkash, B. and Gohain, K. (1993). Facies analysis of the Kosi megafan depos-
its, Sedimentary Geology, 85:€87–113.
Singh, I. B. (2007a). The Ganga River, in Large Rivers:€Geomorphology and management,
A. Gupta (ed.). Chichester:€Wiley, 347–371.
Singh, M. (1996). The Ganga River:€ Fluvial Geomorphology, Sedimentation Processes
and Geochemical Studies Heidelberg: Heidelberger Beiträge zur Umwelt-Geochemie,
Band 8.
Singh, S. K. (2007b). Erosion and weathering in the Brahmaputra River System, in
Large Rivers:€ Geomorphology and management, A. Gupta (ed.). Chichester:€ Wiley,
373–383.
Singh, S. K. and France-Lanord, C. (2002). Tracing the distribution of erosion in the
Brahmaputra watershed from isotopic composition of stream sediments, Earth Planetary
Science Letters, 202:€645–662.
Singh, S. K., Sarin, M. M. and France-Lanord, C. (2005). Chemical erosion in the eastern
Himalaya:€major ion composition of the Brahmaputra and δ13C of dissolved organic car-
bon, Geochimica et Cosmochimica Acta, 69:€3573–3588.
Sinha, R. and Friend, P. F. (1994). River systems and their sediment flux, Indo-Gangetic
plains, northern Bihar, India, Sedimentology, 41:€825–845.
Sinun, W., Wong, W. M., Douglas, I. et al. (1992). Throughfall, stemflow, overland flow and
throughflow in the Ulu Segama rain forest, Sabah, Malaysia, Philosophical Transactions
Royal Society London, B, 335:€389–395.
Smith, H. T. U. (1954). Coast Dunes, Coastal Geography Conference, Office of Naval
Research, 51–56.
Smith, N.D. (1971) Transverse bars and braiding in the Lower Platte River, Nebraska,
Bulletin, Geological Society of America, 82:€3407–3420.
Snow, J. W. (1976). The climate of Northern South America, in Climates of Africa:€World
Survey of Climatology, W. Schwerdtfeger (ed.) vol. 12. Amsterdam:€ Elsevier,
295–403.
Sonu, C. J. (1972). Field observation of nearshore circulation and meandering currents,
Journal of Geophysical Research, 18:€3232–3247.
Stallard, R. F. (1985). River chemistry, geology, geomorphology and soils in the Amazon
and Orinoco Basins, in The Chemistry of Weathering, J. J. Driver (ed.). Dordrecht:€Reidel,
293–316.
368 References

Stallard, R. F. and Edmond, J. M. (1983). Geochemistry of the Amazon:€1. The influence of


the geology and weathering environment on the dissolved load, Journal of Geophysical
Research, 88:€9671–9688.
Stallard, R. F. and Edmond, J. M. (1987). Geochemistry of the Amazon:€3. Weathering chem-
istry and limits to dissolved inputs, Journal of Geophysical Research, 92:€8293–8302.
Stanistreet, I. G. and McCarthy, T. S. (1993). The Okovango fan and the classification of
subaerial fan systems, Sedimentary Geology, 85:€115–133.
Stanley, D. J. and Warne, A. G. (1994). World initiation of Holocene marine deltas by
deceleration of sea-level rise, Science, 265:€228–231.
Starkel, L. (1972). The role of catastrophic rainfall in the shaping of the relief of the lower
Himalaya (Darjeeling hills), Geografica Polonica, 21:€133–147.
Stoddart, D. R. (1969). Ecology and morphology of recent coral reefs, Biological Reviews,
44:€433–498.
Strahler, A. N. (1952a). Dynamic basis of geomorphology, Bulletin, Geological Society of
America, 63:€923–938.
Strahler, A. N. (1952b). Hypsometric (area-altitude) analysis of erosional topography,
Bulletin, Geological Society of America, 63:€1117–1141.
Strahler, A. N. (1964). Quantitative geomorphology of drainage basins and channel net-
works, in Handbook of Applied Hydrology, V. T. Chow (ed.). New York:€McGraw-Hill,
439–476.
Strahler, A. N. (1975). Physical Geography. New York:€Wiley.
Sugden, D. E. and John, B. S. (1976). Glaciers and landscapes:€ A geomorphological
approach. London:€Edward Arnold.
Summerfield, M. A. (1991). Global Geomorphology. Harlow:€ Longman Scientific and
Technical.
Sunamura, T. (1983). Processes of sea cliff and platform erosion, in Handbook of Coastal
Processes and Erosion, P. D. Komar (ed.). Boca Raton, FL:€CRC Press, 233–265.
Sweeting, M. M. (1958). The karstlands of Jamaica, Geographical Journal,
124:€184–199.
Sweeting, M. M. (1972). Karst Landforms. London:€Macmillan.
Syvitski, J. P. M., Vörösmarty, C. J., Kettner, A. J. et al. (2005). Impact of humans on the
flux of terrestrial sediment to the global coastal ocean, Science, 308:€376–380.
Ta, T. K. O., Nguyen, V. L., Tateishi, M., et al. (2005). Holocene delta evolution and deposi-
tional models of the Mekong River Delta, southern Vietnam, in River Deltas: Concepts,
models and examples, L. Giosan and J. P. Bhattacharya (eds.). Tulsa:€SEPM (Society for
Sedimentary Geology), special publication 85, 453–466.
Ta, T. K. O., Nguyen, V. L., Tateishi, M., et al. (2002). Holocene delta evolution and sedi-
ment discharge of the Mekong River, southern Vietnam, Quaternary Science Review,
21:€1807–1819.
Tanabe, S., Saito, Y., Sato, Y., et al. (2003). Stratigraphy and Holocene evolution of the mud-
dominated Chao Phraya delta. Thailand, Quaternary Science Review, 22:€789–807.
Tandon, S. K. and Sinha, R. (2007). Geology of large river systems, in Large
Rivers:€Geomorphology and management, A. Gupta (ed.). Chichester:€Wiley, 7–28.
369 References

Tapponier, P., Peltzer, G., Le Dain, A. Y., et al. (1982). Propagating extrusion tectonics in
Asia, new insights from simple experiments with plasticine, Geology, 10:€611–616.
Tapponnier, P., Peltzer, G. and Armijo, R. (1986). On the mechanics of the collision
between India and Asia, in Collision Tectonics, M. P. Coward and A. C. Ries (eds.).
London:€Geological Survey of London Special Publication 19, 115–157.
Taylor, A. B and Velbel, M. A. (1991). Geochemical mass balances and weathering rates
in forested watersheds of the southern Blue Ridge. II: Effects of botanic uptake terms,
Geoderma, 51:€29–50.
Temple, P. W. and Rapp, A. (1972). Landslides in the Mgeta area, Western Uluguru
Mountains, Tanzania, Geografiska Annaler, 54A, 157–193.
Terry, J. P. (1999). Kadavu Island, Fiji:€Fluvial studies of a volcanic island in the humid
tropical South Pacific, Singapore Journal of Tropical Geography, 20:€86–98.
Thomas, D. S. G. (ed.) (1997). Arid Zone Geomorphology:€Process, form and change in
drylands. Chichester:€Wiley.
Thomas, M. F. (1994). Geomorphology in the Tropics. Chichester:€Wiley.
Thomas, M. F., Nott, J., Murray, A. S. et al. (2007). Fluvial response to late Quaternary
climate change in NE Queensland, Australia, Palaeogeography, Palaeoclimatology,
Palaeoecology, 251:€119–136.
Thompson, L. G., Mosley-Thompson, E., Davis, M. E., et al. (1995). Late glacial stage and
Holocene tropical ice core records from Huascarán, Peru, Science, 269:€46–50.
Thornbury, W. D. (1960). Principles of Geomorphology. New York:€Wiley.
Thornthwaite, C. W. and Mather, J. R. (with Carter, D. B.) (1957). Instructions and
Tables for Computing Potential Evapotranspiration and the Water Balance. Centerton,
NJ:€Drexel Institute of Technology, Publications in Climatology, 10, 3.
Thouret, J. C. and Lavigne, F. (2005). Hazards and risks at Guning Merapi, central Java:€a
case study, in The Physical Geography of Southeast Asia, A. Gupta (ed.). Oxford
University Press, 275–299.
Thouret, J. -C., Suni, J., Eissen, J.-Ph. et al. (1999) Assessment of volcanic hazards in the
area of Arequipa City, based on the eruptive history of the Misti Volcano, southern Peru,
Zeitschrift für Geomorphologie, 114:€89–112.
Thouret, J. -C., Lavigne, F., Suwa, H., et al. (2007) Volcanic hazards at Mount Semuru, east
Java (Indonesia), with emphasis on lahars, Bulletin of Volcanology, 70:€221–244.
Thouret, J. -C., Gupta, A., Lube, G., et al. (2010) The 2006 pyroclastic deposits of Merapi
Volcano, Java, Indonesia:€High-spatial resolution IKONOS images and conplementary
ground-based observations, Remote Sensing of Environment, 114:€1949–1967.
Tooth, S. (1999). Floodouts in central Australia, in Varieties of Fluvial Form, A. J. Miller
and A. Gupta (eds.). Chichester:€Wiley, 219–247.
Torrance, J. D. (1972). Malawi, Rhodesia and Zambia, in Climates of Africa:€World Survey
of Climatology, J. W. Griffiths (ed.) vol. 10. Amsterdam:€Elsevier, 409–460.
Trewartha, G. T. (1954) An Introduction to Climate. New York:€McGraw-Hill.
Trexler, M. C. and Haugen, C. (1994) Keeping it Green:€Tropical forestry opportunities for
mitigating climate change. Washington, DC:€World Resources Institute.
Twidale, C. R. and Campbell, E. M. (1993) Australian Landforms:€Structure, process and
time. Adelaide:€Gleneagle Publishing.
370 References

United Nations (2008). World Urbanization Prospects:€The 2007 revision. New York:€United
Nations, ESA/P/WP/205.
United Nations Environmental Programme (1992). World Atlas of Desertification.
Sevenoaks:€Edward Arnold.
Valdiya, K. S. (2002). Emergence and evolution of the Himalaya:€reconstructing history in
the light of recent studies, Progress in Physical Geography, 26:€360–399.
Van der Hammen, T., Barelds, J., De Jong, H. et al. (1981). Glacial sequence and envir-
onmental history in the Sierra Nevada del Cocuy (Colómbia), Palaeogeography,
Palaeoclimatology, Palaeoecology, 32:€247–340.
Van der Kaars, W. A. and Dam, M. A. C. (1995). A 135,000-year record of vegetational
and climatic change from the Bandung area, west-Java, Indonesia, Palaeogeography,
Palaeoclimatology, Palaeoecology, 117:€55–72.
Varnes, D. J. (1958). Landslide types and processes, in Landslides and Engineering
Practice, E. Eckel (ed.). Highway Research Board Report, Washington DC, 29:€20–47.
Varnes, D. J. (1978). Slope movement types and processes, in Landslides, R. Shuster and
R. Krizak (eds.). Transport Research Board, National Academy of Sciences, Washington
DC, 11–33.
Verstappen, H. Th. (1975). On palaeo-cliamte and landform development in Malesia, in
Modern Quaternary Research in Southeast Asia 5, G. J. Bartstra and W. A. Casparie
(eds.). Rotterdam:€Balkema, 3–35.
Verstappen, H. Th. (1980). Quaternary climate changes and natural environment in SE
Asia, GeoJournal, 4:€45–54.
Verstappen, H. Th. (2005). Volcanic islands, in The Physical Geography of Southeast Asia,
A. Gupta (ed.). Oxford University Press, 142–156.
Vickers, D. O. (1967). Very heavy and intense rainfalls in Jamaica, Proceedings, University
of the West Indies Conference on Climatology and Related Fields, Underground Water
Authority, Kingston, Jamaica, 57–63.
Viles, H. A. (1984). Biokarst:€ review and prospect, Progress in Physical Geography,
8:€523–542.
Viles, H. A. (2005). Microclimate and weathering in the central Namib Desert, Namibia, in
Weathering and Landscape Evolution, A. Turkington, J. Phillips and S. Campbell (eds.).
Amsterdam:€Elsevier, 189–209.
Viles, H. A. and Spencer, T. (1995). Coastal Problems:€Geomorphology, ecology and soci-
ety at the coast. London:€Wm C. Brown Publishers.
Vohra, C. P. (1981). Himalayan glaciers, in The Himalaya:€Aspects of change, J. S. Lal
(ed.). Delhi:€Oxford University Press, 138–151.
Voight, B. (1996). The management of volcano emergencies:€ Nevado del Ruiz, in
Monitoring and Mitigation of Volcanic Hazards, R. Scarpa and R. I. Tilling (eds.).
Heidelberg:€Springer-Verlag, 719–769.
Vörösmarty, C. J., Meybeck, M., Fekete, B., et al. (2003). Anthropogenic sediment reten-
tion:€major global impact from registered river impoundments, Global and Planetary
Change, 39:€169–190.
371 References

Walling, D. E. and Webb, B. W. (1983). Patterns of sediment yield, in Background to


Palaeohydrology, K. J. Gregory (ed.). Chichester:€Wiley, 69–100.
Walling, D. E. and Webb, B. W. (1987). Material transport by the world’s rivers:€evolv-
ing perspective, in Water for Future:€Hydrology in perspective, IAHS Publication 164.
Wallingford:€Institute of Hydrology, 313–329.
Waltham, A. C. (1995). The pinnacle karst of Gunung Api, Mulu, Cave and Karst Science,
22(3):€123–126.
Warne, A. G., Meade, R. H., White, W. A., et al. (2002). Regional control on geomorphology,
hydrology, and ecosystem integrity in the Orinoco Delta, Venezuela, Geomorphology,
44:€273–307.
Warner, R. F. (ed.) (1988). Fluvial Geomorphology of Australia. Sydney:€ Academic
Press.
Wasson, R. J. (1995) The Asian monsoon during the Late Quaternary:€a test of orbital for-
cing and Palaeoanalogue forecasting, in Quaternary Environments and Geoarchaeology
of India:€Essays in honour of Professor S.N. Rajaguru, S. Wadia, R. Korisettar and V. S.
Kale (eds.). Bangalore:€Geological Society of India, Memoir 32, 22–35.
Wasson, R. J. (2003). A sediment budget for the Ganga–Brahmaputra catchment, Current
Science, 84:€1041–1047.
Wasson, R. J., Smith, G. I. and Agarwal, D. P. (1984). Late Quaternary sediments, minerals
and inferred geochemical history of Didwana lake, Thar Desert, India, Palaeogeography,
Palaeoclimatology, Palaeoecology, 46:€345–372.
Webb, R. H., Wegner, D. L., Andrews, A. D., et al. (1999). Downstream effects of Glen
Canyon Dam on the Colorado River in Grand Canyon:€a review, in The Controlled Flood
in Grand Canyon, R. H. Webb, J. C. Schmidt, G. R. Marzolf and R. A. Valdez (eds.).
Washington, DC:€American Geophysical Union, Geophysical monograph 110, 1–21.
Weera, P. and Kittipong, P. (1987). Amount of runoff and soil losses from various land-use
sampling plots in Sakolnakorn Province, Thailand, in Forest Hydrology and Watershed
Management, R. H. Swanson, P. Y. Bernier and P. D. Woodward (eds.). Wallingford:€IAHS
Publication 177, 231–238.
Wells, N. A. and Dorr, J. A., Jr. (1987). Shifting of the Kosi River, northern India, Geology,
15:€204–207.
Wescott, W. A. and Ethridge, F. G. (1980). Fan-delta sedimentology and tectonic set-
ting€– Yallahs fan delta, northeast Jamaica, Bulletin, American Association of Petroleum
Geologists, 64:€374–399.
Whipple, K. X. (2004). Bedrock rivers and the geomorphology of active orogenes, Annual
Review of Earth and Planetary Sciences, 32:€151–185.
White, E. L., Troester, J. W. and White, W. B. (1984). A comparison of sinkhole depth fre-
quency distribution in temperate and tropical karst regions, in Sinkholes:€Their geology,
engineering and environmental impact, B. F. Beck (ed.). Rotterdam:€Balkema, 65–73.
White, S. (1995). Soil erosion and sediment yield in the Philippines, in Sediment and Water
Quality in River Catchments, I. D. L. Foster, A. M. Gurnell and B. W. Webb (eds.).
Chichester:€Wiley, 391–406.
372 References

Wilford, G. E. and Wall, J. R. D. (1965). Karst topography in Sarawak, Journal of Tropical


Geography, 21:€44–70.
Williams, G. P. (1983) Palaeohydrological methods and some examples from Swedish fluvial
environments. I. cobble and boulder deposits, Geografiska Annaler, 65A:€227–243.
Williams, G. P. and Wolman, M. G. (1984). Downstream effects of dams on alluvial rivers,
U.S. Geological Survey Professional Paper 1286, Washington, DC.
Williams, M., Dunkerley, D., De Deckker, P., et al. (1998). Quaternary Environments.
London:€Arnold.
Williams, P. W. (1971). Illustrating morphometric analysis of karst with examples from
New Guinea, Zeitschrift für Geomorphologie, 15:€40–61.
Williams, P. W. (1983). The role of the subcutaneous zone in karst hydrology, Journal of
Hydrology, 61:€45–67.
Williams, P. W. (1987). Geomorphic inheritance and the development of tower karst, Earth
Surface Processes and Landforms, 12:€453–465.
Wilson, L. (1973). Variations in mean annual sediment yield as a function of mean annual
precipitation, American Journal of Science, 273:€335–49.
Wit, M. C. J. de (1999). Post-Gondwana drainage and the development of diamond placers
in western South Africa, Economic Geology, 94:€721–740.
Wohl, E. E. (1992). Bedrock benches and boulder bars:€floods in the Burdekin Gorge of
Australia, Bulletin, Geological Society of America, 104:€770–778.
Wohl, E. E. and Merritt, D. (2005). Prediction of mountain stream morphology, Water
Resources Research, 41, W08410, doi:10.1029/2004 WR003779.
Wohl, E. E., Ogden, F. L. and Goode, J. (2009). Episodic wood loading in a mountainous
neotropcial watershed, Geomorphology, 111:€149–159.
Wolman, M. G. (1955) The natural channel of Brandywine Creek, Pennsylvania, U.S
Geological Survey Professional Paper 271, Washington, DC.
Wolman, M. G. (1967). A cycle of sedimentation and erosion in urban river channels,
Geografiska Annaler, 49A:€385–395.
Wolman, M. G. and Gerson, R. (1978). Relative scales of time and effectiveness of climate
in watershed geomorphology, Earth Surface Processes, 3:€189–208.
Wolman, M. G. and Leopold, L. B. (1957). River flood plains:€some observation on their
formation, U.S. Geological Survey Professional Paper 282-C.
Wolman, M. G. and Miller, J. P. (1960). Magnitude and frequency of forces in geomorphic
processes, Journal of Geology, 68:€54–74.
Wood, A. (1942). The development of hillside slopes, Proceedings, Geological Association,
53:€128–40.
Woodroffe, C. (2003). Coasts:€Form, process and evolution. Cambridge University Press.
Woodroffe, C. D. (2005). Southeast Asian deltas, in The Physical Geography of Southeast
Asia, A. Gupta (ed.). Oxford University Press, 219–236.
Woodward, J. C., Macklin, M. G., Krom, M. D. et al. (2007). The Nile:€ evolution,
Quaternary river environments and material fluxes, in Large Rivers:€ Geomorphology
and management, A. Gupta (ed.). Chichester:€Wiley, 261–292.
World Commission of Dams (2000). Dams and Development:€A new framework for deci-
sion making. London:€Earthscan.
373 References

World Resources Institute (1994). World Resources 1994–95. New York:€Oxford University
Press.
World Resources Institute (1996). World Resources 1996–97:€ The urban environment.
New York:€Oxford University Press.
Yang, D., Li, X., Ke, X., et al. (2001). A note on the troughs in the Three Gorges channel
of the Changjiang River, China, Geomorphology, 41:€137–142.
Young, A. (1972). Slopes. London:€Longmans.
Young, R. W., Wray, R. A. L. and Young, A. R. M. (2009). Sandstone Landform. Cambridge
University Press.
Index

aa, 257 Arabian Peninsula, 3


Aalto, R., 44, 148, 249, 349 Arabian Plate, 23
Aberdares, 234, 291 Arabian Sea, 25, 28, 53, 267, 293
abrasion, 224, 241, 277, 281 Araguaia, 90, 315, 317, 361
Aceh, 17, 184, 193, 361 Arequipa, 339, 369
Adriatic, 281 arête, 241
Africa, 3, 5, 6, 20, 22, 23, 28, 31, 33, 34, 37, 38, 44, Argentina, 236, 250, 290
49, 53, 55, 81, 119, 121, 143, 192, 209–211, 223, Aricha, 205
227, 229, 234, 251, 254, 266, 272, 288, 291, 292, Arichaghat, 159
312, 319, 328, 343, 353, 356, 358, 363, 367, 370 aridity index, 209
Africa, East, see€Africa Armero, 264
African Plate, 13, 23 Arunachal Pradesh, 159
Ahaggar, 23, 212 Ascension, 16, 254
Ahagger, 23 ash cloud, 258, 265
Ahmad, R., 330 ash fall, 259
Alaknanada, 156 Asia, 6, 23–25, 28, 33, 43, 44, 55, 56, 166, 192, 197,
Alice Springs, 215, 217, 219 223, 282, 284, 287, 293, 295, 296, 309, 312, 319,
alluvial fans, 9, 22, 25, 91, 129, 137, 140, 213–216, 351, 352, 354, 355, 357–359, 363–365, 369–370
231, 232, 247, 249, 250, 253, 329, 340, 345, 346, Assam, 80, 159, 161–163, 356
356 Assam Plains, 159, 162
alluvial valleys, 4, 10, 26, 130, 247, 289, 298 Aswan Dam, 143
Altiplano, 21 Atacama, 210, 211, 262
Amazon, v, 3, 6, 9, 20, 21, 27, 52, 56, 79, 104, 109, Atlantic, 6, 16, 22, 27, 28, 34, 39, 40, 42, 52, 144,
129, 131, 138, 143–152, 158, 169, 173, 182, 200, 146, 148, 150, 154, 182, 184, 190, 200, 210, 236,
251, 252, 264, 292, 293, 342, 343, 354, 355, 252, 254, 287, 289, 292, 344, 362
359–364, 368 Atlas Mountains, 22
Amazon Cone, 151 Atlas of Urban Geology, 340, 355
anabranching river, 115, 118 atolls, 190
anastomosing river, see€anabranching river Atrai, 205
Anatolian Plate, 19 Attock, 247
Andes, 6, 7, 17, 20, 21, 27, 52, 55, 56, 79, 83, 85, 92, attrition, 180, 224
96, 98, 109, 129, 144, 146–149, 151, 201, 210, 232, Auranga, 5, 123, 357
234–236, 243, 247, 249–252, 287, 288, 290, 291, Australia, 3, 5, 6, 26, 33, 44, 53, 55, 56, 65, 74,
340, 352, 358, 362 95, 121, 123, 124, 170, 190, 192, 209–211, 215,
andesite, 17, 64, 71, 259 217–223, 226, 227, 230, 249, 250, 291–297, 319,
Annamite Mountains, 165, 166 343, 350, 351, 354, 359, 364, 366, 367,
Annapurna, 88, 235 369–372
annual maximum series, 112 avalanches, 246
Antarctica, 5, 6, 20, 143, 287, 344
antecedent rainfall, 87, 88, 100 backswamps, 136, 137, 151, 167, 198, 199
anthropogenic modification, 5, 11, 29, 309, 347 backwash, 177, 182
antidunes, 133, 134 Bagnold, R.E., 103, 107, 223–225, 349
Antisana, 290 Baja California, 189
Appalachian Plateau, 281 bajada, 215
Appalachians, 66 Bale Mountains, 291
aquaculture, 207 Baltoro, 235

374
375 Index

Bandung, 330, 338, 339 Brahmaputra, 3, 8, 23–25, 27, 28, 80, 106, 109, 137,
Bangkok, 330, 335–337, 340, 355, 362, 364 143–145, 147, 150, 152, 155, 157, 159–163, 169,
Bangladesh, 37, 137, 157, 159, 162, 163, 169, 203, 182, 189, 197–199, 203–208, 247, 251, 252, 293,
205, 206, 350 299, 301, 350–352, 356, 360, 367, 371
bankfull discharge, 112 braided rivers, 116, 137
bar, 116, 134, 136, 137, 142, 150, 158, 163, 168, 202, Brandywine Creek, 111, 372
206, 248, 268, 323, 367 Brazil, 46, 52, 55, 56, 90, 96, 148, 150, 151, 200, 266,
Barbados, 22, 272, 292, 296, 350 314, 315, 317, 354, 355, 361, 362, 364
Barind, 204, 205 Brazilian Shield, 20, 22, 27, 56
Barisan Mountains, 17 Brice, J.C., 116, 163, 351
barrier island, 184 Bruijnzeel, L.A., 55, 247, 351, 354
barrier reefs, 190 Brunsden, D., 246, 252, 351
basal sliding, 238–241 Bryan, K., 216, 351
basalt, 6, 16, 17, 25, 62, 70, 71, 81, 86, 153, 179–181, Buchanan, F., 7, 78, 351
254–256, 267, 268, 270 Büdel, J., 8
base level, 109, 139, 140, 216, 293, 297 Bukit Timah Basin, 332, 352
Baspa, 235 Bukittingi, 260
Bassac, 165 Burbank, D.W., 24, 88, 249, 351, 355
Batoka Gorge, 153 Burdekin, 124, 372
Batura, 235
bauxite, 70, 81 Cahora Bassa, 152–154, 323
Bay of Bengal, 25, 28, 33, 155, 158, 163, 173, 182, Calcite, 67, 69, 270
184, 192, 205, 247, 293, 299 calcrete, 77
beach, 17, 41, 167, 171, 172, 175, 177–179, 181–184, caliche, 69, 78
186, 187, 189, 192–194, 199, 200, 202, 203, 223, California, 19, 89, 351, 362
251, 309, 312, 323 Cambodia, 122, 164–167, 268, 317
beach drifting, 182 Canning Hill, 331
beach ridges, 184, 199, 203 capacity, 107
beach rock, 183 Cape Fold Belt, 22
bed load, 79, 80, 87, 104–107, 113, 116, 118, 127, Cardamom Mountains, 165
132, 143, 158, 203, 251, 268, 311, 324 Caribbean, 9, 13, 19, 22, 28, 40, 42, 55–56, 90, 96,
bed material, 108, 116, 132–134, 158, 169, 316 121, 122, 124, 183, 190, 192, 236, 257, 272, 282,
Beira, 153 288, 333, 349, 355, 361
Belize Chamber, 280 Carmen de Uria, 340
Bengal Basin, 203, 207, 366 caverns, 270, 280
Bengal Deep Sea Fan, 205 caves, 69, 75, 181, 270, 271, 274, 275, 277, 279–281,
Bengal Fan, 207, 300, 301 283, 285, 296, 364
Beni, 44 Cayambe, 290
bergschrund, 241 Central America, 9, 22, 28, 52, 55–56, 210, 215, 249,
berm, 182, 183 312, 319, 343, 355, 364
Best, J. 162, Central Cordillera, 21
Betwani Mountains, 19, 96 CFB, see€flood basalts
Bhagirathi, 156, 157 Chaco Plain, 250, 251
Biafo, 235 Chad, 254
Blackhawk landslide, 89 chambers, 280, 281
blind valley, 277 Chandina, 205
block and ash flows, 261, 265 Chang Jiang, 26, 28, 33, 111, 150, 197
Blue Mountains, 91, 108, 138, 139, 197, 248 channel geometry, 108, 109
Blue Nile, see€Nile channel network, 119
Blum, M. D., 344, 347 channel pattern, 109, 113, 115, 128, 136, 159, 167,
Bohol Island, 282 321
Bolivia, 44, 236, 250, 358 channel shape, 103, 109, 113, 124, 126
Borneo, 52, 234, 283, 301 channel size, 109, 111, 113, 120
bornhardt, 8 channel slope, 101, 108, 109, 345
Bornhardt, 7 channel-forming discharge, 112
Bourke, M., 217–222, 351 channel-in-channel topography, 124, 126
376 Index

Chao Phraya, 26, 197, 335, 369 Cotopaxi, 254, 290


chelation, 66 craton, 4–6, 10, 13, 19, 25, 29, 75, 145, 158, 210, 211,
chemical weathering, 61, 63, 66, 72, 79, 80, 89, 230, 229, 295, 299
see€also€weathering creep, 84, 85, 92, 93, 95, 96, 98, 100, 224, 238, 239
Chézy, 102 crescentic dune, 225
Chézy equation, 102 crevasse splays, 136, 137, 202, 203, 346
Chihuahuan, 210 critical erosion velocity, 105
Chimborazo, 254, 290 critical flow, 133, 134
China, 33, 42, 121, 140, 164–168, 179, 191, 197, 271, crusts, v, 26, 74, 77, 80, 81, 224
275, 278, 279, 282, 295, 299, 323, 350, 358, 359, Cuba, 22, 38, 41, 272, 279
373 Curah Lengkong, 263
Chittagong, 204 Curray, 301, 354
Chocolate Hills, 282 currents, 32, 44, 170, 172, 177–180, 195, 198, 207,
Cimarron River, 137, 366 209, 295–297, 368
cirque glacier, 237, 241 Cvijíc, J., 270
cirques, 241, 290, 291 CWD, see€coarse wood debris
clapotis, see€standing waves
clay minerals, 65, 69–72, 138, 150 d’Arcy-Weisbach equation, 102
climate change, 10–12, 78, 147, 149, 171, 206, dam, 10, 29, 111, 128, 154, 165, 198, 247, 248,
207, 288, 292, 293, 298, 299, 303, 323, 342, 343, 293, 318–324, 346, 347, 360, 372, 373
345–348, 351, 356, 362, 365, 369, 370 dambos, 152
climate proxies, 131, 292, 297 Dana, J.D., 7
climo-morphogenetic landforms, 4 Danum Valley, 55, 95, 312
coarse wood debris, 248 Dardi, 111, 124
coast, 9, 17, 25, 28, 29, 33, 34, 37, 41, 43, 44, 52, 98, Darjeeling, 36, 247, 352, 368
141, 150, 153, 170–173, 175–177, 179–184, 187, Darwin, Charles, 7
189–194, 196, 200, 205, 206, 209, 210, 226, 232, Davis, W.M., 13, 353, 369
233, 251, 257, 263, 281, 285, 308, 309, 312, 319, De Martonne, 8
324, 328, 336–337, 340, 347, 361, 370 Dead Sea, 211
coastal cliff, 170, 179, 180 debris avalanches, 90, 262, 339
Coastal Cordillera, 21 debris fall, 85, 86
coastal dunes, 186 debris flow, 90, 91, 98, 216, 246, 334, 353
cockpit karst, 278 debris slide, 85, 87, 97
Cocos, 6, 20 Deccan Plateau, 20, 254
Coleman, J.M., 8, 137, 162, 163, 352 Deccan Traps, 25
collision edge coasts, 172 DeCelles, P.G., 250, 251, 361
Colombia, 52, 263, 359 deflation, 224
Colorado, 322, 371 deforestation, 10, 29, 37, 58, 99, 147, 149, 246, 308,
Comilla, 205 309, 311–314, 317, 324, 334, 363
competence, 107 delta, 4–6, 28, 150, 153, 155, 157, 158, 163–167,
compressive flows, 239 173, 189, 195–198, 200–208, 245, 251, 288, 297,
conekarst, 278 299–301, 309, 311, 317, 335, 337, 342, 345, 346,
Congo, 28, 53, 143, 144, 150, 152 354, 356, 360, 368, 369, 371
Congo Air Boundary, 154 delta front, 198, 201–203, 205, 206
congruent dissolution, 67 delta plain, 201–203, 206
continental shelf, 20, 150, 195, 203, 206, 297, 360 Deodhar, C. A., 268
convergent boundary, 15 depression, 16, 24, 33, 36, 38, 130, 151, 152, 216,
convergent plate margins, 9, 99, 254, 339 237, 241, 245, 275–277, 283, 335
coral reef, 8, 10, 172, 182, 186, 187, 189–192, 194, desert varnish, 216
295, 309, 312, 354 Devprayag, 156
Cordillera Apolobamba, 236 Dhaulagiri, 235
Cordillera Blanca, 236 Dhubri, 159
Cordillera Huayhuash, 236 Dibang, 159
Cordillera Real, 236 Dietrich, W.E., 138, 353
corestones, 72, 73, 77, 95, 249 Dihang, see€Brahamputra
corrasion, 180 discharge, 24, 28, 29, 81, 101, 105–108, 110, 111,
corrosion, 180, 274, 276, 277, 279, 281 113–117, 119, 120, 122–127, 134, 136, 138,
377 Index

142–145, 147–149, 154–157, 160–163, 166, Ethiopia, 291, 292


195–197, 214, 217, 218, 222, 232, 248–251, 264, Etosha, 210
277, 280, 289, 292, 293, 312, 316–318, 320–322, Eurasia, 5, 6, 13, 19, 23, 165, 167, 364
326, 342, 345–347, 356, 357, 363, 364, 368 Eurasian Plate, 5, 13, 23
dissolution, 66–69, 77, 180, 274, 275, 296 Europe, 11, 112, 287
distributary channels, 153, 158, 195, 198, 202, 206, evapotranspiration, 3, 45–47, 50, 209
346 exfoliation, 64
divergent boundary, 15, 20 experimental flood, 323
dolerites, 86 experimental plot, 310
doline, 275–278, 280–283 extending flows, 239
dolomite, 69, 274
dominant discharge, see€channel-forming discharge fall velocity, 105, 106
Dominican Republic, 277 falls, 23, 32, 33, 35, 38, 39, 43, 47, 49–51, 85, 87, 89,
Dorr, J. A., 251 92, 105, 111, 113, 154–157, 177, 180, 207, 212,
Douglas, I., 8, 51, 58, 247, 312, 314, 337, 338, 354, 227, 258, 263, 264, 314, 316
356, 367 fan-delta, 28, 29, 197, 371
Drakensberg, 254, 266, 291 Faniran, A., 65, 67
drips, 54 feldspar, 65, 67–69, 71, 73, 80, 227
dune, 17, 26, 41, 133, 134, 171, 172, 183, 185–187, fengcong, 282
192, 193, 199, 202, 203, 207, 210, 211, 223, 233, fenglin, 282, 359
251, 252, 261, 292, 293, 295, 325, 350, 360 Ferrel Cell, 4
Dunne, T., 45–48, 82, 84–87, 94, 131, 138, 147–151, ferricrete, 77
313, 314, 324, 325, 349, 354, 362 fetch, 173
Duplessy, J.C., 293, 354 Finke, 217
duricrust, 77, 80 firewood, 58, 311, 312
firn, 236, 241
earthflows, 91, 246 firn line, 236
earthquake, 8, 16, 17, 19, 24, 29, 85, 96, 161, 163, flocculation, 187, 198
197, 232, 247, 251, 262, 328, 333, 334, 337, 366, flood, 19, 22, 31, 34, 35, 37, 38, 41, 42, 44, 47, 98,
367 105–107, 112, 113, 119, 121, 123–127, 132, 133,
East African Rift System, 23 137, 138, 142, 147, 150, 155, 157, 159, 160, 163,
East China Sea, 33 166, 167, 189, 196, 197, 207, 213, 214, 216–220,
easterly waves, 35, 39, 42 222, 223, 240, 246–248, 250, 267, 268, 277, 280,
Eastern Cordillera, 21 283, 287, 293, 294, 322, 326, 327, 331–333, 335,
Eastern Desert, 210 338, 340, 342, 350, 353, 356, 357, 359, 365, 372
Eastern Ghats, 25 flood basalt, v, 20, 71, 254, 266, 269
Eastern Syntaxis, 80, 159, 160, 162, 252 flood diversions, 333
Ecuador, 52, 256 floodout, 218, 223
Egypt, 110, 189, 210, 319 floodplain, 112, 117, 125, 132, 136–142, 147,
Eilat, 211, 213, 214, 356 149–153, 155, 158–160, 167, 219, 221–223, 247,
El Misti, 256 307, 320, 331, 332, 345, 353, 354, 359, 362, 366
El Niño, 27, 34, 43, 44, 315, 349 flows, 19, 24, 27, 33, 44, 45, 47, 55, 63, 66, 69, 80,
El Niño Southern Oscillation, see€El Nino 90–92, 94, 95, 97, 98, 100, 105, 108, 117, 121,
Elephant Hills, 165 123–128, 131, 133, 137, 138, 144, 145, 147, 148,
Elm, 89 151–153, 157–159, 162–167, 177, 178, 195, 198,
Emiliani, C., 288 202, 205, 206, 214, 216, 217, 222, 239–242, 245,
Emmel, F.J., 301, 354 247, 249, 251, 256, 257, 259–268, 270, 272, 281,
England, 281 299, 316, 317, 327, 334, 335, 339, 340, 345, 352,
Enriquillo-Plantain Garden Fault, 19 353
ENSO, see€El Nino fluid threshold, 224
Enzel, Y., 295, 354, 359 Fly River, 138, 203, 353
ergs, see€sand sea Ford, D.C., 270, 271, 273–275, 277, 279, 280, 355
erodibility, 49, 51, 75, 110 forest litter, 66, 68
erosion plot, 309 form ratio, see€width-depth ratio
erosion rate, 19, 24, 162, 309 Fournier, F., 8, 51, 355
erosional threshold, 87 Francis, P., 255, 256, 259–264, 266, 355
erosivity, 49 French Guiana, 200
378 Index

freshwater swamp, 52, 195 Great Rangit, 247


friction factor, 102 Greater Antilles, 19, 22
fringing reefs, 189 greenhouse gases, 343
Froude Number, 133 Greenland, 143, 344
Grodek, T., 283
Gabet, E., 88, 94, 96, 100, 249, 355 groundwater, 45–47, 160, 186, 268, 271, 272, 280,
Galapagos, 256 330, 333, 335, 337, 338, 341, 350
Galunggung, 262 Gua Nasib Bagus, 283
Gandak, 137, 157 Guanxi, 282
Ganga, 3, 20, 25, 27, 28, 57, 99, 109, 125, 137, 138, gufeng, 282
140, 143–145, 147, 150, 152, 155–163, 166, 169, Guilin, 282
182, 189, 197–199, 203–208, 251, 252, 288, 293, Guinea Current, 200
297–301, 337, 342, 344, 353, 356, 367, 371 Guizhou, 278, 359
Ganga Sagar, 158 Gulf of Aden, 23, 292
Ganga–Brahmaputra Delta, 3, 161, 173, 197–199, Gulf of Mexico, 122
203, 204, 208 Gulf of Thailand, 336
Gangamopteris, 6 Gunung Api, 283, 371
Gangotri Glacier, 156 Gunung Sewu, 282
Garden of Eden, 283 Gupta, A., 8, 10, 40, 58, 84, 110, 115, 117, 119,
Gardner, T.W., 9, 156 121–125, 131, 144–146, 148, 149, 153, 155, 157,
Gaumukh, 156 158, 161, 164–167, 193, 199–201, 218, 220–222,
Gendol, 265, 266 224, 247, 248, 266–268, 284, 296, 298, 302, 309,
Ghaghara, 157 313, 315–317, 323, 330, 332, 336, 349–351,
Gibbs, R. J., 157 353–359, 361–365, 367, 369, 370, 373
gibbsite, 65, 69, 71, 81 Guyana Shield, 21, 27
Gilbert, G.K., 202 Gyala Peri, 162
Gillieson, D., 282, 284, 355
glacial mass balance, 237 Ha Long Bay, 284, 285
glacial mass budget, see€glacial mass balance Hadley Cell, 4, 32
glacial outwash, 118, 245 Haiti, 19, 296, 354, 366
glaciated trough, see€glaciated valley Hawaii, 70, 256, 257
glaciated valley, 242, 243 Hell, 282, 355
glacier, 85, 235–244, 288–291, 299, 300, 342, 344, Himachal Pradesh, 140, 235, 365
345, 364, 371 Himadri, see€Great Himalaya
glacier surge, 240 Himalaya, 5, 6, 11, 16, 20, 23–25, 28, 33, 52, 80, 83,
Glen Canyon Dam, 322, 371 85, 88, 96, 98–100, 109, 144, 155–163, 165, 168,
Glen’s Flow Law, 238, 239 197, 232, 234–236, 243, 246, 247, 249–253, 287,
global warming, 342–344, 348 288, 290, 291, 299, 345, 349, 351–353, 355, 366,
Globorotalia menardii, 288 368, 370–371
Glossopteris, 6 Himalayan foreland, 156, 299
Godavari, 25, 115, 140, 198 Himalayan Frontal Fault, 23
Gomati, 157 Hispaniola, 22
Gondwana, xi, 5, 6, 20, 23, 26, 99, 152, 154, 353, 363 Hispar, 235
Gondwanaland, see€Gondwana Hjulstrom, F., 105, 106
Good Luck Cave, see€Gua Nasib Bagus Holocene, 5, 57, 85, 147, 151, 167, 197, 200, 203,
Goodbred, S.L., Jr., 145, 163, 197, 204, 206, 207, 205, 207, 285, 289, 290, 292–297, 301–303, 311,
293, 297, 299–301, 344, 345, 356, 360 350, 354, 356, 359, 361, 363, 364, 368, 369
Goswami, D.C., 162, 163, 246, Hong Kong, 8, 72, 96, 338, 340, 350, 360, 365
Goudie, A., 26, 77, 211 hook, see€recurved spit
graded profile, see€profile of equilibrium Hope, G., 291, 334, 358
Grand Canyon, 321, 322, 371 horns, 187, 242, 243
granite, 8, 17, 27, 63, 66, 72, 73, 81, 179, 180, 212, Horton, R.E., 8, 21, 93, 119, 120, 250, 251, 358
270, 333, 365 Horton overland flow, 93
granodiorite, 71 Huang He, 198
Great Barrier Reef, 190, 296, 352, 364 Huascaran, 247
Great Escarpment, 22, 210 Hughli, 157, 158
Great Himalaya, 23, 24 Humboldt, Alexander von, 7
379 Index

hurricane, 9, 22, 36, 39, 42, 58, 89, 90, 97, 98, 100, Java, 17, 57, 81, 92, 95, 183, 190, 196, 232, 233, 247,
124, 139, 194, 196, 197, 248, 333, 334, 337, 344 252, 254, 257, 258, 262, 263, 265–267, 271, 278,
Hurricane Flora, 38, 41, 139 282, 291, 292, 301, 313, 328, 339, 361, 369, 370
hydration, 68, 75 Java Trench, 17
hydraulic action, 180 Jeje, K. K., 22, 210
hydraulic geometry, 113–115, 128, 361 jet streams, 34
hydrolysis, 68 Johore, 330
hyperconcentrated flow, 84, 216
Kafue, 323
Iceland, 16 Kailash, 159
ICOLD, see€International Commission on Large Kailas-Mansarovar, 24
Dams Kalahari, 22, 210, 226, 227
ignimbrite, 260 Kale, V.S., 22, 210
IKONOS, 57, 112, 193, 208, 257, 265, 267, 285, 286, Kaliadem, 265, 266
318, 369 Kalimantan, 202, 354
impact threshold, 224 kame delta, 245
imperviousness, 325 kames, 245
impoundments, 23, 143, 147, 308, 319, 323, 347, 371 Kanchenjunga, 235
incongruent dissolution, 67 Kangning, 278, 359
India, 5, 6, 20, 23, 28, 33, 36–38, 50, 53, 70, 77, 80, kankar, 77
110, 111, 115, 124–126, 140, 142, 157, 159–161, kaolinite, 65, 68, 69, 71, 80, 81
163, 169, 173, 184, 188, 203, 205, 210, 249, 254, Kar, A., 24
261, 266, 267, 272, 292, 295, 330, 351, 353, 354, Kariba, 153, 154, 323
356, 358–360, 364, 366, 367, 369, 371 Karnali, 24
Indian Ocean, 5–7, 13, 17, 23–25, 28, 33, 34, 40, 42, karren, 275
43, 56, 121, 123, 152, 153, 157, 158, 167, 181, 183, karst, 9, 22, 270–279, 281–286, 337, 359, 371, 372
204, 205, 211, 232, 250, 293, 299, 350, 354, 360, karst gorge, 277
363–365 karst, tropical, 8
Indian Peninsula, 5, 6, 23, 25, 33, 56, 123, 157, 158, Kashmir, 234, 235
181 Kathmandu, 338, 351
Indian Plate, 165 Kaveri, 25, 126
Indian–Australian Plate, 17 Kelantan, 184
Indochina, 33, 34 Kelud, 263
Indo-Gangetic Plain, 23–25 Kerala, 184
Indonesia, 17, 26, 28, 33, 43, 57, 64, 70, 81, 92, 202, Kershaw, A.P., 295, 297, 360, 372
234, 249, 257, 258, 261, 262, 266, 267, 271, 278, Khartoum, 122, 350
291, 295, 313, 314, 349, 350, 361, 369, 370 Khudi Khola, 88
Indus, 23–25, 108, 143, 144, 147, 247, 249 Kilauea, 256
Indus–Tsangpo suture, 159 Kilimanjaro, 52, 234, 254, 291, 364
inselberg, 8, 22, 73, 216, 307 King, L.C., 8, 364
Inter Tropical Convergence Zone, 32, 147, 154, 232 Kingston, 328, 330, 333–335, 337, 338, 359, 361, 370
internal deformation, 238, 239 Kinta Valley, 282, 285
International Commission on Large Dams, 319 Kittipong, 314, 371
IPCC, 342, 343 knickpoint, 111
Ipoh, 282 Kolkata, 328, 329, 337
Irrawaddy, 26, 28, 109, 116, 117, 126, 128, 143–145, Kolohai, 235
197, 207, 208 Komar, P.D., 172, 173, 175, 184, 360, 368
ITCZ, see€Inter Tropical Convergence Zone Kong, 72, 166, 350
Itezhi-tezhi, 323 Köppen, W., 31
Köppen–Geiger classification, 31
Jacks Hill, 334 Korat, 165, 166
Jakarta, 190, 328, 329, 337 Kosi, 137, 157, 250, 356, 367, 371
Jamaica, 22, 36, 38, 42, 91, 97, 108, 115, 124, 138, Kota Gadang, 260
139, 192, 197, 248, 272, 278, 328, 330, 333, 335, Krakatau, 25, 140
349, 357, 359, 361, 362, 366, 368, 370, 371 Krishna, 25, 140
Jamuna, 137, 142, 159, 163, 206, 350 Krynine, P.D., 8, 360
Japan, 295 Kuala Lumpur, 337, 338, 354
380 Index

Kuehl, S.A., 150, 197, 203–207, 293, 356, 360, 363 Lumajang, 263
Kuta Tjuta, 34, 44 Lunkaransar, 295
Luquillo Mountains, 248
La Niña, 34, 44 Luzon, 43, 92, 259, 261
Ladakh, 235
lag time, 212, 325–327 maars, 255
lagoon, 182, 184, 185, 190 MacDonnell Ranges, 217
lahar, 92, 98, 246, 247, 251, 259, 261–265, 269, 339, Mackin, J.H., 109, 362
363, 369 Madagascar, 6, 23, 34, 36, 43, 53, 121, 272, 356
Lake Eyre, 217, 218 Madeira, 44, 149, 150
Lake Powell, 323 Madhupur, 204–206
Lake Tanganyika, 23 Magdalena, 27, 143, 144
Lake Titicaca, 290 magnitude-frequency concept, 120, 126–128
Lancaster, N., 224, 225, 227–229, 360 Mahakam Delta, 202
land use, 10, 11, 29, 119, 120, 147, 207, 213, 214, Main Central Thrust, 23
307–309, 311–313, 317, 324, 326, 331, 332, 352, Main Range, 20, 232
354, 356, 357 Majuli Island, 162
landslide, 8, 81, 96, 309, 337, 361, 362, 366 Makalu, 235
Langkawi, 285 Makgadikgadi, 153, 210
Lao PDR, 112, 122, 164–168, 268, 271, 315, 317, 318 Malacca Strait, 302, 354
lapiés, 275 Malay Peninsula, 20, 37, 182, 184, 232, 250, 285,
Larsen, M.C., 19, 89, 97, 98, 100, 340, 361, 367 301, 331
Last Glacial Maximum, 150, 290 Malaysia, 37, 55, 57, 95, 179, 181, 279, 280,
laterite, 7, 78, 299, 366 282–284, 288, 302, 312–314, 349, 350, 354, 358,
Latrubesse, E., 152, 315, 361 360, 367
lava, 16, 25, 256, 257, 260, 261, 266, 268 Mameyes Basin, 19
Lavigne, F., 339, 369 Mameyes River, 97
Leier, A. L., 249, 361 Mana Pools, 153
Lembang, 339 Manasalu, 235
Leopold, L.B., 8, 45–48, 82, 84–87, 94, 110, 112–114, mangroves, 10, 52, 172, 186–190, 194, 195, 198, 199,
116, 117, 120, 137, 138, 214, 325–327, 354, 361, 202, 203, 206–208, 285, 295, 309, 312, 336
372 Manila, 261, 329
Lesser Antilles, 22 Manizales, 340, 352
Lesser Himalaya, 23, 24 Manning’s equation, 102, 103
levees, 91, 124, 136, 137, 151, 195, 198, 199, 202, Manning’s n, 102, 103, 134
203, 337 Marajó Rift, 147
LGM, see€Last Glacial Maximum marginal sea coasts, 171
Lichty R. W., 137, 366 marine oxygen isotope series, 289
Liew, S.C., 17, 119, 124, 131, 167, 193, 266–268, Marmore, 44
358, 361, 369 Marsh, G. P., 307, 362
Liguanea, 334, 335 Martinique, 22
limestone, 12, 69, 75, 81, 180–182, 190, 268, 270, Maryland, 137, 357
271, 273–276, 278–280, 282–286, 296, 351, 364 mass movement, 83, 85, 246, 307
limestone pavements, 275 mass wasting, see€mass movement
Limpopo, 28, 154 Mauna Kea, 256
linear dune, 16, 20, 37, 82, 130, 136, 172, 176, 184, Mauna Loa, 256
187, 188, 201, 225, 228, 245, 275 Mauritius, 62, 254
lined canal, 333 Mavis Bank, 97, 138
llanos, 55 Meade, R.H., 27, 28, 109, 129, 138, 143, 146,
logging, 51, 311–313, 354 148–150, 154, 166, 182, 232, 251, 252, 354, 359,
Lohit, 159 362, 363, 371
Lombok, 64, 268, 291 mechanical weathering, see€physical weathering
long profile, see€longitudinal profile Mediterranean, 51, 66, 212
longitudinal profile, 109–111, 242, 320, 324 Medlicott, H.B., 6
longshore currents, 177 megacities, 328
longshore drifts, 184 megafans, 157, 232, 249–251, 253, 299, 300, 358,
lower flow regime, 134 361
381 Index

megalakes, 293 Nahal Yael, 211–214


Meghna, 156, 157, 205, 299, 350 Nairobi, 328
Megi, 37 Nam Ngum, 166
Mekong, 5, 26–28, 109, 111, 112, 119, 122, 124, 127, Nam Theun, 166
128, 131, 142–145, 152, 164–169, 197, 199, 200, Namche Barwa, 23, 80, 159, 162, 235, 252
202, 268, 315–319, 323, 357, 358, 360, 362, 363, Namib, 22, 64, 209, 210, 226, 229, 370
368 Nanga Parbat, 23, 24, 235, 236, 247, 249
Mekong River Commission, 122, 166, 362 Nankun, 235
Merapi, 92, 183, 232, 233, 247, 252, 261, 262, 264, Nargis, 207, 208
265, 267, 269, 339, 369 Narmada, 25, 109, 111, 115, 124, 125, 133, 138, 293,
Merritt, D., 248, 372 358, 359
Mertes, L.A.K., 131, 138, 147–151, 354, 362 Nazca, 6, 13, 20, 147
Mexico, 22, 34, 35, 39, 52, 55, 189, 210, 215, 312, nebkhas, 228
329, 330, 337, 338, 343 Negev, 212
migration, 138, 151, 163, 296, 297, 299, 308, 311, Negro, 150, 152, 355
312 Neogene, 23, 287, 333
Miller, J. P., 120, 125, 128, 214 neotectonics, see€tectonism
Milliman, J. D., 158 Nepal, 88, 100, 235, 246, 249, 252, 351, 355, 365
Milliman and Meade, 9, 27, 28 nested discharges, 113
Milliman and Syvitski, 9, 12, 27, 28, 30, 146, 158, Nevado del Ruiz, 263, 264, 341, 371
232 New Guinea, 8, 19, 28, 52, 53, 56, 96, 138, 249, 271,
MIS, see€marine oxygen isotope series 274, 275, 283, 291, 295–297, 353, 359, 367, 372
Mississippi, 137, 145, 150, 198 New Zealand, 11, 254, 349
mogotes, 279 Ngonye Falls, 153
Mohr and Van Baren, 7 Niekerk, A.W., 119, 363
Molengraff, G. A. F., 301, 363 Niger, 28, 143, 144, 169, 202, 226
monsoon, 5, 24, 25, 33–35, 37, 47, 48, 50, 55, 88, Niger Delta, 196
100, 122, 123, 125, 145, 154, 156–160, 166, 167, Nigeria, 50, 360
184, 196, 198, 203, 205, 207, 250–252, 262, 267, Nile, 23, 28, 110, 122, 143, 144, 169, 198, 252, 293,
282, 287, 293, 295, 299, 300, 315, 317, 335, 337, 302, 319, 350, 362, 373
343, 353, 356, 357, 359, 365, 367, 371 nipa palm, 202
montane streams, 248, 253, 349 Noakhali, 205
Montserrat, 268 non-equilibrium rivers, 217, 230
Moore, A. E., 90, 244, 245 non-stratified drift, 244
moraine, 90, 244, 245 North American Plate, 8, 13, 19, 75, 226, 333
moraine ridge, 244 North Anatolian Fault, 19
Morakot, 38, 41 North Sunda River, 301
Mosi-a-tunya, see€Victoria Falls Northern Australia, see€Australia
Mount Badda, 291 Nossin, J.J., 92, 259, 339, 349, 364
Mount Elgon, 234, 291 Nyamuragira, 254
Mount Everest, 16, 23 Nyiragongo, 254
Mount Jaya, 234, 291 Óbidos, 150
Mount Kenya, 234, 291
Mount Kinabalu, 234, 288, 291 Okavango Delta, 153, 154, 210, 362
Mount Mayan, 211 Old Alluvium, 302, 358
Mount Sodom, 211 organic acids, 66, 68, 271, 282, 283
mountain-building, 19, 20 Orinoco, 21, 27, 28, 52, 56, 79, 80, 104, 143–146,
Mozambique, 152 148, 150, 182, 200, 201, 252, 359, 362, 368, 371
mudcapes, 201 overland flow, 45, 55, 95, 150, 335, 367
mudflows, 91, 92, 98, 246, 259, 262, 263
mudslide, 22, 85, 87 Pacific, 7, 13, 17, 19, 21, 28, 33, 40, 41, 43, 44, 48,
Mulu Hills, 283 52, 55, 56, 183, 190, 191, 256, 287, 292, 353, 366,
Mulu National Park, 280, 283, 284, 351, 364 369
Murray Darling, 293 Pacific Plate, 13, 19
Myanmar, 33, 56, 116, 117, 121, 128, 160, 164, 208, Padma, 143, 157
271 pahoehoe, 257, 258
382 Index

Pakistan, 46, 210, 249, 366 quartz, 65, 68, 71, 73, 75, 80, 138, 142, 150, 158,
palaeokarst, 270 223, 359
Palmer, A. N., 248, 287 Quaternary, 10, 12, 23, 24, 131, 140, 151, 152, 163,
Panama, 248, 287 165, 167, 168, 186, 191, 207, 209, 229, 230, 283,
Pangaea, 5, 6, 23 287–290, 293, 298, 299, 302, 303, 335, 336, 338,
pans, 77, 81, 89, 152 344, 345, 347, 350, 352, 353, 356, 358, 359, 362,
Pantanal, 56 364, 366, 368–373
Papua New Guinea, 96, 291 Queensland, 53, 295, 352, 360, 369
parabolic dune, 228 Quelimane, 153
Paraná, 143–145, 266
Parangtritis, 183, 252 Raikot Glacier, 236
Passarge, 7 raindrop, 49, 54, 55, 93
Patton, P. C., 222 rainfall intensity, 35, 49, 94, 97
peat, 52, 203, 206, 293 Rajasthan, 77, 287, 295
pediment, 82, 215, 216, 223, 307 Rajmahal, 157
Peltier, L., 8 ranching, 312
Peru, 44, 181, 256, 340, 369 rapids, 131, 132, 134, 153, 165, 167, 322
Philippines, 26, 28, 36, 38, 42, 43, 92, 121, 249, 256, Rapp, A., 37, 307, 308, 365, 369
257, 259, 261, 262, 278, 282, 315, 355, 363, 372 Raub al Khali, 3
Phnom Penh, 165 recurved spit, 184
phyllosilicates, 64 Red River, see€Sông Hóng
physical weathering, 61, 63, 64, see€also€weathering Red Sea, 16, 23, 189, 212, 292
Pickup, G., 112, 123, 217, 218, 220–222, 301, 350, regolith, 73, 76, 79, 80, 82, 84, 87, 88, 91, 95, 96, 99,
351, 364 275, 276, 302, 309, 314, 326, 338
Pinatubo, 43, 92, 259, 262, 363 reptation, 223, 225, 228
pinnacles, 27, 283, 296 Réunion, 254
Pitts, J., 88, 100, 333, 357, 358, 364 Riau, 330
plagioclase feldspar, 68–71, 81 Richards, K. S., 183
planar slide, see€translational slide ridges and runnels, 183
plane beds, 134 Rieger, W.A., 112, 123, 364
plantation agriculture, 312 riffles, 116, 132, 133
plate tectonics, 4, 10, 13, 19, 27, 28, 145, 147, 172, rift valley, 3, 6, 16, 20, 144, 210, 247, 254
254 ring of fire, 17
platform, 19, 179, 180, 190, 351 Rio Chagres, 248
Platte River, 123, 367 Rio Cubuy, 97
playa, 215, 216 Rio Grande, 250
Pleistocene, 5, 10, 24, 105, 147, 190, 191, 203, Rio Parapeti, 250
205–207, 217, 234, 235, 243, 252, 285, 288–290, Rio Pilcomayo, 250
292, 295, 296, 298, 303, 350, 354, 359 ripples, 133, 134, 136, 217, 225, 227–230, 283, 366
Plinian eruption, 258, 259 River Red Gum, 221
plucking, 241 rock avalanche, 90, 260
plunging breaker, 177 rock desert, 210, 216
pocket valley, 278 rock fall, 85
polje, 277, 283 rock glide, see€rock slide
pollen, 292, 352 rock rib, 131, 142, 167, 168
polyzonal, 141, 142, 144, 298, 345 rock slide, 85, 87
pool, 111, 116, 132, 134, 136, 167, 248 rotational slide, 87, 89, 96
porous pavement, 333 Rufiji, 28
Port Royal, 328 Ruwenzori, 234
Potter, P.E., 143, 147, 232, 353, 365 Ruxton and Berry, 8, 72, 76
prodelta, 202, 203
profile of equilibrium, 109 Sabah, 55, 95, 234, 272, 288, 291, 312–314, 354, 363,
protalus rampart, 241 367
pseudokarsts, 270 Sabie, 119, 363
Puerto Rico, 19, 22, 43, 55, 71, 81, 89, 95–97, 100, sabkhas, see€salts flats
248, 272, 277, 279, 349, 361, 363, 367 SACZ, see€South Atlantic Convergence Zone
pyroclastic surge, 261 Saffir–Simpson Damage Potential Scale, 41
383 Index

sag ponds, 20 Senegal, 28, 198


Sagaing Fault, 116, 117 shear stress, 103, 107, 121, 124, 126, 127, 166, 238,
Sahara, 22, 34, 44, 210, 211, 223, 225, 226, 287, 239, 242, 264
291–293, 362 sheetflood, 94
Sahel, 44, 343 sheetwash, 94
Sahul Shelf, 295–297 shield, 19, 79
saline water intrusion, 337 shifting agriculture, 312, 314, 315
salt flats, 23, 172, 186, 187, 189, 207 shifting cultivation, see€shifting agriculture
saltation, 223–225, 228 Shillong Plateau, 161
Salween, 26, 28, 144, 168, 293 shore, 170–172, 176–179, 181–183, 187, 190,
Samosir, 261 192–194
San, 19, 46, 85, 86, 166, 350, 354, 361 shore platform, 180, 181
San Andreas Fault, 19 Shreve, R.L., 89, 119, 120, 366
sand sea, 211, 223, 225, 226, 228, 229 Shroder, J.F., 236, 246, 253, 366
sand sheet, 226, 228, 250 shutter ridge, 20
sandur, 245 Siachen, 235
Sâo Francisco, 198, 202 Sierra Madre Occidental, 210
Sâo Paulo, 328, 329 Sierra Madre Oriental, 210
Sapper, 7 Sierra Nevada de Santa Marta, 236
Sarawak, 272, 275, 278, 280, 283, 284, 302, 314, 351, Sierra Nevada del Cocuy, 236, 370
358, 364, 372 Sihang, see€Brahmaputra
Sarawak Chamber, 280, 283 Sikkim, 235
satellite images, 39, 193, 264, 265, 309, 315, 317 silcrete, 77, 89
Satluj, 24 Simen, 291
saturation overland flow, 93 Simonett, D.S., 8, 19, 96, 367
savanna, see€tropical grassland Simpang Formation, 302
Schick, A.P., 211–214, 217, 356, 365 Simpson Desert, 217
Schlesinger, W. H., 93 Simpson–Strzelecki Desert, 227
Schumm, S.A., 8, 50, 87, 113, 116, 117, 119, 130, Sinai, 212, 217, 226, 365
137, 361, 366 Singapore, 37, 88, 100, 265, 302, 330–333, 338, 341,
scoria, 255 352, 354, 357, 358, 363, 364, 369
scoria cones, 255 Singapore Meteorological Service, 331
sea arch, 181 Singh, S. K., 217
sea-floor cores, 288, 292 Sinha, R., 145, 158, 367, 369
sea level, 5, 17, 43, 111, 139, 144, 145, 150, 168, 171, sinkhole, see€doline
173, 181, 184, 186, 189–191, 195, 197, 200, 204, sinuosity Index, 116
207, 210, 282, 283, 285, 288, 289, 293, 295–299, Siwalik Hills, 23, 24
301, 303, 335, 336, 342, 344, 346, 354 slackwater deposit, 293, 294
sea surface temperature, 39, 292, 297, 343, 350 slides, 87, 90
seasonality, 52, 77, 113, 122, 123, 125, 127, 155, 157, slope failure, 19, 24, 31, 34, 37, 38, 42, 44, 76, 80, 84,
247, 273, 287, 344, 345 86, 88, 89, 96–100, 165, 180, 246, 249, 252, 253,
sediment, 3, 5, 6, 9–13, 17, 19–22, 24, 27–30, 34, 50, 291, 297, 312, 317, 323, 330, 333, 334, 339, 345
51, 58, 79–82, 87, 89–92, 94–102, 105–113, slope profile, 82
115–117, 119–122, 124–139, 141–152, 154, slumps, see€rotational slide
156–159, 162, 163, 165–167, 169–173, 177–179, smectite, 65, 69, 71, 81
182–184, 186, 189, 192, 194–198, 200–203, snow line, 236, 290
205–208, 215–217, 219, 232–234, 236, 238, Socompa, 262
239, 243–252, 260, 263, 264, 268, 270, 274, 277, soil, 6–8, 11, 29, 34, 38, 44–47, 49, 50, 53–55, 58, 61,
281, 288, 289, 291–295, 297–300, 307, 309–322, 62, 64, 66–68, 70, 71, 75–80, 88, 93–97, 122, 126,
324–328, 334, 345–347, 349, 350, 352–354, 356, 188, 207, 212, 268, 272, 273, 275, 307, 309, 314,
357, 359–363, 366–368, 371, 372 317, 325, 335, 351, 358, 367, 371
sediment plume, 10, 309 soil creep, 92
sediment splays, 198 soil pipe, 54, 95
sediment yield, 9, 10, 28, 29, 50, 51, 95, 247, 263, Solimões, see€Amazon
309, 310, 313–315, 324, 354, 357, 360, 371, 372 Solo, 196
seiches, 175 solution load, 104, 157
semiblind valley, 277 solution pit, 275
384 Index

Son, 157 subduction, 6, 16–18, 26, 71, 147, 181, 247, 251, 256,
Sông Hóng, 26, 168, 197 261
Sonoran, 210 subsidence, 32, 275, 328, 335–338, 362, 364
South Africa, 254, 266 sudd, 252
South America, 5–7, 20, 21, 27, 43, 55, 143, 181, 182, Sulawesi, 271, 291
210, 232, 249, 272, 290, 312, 364, 367 Sumatra, 3, 17, 18, 37, 57, 184, 193, 258, 260, 271,
South American Plate, 6, 13, 20 291, 301, 309, 331, 361
South Asia, 9, 23, 33, 44, 55, 121, 123, 192, 298 Sumbawa, 43, 261
South Atlantic Convergence Zone, 147 Sunda Plate, 17
South Atlantic Ocean, 6, 144 Sunda Shelf, 295, 296, 298, 299
South China Sea, 164, 166, 168, 179, 191, 197, 295, Sundarbans, 206
299, 350, 358 supercritical flow, 133, 214
South Equatorial Current, 297 surging breaker, 177
Southeast Asia, 5, 7, 10, 25, 28, 33, 38, 53, 55, 57, 73, Surinam, 200
95, 121, 124, 143, 164, 186, 191, 192, 196, 262, suspended load, 79, 80, 104–107, 113, 121, 127, 128,
264, 271, 282, 284, 288, 291, 292, 295–299, 302, 132, 158, 187, 203, 251, 252, 268, 311
309, 312–314, 317, 319, 336, 343, 350, 354–358, swash, 177, 182, 183
363, 364, 369–370, 373 Swatch of No Ground, 205, 206, 299, 301
specific surface, 63, 81 Sweetings, M., 271, 281
spilling breakers, 177 Sylhet, 205
spit, 182, 184 Sylhet Basin, 205
Spiti Valley, 140, 365 Syvitski, J.P.M., 9, 27, 196, 347, 363, 368, 371
splash erosion, 49, 54, 93–95
SPOT, 117, 315, 317 Taal, 261
spring, 20, 152, 268, 270 Taiwan, 36–38, 41, 43, 96, 122, 295
squall, 37 Tambora, 43, 261, 264
Srepok, 166 Tandon, S.K., 145, 369
SST, see€sea surface temperature Tangkuban Perahu, 339
stack, 179, 181, 192 Tanzania, 37, 307, 308, 365, 369
stalactite, 281 Tapi, 25, 115, 359
stalagmite, 281 Tapponier, P., 26, 165, 369
Stallard, R. F., 281 Tasikmalaya, 262
standing wave, 175 tectonic belts, 4, 10
Stanley, D.L., 197, 297, 368 tectonic terraces, 139
star dune, 225, 228 tectonics, see€tectonism
Starkel, L., 247, 368 tectonism, 9, 19, 27, 48, 149, 356, 362
steephead, see€pocket valley tephra, 258, 259, 262
stemflow, 54, 55, 95, 367 terraces, 129, 139–141, 151, 158, 165, 218, 246, 247,
Stevaux, J., 315, 361 249, 251, 268, 290
stone line, 78 Thailand, 112, 122, 131, 164–166, 271, 285, 295, 302,
storm sewer, 325, 326 309, 314, 355, 369, 371
storm surge, 41, 173 Thar, 210, 226, 295, 354, 371
storm, 29, 34, 37–39, 42, 47, 50, 63, 95, 97, 121, thermokarst, 270
126, 171, 173, 183, 184, 190, 192, 202, 212, 217, Thomas, D.S.,
247–250, 263, 272, 327, 331, 337, 338, 344, Thomas, M.,
345 Thornthwaite, C.E., 44–46, 48, 369
Strahler, A.H., 8, 46, 119, 120, 128, 368 Thouret, J.C., 262, 265, 266, 339, 369
straight channel, 116, 136, 162, 250 Three Gorges Dam, 111, 319
stratified drift, 244 throughfall, 54, 55, 95
stratovolcano, 232, 256, 264, 268 thunderstorm, 37, 49, 50
stream power, 103, 107, 121, 124, 126, 127, 149, 166, Tibesti, 23, 254
169 Tibet, 24, 159, 160, 162, 352, 364
streamflow, see€discharge Tibetan Plateau, 24, 25, 33, 159, 160, 162, 164, 166,
Strombolian eruption, 258 197, 287, 349, 355
sturtzstrom, 89 tidal bore, 173
subcritical flow, 133 tidal mudflats, 184
385 Index

tides, 154, 172, 173, 179, 183, 195, 198, 201, 203, Venezuela, 21, 75, 98, 340, 352, 359, 361,
205, 335 371
till, see€non-stratified drift ventifacts, 225
tillite, 6 Verstappen, H. Th., 255, 264, 268, 299, 370
Timor Leste, 296 Victoria Falls, 152, 153, 155
Tista, 163, 247, 251, 352 Vientiane, 127, 128, 165, 167
Toba, 261 Viet Nam, 121, 165, 271, 278, 279, 283–285, 302,
Todd, 217–223 317
Tombolo, 184 Viles, H.A., 66, 181, 282, 370
Tonlé Sap, 165 Virgin Islands, 38
Tooth, S., 218, 221, 370 Virunga Mountains, 254
TOPEX/Poseidon, 344 Voight, B., 263, 264, 341, 371
topple, 85, 86 volcanism, 10, 19, 23, 28, 30, 98, 254, 256, 268
Toricelli Mountains, 19, 96, 367 volcano, 17, 22, 43, 183, 190, 255–259, 261, 262,
Torres Sánchez, A.J., 19, 97, 98, 100, 361 264, 265, 268, 339, 371
tower karst, 278, 285 Vörösmarty, C.J., 319, 347, 368, 371
trade winds, 32, 33, 40, 42, 70, 154, 279, 293 Vulcanian eruption, 258
trailing-edge coast, 172
transform boundaries, 15 Wadi El Arish, 217
transform fault boundary, see€transform boundaries Wadi Mikeimin, 217
translational slides, 87 Wadi Watir, 217
transport-limited slopes, 79 Wallace, Alfred Russel, 7
Trengganu, 184 Walvis Bay, 266
trigger rainfall, 87, 88, 100 Warne, A., 197, 201, 297, 368, 371
tropical cyclone, 28, 35, 38–43, 47, 48, 84, 88, 95, 96, Wasson, R.J., 158, 295, 371
99, 105, 107, 121, 171, 173, 192, 194, 205, 207, water balance, 44–46
208, 247, 262, 315, 330, 331, 343, 346, 347 water budget, see€water balance
tropical disturbances, 34, 35, 37 Watts Branch, 137
Tropical Geomorphology Newsletter, 9 wave diffraction, 175, 176
tropical grassland, 55, 57 wave refraction, 175, 183, 184
tropical rain forest, 10, 50, 51, 53–55, 57, 58, 66, 93, wave-cut platform, 180
95, 100, 190, 311, 312, 354 waves, 41, 42, 89, 133, 134, 170–180, 183–185, 187,
tropical storm, 10, 22, 38, 40, 47, 98, 99, 121–123, 188, 190, 194, 195, 198, 201, 202
157, 166, 185, 186, 190, 197, 205, 207, 248, 333, weathering, 6, 8, 10, 11, 20, 31, 47, 61–73, 75–81, 87,
334 94, 98, 124, 138, 150, 157, 161, 170, 180, 223, 230,
Tsangpo, 24, 159, 162 267, 333, 334, 338, 352, 362, 367–370
tsunamis, 17, 171, 190 weathering-limited slope, 82
tuff rings, 255 Wells, N. A., 82
Tun Kul Cave, 280 Western Cape, 291
Turkey, 19 Western Cordillera, 21
typhoon, 40, 42, 196, see€also€hurricane Western Ghats, 25, 53, 267
White Nile, see€Nile
Uluru, 74 width-depth ratio, 112, 113, 117, 119, 124, 127
United States, 7, 34, 50, 112, 123, 184, 202, 215, 259, Williams, G.P., 8
277, 320, 325, 353 Williams, M., 8, 291, 292, 295, 297, 320, 372
upper flow regime, 134 Williams, P.W., 270–275, 277, 279, 280, 287, 372,
urban geomorphology, 325, 330, 339 373
urbanisation, 10, 51, 57, 307–309, 325–328, 330–332, Wilson, T.J., 18
335, 337, 338 Wohl, E.E., 124, 248, 358, 365, 372
uvalas, 277 Wolman, M.G., 111, 116, 117, 120, 125, 128, 137,
217, 320, 321, 325, 326, 361, 372
valley glacier, 237, 240 Woodroffe, C.D., 190, 191, 197, 202, 373
velocity, 39, 40, 50, 54, 87, 89, 90, 93, 101–107,
109–111, 113, 114, 123, 124, 127, 129, 133, 134, Xiaowan, 319
159, 173, 174, 176–178, 214, 224, 228, 239, 240,
242, 260, 261, 264, 272, 294, 316, 327, 364 Yallahs, 97, 115, 124, 138, 139, 371
386 Index

Yamuna, 157 Yucatan Peninsula, 22


Yangtze, see€Chang Jiang
yardangs, 225 Zambezi, 28, 143, 144, 152–155, 169, 323, 363
yazoo streams, 137 Zhujiang, 28, 144
Yogyakarta, 265, 338, 339 zibars, see€sand sheet
Fig 2.6 The Himalaya Mountains and Ganga Plains:€a tectonically active high relief environment. From NASA/MODIS
Fig 2.7 Northern Australia, an old arid landscape. From NASA/MODIS
Fig 4.8 Vegetation clearance for oil palm plantation in eastern Sumatra, Southeast Asia. Note lines of young oil palms towards
the upper left corner, canals put in for drainage and burning of vegetation in the centre. IKONOS satellite image
© Centre for Remote Imaging, Sensing and Processing, National University of Singapore (2006), reproduced with
permission

Fig 7.8 The Mekong eroding into quartzitic rocks along the border of Thailand with Lao PDR, upstream of Pakse, Lao, PDR.
Scale indicated by trees at the bottom left. IKONOS satellite image © Centre for Remote Imaging, Sensing and
Processing, National University of Singapore (2003), reproduced with permission
Fig 9.10 MODIS/NASA image of the Brahmaputra Valley and the Ganga–Brahmaputra Delta
Fig 10.18 Tsunami erosion and recovery, northern Aceh coast. (a) location of the image used for the Aceh coast; (b) coast on January 2003; (c) December 2004 (tsunami); (d)
February 2006; (e) January 2007; (f) April 2008. Note removal of the beaches, destruction of the wetland and erosion along swales. Within 13 months, beaches have
returned, although the wetland in the centre is under water. Vegetation has returned to the swales. The effect of the tsunami is not generally perceptible even in the
2006 image. The next two images show the continuation of the rebuilding process. IKONOS satellite images © Centre for Remote Imaging, Sensing and Processing,
National University of Singapore, reproduced with permission. From Liew et al., 2010
(a) (b)

Fig 11.9 The effect of a tropical cyclone on a major delta. A. Part of the Irrawaddy Delta. B. Same area after tropical cyc-
lone Nargis. IKONOS satellite image. © Centre for Remote Imaging, Sensing and Processing, National University of
Singapore (2009), reproduced with permission
Fig 14.2 The crater of the stratovolcano of Raung in east Java is about 2 km across. The volcano is high (3332 m) and erupts
regularly; it has already been active several times in the twenty-first century. Flow and explosive deposits and screes of
volcanic material mark the crater rim and the caldera floor. A second crater with a deep volcanic vent rises asymmetrically
from the floor of the caldera. IKONOS satellite image© Centre for Remote Imaging, Sensing and Processing, National
University of Singapore (2009), reproduced with permission

Fig 14.6 DEM of Merapi draped by the IKONOS image of 16 June 2006, showing the southern slope of the volcano and the
Gendol River Basin. A pyroclastic flow is emerging from below the billowing ash cloud. The river channel is marked by
deposition of earlier pyroclastic flows. From Thouret et al., 2010. By permission of Elsevier
Fig 14.7 Lower slopes of the DEM of the Gendol Valley showing details of a volcanic flow. Segment 1: upper section of the
valley; Segment 2: widening of the valley and a fan of volcaniclastic material; Segment 3: a wider valley entrenched
in a volcaniclastic fan. A: the moving pyroclastic flow from upstream; B: a ridge built by a previous flow; C: an
area stripped by an ash-cloud surge; D: spilling and re-routing of flow at a bend; E: avulsion of re-routed flow to a
neighbouring channel; F: lobes of flow reaching overbank which buried the village of Kaliadem (KA). From Thouret
et€al., 2010. By permission of Elsevier

Fig 15.7 IKONOS image of coast near Phang Nga, southern Thailand. Limestone towers are partially inundated by a Holocene
rise in sea level and surrounded by mangroves. Boats provide scale. Seawater frequently invades sinkholes in karst
giving rise to a spectacular landscape. IKONOS satellite image © Centre for Remote Imaging, Sensing and Processing,
National University of Singapore (2007), reproduced with permission

S-ar putea să vă placă și