Mass Transport, Gravity Flows, and Bottom Currents: Downslope and Alongslope Processes and Deposits 0128225769, 9780128225769

Mass Transport, Gravity Flows, and Bottom Currents focuses solely on important downslope and alongslope processes. It pr

912 92 145MB

English Pages 340 [601] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Mass Transport, Gravity Flows, and Bottom Currents: Downslope and Alongslope Processes and Deposits
 0128225769, 9780128225769

Table of contents :
Title-page_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Mass Transport, Gravity Flows, and Bottom Currents
Copyright_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Copyright
Dedication_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Dedication
Contents_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Contents
About-the-author_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
About the author
Professional preparation
Employment
Research
Publications
Global workshops on deep-water sandstone petroleum reservoirs
Awards, recognitions, and nomination
Philanthropy
Online resources for his publications
Transformation from a local science teacher to a global petroleum geologist
Science teacher
Motivations by TNM
Nature photographer
Reference
Preface_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Preface
Acknowledgments_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Acknowledgments
Parents, teachers, and benefactors
Early research (1965–78)
Mobil research on mass transport, gravity flows, and bottom currents (1978–2000)
Consultant research (2000–present)
Tsunamite research (2004–present)
Contourite research (1974–present)
Density plumes and hyperpycnite research (2002–present)
Sediment deformation and seismite research (1978–present)
Academic events (1968–present)
Editorial board member (2018–present)
Photographs
Elsevier
Copyrights and permissions
Editors and reviewers (1978–2020)
Logistics
ResearchGate (2014–present)
Wife and friend (1975–present)
The rocks (1962–present)
Chapter-1---Introductio_2021_Mass-Transport--Gravity-Flows--and-Bottom-Curre
1 Introduction
1.1 Why this book?
1.2 History
1.3 Universal case studies
1.4 Environments and processes
1.5 Objectives
1.6 Organization
1.7 Other aspects of the book
1.8 Synopsis
Chapter-2---Mass-transport--slides--slu_2021_Mass-Transport--Gravity-Flows--
2 Mass transport: slides, slumps, and debris flows
2.1 Introduction
2.1.1 Cosmic congruity
2.1.2 Background information
2.2 International projects and symposiums
2.3 Mechanics of sediment failure and sliding
2.4 Soil strength and slope stability
2.5 The role of excess pore-water pressure
2.6 Nomenclature and classification
2.6.1 Landslide versus mass transport
2.6.2 Subaerial processes based on the types of movement and material
2.6.3 Subaqueous processes based on mechanical behavior
2.6.4 Processes based on transport velocity
2.6.5 Excessive synonyms
2.7 Recognition of the three basic types of mass-transport deposits
2.7.1 Process sedimentology
2.8 Slides
2.8.1 Definition
2.8.2 Origin
2.8.3 Identification
2.8.4 Case studies
2.8.5 Facies models
2.8.6 Problems
2.9 Slumps
2.9.1 Definition
2.9.2 Origin
2.9.2.1 Submarine channel deposition
2.9.2.2 Earthquake-induced gravity tectonics
2.9.3 Identification
2.9.4 Case studies
2.9.5 Facies models
2.9.6 Problems
2.10 Debris flows: a prelude
2.11 Long-runout mechanisms
2.11.1 Basic concept
2.11.2 Subaerial environments
2.11.3 Submarine environments
2.11.4 Extraterrestrial environments
2.11.5 H/L ratio problems
2.12 Reservoir characterization
2.13 Synopsis
Chapter-3---Gravity-flows--debris-flows--grain-flows_2021_Mass-Transport--Gr
3 Gravity flows: debris flows, grain flows, liquefied/fluidized flows, turbidity currents, hyperpycnal flows, and contour c...
3.1 Introduction
3.2 Gravity (i.e., density) flows
3.3 Gravity-driven downslope processes
3.3.1 Mass transport versus turbidity currents
3.3.2 Sediment-gravity flows
3.3.3 Newtonian versus plastic fluid rheology
3.3.4 Turbulent versus laminar flow state
3.3.5 Sediment concentration
3.4 Debris flows
3.4.1 Definition
3.4.2 Origin
3.4.3 Identification
3.4.4 Case studies
3.4.4.1 Experimental sandy debris flows
3.4.4.2 Krishna-Godavari Basin, Bay of Bengal, India
3.4.5 Facies models
3.4.6 Problem
3.5 Liquefied/fluidized flows
3.5.1 Definition
3.5.2 Origin
3.5.3 Identification
3.5.4 Case studies
3.5.5 Facies models
3.5.6 Problem
3.6 Grain flows
3.6.1 Definition
3.6.2 Origin
3.6.3 Identification
3.6.4 Facies models
3.6.5 Problem
3.7 Turbidity currents
3.7.1 Definition
3.7.2 Origin
3.7.3 Identification
3.7.4 Case studies
3.7.5 Facies models
3.7.6 Problem
3.8 Hyperpycnal flows: a prelude
3.8.1 Definition
3.9 Thermohaline contour currents: a prelude
3.9.1 Definition
3.9.2 Downslope initiation
3.10 Synopsis
Chapter-4---A-paradigm-shi_2021_Mass-Transport--Gravity-Flows--and-Bottom-Cu
4 A paradigm shift
4.1 Introduction
4.2 Amazon Fan, Equatorial Atlantic
4.2.1 Sinuous channels
4.2.2 High-amplitude reflection packet (HARP) units
4.2.3 Lower-fan lobes
4.3 Mississippi Fan, Gulf of Mexico
4.3.1 Channelized lobes
4.4 Monterey Fan, North Pacific
4.4.1 Monterey Canyon
4.4.2 Depositional lobe
4.5 Krishna-Godavari (KG) Basin, Bay of Bengal, India
4.6 The Annot Sandstone (Eocene–Oligocene), Peira-Cava Area, Maritime Alps, SE France
4.7 The Jackfork Group, Pennsylvanian, Ouachita Mountains
4.8 Basin-floor fan model, Tertiary, North Sea
4.9 Mass-flow lobes, Ulleung Basin, East Sea, Korea
4.10 Upper Triassic Yanchang Formation, Ordos Basin, central China
4.11 Supercritical and subcritical fans
4.12 Synopsis
Chapter-5---Density-plumes--types--deflec_2021_Mass-Transport--Gravity-Flows
5 Density plumes: types, deflections, and external controls
5.1 Introduction
5.2 Dataset
5.3 General types of density plumes
5.4 Deflected sediment plumes and their control
5.4.1 Elwha sediment plume, Strait of Juan de Fuca, United States
5.4.2 Connecticut River, New England region, United States, Long Island Sound, United States
5.4.3 Eel River, California, Pacific Ocean, United States
5.4.4 Mississippi River, Gulf of Mexico, United States
5.4.5 Río de la Plata Estuary, South Atlantic Ocean, Argentina, and Uruguay
5.4.6 Rhone River, Gulf of Lions, Mediterranean Sea, France
5.4.7 Ebro Delta, Mediterranean Sea, Iberian Peninsula
5.4.8 Guadalquivir River, Gulf of Cádiz, Southern Spain
5.4.9 Tiber River, Tyrrhenian Sea, Italy
5.4.10 Mornos and Fonissa Rivers, Gulf of Corinth, Greece
5.4.11 Congo (Zaire) River, Atlantic Ocean, West Africa
5.4.12 Yellow River, Bohai Bay, China
5.4.13 Yangtze River, East China Sea, China
5.4.14 Pearl River, South China Sea, China
5.4.15 Krishna-Godavari Rivers, Bay of Bengal, India
5.4.16 Brisbane River, Moreton Bay, Australia
5.4.17 Dart River, Lake Wakatipu, South Island, New Zealand
5.5 Global significance of wind forcing on sediment plumes
5.6 Implications for sediment transport
5.7 Implications for provenance
5.8 Synopsis
Chapter-6---Hyperpycnal-fl_2021_Mass-Transport--Gravity-Flows--and-Bottom-Cu
6 Hyperpycnal flows
6.1 Definition
6.2 Origin
6.3 Identification
6.4 Hyperpycnites and related issues
6.4.1 The incentive
6.4.2 The history
6.4.3 The hyperpycnite problem
6.4.4 The objective
6.5 Basic concepts
6.5.1 Hyperpycnite
6.5.2 Continental margin
6.5.3 Plunge point
6.5.4 Plume versus flow
6.5.5 Types of river-mouth flows
6.5.6 River currents versus turbidity currents
6.5.7 Transformation of river currents into turbidity currents
6.5.8 Fine-grained deltas versus coarse-grained deltas
6.6 The Yellow River, China: a case study
6.6.1 Delta versus estuary
6.6.2 Bathymetry
6.6.3 River-mouth processes
6.6.4 Bottom-turbid layers
6.6.5 Multilayer hyperpycnal flows
6.6.6 Tide-modulated hyperpycnal flows
6.6.7 Internal waves
6.6.8 Velocity measurements
6.6.9 Tidal shear front
6.6.10 M2 tidal dynamics in Bohai and Yellow Seas
6.7 The Yangtze River, China: a case study
6.7.1 Hyperpycnal and hypopycnal plumes
6.7.2 Ocean currents
6.7.3 Tidal river dynamics
6.8 External controls
6.9 Recognition of ancient hyperpycnites
6.9.1 The hyperpycnite facies model
6.9.2 Inverse to normal grading
6.9.3 Internal erosional surface
6.9.4 Traction structures
6.9.5 Massive sandstones
6.9.6 Lofting rhythmites
6.9.7 Plant remains
6.9.8 Hyperpycnite fan models
6.9.9 Flawed principles
6.9.10 Grain size
6.9.11 Modern analogs
6.9.12 Submarine canyons
6.9.12.1 Origin
6.9.12.2 Classification
6.10 Cyclone-induced hyperpycnal turbidity currents in canyons
6.11 Configurations of density plumes
6.12 Global case studies
6.12.1 Dissipating plume with irregular front: the Río de la Plata Estuary, Argentina and Uruguay, South Atlantic Ocean
6.12.2 Coalescing lobate plume: Zambezi River, Indian Ocean
6.12.3 Tidal lobate plume: San Francisco Bay, Pacific Ocean
6.12.4 Swirly cyclonic plume: Tropical Storm Ida, Northern Gulf of Mexico
6.12.5 Whitings plume: the Great Bahama Bank, North Atlantic Ocean
6.12.6 Ring plume: South Pacific Ocean
6.12.7 Tendril plume: Lake Michigan, United States
6.12.8 Swirly plume: Lake Erie, United States
6.13 Challenges
6.14 Future research directions
6.15 Academic discussions
6.15.1 Stages of scientific development
6.15.2 Hyperpycnites and the remaining unresolved issues
6.16 Synopsis
Chapter-7---Triggering-mechanisms-of-_2021_Mass-Transport--Gravity-Flows--an
7 Triggering mechanisms of downslope processes
7.1 Definition
7.2 Origin
7.2.1 Earthquakes
7.2.2 Meteorite impact
7.2.3 Volcanic activity
7.2.4 Tsunami wave
7.2.5 Rogue waves
7.2.6 Cyclonic waves
7.2.7 Internal waves and tides
7.2.8 Ebb-tidal currents
7.2.9 Monsoon flooding
7.2.10 Groundwater seepage
7.2.11 Wildfire
7.2.12 Human activity
7.2.13 Tectonic oversteepening
7.2.14 Glacial maxima and loading
7.2.15 Salt movements
7.2.16 Depositional loading
7.2.17 Hydrostatic loading
7.2.18 Ocean-bottom currents
7.2.19 Biological erosion
7.2.20 Gas-hydrate decomposition
7.2.21 Sea-level lowstand
7.3 Synopsis
Chapter-8---Bottom-curren_2021_Mass-Transport--Gravity-Flows--and-Bottom-Cur
8 Bottom currents
8.1 Introduction
8.2 Vertical continuum: surface currents, deep-water masses, and bottom currents
8.3 The thermohaline circulation
8.4 Four types of bottom currents
8.5 Thermohaline-induced geostrophic bottom currents (i.e., contour currents)
8.5.1 Origin of the Antarctic bottom water
8.5.2 Seismic geometries of contourites
8.5.3 Current velocity
8.5.4 Traction structures
8.6 The contourite problem
8.6.1 Dual forcing of global ocean circulation
8.6.2 Continuum between turbidity currents and contour currents
8.6.3 Revision of the basic principle of contour currents
8.6.4 Hiatuses in contourites
8.6.5 Origin of erosional features
8.6.6 Gulf of Cadiz as the type locality
8.6.6.1 Channel-current stage
8.6.6.2 Mixing and spreading Stage
8.6.6.3 Contour-current Stage
8.6.7 The contourite facies model
8.6.7.1 Five internal divisions
8.6.7.2 Current velocities
8.6.7.3 Internal hiatuses
8.6.7.4 Bioturbation
8.6.7.5 Multiple interactive processes
8.6.7.6 Grain-size data
8.6.8 Traction structures and shale clasts
8.6.9 Bedform-velocity matrix
8.6.10 Abyssal plain contourites
8.6.11 Sandy intervals of contourite facies models
8.7 Wind-driven bottom currents
8.7.1 The Gulf Stream and the Loop Current
8.7.2 Current velocity
8.7.3 Ewing Bank 826 Field: a case study
8.8 Tidal bottom currents in submarine canyons
8.8.1 Types of submarine canyons
8.8.2 Current velocity
8.8.3 Identification
8.9 Baroclinic currents (internal waves and internal tides)
8.9.1 Basic concept
8.9.2 Empirical data
8.9.3 Depositional framework
8.10 Sediment provenance
8.10.1 Current directions
8.10.2 Detrital composition
8.11 Reservoir quality
8.12 Synopsis
Chapter-9---Soft-sediment-deformati_2021_Mass-Transport--Gravity-Flows--and-
9 Soft-sediment deformation structures
9.1 Introduction
9.2 Datasets
9.3 Definition
9.4 Origin
9.5 Classification
9.6 Advances
9.7 Geological implications based on case studies
9.7.1 Breccias
9.7.2 Lateral extent
9.7.3 Ocean bottom currents
9.7.4 Mass-transport deposits
9.7.5 Future research
9.7.6 Unresolved issues
9.7.6.1 The seismite problem
9.7.6.2 The breccias problem
9.7.6.3 The tombolith problem
9.8 Synopsis
Chapter-10---Epilogue--lessons-_2021_Mass-Transport--Gravity-Flows--and-Bott
10 Epilogue: lessons learned
10.1 Lessons learned
Appendix-A---Concepts--glossary--an_2021_Mass-Transport--Gravity-Flows--and-
Appendix A Concepts, glossary, and methodology
Appendix-B---Video-of-flume-experiments_2021_Mass-Transport--Gravity-Flows--
Appendix B Video of flume experiments on Sandy debris flows
B.1 Composition of slurries used in experiments
B.2 Video content
B.3 Deposits
Bibliography_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Bibliography
Author-Index_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Author Index
Subject-Index_2021_Mass-Transport--Gravity-Flows--and-Bottom-Currents
Subject Index

Citation preview

MASS TRANSPORT, GRAVITY FLOWS, AND BOTTOM CURRENTS

MASS TRANSPORT, GRAVITY FLOWS, AND BOTTOM CURRENTS Downslope and Alongslope Processes and Deposits G. SHANMUGAM Department of Earth and Environmental Sciences, The University of Texas at Arlington, Arlington, Texas, United States

Elsevier Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom 50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States Copyright © 2021 Elsevier Inc. All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions. This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein). Notices Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary. Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility. To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein. British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress ISBN: 978-0-12-822576-9 For Information on all Elsevier publications visit our website at https://www.elsevier.com/books-and-journals

Publisher: Candice Janco Acquisitions Editor: Louisa Munro Editorial Project Manager: Andrae Akeh Production Project Manager: Kumar Anbazhagan Cover Designer: Matthew Limbert Typeset by MPS Limited, Chennai, India

Dedication Dedicated to five sedimentologic and oceanographic pioneers of the 20th century: R.A. Bagnold, J.E. Sanders, G.D. Klein, F.P. Shepard, and C.D. Hollister

Contents 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9

Gravity (i.e., density) flows 92 Gravity-driven downslope processes 92 Debris flows 98 Liquefied/fluidized flows 126 Grain flows 131 Turbidity currents 133 Hyperpycnal flows: a prelude 140 Thermohaline contour currents: a prelude 145 3.10 Synopsis 146

About the author ix Preface xxiii Acknowledgments xxvii 1. Introduction 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8

Why this book? 1 History 2 Universal case studies 2 Environments and processes 3 Objectives 3 Organization 5 Other aspects of the book 5 Synopsis 6

4. A paradigm shift 149 4.1 4.2 4.3 4.4 4.5

2. Mass transport: slides, slumps, and debris flows 7 2.1 Introduction 8 2.2 International projects and symposiums 34 2.3 Mechanics of sediment failure and sliding 35 2.4 Soil strength and slope stability 35 2.5 The role of excess pore-water pressure 37 2.6 Nomenclature and classification 38 2.7 Recognition of the three basic types of masstransport deposits 49 2.8 Slides 50 2.9 Slumps 63 2.10 Debris flows: a prelude 73 2.11 Long-runout mechanisms 74 2.12 Reservoir characterization 83 2.13 Synopsis 87

4.6

4.7 4.8 4.9 4.10 4.11 4.12

5. Density plumes: types, deflections, and external controls 181

3. Gravity flows: debris flows, grain flows, liquefied/fluidized flows, turbidity currents, hyperpycnal flows, and contour currents 89 3.1 Introduction

Introduction 149 Amazon Fan, Equatorial Atlantic 150 Mississippi Fan, Gulf of Mexico 155 Monterey Fan, North Pacific 156 Krishna-Godavari (KG) Basin, Bay of Bengal, India 160 The Annot Sandstone (Eocene Oligocene), Peira-Cava Area, Maritime Alps, SE France 161 The Jackfork Group, Pennsylvanian, Ouachita Mountains 169 Basin-floor fan model, Tertiary, North Sea 171 Mass-flow lobes, Ulleung Basin, East Sea, Korea 176 Upper Triassic Yanchang Formation, Ordos Basin, central China 176 Supercritical and subcritical fans 177 Synopsis 179

5.1 5.2 5.3 5.4

90

vii

Introduction 182 Dataset 187 General types of density plumes 188 Deflected sediment plumes and their control 189

viii

Contents

5.5 Global significance of wind forcing on sediment plumes 205 5.6 Implications for sediment transport 207 5.7 Implications for provenance 210 5.8 Synopsis 211

6. Hyperpycnal flows 6.1 6.2 6.3 6.4 6.5 6.6 6.7 6.8 6.9 6.10 6.11 6.12 6.13 6.14 6.15 6.16

213

Definition 214 Origin 214 Identification 214 Hyperpycnites and related issues 215 Basic concepts 218 The Yellow River, China: a case study 229 The Yangtze River, China: a case study 237 External controls 239 Recognition of ancient hyperpycnites 241 Cyclone-induced hyperpycnal turbidity currents in canyons 253 Configurations of density plumes 254 Global case studies 255 Challenges 262 Future research directions 262 Academic discussions 263 Synopsis 270

7. Triggering mechanisms of downslope processes 273

8.4 Four types of bottom currents 320 8.5 Thermohaline-induced geostrophic bottom currents (i.e., contour currents) 321 8.6 The contourite problem 327 8.7 Wind-driven bottom currents 349 8.8 Tidal bottom currents in submarine canyons 356 8.9 Baroclinic currents (internal waves and internal tides) 363 8.10 Sediment provenance 366 8.11 Reservoir quality 370 8.12 Synopsis 373

9. Soft-sediment deformation structures 377 9.1 9.2 9.3 9.4 9.5 9.6 9.7

Introduction 377 Datasets 395 Definition 395 Origin 396 Classification 401 Advances 402 Geological implications based on case studies 405 9.8 Synopsis 437

10. Epilogue: lessons learned 441 10.1 Lessons learned

7.1 Definition 273 7.2 Origin 274 7.3 Synopsis 307

8. Bottom currents 309 8.1 Introduction 310 8.2 Vertical continuum: surface currents, deepwater masses, and bottom currents 310 8.3 The thermohaline circulation 313

441

Appendix A: Concepts, glossary, and methodology 449 Appendix B: Video of flume experiments on Sandy debris flows 477 Bibliography 479 Author Index 547 Subject Index 557

About the author G. (Shan) Shanmugam is an adjunct professor of Earth and Environmental Sciences at the University of Texas at Arlington, Arlington, TX, United States. He is also an adjunct professor of Earth and Planetary Sciences at the University of Tennessee, Knoxville, TN, United States. He is a person of Indian origin (Fig. 1). He emigrated to the United States in 1970 and became a naturalized US citizen in 1990. He has been married to his American wife, Jean, since 1976. He is a pragmatic and an iconoclastic deep-water process sedimentologist. His primary contributions are aimed at documenting the volumetric importance of sandy mass-transport deposits and bottom-current reworked sands in deep-water petroleum reservoirs worldwide and at dispelling the popular myth that most deep-water sands are turbidites.

Professional preparation 1978: Ph.D., Geology, University of Tennessee, Knoxville, TN, United States 1972: M.S., Geology, Ohio University, Athens, OH, United States 1968: M.Sc., Applied Geology, Department of Civil Engineering, Indian Institute of Technology (IIT) Bombay, India 1965: B.Sc., Geology and Chemistry, Annamalai University (AU), Tamil Nadu, India Note: He served as a research scholar under the Council of Scientific and Industrial Research (CSIR), Government of India, at IIT Bombay during 1968–1970.

Employment He joined Mobil Research and Development Corporation in Dallas, TX, United States as a Research Geologist in 1978 and retired from Mobil (now ExxonMobil) as a Geological Scientist in 2000 (Fig. 2).

Research He conducted outcrop studies of deepwater deposits in the Southern Appalachians (Tennessee, United States), Ouachita Mountains (Arkansas and Oklahoma, United States),

ix

x

FIGURE 1

About the author

G. Shanmugam was born in Sirkazhi, Tamil Nadu, India.

and Peira Cava area (French Maritime Alps, SE France). He described deep-water strata using conventional cores and outcrops (1:20 to 1:50 scale), which include 32 deepwater sandstone petroleum reservoirs worldwide, totaling over 10,000 m in cumulative thickness during 1974 2011. He also conducted field studies of coal deposits in Victoria (Australia), coniferous rain forests in the North Island (New Zealand), limestone karst in Guilin (China), fluvial deposits in Gujarat (India), 2004 Indian Ocean Tsunami-related coastal deposits in Tamil Nadu (India), shallow-marine deposits in Qassim area (Saudi Arabia), and estuarine deposits in the Oriente Basin (Ecuador).

Publications He has over 380 published works, including two volumes of Elsevier’s Handbook of Petroleum Exploration and Production (2006 and 2012) and their Chinese editions (Fig. 3).

About the author

xi

FIGURE 2 Three Mobil research laboratories in Texas, United States, where G. Shanmugam conducted research during 1978-2000. Top image: Research includes oil generation from coal in the Gippsland Basin, Australia (Shanmugam, 1985a) and porosity enhancement from chert dissolution beneath Neocomian unconformity in the Prudhoe Bay Field, Alaska (Shanmugam and Higguns, 1988). Middle image: Research includes the Ouachita flysch in the USA (Shanmugam and Moiola, 1995) and basin-floor fans in the North Sea (Shanmugam et al., 1995a). It is worth noting that this Mobil Dallas Research Laboratory was designed by a world-renowned architect I. M. Pei, who also designed the Louvre Pyramid in Paris, France. Bottom image: Research includes bottom-current reworked sands by hybrid flows in the Gulf of Mexico (Shanmugam et al., 1993a), tide-dominated estuarine facies in the Oriente Basin, Ecuador (Shanmugam et al., 2000), and the Annot Sandstone in the Peira Cava area, Maritime Alps, SE France, which served as the type locality for developing the "Bouma Sequence" (Shanmugam, 2002a).

Global workshops on deep-water sandstone petroleum reservoirs He organized deep-water sandstone workshops for: • • • •

the UK Department of Trade and Industry (DTI) in Scotland (1995 and 1997); Petrobras, Mobil, and Unocal in Brazil and Dallas, Texas (1998 and 1999); Oil and Natural Gas Corporation (ONGC) in India (2002 and 2004); Reliance Industries Ltd. in India (2006 09) (Fig. 4);

xii

FIGURE 3

About the author

Images of Elsevier books published by G. Shanmugam in English and Chinese.

• Research Institute of Petroleum Exploration and Development (RIPED), PetroChina in Beijing (2009 10) (Fig. 5); • Yanchang Oilfield Exploration and Development, Research Institute of Yan’an Branch (China) (2014); • China University of Petroleum, Qingdao, China (2014) (Fig. 6).

Awards, recognitions, and nomination • 1968: IIT Medal for the top-ranking student in Applied Geology, Civil Engineering Department, IIT Bombay, India (Fig. 7). • 1995: Best paper award from NAPE (Nigerian Association of Petroleum Explorationists) for his paper “Deepwater Exploration: Conceptual Models and their Uncertainties.” • His paper “High-density turbidity currents: are they sandy debris flows?” published in the Journal of Sedimentary Research in 1996, has achieved the status of the single most cited paper in sedimentological research published in three world-renowned periodicals - Journal of Sedimentary Research, Sedimentology, and Sedimentary Geology during the survey period of 1996 2003 (Source: International Association of Sedimentologists Newsletter, August 2003).

About the author

xiii

FIGURE 4 Images of core workshop conducted at Reliance Industries, India. See a core-seismic based sedimentologic study of sandy debrites and tidalites in submarine canyons from the offshore Krishna-Godavari Basin, Bay of Bengal (India) by Shanmugam, Shrivastava and Das (2009).

• He was interviewed by the SUN TV, Chennai, India (televised on December 30th, 2003) on his controversial research papers on turbidite sedimentation and their implications for petroleum reservoirs (Fig. 8). • He is an Emeritus Member of SEPM (Society for Sedimentary Geology); member since 1970. • 2010 11: Scientific Advisor: Research Institute of Petroleum Exploration and Development (RIPED) of PetroChina, Beijing, China. • 2018: He is the recipient of FeTNA 2018 “Tamil American Pioneer Award” for his extraordinary professional achievements in academia. FeTNA: Federation of Tamil Sangams of North America. Award Date: June 30, 2018. Frisco, Texas (Fig. 9). http:// tap.fetna.org/category/2018/. • 2018: He is the recipient of the University of Tennessee College of Arts & Sciences 2018 Professional Achievement Award. Award Date: September 21, 2018. Knoxville, Tennessee (Fig. 10). https://artsci.utk.edu/dialogue/honor-college-alumni/. • 2019 21: He was nominated for the SEPM 2020 William F. Twenhofel Medal, which is the top award given every year for contributions in sedimentary geology.

xiv

FIGURE 5

About the author

Image of workshop participants at RIPED in Beijing.

Philanthropy • He has two Endowed Graduate Fellowships in Sedimentary Geology and Petroleum Geology ($60,000) at his alma mater, Department of Earth and Planetary Sciences (EPS), The University of Tennessee, Knoxville, Tennessee. • I am thankful to Larry McKay (Associate Dean) and Andrew Sheehy (Senior Director of Development) for help with establishing fellowships.

Online resources for his publications • Blog (deep-water processes): http://g-shanmugam.blogspot.com/ • ResearchGate: http://www.researchgate.net/profile/G_Shanmugam/publications • UTA profile: http://www.uta.edu/profiles/Ganapathy-Shanmugam

About the author

FIGURE 6 Image of workshop participants at China University of Petroleum, Qingdao.

FIGURE 7 1968 IIT Bombay Medal.

xv

xvi

About the author

FIGURE 8

Appearance at SUN TV, Chennai, Tamil Nadu, India.

Transformation from a local science teacher to a global petroleum geologist In celebrating the 90th anniversary of AU, my alma mater, in 2020, Dr. K. Muthuraman, Dean of Faculty of Fine Arts and the Convener of Souvenir Committee, invited me to write a brief article on my reminiscence about the AU. Upon receiving this letter, my instinctual response was one of exuberance. The very first thought came to my mind was the late Professor T.N. Muthuswami Iyer (Fig. 11). He was popularly known as “TNM.” He was an internationally known mineralogist and a pioneer in the study of “Madras charnockites” in the early 1950s. He was the first Head of the Department of Geology at AU since its founding in 1953 and also during my B.Sc. years (1962 65). TNM was solely responsible for my successful career as a global petroleum geologist. I would like to share his motivational story with colleagues and students on this historic and auspicious occasion (Shanmugam, 2020).

Science teacher I was born in 1944 in a town called Sirkazhi, which is located 23 km south of AU near Chidambaram in Tamil Nadu, India (Fig. 1). I attended AU as a train student, commuting everyday from Sirkazhi to Chidambaram. Although my parents, K. Ganapathy Mudaliar and G. Sambooranam Ammal, were not educated, they were keen on my education. They

About the author

xvii

FIGURE 9 2018 Tamil American Pioneer Award.

knew that my education was the only escape out of their poverty. Our family consisted of my parents, myself, and my four younger sisters, namely Dhanalaxmi, Saraswathi, Chandra, and Savithri (deceased). My parents’ primary concern was dowries associated with my sisters’ forthcoming weddings. This financial background is important to this story. I earned my B.Sc. degree in geology with a first class (equivalent to “A” grade in the United States). It is worth noting that my degree was signed by Sir C.P. Ramaswami Iyer, then Vice Chancellor of AU. In the summer of 1965, I secured a position as a science teacher at Krishnamoorthy Arunachala Mudaliar High School, located a few kilometers from my home. My parents were ecstatic because for the first time they will have a monthly income from my salary. Normally, my story would have ended as a science teacher, but the story took a drastic turn and has continued as a petroleum geologist because of TNM.

Motivations by TNM During my employment as a science teacher in 1965, I received a postcard from TNM. The card simply read “Come see me.” His postcard was a surprise to me. Anyway, I went to AU and met with TNM in his office. This was my first face-to-face meeting with the Head of the Department. I was rather nervous, not knowing what to expect.

xviii

FIGURE 10

About the author

2018 Professional Achievement Award.

TNM greeted me with great affection like a grandfather. He first congratulated me in passing my exams with a First Class. He wanted to know what my plans were for the future. I explained to him that my life is settled in Sirkazhi as a science teacher. He said, “You are one of our top students, you have unlimited potential to become a successful geologist, and you should pursue your graduate studies in Applied Geology at IIT Bombay”. At that time, I did not know anything about IIT Bombay. Given my family’s financial challenges, I knew that TNM’s proposal was impossible. I explained my family situation to TNM. He said in a rather commanding tone, “I cannot let you miss this rare opportunity. Do something to get a loan and go to IIT.” At that point, I realized that I should do something to resolve the financial problem. Although I did not know how to resolve the problem, I told TNM that I will resolve the problem and will attend IIT. He was pleased with my determination and with my positive response. I did resolve the financial problem by obtaining a long-term loan from a local businessman, Sri. D. Sambandam, who was an elder brother of my childhood friend, Sri. D. Arumugam. At IIT Bombay, I studied under the supervision of Professor A. Parthasarathy (his DIC and PhD degrees were from the Imperial College and London University, respectively, 1954). My M.Sc. thesis at IIT was on fluvial sedimentology and statistics. I received the Institute Medal for the top-ranking student in Applied Geology (1968). As part of the

About the author

xix

FIGURE 11 Professor T. N. Muthuswami Iyer.

curriculum, I received my first field training from Oil and Natural Gas Corporation (ONGC) in the Great Rann of Kutch in the Thar Desert under Dr. S.K. Biswas and laboratory training in the Ahmadabad office (Gujarat). IIT Bombay not only prepared me for my sedimentology and petroleum geology career but also led me to pursue graduate studies in the United States. Throughout my studies at IIT Bombay, I kept TNM informed of my progress. Finally, I informed him about my plans to go to the United States in all of 1970. At this point in time, TNM had retired from AU and settled in Madras. In his response, he wrote me a letter in August 1970. After 50 years, I still have his letter in my possession (Fig. 12). Because of poor resolution of the scanned copy of letter, I have transcribed the letter content below: “Raja Annamalaipuram-Madras: 10-8-70 My Dear Shanmugam, Very happy to see your kind letter. I am sure you will have a very successful and bright career in Ohio University. It is very good of you to think of me. Few people have this affection and regard. I have permanently shifted to Madras. I am staying with my son Naganathan, who is Area Manager-ALITALIA. If you have to book your passage to Ohio he will do everything for you. His office address. . .

xx

About the author

FIGURE 12

1970 Post card to G. Shanmugam by Professor T. N. Muthuswami Iyer.

Hope to meet you when you go next to Madras. Yours sincerely, T.N. Muthuswami (Signature)” Indeed, his son did book my passage to America by ALITALIA Airlines. Of course, I did go to his house in Madras and did meet him in person before my departure to the United States. He was very proud of my achievements. He would have been even more thrilled to witness my achievements since then, including the arrangements of weddings of all my three sisters. He is in Heaven and smiling down on me, I am sure!

Nature photographer Shanmugam has published numerous photographs of outcrops and cores showing unique geological features on the covers of international geological journals (Fig. 13, 14 and 15).

About the author

FIGURE 13 (A) Pulpit Rock, Norway; (B) Basin-plain turbidites Zumaya, Spain.

FIGURE 14 (A) Karst topography, China; (B) Granitix monolith, near Chennai, India.

xxi

xxii

FIGURE 15

About the author

(A) Ganges River Haridwar, India; (B) Coropaxi Volcano, Ecuador.

Reference Shanmugam, G., 2020. Professor T.N. Muthuswami Iyer and his momentous motivations at Annamalai University (1965): A personal story of transformation from a local science teacher to a global petroleum geologist. Souvenir, 90th anniversary of Annamalai University, Annamalai Nagar, Tamil Nadu, India, 7 p.

Preface This book is a follow-up to my earlier two books on deep-water processes and deposits published by Elsevier (Shanmugam, 2006a, 2012a). The two previous volumes were aimed primarily at petroleum geoscientists, whereas this one is of a broader scope. There are many empirical reasons for my undertaking this assignment: 1. Literature gap: There are no books in the market dedicated solely to downslope and alongslope processes. 2. Publishing opportunity: Research-based invitation by Louisa Munro, Acquisitions Editor, Elsevier Limited, Oxford, United Kingdom, based on my 2019 Encyclopedia chapter titled Slides, Slumps, Debris Flows, Turbidity Currents, Hyperpycnal Flows, and Bottom Currents (Shanmugam, 2019a) published in the Elsevier’s Encyclopedia of Ocean Sciences (third edition, edited by Kirk Cochran, J., Bokuniewicz, H., and Yager, P.), to write a new book expanding the content for a broader readership. 3. Global readership: The need for such a book has been revealed by the readers’ response worldwide, who have frequently read/downloaded the above Encyclopedia chapter from my online ResearchGate webpage. 4. Interpretation of hyperpycnites in deep-water basinal settings based on outcrop studies. Perhaps the compelling reason for writing this book in 2019 is because of the ongoing academic discussions on hyperpycnal flows (Shanmugam, 2019c). The current trend is to promote an unsubstantiated

notion, based on study of ancient rock record, that hyperpycnal flows transport sand into the deep sea (Mulder et al., 2003; Mutti, 2019). For example, Mutti (2019) states that “Field observations suggest that hypopycnal plumes can generate thinbedded sand/mud couplets, here termed plumites, that are virtually ubiquitous in turbidite systems. This close association is the best and most direct evidence of the relationship between turbidite and fluvial sedimentation. Plumes propagate in seawater as dilute surface flows and, depending upon their original volume, sediment concentration and basin size, may mantle the basin floor with their fine-grained deposits from shelfal to deep basin plain regions. They may trigger major hyperpycnal flows and deposit thick sand beds in basinal regions, but most commonly form thin beds displaying a spectrum of highly diagnostic facies. Much care must be taken not to mistake these facies for the distal or overbank sediments of turbidity currents. Most of the fine details of plumites are certainly better observed in cores; most cores should be therefore re-analyzed in the light of these new data.” The problem is that no one could distinguish hyperpycnal flows from turbidity currents in terms of fluid rheology, flow state, and sediment concentration. In addition, there are 16 types of hyperpycnal flows (Shanmugam, 2018b). Mutti (2019) did not explain which one of the 16 types had caused the sand/mud couplets in hyperpycnites in deep-water basinal settings. Satellite images of modern systems do not support such views of sand

xxiii

xxiv

Preface

transport across the shelf into the deep sea (Chapter 5: Density plumes: types, deflections, and external controls and Chapter 6: Hyperpycnal flows). Also, available books on deep-water sedimentation do not address the link between satellite images and sediment transport with implications for deep-sea sedimentation, which is addressed in Chapter 5, Density plumes: types, deflections, and external controls. In interpreting ancient sedimentary record of deep-water origin, it is much more meaningful to understand a process and its deposits from modern settings and from laboratory experiments and then apply that knowledge in the field. Mutti’s (2019) approach of interpreting ancient deposits in outcrops without knowing the fluid mechanics of hyperpycnal flows is misguided. This basic flaw is often overlooked in publications on deep-water sedimentation. This book is a cautionary account of this problem associated with deconstructing depositional origin of ancient rock record by reverse engineering. In addition to the above reasons, my previous contributions on the subject matter facilitate an ideal scenario for writing this book. For example, my contributions include the following: 1. documenting the volumetric importance of sandy mass-transport deposits and bottom-current reworked sands in deep-water petroleum reservoirs worldwide; 2. dispelling the popular myth that most deep-water sands are turbidities (Shanmugam, 2006a, 2012a, 2019a); 3. debunking the myths of facies models on high-density turbidities (Shanmugam, 1996a,b, 2000), tsunamites (Shanmugam, 2006b), landslides (Shanmugam, 2015a), submarine fans (Shanmugam, 2016a), contourites (Shanmugam, 2016b, 2017b),

seismites (Shanmugam, 2016c), SSDS (Shanmugam, 2017a), and hyperpycnites (Shanmugam, 2018b); 4. participating in 38 academic debates, both written and oral, during the past 36 years (1983 2019) (see Chapter 6: Hyperpycnal flows). However, ultimately the objective of any book is to offer a convincing story. This is difficult in dealing with deep-water processes and deposits. This is because of the prevailing plethora of conflicting concepts and models. In this situation, it is imperative to take a fresh look at available data. Therefore the emphasis of this book is to gather a multitude of empirical data universally. I have used a total of 540 case studies or datasets. Images of important examples are included. Whether one agrees or disagrees with my views on a given issue (e.g., high-density turbidity currents or hyperpycnal flows), at least, both sides of an issue will have an opportunity to examine the same data. Although I offer my views on various issues, the reader will have the ultimate say on any given issue. Hopefully, this case study driven approach will yield a more harmonious outcome on controversies dealing with deep-water processes and deposits. Although this volume is intended for a wide range of knowledge levels, including students, teachers, and researchers of gravity-driven sedimentary phenomena, and practitioners in the petroleum industry, it is written mostly with the student in mind. Therefore I have (1) adopted bulleted or numbered text format, (2) included copious number of color images of modern and ancient examples in an atlas format, (3) offered solutions to lingering nomenclatural and conceptual problems, (4) explained the practical implications of downslope and alongslope processes from

Preface

a petroleum reservoir viewpoint, and (5) included an illustrated appendix on concepts, glossary, and methodology and a video on experimental sandy debris flows. This book contains 540 case studies, 344 figures, 28 tables, and a Bibliography with 1,612 published works, which include 106 self-citations as first author and two video lectures, one on the beauty of the rocks by Mutti (2020) and the other on

xxv

the orthodoxy versus empirical evidence behind the “Bouma Sequence” by Shanmugam (2020a). Thank you for your interest on this topic. G. Shanmugam, Ph.D. Email: [email protected] August 21, 2020

Acknowledgments This book is the culmination of my learning and conducting research for over 50 years. And I take this singular opportunity to account for a comprehensive acknowledgment of people, institutions, and events that have made significant impacts in my research career.

taught me clastic sedimentology and introduced me to the Jackfork Group in the Ouachita Mountains. Third, I am grateful to the late D. Sambandam, the late D. Arumugam, and K. Swaminathan for their timely financial and logistical support during my college days in India.

Parents, teachers, and benefactors

Early research (1965 78)

First, I would like to acknowledge my late parents’ (K. Ganapathy and G. Sambooranam) limitless enthusiasm for my education that propelled me out of a remote village in southern India. Second, my sincere thanks to my teachers of geology, in particular, the late Prof. T.N. Muthuswami (Annamalai University, southern India) who persuaded me to go to graduate school at Indian Institute of Technology, Bombay (IIT Bombay); the late Prof. A. Parthasarathy (IIT Bombay) who trained me on engineering aspects of modern mass-transport deposits (MTDs) and who supervised my MSc thesis on fluvial deposits and statistical analysis; the late Prof. Stanley P. Fisher (Ohio University, Athens) who supervised my MS thesis on sandstone diagenesis of the Ordovician Simpson Group in Southern Oklahoma; Prof. Kenneth R. Walker (University of Tennessee, Knoxville) who supervised my PhD dissertation on tectonics and sedimentation of the Middle Ordovician Sevier Basin in the Southern Appalachians in Tennessee; and Prof. Garrett Briggs (University of Tennessee, Knoxville) who

• In compiling this book with emphasis on MTDs (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows), my educational background at the Civil Engineering Department of the IIT Bombay is crucial. This is because of graduate-level courses on soil mechanics, engineering geology, metallurgy, and field measurements of engineering properties of MTDs triggered by monsoonal floods every year. I am very grateful to my supervisor, the late Prof. A. Parthasarathy, who was an engineering geologist by training and who earned his DIC and PhD from the Imperial College and London University, respectively, in 1954 in the United Kingdom. He authored a textbook on Engineering Geology (Parthasarathy et al., 2013). • I was greatly benefited from graduatelevel courses on fluid mechanics taught by Prof. Iaakov Karcz and Fluvial Geomorphology taught by Prof. Marie Morisawa at SUNY Binghamton, New York (1972 73). I also learned the nuances of flume experiments from

xxvii

xxviii

Acknowledgments

Iaakov Karcz. A flume study resulted in my first paper on flume experiments published in the Proceedings of the American Society of Civil Engineers (ASCE), Journal of the Hydraulics Division (Karcz and Shanmugam, 1974). • I am thankful to my two fellow students G.L. Benedict and the late S.C. Ruppel for their help during my PhD research at the University of Tennessee (1974 78). Our publications include Shanmugam and Benedict (1978) on fine-grained carbonate debris flows, Shanmugam and Benedict (1983) on Manganese distribution in carbonate fractions, and Walker et al. (1984) on a model for carbonate to terrigenous clastic sequences.

Mobil research on mass transport, gravity flows, and bottom currents (1978 2000) My tenure with Mobil Oil Company (1978 2000) is the primary source of data for this book. However, my research at Mobil also included oil from coal in Australia and New Zealand (Shanmugam, 1985c), porosity enhancement beneath erosional unconformities in the Prudhoe Bay reservoir, Alaska (Shanmugam and Higgins, 1988), tidedominated estuaries in Ecuador (Shanmugam et al., 2000), among others. • During my employment with Mobil (1978 2000), I worked in three Mobil research facilities (Fig. 2 under "About the Author" section): 1978 83: Mobil Field Research Laboratory, Duncanville, Texas. 1983 92: Mobil Dallas Research Laboratory, Farmers Branch, Texas. It is worth noting that this building was designed by a world-renowned architect I.M. Pei who also designed

the Louvre Pyramid for the Muse´e du Louvre in Paris, France. 1992 2000: Mobil Technology Company, Dallas, Texas. Mobil Oil Corporation was extremely generous in granting me permission to publish over 100 journal articles and 80 abstracts. Several of the petroleum-related case studies used in this book were formally reviewed and approved for external publication by Mobil management and partners (1978 2000). Those case studies that were published in peer-reviewed journals and those that were presented at national and international conferences are the primary source of data on core and outcrop, sandbody geometry, wireline logs, seismic profiles, and measured porosity and permeability values used in this book. I extend my sincere gratitude to Mobil Technology managers (1978 2000): the late E.L. Jones, the late N.J. Guinzy, J.J. Wise, M.P. Ramage, M.G. Bloomquist, E.C. Griffiths, S.J. Moncrieff, R.P. Nixon, the late A.J. Koch, R.J. Moiola, D.M. Summers, S.E. Sommer, M.A. Northam, G.K. Baker, and J. E. Krueger. I am thankful to Mobil Vice President P.E. Luttrell for her constant support of my studies on deep-water systems and her enthusiasm for organizing deepwater sandstone workshops for Mobil affiliates and partners. My special thanks to R.J. Moiola, who guided my career in Mobil as my manager, mentor, colleague, coauthor, and friend. I am grateful to D.W. Kirkland who has been an inspiration throughout my career in Mobil. I thank Mobil colleagues J.E. Damuth, J.G. McPherson, S.B. Famakinwa, J.B. Wagner, R.D. Kreisa, J.W. Snedden, the late M.H. Link, P. Weimer, S. Gabay, J.F. Sarg, J.M. Armentrout, J. Helwig, J.K. Sales, and J.S. Wickham for stimulating discussions.

Acknowledgments

I am grateful to L.J. Aucrermann, B.K. Bowlin, S. Limerick, J. Zeng, and D. Prose who assisted me under the Mobil intern program on deep-water systems. My special thanks to M.K. Lindsey, who drafted most of my illustrations, for his creativity and patience. Mark Lindsey depicted my geologic perspectives of hybrid flows beautifully for my 1993 AAPG Bulletin article, which is reproduced on this book cover. I thank A.F. Long, N.D. Pine, J. Livermon, R. Gilcrese, C. Branson, and A. Gonzales for drafting; S.A. Kizer and D.L. Miller for photography; N. Houghton for petrography, B.J. Phillips, T.A. Allison, F.B. Roof, and C.M. Wall for assistance in the field and in laboratory work. I am indebted to Iaakov Karcz who introduced me to flume experiments. I thank John Sales for developing experimental small-scale duplex structure in soft plaster that was used in explaining the origin of sigmoidal deformation. I thank H.T. Mullins for providing samples of calciclastic sandy contourites from the northern Straits of Florida. • For the first time, to understand mechanics of sandy debris flows and their deposits, a Mobil-funded experimental flume study was carried out at St. Anthony Falls Laboratory (SAFL), University of Minnesota (1996 98) under the direction of Prof. G. Parker. Results were published in two major articles (Shanmugam, 2000a,b; Marr et al., 2001). I am grateful to G. Parker, J.G. Marr, and P. A. Harff for impressive experimental runs of sandy debris flows and related discussions. • I am thankful to the following colleagues from Mobil Research and Development Corporation, other companies, and universities who were involved in the

xxix

description of sediments and sedimentary rocks during 1974 2011 period: • United Kingdom: C.E. Shepard, C.F. Stephens, P.H. Naylor, K.P. Dean, S.-J. Kelland, J. Mathews, F. Longworth, A. Turner, M. Slatford, G.W.J. Beamish, S.M. Mitchell, and J.E. Damuth; • Norway: L.R. Lehtonen, T. Straume, S. E. Syevertsen, R.J. Hodgkinson, and R. J. Fife; • Nigeria: W.E. Hermance, B.J. Welton, J. O. Olaifa, and U. Ewherido; • Equatorial Guinea: S.B. Famakinwa, W. E. Hermance, and R.J. Hodgkinson; • Gabon: T.D. Spalding, S.B. Famakinwa, and E. Delbos; • France: R.J. Moiola and R.B. Bloch; • Brazil: S.H. Gabay, A.E. Cunningham, W.B. Gardiner, D.O. Hurtubise, Celso Guirro, Paola Fontanelli, and the late Luiz Caddah; • Gulf of Mexico: G. Zimbrick, T.D. Spalding, J. Fouts, J.M. Armentrout, K. Schindler, J. Caravella, T. Scott, R.D. Kreisa, and D.H. Rofheart; • California: C.A. Clayton and R.D. Kreisa; • Arkansas and Oklahoma: R.J. Moiola, H. Jamieson, R. Edington, C. Knutson, D. Prose, T. Stolan, and M. Barrett; • Lamont-Doherty Earth Observatory in New York: J.E. Damuth and S. O’Connell for helping with the description of DSDP Leg 96 cores (Mississippi Fan); • Ocean Drilling Program’s (ODP) Gulf Coast Repository in College Station, Texas: J.E. Damuth for assistance with core description at the Ida Green Cruise cores (Gulf of Mexico); • China: Examination of karst topography near Guilin and Li River in understanding karst dissolution and breccia formation (Shanmugam, 2017d);

xxx

Acknowledgments

• Ecuador: M. Poffenberger, and J. Toro Alava on tide-dominated estuarine facies; • Australia: R.J. Gaulton on the Yallourn and Morwell open cut coal mines; • New Zealand (North Island): R.A. Cook and K. Robinson on modern Kauri rain forests; • Alaska: G. Zimbrick on fluvio-deltaic facies of the Prudhoe Bay oil field; • Azerbaijan: Government and academic personnel. I am deeply indebted to R.J. Moiola for his support and collaboration with all my studies of submarine fans throughout my career with Mobil (1978 2000). My field experience, gained from studies of ancient submarine fans with R.J. Moiola (Spain, Italy, the Ouachita Mountains, and the Annot Sandstone, SE France), with Garrett Briggs (the Ouachita Mountains), with Kenneth Walker (the Southern Appalachians, my PhD work at the University of Tennessee, Knoxville, United States), and on modern submarine fans with J.E. Damuth at LamontDoherty Earth Observatory (Mississippi Fan, DSDP Leg 96 cores), has greatly enhanced my understanding of modern and ancient submarine fans. My sincere gratitude to the late Prof. George Devries Klein, Emeritus Professor at University of Illinois at UrbanaChampaign, who served as the Editor of Earth-Science Reviews and who handled my first review article on submarine fans for the journal (Shanmugam and Moiola, 1988). His publications have been a great source of inspiration for my research on deep-water sedimentation. Tributes to Klein were published in a “Special Issue dedicated to George Devries Klein by the Journal of the Indian Association of Sedimentologists (JIAS)” (Shanmugam, 2018d).

I must thank Prof. Emiliano Mutti for organizing a special field trip to the Eocene Hecho Group in the South-Central Pyrenees (Spain) in March 1981, as a consultant, for two Mobil geologists (R.J. Moiola and G. Shanmugam). Every night at dinner time, after a long, fruitful, and grueling day in the field, we would embark on lively and often heated debates on deepwater sedimentation that would last until the wee hours of the morning. These debates were the root cause of my passion for investigating the very foundation of the turbidite paradigm and submarine fans. As a consequence, my first two critical papers on submarine fans were: (1) “Is the turbidite facies association scheme valid for interpreting ancient submarine fan environments?” (Shanmugam et al., 1985a) and (2) “Submarine fan models: problems and solutions” (Shanmugam and Moiola, 1985a, b). Despite our professional feud, Emiliano and I have exchanged cordial emails to this day. Mutti’s most recent communication was on November 17, 2019, by which he sent me his recent reprint on Plumites (Mutti, 2019). I am grateful to J.E. Warme for organizing field trips to study deep-water deposits in the San Diego area (La Jolla) in February 1980 for Mobil sedimentologists (R.J. Moiola and G. Shanmugam). I also thank the late T. H. Nilsen for organizing a special field trip to study deep-water “turbidites” in California in October 1992, as a consultant, for a Mobil geologist (G. Shanmugam).

Consultant research (2000 present) 1. India (Reliance-core): S.K. Shrivastava, B. Das, M. Acharya, M. Chowdhury, M. Santra, S.S. Roy, S. Gupta, A. Soman, S. Sharma, R. Das, S. Mushnuri, A. Kumar,

Acknowledgments

2.

3. 4. 5.

and V. Yesudian for their assistance during core description (2004 08); India (Reliance-field and seismic): Our field investigation of Godavari estuary near Yanam and Antarvedi (Andhra Pradesh) in August 2007 was assisted by S. Sharma and S.I. Arsalan, and of the Kakinada Bay in January 2008 was assisted by Sandeep Sharma, Chakradhar Rao Basa, Jyoti Rout, Amit Sinha, Sandeep Rawat, Hema Sharma, and Mahendra Thame. Sandeep Sharma also assisted in our study of RMS amplitude maps; India (Oil and Natural Gas Corporation): S. Prabakaran; India (Hardy): Ravi T. Venkatesawaran; China [Research Institute of Petroleum Exploration and Development (RIPED) PetroChina]: Coining Zoo, Wang Land, Li Ying, Sonata Wu, and Xiamen Zhao on deep-lacustrine facies (Zoo et al., 2012).

Tsunamite research (2004 present) My personal interest on tsunamis was accelerated by the 2004 Indian Ocean tsunami, which hit the coast of Tamil Nadu in southeastern India on 26 December. My hometown (Sirkazhi), which is located about 12 km from the tsunami-devastated coast, provided immediate shelter for tens of thousands of tsunami victims. Color videos of the tsunami, shown on BBC and CNN Television from December 27, 2004 to January 5, 2005, were used for inferring flow transformation near the coasts (offshore) of Thailand, India, and Sri Lanka. I am grateful to Brock Adam McCarty of DigitalGlobe for granting permission to use an aerial image of Kalutara Beach in Sri Lanka. I thank many local individuals who narrated their eyewitness accounts of

xxxi

tsunami waves and who helped in digging trenches along the coast of Tamil Nadu to study the effects of tsunami on coastal sedimentation. I dedicate this paper to all those who perished (over 265,000) in 15 countries in the 2004 Indian Ocean tsunami. In an attempt to trace the source of the genetic term “tsunamite,” I sought information from many colleagues worldwide. I would like to thank the following who responded promptly to my email queries: G. Racki, J. Bourgeois, T. Shiki, M. Cita, B. Pratt, G. van den Bergh, K. Rodolfo, M. Simms, S. Barnett, J. Morrow, D. Stow, A. Hurst, D. Gorsline, and G. Klein. In publishing my paper “The tsunamite problem” (Shanmugam, 2006b), I thank the JSR reviewers from Australia (Ron Boyd and Brian Jones), New Zealand (Scott Nichol), and the editorial crew from Scotland (Colin North, Co-Editor), Canada (David Piper and Martin Gibling, Associate Editor), and the United States (John Southard and Melissa Lester) for their comments and help. I thank the late N. Swedaranyam, T. Saraswathi (my sister), S. Thambidurai, and S. Murugan for their assistance during my 2005 field study of coastal deposits of the 2004 Indian Ocean Tsunami in Tamil Nadu. In publishing my paper “Process-sedimentological challenges in distinguishing paleo-tsunami deposits” (Shanmugam, 2012b), I thank Guest Editor Arun Kumar for inviting me to contribute this article. I wish to thank Journal Editor T. Murty for his suggestions on content during early stages of manuscript preparation in 2009. My sincere thanks to two anonymous reviewers for their detailed, critical, and helpful comments on the manuscript. I am grateful to Cliff Frohlich, The University of Texas at Austin, for providing photographs of tsunami emplaced

xxxii

Acknowledgments

boulders in Tongatapu Island, south-west Pacific.

Contourite research (1974 present) I thank Rajat Mazumder, the volume editor of “Sediment Provenance,” for encouraging me to contribute a review chapter on contourites (Shanmugam, 2016b). I also thank both Tasha Frank and Marisa LaFleur, Associate Acquisition Editors (Elsevier), for their enthusiastic help with various issues. I am deeply indebted George Devries Klein, a sedimentologic pioneer on contourites and tidalites, for his total endorsement of science in this chapter and for his helpful editorial comments. I also thank A.J. (Tom) van Loon, who served as the Series Editor for Elsevier’s Developments in Sedimentology 60 on “Contourites” (Rebesco and Camerlenghi, 2008) for his meticulous editing of the manuscript. Chapter 8, Bottom Currents, of this book addresses the contourite problem. I acknowledge with gratitude the following organizations and colleagues involved in various academic activities that are of relevance in my contourite research: • My interest on provenance began with my research on sandstone reservoirs at Mobil Oil Company in 1978. As a consequence, I was an invited speaker at the NATO Advanced Study Institute Conference on “Reading Provenance from Arenites” held in Calabria, Italy (1984) by G.G. Zuffa. In a related conference volume edited by Zuffa (1985), my contribution dealt with “Types of porosity in sandstones and their significance in interpreting provenance” (Shanmugam, 1985b,c).

• My sedimentological research on deepwater bottom currents began in 1974 as a part of my PhD work on the Middle Ordovician of the Southern Appalachians in the United States (Shanmugam, 1978; Shanmugam and Walker, 1978, 1980) and has continued through my employment with Mobil Oil Company (Shanmugam and Moiola, 1982, 1983; Shanmugam, 1990a,b,c; Shanmugam et al., 1993a,b) to the present as an adjunct professor and as a consultant (Shanmugam, 2006a, 2008a, 2012a, 2013a, 2014a). • As my manager and coresearcher, R.J. Moiola provided enthusiastic support for my contourite research throughout my employment with Mobil (1978 2000). As a Mobil colleague, J.E. “Jed” Damuth provided me historical information on contourite research at Lamont-Doherty Earth Observatory of Columbia University (New York) where he received his PhD under Bruce Heezen. I am indebted to numerous colleagues at Mobil and other oil companies, petroleum-related service companies, academic institutions, and government agencies for assisting me in core and outcrop descriptions worldwide during the past 40 years (Table 9.2). • My first major paper on process sedimentology and reservoir quality of “sandy contourites,” which focused on the significance of traction structures in “contourites” following Heezen’s (1959) pioneering concept, was peer-reviewed by Charles Hollister for the AAPG Bulletin (Shanmugam, 1993a). I dedicate this paper to the late Charles Davis Hollister (1936 99), considered to be “the father of Contourites” (McCave, 2002), who died in a climbing accident while on vacation in Wyoming with his family at an untimely age of 63. His pioneering publications

Acknowledgments













have greatly influenced my research during the past 40 years. In response to an invitation from R.D. Winn, Jr. and J.M. Armentrout, I (Shanmugam et al., 1995b) participated in the 1995 SEPM Core Workshop held in Houston, Texas. This study dealt with core examination of traction sedimentary structures indicating bottom-current reworking in the Gulf of Mexico. In response to an invitation from the UK Department of Trade and Industry (DTI), I organized a deep-water sandstone workshop in Edinburgh, Scotland, for petroleum geoscientists from various countries in Europe in 1995 (October). This workshop utilized cores from the U. K. Atlantic Margin (Table 9.2, item 7) that contain deposits of sandy MTDs and bottom-current reworked sands (Shanmugam et al., 1995a). In response to an invitation from M. Rebesco, I contributed Chapter 5 (Shanmugam, 2008a) entitled “Deepwater bottom currents and their deposits” to the thematic volume on “Contourites” (Rebesco and Camerlenghi, 2008). In response to an invitation from A.J. (Tom) van Loon, I reviewed a book (Shanmugam, 2008d) entitled Economic and Palaeoceanographic Significance of Contourite Deposits, edited by Viana and Rebesco (2007), for Geologos (republished in Journal of Sedimentary Research). I also reviewed a book (Shanmugam, 2011a,b) entitled Deep-sea sediments, edited by Hu¨neke and Mulder (2011) with Chapter 3 on “Contourites”, for Geologos (republished in Journal of Sedimentary Research). Since the 1970s, D.J.W. Piper and D.A.V. Stow have been helpul in my research on both fine-grained turbidites and contourites.

xxxiii

• In 2020, F.J. Herna´ndez-Molina and S. de Castro, helped me with research material from IODP Expedition 339, Gul of Cadiz.

Density plumes and hyperpycnite research (2002 present) I thank Prof. G.N. Nayak, CSIR Emeritus Scientist at School of Earth, Ocean and Atmospheric Sciences, Goa University, Goa, India, who is also the President of the Indian Association of Sedimentologists, for a thorough and helpful review of my paper on deflected sediment plumes (Shanmugam, 2019d). I also thank the second reviewer Dr. Mayla Ramos-Vazquez for helpful review. I am grateful to Dr. John S. Armstrong-Altrin, National Autonomous University of Mexico, for handling my paper. I thank both Managing Editors of the journal, Prof. G.M. Bhat and Dr. Bashir Ahmad Lone, both of Jammu University, India, for their help with editorial matters. I thank Dr. Huaixian Xu, Executive Chief Editor of Petroleum Exploration and Development (PED) for inviting me to contribute my article (Shanmugam, 2018c) for the Special Issue of PED in celebrating the 60th anniversary of RIPED. The year 2018 also marks the 10th anniversary of my association with RIPED. I sincerely thank Prof. B. Charlotte Schreiber (University of Washington, Seattle) for her no-nonsense, critical, and helpful review comments on my paper “The hyperpycnite problem” (Shanmugam, 2018b). I am grateful to the U.S. National Aeronautics and Space Administration (NASA) for their excellent collection of satellite and other images of sediment plumes triggered by river, tide, glacier, volcanic eruptions, cyclones, etc. I would like to acknowledge the following publishers,

xxxiv

Acknowledgments

governmental and nongovernmental agencies, journals, and colleagues, for their help in granting permissions either formally or through their publication policies to reuse images: 1. American Association of Petroleum Geologists (AAPG), 2. American Geophysical Union (AGU), 3. Cambridge Core and Global-Science Press, 4. Copyright Clearance Center (CCC), Rightslink, 5. Elsevier, 6. European Geosciences Union (EGU), 7. Geological Society of America (GSA), 8. Geological Society of London (GSL), 9. Indian Journal of Geo-Marine Sciences (IJMS), 10. J.G. McPherson, 11. John Wiley and Sons, 12. The U.S. National Aeronautics and Space Administration (NASA), 13. The U.S. National Geophysical Data Center (NGDC), 14. R.D. Kreisa, 15. Society for Sedimentary Geology (SEPM), 16. Springer Nature, 17. The U.S. Army Corps of Engineers, 18. The U.S. Geological Survey (USGS), and 19. Wikipedia.

Sediment deformation and seismite research (1978 present) I am thankful D.W. Kirkland for lending core slabs from the Permian Castile Formation, New Mexico and for his valuable discussion on the origin of microfolds. The Journal of Palaeogeography reviewers Prof. Yuan-Sheng Du (State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Wuhan), Prof. Tian-Rui Song (Institute of Geology,

Chinese Academy of Geological Sciences, Beijing), Dr. D.W. Kirkland, and an anonymous reviewer are thanked for their detailed and helpful comments that considerably improved the quality and clarity of my paper “The seismite problem” (Shanmugam, 2018b) paper.

Academic events (1968 present) I benefited from my participation in the following academic events: • As an invited coauthor of a book chapter (Shanmugam and Moiola, 1985) to the COMFAN 1 volume edited by Bouma et al. (1985a); • As an invited participant in the NATO Advanced Study Institute Conference on “Reading Provenance from Arenites,” Calabria, Italy (1984); • As an invited participant in the COMFAN II Meeting, Parma, Italy (1988); • As an invited lecturer in the SEPM Pacific Section Short Course (Shanmugam, 1990b) held in San Francisco as part of the 1990 AAPG Convention on “Deep-Marine Sedimentation, Depositional Models and Case Histories in Hydrocarbon Exploration & Development.” Course organizers: G.C. Brown, D.S. Gorsline, W.J. Schweller. • As an invited panelist in the 1997 (April) AAPG/SEPM Convention Debate, Dallas, Texas Topic: Processes of Deep-Water Clastic Sedimentation and Their Reservoir Implications: What Can We Predict? Moderator: H. E. Clifton. Panelists: A.H. Bouma, J.E. Damuth, D.R. Lowe, G. Parker, and G. Shanmugam.

xxxv

Acknowledgments

• I presented an invited lecture on John Sanders’ pioneering contributions at a conference in Troy, New York (Shanmugam, 2000b). I was invited by Prof. G.M. Friedman and George Devries Klein. • Academic lecture tour of India on deepwater processes and sediment deformation (2016) • Institutions: RIL: Reliance Industries Ltd. ISI: Indian Statistical Institute IITB: Indian Institute of Technology Bombay ITM: Indian Institute of Technology Madras Annamalai University, Chidambaram, Tamil Nadu • Contacts: Mr. Bhagaban Das, Manager, Reservoir Characterization, RIL Prof. Sarbani Patranabis-Deb, ISI, Geological Studies Unit Prof. M. Radhakrishna, IITB, Earth Sciences Prof. Santanu Banerjee, IITB, Earth Sciences Prof. P. Shanmugam, IITM, Ocean Engineering Prof. T. Ramkumar, Annamalai University, Earth Sciences I thank the abovementioned Indian colleagues for organizing my lectures. I am grateful to RIL and ISI for absorbing domestic airline, ground transportation, and lodging expenses. I also thank students and participants for lively discussions. • 2018: I presented a two-part lecture entitled “Deep-Water Turbidites and Density Plumes” at the Dallas Geological Society, International Dinner Event, Brookhaven Country Club, Dallas, Texas, November 14, 2018. https://www.dgs.

org/events/2018/11/14/ Accessed August 17, 2020 • 2020. I presented a virtual lecture entitled “The turbidite – contourite –tidalite – hybridite problem: Orthodoxy Vs Empirical Evidence behind the Bouma Sequence”, organized by the indian Association of Sedimentologists. Virtual Lecture on Google Meet Platform. July 2, 2020 at 10:00 am (Indian Standard Time). • 2020. I presented a zoom lecture entitled "The turbidite – contourite –tidalite – hybridite problem: Orthodoxy Vs Empirical Evidence behind the “Bouma Sequence”. The Drifters VGT (Virtual Get-Together) Zoom Lecture organized by F. J. Hernandez-Molina, Dept. Earth Sciences, Royal Holloway, University of London (UK), July 27, 2020, Monday, 2.30 PM London (UK) Time. PowerPoint presentation. • 2000-2020: I have been associated with the Department of Earth and Environmental Sciences, The University of Texas at Arlington, USA, as an adjunct professor since 2000. I taught undergraduate and graduate courses in sedimentology and stratigraphy during 2003-2004 period. I have routinely presented my research results in department seminars. I am grateful to Professor Emeritus John Wickham, Professor Emeritus Asish Basu, and Professor and Chairman Arne Winguth, for their support.

Editorial board member (2018 present) As an editorial board member of the following three journals, I have benefited immensely on recent developments, and

xxxvi therefore I following:

Acknowledgments

would

like

to

thank

the

1. Journal of Palaeogegraphy (JOPG) Prof. Zeng-Zhao Feng, Editor-inChief, Beijing, China Dr. Yuan Wang, Editor, Beijing, China Dr. Min Liu, Editor, Beijing, China Dr. Xiu-Fang Hu, Editor, Beijing, China 2. Petroleum Exploration and Development (PED) Dr. Huaixian Xu, Executive Chief Editor of Petroleum Exploration and Development, Beijing, China Dr. C. Zou, Editor, Beijing, China Jesse (Song Lichen), Deputy Director, PED, Beijing, China 3. Journal of the Indian Association of Sedimentologists (JIAS) Prof. Abhijit Basu, Indiana, United States, Editors-in-Chief Prof. G.M. Bhat, Jammu University, India, Managing Editor Dr. Bashir Ahmad Lone, Jammu University, India, Managing Editor JIAS dedicated a special issue to the memory of Prof. George Devries Klein with the following contributions: 1. “Preface” by G. Shanmugam 2. “Post-modernism and climate change” by Van der Lingen 3. “Bioturbation and trace fossils in deepwater contourites, turbidites, and hyperpycnites: A cautionary note” by G. Shanmugam 4. “Petroleum potential of the West Coast of India” by Naresh Kumar 5. “Diagenetic evolution of onshore Campanian Sandstone, AriyalurCauvery Basin” by R. Nagendra 6. “Reflections from the heavy mineral distributions of some Gondwana basins of extra-peninsular India” by Hrishikesh

Baruah, Ranjeeta Kar, Sarat Phukan, Pradip Kumar Das, Manab Deka, and Tulika Dey 7. “Professor Virendra Kumar Srivastava” by S.M. Casshyap and M. Raza 8. “Robert Louis Folk” by Kitty Milliken, Earle McBride, and Lynton Land 9. “Robert Henry Dott, Jr. (June 2, 1929 to February 27, 2018)” by Marjorie A. Chan and Steven G. Driese

Photographs I thank Tom Roorda, Roorda Aerial, Port Angeles, Washington for aerial photo of Elwha sediment plume in the Strait of Juan de Fuca and Professor Emeritus R.D. Hatcher, Jr., Department of Earth and Planetary Sciences, The University of Tennessee, Knoxville, for outcrop photo of SSDS from Israel. I thank John G. McPherson for providing aerial photographs of Dart River braid delta from New Zealand, and for photographs of alluvial fans from Death Valley, California.

Elsevier This book is built on the foundation of datasets used in my two previous books by Elsevier (2006 and 2012) coupled with new data published in various journals and other publications since 2012 by myself and by other authors. Because I have used the same core, outcrop, seismic, petrophysical, and satellite database that I used in my previous two books published by Elsevier in 2006 and 2012. My sincere thanks to John Cubitt, Editor-in-Chief of Elsevier’s Handbook of Petroleum Exploration and Production series for valuable suggestions. I thank Susan Dennis, Associate Acquisitions Editor | Series (Elsevier

Acknowledgments

Science Ltd., Oxford, United Kingdom), Derek Coleman, Senior Editorial Project Manager, Elsevier (Amsterdam), and the Elsevier production team (Chennai, Tamil Nadu, India) for their help with the 2012 edition. In publishing this third book for Elsevier, I thank Louisa Munro, Senior Acquisitions Editor, Aquatic Sciences Elsevier Limited, Oxford, United Kingdom, for inviting me to write this book and for providing logistical support. I would like to thank Nicholas Christie-Blick, Professor of Earth and Environmental Sciences, Lamont-Doherty Earth Observatory of Columbia University, New York, and two other Elsevier reviewers of my book proposal for their constructive comments. I also thank Andrae Akeh, Senior Editorial Project Manager, Elsevier for help during the production of this book. In particular, I thank him for his organization and swift transfer of files to the production department. Importantly, both Louisa and Andrae have been very helpful with the necessary budgeting for 300 color figures. This book was produced amid the challenging times of the coronavirus (COVID19) global pandemic in Chennai, Tamil Nadu, India (February October 2020). My special thanks to Elsevier Production Team in Chennai that consisted of R. Vijay Bharath and Kumar Anbazhagan— Production Managers, Sathya Narayanan— Copyrights coordinator, Vinod Kumar– Graphics, and M. Bhuvanaraj. I am deeply indebted to their dedication and meticulous attention to details.

Copyrights and permissions I thank the following individuals for granting permission to reproduce figures:

xxxvii

G.C. Brown, J.V. Gardner, S. Krastel, D.J.M. Macdonald, J.G. Marr, C.K. Paull, E. Mutti, A.R. Viana, C.S.L., Duarte, and A. Solheim. I acknowledge the following publishers and entities for help during the course of acquiring permission to use figures, tables, and data: 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25.

Copyright Clearance Center Elsevier Wiley-Blackwell John Wiley & Sons Ltd. Springer Cambridge University Press Annual reviews, Inc. Rand McNally & Company Gulf Publishing Company American Association of Petroleum Geologists (AAPG) Society for Sedimentary Geology (SEPM) Geological Society of America The Geological Society (London) Gulf Coast Association of Geological Societies Minerals Management Service of the U. S. Department of the Interior Atlantic Oceanographic and Meteorological Laboratory The Cooperative Institute for Marine and Atmospheric Studies U.S. National Oceanic and Atmospheric Administration (NOAA) U.S. National Aeronautics and Space Administration (NASA) U.S. National Geophysical Data Center (NGDC) U.S. National Hurricane Center (NHC) U.S. Geological Survey (USGS) Ocean Drilling Program (ODP) The American Meteorological Society Planetary and Space Science Centre (PASSC), University of New Brunswick, Fredericton, New Brunswick, Canada

xxxviii

Acknowledgments

26. Dave’s Landslide Blog 27. The Landslide blog (AGU) Blogosphere 28. “Landslide” database center at the University of Durham 29. Wikipedia, the free encyclopedia

Editors and reviewers (1978 2020) I would like to acknowledge a select group of world-renowned editors, associate editors, and reviewers who evaluated my contributions during the past 40 years: 1. J. Southard (Journal of Sedimentary Research) 2. P. McCarthy (Journal of Sedimentary Research) 3. C. North (Journal of Sedimentary Research) 4. P.J. Talling (Journal of Sedimentary Research) 5. G.A. Smith (Journal of Sedimentary Research) 6. G. Postma (Journal of Sedimentary Research) 7. D.J.W. Piper (Journal of Sedimentary Research) 8. Martin Gibling (Journal of Sedimentary Research) 9. O.H. Pilkey (Journal of Sedimentary Petrology) 10. Jean Lajoie (Journal of Sedimentary Petrology) 11. G. Kelling (Sedimentary Geology) 12. A.D. Miall (Sedimentary Geology and Earth-Science Reviews) 13. G.D. Klein (Earth-Science Reviews) 14. G.M. Friedman (Earth-Science Reviews and History of Geologic Pioneers)

15. Andre´ Strasser (Earth-Science Reviews) 16. R. Steinmetz (AAPG Bulletin) 17. J.A. Helwig (AAPG Bulletin) 18. S.A. Longacre (AAPG Bulletin) 19. K.T. Biddle (AAPG Bulletin) 20. N.F. Hurley (AAPG Bulletin) 21. E.A. Mancini (AAPG Bulletin) 22. G.M. Gillis (AAPG Bulletin) 23. M. Sweet (AAPG Bulletin) 24. Barry J. Katz (AAPG Bulletin) 25. D.G. Roberts (Marine and Petroleum Geology) 26. E.M. Moores (Geology) 27. H.T. Mullins (Geology) 28. R.E. Arvidson and M.E. Bickford (Geology) 29. R.D. Hatcher, Jr. and W.A. Thomas (GSA Bulletin) 30. J.D. Collinson (Sedimentology) 31. P. Carling (Sedimentology) 32. B.W. Flemming and M.T. Delafontaine (Geo-Marine Letters) 33. Kuldeep Chandra (Indian Journal of Petroleum Geology) 34. A.J. Michael (Bulletin of the Seismological Society of America) 35. A.J. (Tom) van Loon (Geologos, Journal of Sedimentary Research, Journal of Palaeogeography, and Series Editor for Elsevier’s Developments in Sedimentology 60 on “Contourites”). 36. Z.-Z. Feng (Journal of Palaeogeography) 37. G.M. Bhat and Bashir Ahmad Lone (Journal of the Indian Association of Sedimentologists) 38. J. Rodgers, J.H. Ostrom, and P.M. Orville (American Journal of Science)

xxxix

Acknowledgments

39. K.R. Walker and D. Roeder (Appalachian Geodynamic Research: American Journal of Science) 40. A.H. Bouma, W.R. Normark, and N.E. Barnes (Submarine Fans and Related Turbidite Systems) 41. G.G. Zuffa (Provenance of Arenites) 42. G.C. Brown, D.S. Gorsline, and W.J. Schweller (Deep-Marine Sedimentation: Depositional Models and Case Histories in Hydrocarbon Exploration and Development) 43. K.L. Kleinspehn and C. Paola (New Perspectives in Basin Analysis) 44. E.M. Moores and F. Michael Wahl (The Art of Geology) 45. S.P. Hesselbo and D.N. Parkinson (Sequence Stratigraphy in British Geology) 46. R.D. Winn, and J.M. Armentrout (Turbidites and Associated Deep-Water Facies) 47. D.A.V. Stow and M. Mayall (Deepwater Sedimentary Systems) 48. J.H. Steele, K.K. Turekian, and S.A. Thorpe (Encyclopedia of Ocean Sciences, Second Edition) 49. M. Rebesco and A. Camerlenghi (Contourites) 50. A. Kumar and I. Nister (Paleotsunamis: Natural Hazards) 51. J. Cubitt (Handbook of Petroleum Exploration and Production series) 52. Rajat Mazumder (Sediment Provenance) 53. Scott Elias (Reference Module in Earth Systems and Environmental Science) 54. J. Kirk Cochran, H. Bokuniewicz, and P. Yager (Encyclopedia of Ocean Sciences, Third Edition) 55. R.N. Ginsburg (Episodes)

56. W. Nemec and R.J. Steel (Fan Deltas) 57. Scott Elias and David Alderton (Editors) and Nick Lancaster (Section Editors) (Encyclopedia of Geology, Second Edition). Of over 300 reviewers who reviewed my papers, I would like to single out (1) the late Charles Hollister for his review of my paper (Shanmugam et al., 1993a) on bottom-current reworked sands, and (2) the late John Sanders for his review of my paper (Shanmugam, 1997a) on the Bouma Sequence.

Logistics I am grateful to S. Vaideeswaran, Geetha Vaideeswaran, and Divya Sudharsan for their help, including computer-related issues, during preparation of this book. I thank Mr. T.A. Vetriselvan for videotaping and editing of my Zoom Lecture for the Royal Holloway, University of London (UK) on July 27, 2020.

ResearchGate (2014 present) I have benefited from interactions with members of the ResearchGate community. In the following figure compiled from ResearchGate statistics, an unusually high number of 107,759 reads for my articles may be attributed to the following factors: 1. Most of my articles are single-authored. 2. These articles represent diverse academic domains, such as as organic geochemistry, landslides, tsunamis, tropical cyclones, seismicity, deep-water

xl

Acknowledgments

sedimentology, tectonics, petroleum reservoirs, etc. 3. Most of these articles are dedicated to identifying unresolved problems and related controversies that are of critical value to future research. 4. Most readers are Ph.D.-level students.

Wife and friend (1975 present) I am eternally grateful to my wife and friend, Jean Shanmugam, who has helped me since 1975 in the field, in the laboratory, and in typing and editing manuscripts. Although Jean is not a geologist, she has commented on every word that I have ever written for earth and planetary science publications.

The rocks (1962 present) Most importantly, I thank the rocks with hidden clues for giving me the opportunity to examine and interpret them, hopefully correctly!

C H A P T E R

1 Introduction O U T L I N E 1.1 Why this book?

1

1.5 Objectives

3

1.2 History

2

1.6 Organization

5

1.3 Universal case studies

2

1.7 Other aspects of the book

5

1.4 Environments and processes

3

1.8 Synopsis

6

1.1 Why this book? Who knew that satellite images of sediment plumes at river mouths that reveal plume deflections by wind forcing away from the normal course of downslope sediment transport could be useful in understanding deep-water sediment transport and provenance? I did not until recently (Shanmugam, 2019d). Surprisingly, there are no other published accounts of this important phenomenon. This raises fundamental questions about our current knowledge and about the orthodoxy of studying deep-water sedimentation. In this regard, the motivation behind this book is the necessity to go back to the basics on gravity-driven processes and re-evaluate our current understanding of deep-water sedimentation. In accomplishing this objective, I have used 540 case studies or datasets worldwide. Some of the recent developments are the primary focus of this book, which include mass-transport deposits (MTDs, Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows), gravity flows (Chapter 3: Gravity Flows: Debris Flows, Grain Flows, Liquefied/Fluidized Flows, Turbidity Currents, Hyperpycnal Flows, and Contour Currents), density plumes (Chapter 5: Density Plumes: Types, Deflections, and External Controls), hyperpycnal flows (Chapter 6: Hyperpycnal Flows), bottom currents (Chapter 8: Bottom Currents), and sediment deformation (Chapter 9: Soft-Sediment Deformation Structures). Despite its history of over 135 years, since the first description

Mass Transport, Gravity Flows, and Bottom Currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00001-1

1

© 2021 Elsevier Inc. All rights reserved.

2

1. Introduction

of density plumes in Swiss lakes (Forel, 1885), the domain of deep-water sedimentation and related gravity-driven processes is still an evolving field. This book is a just a momentary stopover in a long journey.

1.2 History The domain of deep-water sedimentation has a long tradition of timely literature since the 1930s (Kuenen, 1937, 1951; Bouma, 1962; Bouma and Brouwer, 1964; Middleton and Hampton, 1973; Lowe, 1982; Cook and Enos, 1977; Stanley and Kelling, 1978; Doyle and Pilkey, 1979; Stanley and Moore, 1983; Stow and Piper, 1984; Allen, 1985; Bouma et al., 1985; Postma et al., 1988; Pickering et al., 1989; Mutti, 1992; Piper et al., 1997; Stow and Fauge`res, 1998; Shanmugam, 2006a,b, 2008a,b, 2012a,b,c, 2013a,b, 2014a, 2015a,b, 2016a,b,c, 2017a,b,c,d, 2018a,b,c,d,e,f, 2019a,b,c,d, 2020b; Rebesco and Camerlenghi, 2008; Kneller et al., 2009; Mosher et al., 2010; Mulder, 2011; Pickering and Hiscott, 2015; Gordon, 2019; de Castro et al., 2020; Fonnesu et al., 2020; Fuhrmann et al., 2020, Hu¨eneke et al., 2020, among many others). However, there are still no published books focusing solely on downslope and alongslope processes and deposits. This book is an attempt to fill this void. By design, this volume is aimed at gathering and displaying the complex nature of processes and deposits. Also, by design, this book does not offer new facies models. The purpose here is to erect a firm foundation based strictly on empirical data, without the distractions of facies models. In short, this book is taking inventory of what we know and what we do not know.

1.3 Universal case studies The most important part of this book is its strength derived from 540 universal case studies. Images of most of these case studies are included. The following is a list of case studies used. Extraterrestrial • Planet Mars

3 (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows and Chapter 9: Soft-Sediment Deformation Structures)

• Planet Jupiter’s Moon Callisto

1 (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows)

• Dwarf Planet Ceres (Occator Crater)

1 (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows)

Terrestrial • Earth: submarine MTD sites

29 (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows)

• Earth: subaerial MTD sites

22 (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows)

• Earth: lacustrine environments

17 (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows, Chapter 5: Density Plumes: Types, Deflections, and External Controls, Chapter 6: Hyperpycnal Flows, and Appendix A)

• Earth: sandy debris flows (experiment)

1 (Chapter 3: Gravity Flows: Debris Flows, Grain Flows, Liquefied/ Fluidized Flows, Turbidity Currents, Hyperpycnal Flows, and Contour Currents)

Mass Transport, Gravity Flows, and Bottom Currents

1.5 Objectives

• Earth: COMFAN I (modern fans)

21 (Chapter 4: A Paradigm Shift)

• Earth: COMFAN I (ancient fans)

10 (Chapter 4: A Paradigm Shift)

• Earth: modern river mouths

29 (Chapter 5: Density Plumes: Types, Deflections, and External Controls)

• Earth: density plumes

45 (Chapter 6: Hyperpycnal Flows)

• Earth: cyclonic bottom flows

22 (Chapter 7: Triggering Mechanisms of Downslope Processes)

• Earth: deep-water masses

28 (Chapter 8: Bottom Currents)

• Earth: deep-water bottom currents

68 (35 1 33; Chapter 8: Bottom Currents)

• Earth: internal waves and tides

51 (Chapter 8: Bottom Currents)

• Earth: submarine canyons

15 (Appendix A)

• Earth: SSDS (modern and ancient)

110 (Chapter 9: Soft-Sediment Deformation Structures)

• Earth: scientific drilling sites

30 (Chapter 9: Soft-Sediment Deformation Structures)

• Earth: impact cratering (experiment)

1 (Chapter 9: Soft-Sediment Deformation Structures)

• Earth: G. Shanmugam’s case studies

36 (all chapters)

Total number of case studies

540

3

Note: Additional case studies are mentioned in the text with relevant references but they are not counted in the above list.

1.4 Environments and processes Based on the available empirical data, a general distribution of processes is illustrated in Fig. 1.1. Clearly, mass-transport processes, composed of slides, slumps, and debris flows, occur on the entire spectrum of all environments. Gravity flows and bottom currents are somewhat restricted (Fig. 1.1). I have documented this distribution using case studies from modern environments.

1.5 Objectives The primary objective is to combine process sedimentology with physical oceanography in explaining downslope and alongslope processes, their deposits, and related problems. Specific objectives are: • to provide basic definitions of individual process types, and discuss identification criteria of deposits; • to document the remarkable cosmic congruity in geometry of MTDs (slides, slumps, and debris flows) in Mars, Jupiter, and Earth (i.e., subaerial, sublacustrine, and submarine environments) (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows); • to group hyperpycnal flows and contour currents along with other gravity flows, such as debris flows, liquefied/fluidized flows, grain flows, and turbidity currents (Chapter 3: Gravity Flows: Debris Flows, Grain Flows, Liquefied/Fluidized Flows, Turbidity Currents, Hyperpycnal Flows, and Contour Currents); Mass Transport, Gravity Flows, and Bottom Currents

4

1. Introduction

FIGURE 1.1 Generalized distribution of processes (horizontal arrows) in various sedimentary environments. Note debris flows are considered under both ’mass transport’ and ’gravity flows’ categories. Also note that hyperpycnal flows, a type of gravity lows, is shown as a separate type because of its emphasis in this book (Chapter 6). Additional labeles on processes and related arrows are added by G. Shanmugam. Credit: Principales medios sedimentarios.svg. Wikipedia.

• to illustrate a paradigm shift from turbidites to MTD and bottom-current reworked sands based on empirical data (Chapter 4: A Paradigm Shift); • to document deflections of sediment plumes and their implications for understanding sediment transport and provenance (Chapter 5: Density Plumes: Types, Deflections, and External Controls); • to emphasize the basic problems surrounding the concept of hyperpycnal flows (Chapter 6: Hyperpycnal Flows); • to identify controversies based on 38 published academic discussions associated with processes and deposits (Chapter 6: Hyperpycnal Flows); • to review 21 triggering mechanisms of downslope processes (Chapter 7: Triggering Mechanisms of Downslope Processes); • to discuss four basic types of bottom currents and their deposits (Chapter 8: Bottom Currents);

Mass Transport, Gravity Flows, and Bottom Currents

1.7 Other aspects of the book

5

• to document 110 types of soft-sediment deformation structures (SSDS) in various environments and to discuss related problems (Chapter 9: Soft-Sediment Deformation Structures).

1.6 Organization Research suggests that a consistent structure will become increasingly important as readers search for content online. For this reason, I have organized each chapter in this book with a consistent structure and content subheadings for each process or topic as follows: 1. 2. 3. 4. 5. 6. 7.

definition origins identification case studies facies models problems synopsis

However, this organization may differ in some cases depending on availability of data and publications.

1.7 Other aspects of the book • In covering divergent topics, such as environments (i.e., extraterrestrial impact craters, subaerial alluvial fans, river-mouth plumes, shelf, carbonate platform, slope, basin, and lake), processes (i.e., mass transport, gravity flows, and bottom currents), triggering mechanisms, and laboratory experiments, I have relied on publications of different vintages. These constraints do not lend themselves into conveying a coherent story on downslope and alongslope processes. Therefore this book is organized, somewhat unorthodoxly and disjointedly, into 10 chapters. • In minimizing tedious text, main points are listed using bulleted or numbered format. • Important points are illustrated profusely with color images. • Self-citations are necessary in order to cover my recent contributions, particularly since 2012 when my previous book was published (Shanmugam, 2012a), on key topics such as MTD (landslides), submarine fans, contourites, seismites, SSDS, breccias, and hyperpycnites. • Unlike other books on deep-water sedimentation (e.g., Hu¨eneke and Mulder, 2011; Pickering and Hiscott, 2015), an important contribution of this book is to identify and discuss unresolved issues and related controversies. In this regard, I am uniquely qualified because of my participation in 38 academic discussions during a period of 36 years (1983 2019) (Chapter 6: Hyperpycnal Flows). By necessity, I have repeated key information from these debates (e.g., references, brief text sections, one figure, and one data table) for cohesive flow of thoughts and continuity.

Mass Transport, Gravity Flows, and Bottom Currents

6

1. Introduction

• Published academic discussions on many topics are routinely integrated in each chapter. A complete list is included in Chapter 6, Hyperpycnal Flows. • Individual chapters on density plumes, hyperpycnal flows, and SSDS are unique to this book because these topics are not covered in great detail in textbooks on deep-water sedimentation. • A comprehensive list of references, 1,612 in number, is included. A large number of color images (320), which include satellite images, radar images, aerial photographs, seismic profiles, bathymetric images, snapshots of high-speed movies of experiments, outcrop and core photographs, are included in illustrating subtle differences in color of water due to sediment concentration, bathymetry, etc. • Appendix A is included with a glossary of important terms, explanation of selected concepts, and basic methods of sediment/rock description. In providing conceptual clarity to confusing application of hybrid flows to flow transformation, the basic term “Hybrid” is discussed using a dictionary definition, etymolological context, common examples, conceptual diagram, and recent references. • Appendix B is included with a link to YouTube video of flume experiments on sandy debris flows.

1.8 Synopsis This book is a one-stop knowledge source on deep-water processes and their deposits. It is a compilation of empirical data on gravity-induced sediment movements in both downslope and alongslope environments. A total of 540 case studies are used. Although the primary focus is on deep-water settings, other environments covering terrestrial, shallow-water, lacustrine, and extraterrestrial are considered. This book does not promote genetic facies models because available data suggest that most processes are complex transitional and hybrid kinds rather than end-member types. There are no shortcut means (i.e., facies models) to interpreting deep-water processes. I am hopeful that this universal case study based approach has the potential to minimize confusion and to enhance clarity on gravity-driven processes.

Mass Transport, Gravity Flows, and Bottom Currents

C H A P T E R

2 Mass transport: slides, slumps, and debris flows O U T L I N E 2.1 Introduction 2.1.1 Cosmic congruity 2.1.2 Background information

8 8 14

2.2 International projects and symposiums

34

2.3 Mechanics of sediment failure and sliding

35

2.4 Soil strength and slope stability

35

2.5 The role of excess pore-water pressure

37

2.6 Nomenclature and classification 2.6.1 Landslide versus mass transport 2.6.2 Subaerial processes based on the types of movement and material 2.6.3 Subaqueous processes based on mechanical behavior 2.6.4 Processes based on transport velocity 2.6.5 Excessive synonyms

2.8.1 2.8.2 2.8.3 2.8.4 2.8.5 2.8.6

38 38 45 45 48 49

2.7 Recognition of the three basic types of mass-transport deposits 49 2.7.1 Process sedimentology 50 2.8 Slides

Mass Transport, Gravity Flows, and Bottom Currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00002-3

Definition Origin Identification Case studies Facies models Problems

50 54 55 63 63 63

2.9 Slumps 2.9.1 Definition 2.9.2 Origin 2.9.3 Identification 2.9.4 Case studies 2.9.5 Facies models 2.9.6 Problems

63 63 63 71 72 72 72

2.10 Debris flows: a prelude

73

2.11 Long-runout mechanisms 2.11.1 Basic concept 2.11.2 Subaerial environments 2.11.3 Submarine environments 2.11.4 Extraterrestrial environments 2.11.5 H/L ratio problems

74 74 79 79 80 81

2.12 Reservoir characterization

83

2.13 Synopsis

87

50

7

© 2021 Elsevier Inc. All rights reserved.

8

2. Mass transport: slides, slumps, and debris flows

2.1 Introduction 2.1.1 Cosmic congruity The following examples reveal a remarkable similarity in geometries of mass-transport deposits (MTDs) that occur on Mars, Jupiter, and Earth (Fig. 2.1). 1. Extraterrestrial environment: Tongue and slab geometries of MTD have been observed on the slopes of Mars (Fig. 2.2). 2. Submarine environment: Tongue geometry of MTD has been observed on the slopes of the US Pacific Margin (Shanmugam, 2012a) (Fig. 2.3). FIGURE 2.1 Diagram showing four planets (i.e., Venus, Earth, Mars, and Jupiter’s moon) with observed mass-transport deposits (MTD). Source: NASA. Heading and arrows by G. Shanmugam.

FIGURE 2.2 Image showing tongue and slab geometries of MTD on the Mar’s Cerberus Fossae, which is a steep-sided set of troughs cutting volcanic plains to the east of Elysium Mons. Source: NASA/JPLCaltech/University of Arizona. Labels added by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

2.1 Introduction

9

FIGURE 2.3 Perspective view of Los Angeles Margin bathymetry showing distribution of mass-transport deposits (MTDs) and submarine canyons. The Palos Verdes MTD (dashed line) with tongue geometry covers about 50 km3 (see details in Dartnell and Gardner, 1999), and it has been dated to be about 7500 years BP (Normark et al., 2004). Earthquakes are the major cause of MTD on the U.S. Pacific margin. Vertical exaggeration is 6%. Source: Gardner, J.V., Dartnell, P., Christopher Stone, J., Mayer, L.A., Hughes Clark, J.E., 2002. Bathymetry and selected perspective views Offshore Greater Los Angeles, California. U.S. Geological Survey, Water-Resources Investigations Report 02-4126 and U.S. Geological Survey. http://wrgis.wr.usgs.gov/dds/dds-55/pacmaps/la_pers2.htm (accessed 15.06.04.). USGS. Labels added by G. Shanmugam.

3. Submarine environment: Tongue geometry of MTD has been observed on the slopes of the US Atlantic Margin (Embley, 1980; Shipboard Scientific Party, 1994) (Fig. 2.4). 4. Submarine environment: Tongue geometry of MTD has been mapped on the slopes of East Breaks area, Gulf of Mexico (Fig. 2.5) with core information (Rothwell et al., 1991; Woodbury et al., 1978) (Fig. 2.6). 5. Submarine environment: Tongue geometry of MTD has been interpreted on the Norwegian-Barents Sea Margin (Elverhøi et al., 1997) (Fig. 2.7). 6. Sublacustrine environment: Tongue geometry of MTD has been observed on the floor of Lake Tahoe, California and Nevada (Fig. 2.8). 7. Subaerial environment: Tongue geometry of MTD has been observed in 2016 on the slopes of Su Village, Suichang County, Zhejiang Province, China (Ouyang et al., 2019) (Fig. 2.9).

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 2.4 (A) Map showing the distribution of MTD with tongue geometry on the U.S. Atlantic Continental Margin. Sediment cores from the tongue deposits show true muddy debrites. Note position of ODP Leg 150, Site 905 (filled red circle) added in this study. (B) Core photograph showing brecciated chalk clast. (C) Core photograph showing brecciated chalk clast (Eocene) in sandy clay matrix. Note that the upper sharp edge of the chalk clast was originally interpreted as fault boundary by the Shipboard Scientific Party (1994). (D) Core photograph showing folds and flow structures in sandy clay matrix. Core features from Site 905 are typical of MTD. Red arrow points to site location. Source: (A) Embley, R.W., 1980. The role of mass transport in the distribution and character of deep-ocean sediments with special reference to the North Atlantic. Mar. Geol. 38, 2350, with permission from Elsevier. (B) From Shipboard Scientific Party, 1994. Site 905. In: Mountain, G.S., Miller, K.G., Blum, P., et al., Proceedings of Ocean Drilling Program, Initial Reports, 134. College Station, Texas, pp. 255308, https://doi. org/10.2973/odp.proc.ir.150.109.1994, with additional labels by G. Shanmugam. (C and D) Color photograph courtesy of J.E. Damuth. (BD) From Shipboard Scientific Party (1994), with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251320. Elsevier. FIGURE 2.5 Bathymetric image showing tongue geometry of MTD on the slopes of East Breaks area, Gulf of Mexico. Source: National Geophysical Data Center (NGDC), 2004. NGDC Coastal Relief Model, vol. 4. Central Gulf of Mexico. https://www.ngdc. noaa.gov/thredds/catalog/crm/catalog.html?dataset 5 crmDatasetScan/ crm_vol4.nc. Labels by G. Shanmugam.

2.1 Introduction

11

FIGURE 2.6 (A) Two debris-flow tongues of the “East Breaks slide.” (B) Core taken from the western tongue is composed of contorted sand and mud (Facies 1 and 2). Note that the Ida Green Cruise cores also show Facies 1 and 2. Position of core 146 is shown by a solid dot on the western tongue. Source: (A) Modified after Rothwell, R. G., Kenyon, N.H., McGregor, B.A., 1991. Sedimentary features of the south Texas continental slope as revealed by side-scan sonar and high-resolution seismic data. AAPG Bull., 75, 298312. (B) From Woodbury, H.O., Spotts, J.H., Akers, W.H., 1978. Gulf of Mexico continental slope sediments and sedimentation. In: Bouma, A.H., Moore, G.T., Coleman, J.M. (Eds.), Framework, Facies, and Oil-Trapping Characteristics of the Upper Continental Margin, AAPG Studies in Geology 7, pp. 117137. Reprinted with permission of the American Association of Petroleum Geologists whose permission is required for further use.

8. Subaerial environment: Slab (slide) and tongue (debris flow) geometries of MTD have been observed on the slopes of La Conchita area during the 2005 event (NOAA-USGS Debris Flow Task Force, 2005) (Fig. 2.10). 9. Subaerial environment: Tongue geometry of MTD has been observed in 2006 on the slopes of Southern Leyte, Philippines (M.D. Kennedy, US Navy) (Fig. 2.11). 10. Laboratory experimental setting: Tongue geometry of MTD has been reproduced in flume experiments (Shanmugam, 2012a) (Fig. 2.12). 11. Extraterrestrial environment: Fan geometry of MTD has been observed within impact craters of Jupiter’s moon Callisto (Fig. 2.13). 12. Extraterrestrial environment: Fan geometry of MTD has been observed on the Thaumasia Plateau of Mars (Montgomery et al., 2009) (Fig. 2.14). 13. Submarine environment: Fan geometry of MTD has been photographed undersea at the mouth of the Los Frailes Canyon, Baja California (Shepard and Dill, 1966) (Fig. 2.15).

Mass Transport, Gravity Flows, and Bottom Currents

12

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.7 Map view of the “Bear Island Fan” showing modern debris tongues on the Norwegian-Barents Sea Continental Margin. These features were originally described as ‘debris lobes’ by Elverhoi et al. (1997). In this book, these features are referred to as “tongues” in order to distinguish them from “lobes.” Source: Modified after Elverhøi, A., Norem, H., Anderson, E.S., Dowdeswell, J.A., Fossen, I., Haflidason, H., et al., 1997. On the origin and flow behavior of submarine slides on deep-sea fans along the Norwegian-Barents Sea continental margin. Geo-Mar. Lett., 17, 119125. After Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476. Elsevier.

FIGURE 2.8 Oblique view of western side of Lake Tahoe and surrounding mountains looking towards the northwest (California). Underwater section is in blue tones and land is in brown tones. View shows large failure on the western lake margin, off McKinney Bay. Debris tongue (dashed line) is 7.5 km wide and 9 km long. Thin distal area is about 15 m thick, large blocks within the debris tongue are up to 20 m high. Top of the failure is at about 1635 m and toe of the debris tongue is at 1434 m elevation above sea level. Vertical exaggeration is 5%. See Gardner et al. (2000) for details. Source: U.S. Geological Survey. http://wrgis.wr.usgs.gov/dds/ dds-55/ pacmaps/lt_persp.htm (accessed 15.06.04). USGS. Labels by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

2.1 Introduction

13

FIGURE 2.9 (A) Outcrop photograph showing tongue geometry of MTD on the slopes of Sur village, China. This sediment failure, which occurred on September 28, 2016, killed 27 people and destroyed 20 house. (B) Image showing long-distance travel of MTD under cohesionless conditions. Source: After Quyang et al., 2019. Additional labels by G. Shanmugam. Springer. Public Domain.

14. Subaerial environment: Alluvial fan geometry of MTD has been observed in Borrego Springs, San Diego County, California (Lancaster et al., 2015) (Fig. 2.16). 15. Subaerial environment: Talus cone geometry of MTD has been observed at Isfjord, Svalbard, Norway (Fig. 2.17). 16. Submarine environment: Fan/lobe geometry of MTD has been observed in deep-marine Offshore Santa Barbara, California (Greene et al., 2006) (Fig. 2.18) with three distinct lobes (Fig. 2.19). 17. Submarine environment: Fan/lobe geometry of MTD has been observed in deep-marine Offshore Krishna-Godavari Basin, Bay of Bengal, India on root meansquare seismic amplitude map (Shanmugam et al., 2009) (Fig. 2.20) and on bathymetric image (Fig. 2.21). 18. Extraterrestrial environment: Slab geometry of MTD has been observed along Occator crater wall on Ceres, a dwarf planet (Fig. 2.22). 19. Subaerial environment: Slab (slide) geometry of MTD has been observed by G. Shanmugam near Gubbio, Italy (Fig. 2.23). The above examples clearly demonstrate that MTD as a collective geologic phenomenon is capable of generating similar geometries both on Earth and on Mars. However, MTD is

Mass Transport, Gravity Flows, and Bottom Currents

14

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.10 Outcrop photograph showing scar, slab (i.e., slide) and tongue (i.e., debris flow) geometry of MTD on the slopes of La Conchita area, California. This sediment failure, which occurred on January 10, 2005, killed 10 people and damaged dozens of houses (NOAA-USGS Debris Flow Task Force, 2005). Source: Photo: Mark Reed, USGS. All labels by G. Shanmugam.

composed of three end-member processes, namely slides, slumps, and debris flows. It is unclear as to what is the principal cause of the MTD geometry. Is it a product of one of three processes or a product of all three processes. In answering these questions, first we need to understand each process individually. Therefore, the purpose of this chapter is to define and distinguish three end-member processes and their deposits. This cosmic congruity necessitates an in-depth evaluation of the basics of masstransport processes and their deposits.

2.1.2 Background information Since the birth of modern deep-sea exploration by the voyage of H.M.S. Challenger (December 21, 1872 to May 24, 1876) (Fig. 2.24), organized by the Royal Society of London and the Royal Navy (Murray and Renard, 1891), oceanographers have made considerable progress in understanding the world’s oceans. Similarly, phenomenal improvements in our understanding of sedimentological processes have been made through theoretical (e.g., Varnes, 1958; Bagnold, 1962; Dott, 1963; Sanders, 1965; Nardin et al., 1979; Lowe,

Mass Transport, Gravity Flows, and Bottom Currents

2.1 Introduction

15

FIGURE 2.11 Aerial photograph showing tongue geometry of MTD on the slopes of the Philippine province of Southern Leyte. This sediment failure, which occurred on February 17, 2006, killed 1127 people. Source: Photo was taken from a U.S. Navy helicopter by Michael D. Kennedy, Travel distance of MYD is about 4 km. Posted by Wikipedia. Additional labels by G. Shanmugam. FIGURE 2.12 Flume experiments of debris flows showing tongue-like sediment in front of the channel mouth. Note sharp and irregular fronts (i.e., snouts). Flow from upper right to lower left. Source: Photo courtesy of J. Marr (St. Anthony Falls Laboratory, University of Minnesota). After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 2.13 NASA’s Galileo image showing fan geometry (arrows) of MTD has been observed within two large impact craters of Jupiter’s moon Callisto. This image was acquired on September 16th, 1997 by the SolidState Imaging (CCD) system on NASA’s Galileo spacecraft, during the spacecraft’s tenth orbit around Jupiter. North is to the top of the image, with the sun illuminating the scene from the right. The center of this image is located near 25.3 degrees north latitude, 141.3 degrees west longitude. The image, which is 55 km (33 miles) by 44 km (26 miles) across, was acquired at a resolution of 100 m per picture element. Source: NASA/ JPL. Labels by G. Shanmugam.

FIGURE 2.14 Shaded-relief map of the Thaumasia Plateau [Thermal Emission Imaging System infrared (THEMIS IR) base] showing the broad-scale, inferred sense of translation during gravity spreading, and the boundaries of “microplates” (dashed lines) and outflow channels (open circles). Source: After Montgomery, D.R., Som, S.M., Jackson, M.P.A., Schreiber, B.C., Gillepsie, A.R., 2009. Continental-scale salt tectonics on Mars and the origin of Valles Marineris and associated outflow channels. GSA Bull. 121, 117133. GSA.

2.1 Introduction

17

FIGURE 2.15 Underwater photograph showing a small submarine fan built out in one of the main branches of Los Frailes Canyon at a depth of 29 m (95 ft.), Baja California. This fan was built by deposition from sand flows (i.e., sandy debris flows), and it is analogous to subaerial alluvial fans. See also Dill (1964). Source: Photo by R.F. Dill. After Shepard, F.P., Dill, R.F., 1966. Submarine Canyons and Other Sea Valleys. Rand McNally & Co., Chicago, p. 381. Rand McNally & Company. After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524. Elsevier. Labels by G. Shanmugam.

FIGURE 2.16 Photograph showing a classic alluvial fan geometry of MTD. Borrego Springs, San Diego County, California (Lancaster et al., 2015). Labels by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 2.17 Photograph showing talus cone geometry of MTD, Isfjord, Svalbard, Norway. Source: Photo by Mark A. Wilson, Wikipedia. Labels by G. Shanmugam.

FIGURE 2.18 (A and B) Submarine environment: Fan/lobe geometry of an MTD complex has been observed in deep-marine Offshore Santa Barbara, California (Greene et al., 2006). See Figure 2.19 for three lobes. Source: EGU.

FIGURE 2.19 EM300 multibeam bathymetric image of “Goleta slide” showing three planform lobate forms, offshore Santa Barbara, California. They are composed of slumps and mud flows. Samples from this failure contain gravel, sand, and mud lithofacies. Age: 300 years BP. Source: Simplified after Greene, H.G., Murai, L.Y., Watts, P., Maher, N.A., Fisher, M.A., Paull, C.E., et al., 2006. Submarine landslides in the Santa Barbara Channel as potential tsunami sources. Nat. Hazards Earth Syst. Sci. 6, 6388., image courtesy of H.G. Greene. EGU. Additional labels by G. Shanmugam.

FIGURE 2.20 (A) Sedimentological log showing an amalgamated massive sand interval with floating quartz granules and floating mudstone clasts indicating deposition from sandy debris flows. This sandy interval corresponds to the lobate form 1, which is bright red in seismic amplitude map. Hence, bright-red amplitude areas are inferred to be gas-charged sand in our study area. (B) RMS seismic amplitude map (25 ms time window) showing sinuous planform geometries and canyon-mouth lobate forms. SF 3, well-developed sinuous form, with 90_deflections (deflected arrow), is at least 22 km long along its thalweg. Continuous bright-red amplitude in sinuous forms suggests continuous distribution of sand along the entire length of the sinuous canyon. The lobate form 1, which is 3 km long and 2.5 km wide, corresponds to the cored interval of amalgamated sandy debrites (more than 10 m thick). Source: After Shanmugam, G., Shrivastava, S.K. Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, Offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736756, with permission from SEPM.

20

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.21 Bathymetric image of our study area showing locations of three cored wells (red dots), widespread distribution of mass-transport deposits (i.e., slides, slumps, and debrites), and incipient submarine canyons on the modern upper-slope setting just seaward of the shelf edge. Linked occurrences of headwall scarps (slide scars) near the shelf edge, chutes immediately downslope of slide scars, and slide blocks immediately downslope of chutes are evident. Mass-transport deposits show slope-confined lobate forms in intercanyon areas. Background scale (0, 500, 1000, 1500, 2000, and 2500 m) represents present-day water depths. Source: After Shanmugam, G., Shrivastava, S.K. Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, Offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736756, with permission from SEPM.

1982; Shanmugam, 1996a, 1997a), experimental (e.g., Bagnold, 1962; Kuenen, 1966; Middleton, 1970; Hampton, 1972; Postma et al., 1988; Shanmugam, 2000a; Marr et al., 2001), and modern and ancient field observations (Shepard and Dill, 1966; Mutti and Ricci Lucchi, 1972; Damuth and Embley, 1981; McPherson et al., 1987; Maltman, 1994c; Hampton et al., 1996; Elverhøi et al., 1997; Piper et al., 1997; Moscardelli et al., 2006; Montgomery et al., 2009; Mosher et al., 2010; Alsop and Marco, 2013; Madrussani et al., 2018; Palladino et al., 2019, among others). Nevertheless, the physical processes that are responsible for transporting sediment downslope into the deep sea are still poorly understood. This is simply because the physics and hydrodynamics of these processes are difficult to observe and measure directly in deep-marine environments. This observational impediment has created an enormous challenge for understanding and communicating the mechanics of gravity-driven downslope processes with clarity. Furthermore, deepmarine environments are known for their complexity of processes and their deposits

Mass Transport, Gravity Flows, and Bottom Currents

2.1 Introduction

21

FIGURE 2.22

Image showing slab geometry of MTD has been observed along Occator crater wall on Ceres, a dwarf. This image of MTD along Ceres’ Occator Crater’s eastern rim was obtained by NASA’s Dawn spacecraft on June 9, 2018 from an altitude of about 27 miles (44 km). Source: NASA/JPL-Caltech/UCLA/MPS/DLR/IDA. Labels by G. Shanmugam.

FIGURE 2.23 Photograph showing downslope subaerial slabs (i.e., slide) in step-like segments. One could apply the term “creep” to this slow-moving mass, but this velocity-based term is not practical for interpreting ancient deposits. Note transformation of slide into debris flows with angular clasts in the frontal zone (bottom left). Tree in the center is about 3 m tall. Near Gubbio, Italy. Source: Photo by G. Shanmugam. After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

22

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.24 The first deep-sea exploration carried out by the H.M.S. Challenger and its discovery of the “Challenger Deep,” which is the world’s deepest known point (B11 km), in the Pacific Ocean at the southern end of the Mariana Trench. Source: Wikipedia.

(Fig. 2.25), dominated by MTDs and bottom-current reworked sands (Shanmugam, 2006a, 2012a). Thus, a plethora of confusing concepts and classifications exist. Second, perhaps the most striking challenge is in distinguishing soft-sediment deformation structures (SSDS) formed by mass-transport deposition (MTD) from those associated with tectonics (see Chapter 9: Soft-Sediment Deformation Structures). In this regard, Palladino et al. (2019) state that “. . . distinguishing structures formed by soft-sediment deformation during mass transport from those produced by contractional tectonics can be problematic. In fact, deformation occurring along detachment levels may completely obliterate the original sedimentary fabric. Although a number of advances have been made during recent decades, field criteria for discriminating structures within MTDs that are overprinted by later regional contraction are not readily applicable to all the exposed examples. We address some of these general issues through a detailed case study of the Monte Facito Formation in Italy.” This problem was addressed by pioneering workers, such as Helwig (1970) and Waldron and Gagnon (2011). Third, MTDs constitute major geohazards in subaerial environments (Glade et al., 2005; Jakob and Hungr, 2005; Geertsema et al., 2009; Kirschbaum et al., 2009). They are ubiquitous on submarine slopes (Fig. 2.3) and destructive (Hampton et al., 1996). Submarine mass movements may bear a tsunamigenic potential and are capable of methane gas release into the seawater and atmosphere (Urgeles et al., 2007). The U.S. Geological Survey (USGS, 2010) has compiled data on worldwide damages caused by large subaerial and

Mass Transport, Gravity Flows, and Bottom Currents

2.1 Introduction

23

FIGURE 2.25 Schematic diagram showing complex deep-marine sedimentary environments occurring at water depths deeper than 200 m (shelf-slope break). In general, sediment transport in shallow-marine (shelf) environments is characterized by tides and waves, whereas sediment transport in deep-marine (slope and basin) environments is characterized by gravity-driven downslope processes, such as mass transport (i.e., slides, slumps, and debris flows), and turbidity currents. Bottom currents, composed of thermohaline contour-following currents, wind-driven currents (circular motion), up and down tidal bottom currents in submarine canyons (opposing arrows), and baroclinic currents (not shown) related to internal waves/tides (Shanmugam, 2013a). Source: From Shanmugam, G., 2003a. Deep-marine tidal bottom currents and their reworked sands in modern and ancient submarine canyons. Mar. Pet. Geol. 20, 471491. Elsevier.

submarine MTDs in the 20th and 21st centuries (Table 2.1). Annual losses associated with MTDs have been estimated to be about 12 billion dollars in the United States (Schuster and Highland, 2001). During a 7-year global survey (200410), a total of 2,620 MTDs had caused a loss of 32,322 human lives (Petley, 2012). Fourth, the world’s largest submarine MTD is the Agulhas Slump in SE Africa, which is 20,331 km3 in size (Table 2.2). This submarine MTD is 1,000 times volumetrically larger than the world’s largest subaerial MTD (Saidmarreh in Iran), which is 20 km 3 in size (Table 2.2). On Mars, MTDs of immense dimensions (e.g., 3,000-km wide) have been studied (Montgomery et al., 2009, their Fig. 2.9) (see Fig. 2.14). Large submarine MTDs have important implications for developing deep-water petroleum reservoirs. In fact, many petroleum reservoirs currently produce oil and gas from sandy mass-transport deposits (SMTD) reservoirs worldwide (Shanmugam, 2006a, 2012a) (Table 2.3). Because the petroleum industry is moving exploration increasingly into the deep-marine realm to meet the growing demand for oil and gas, a clear

Mass Transport, Gravity Flows, and Bottom Currents

24

2. Mass transport: slides, slumps, and debris flows

TABLE 2.1 Worldwide large subaerial and submarine mass-transport deposits (MTDs), their sizes (volume), triggering mechanisms, and damages in the 20th and 21st centuries. Triggering mechanism

Size, damage, and loss of human life

Usoy earthquake Magnitude 7.4 Failure of ancient MTD dam

2,000,000,000 m3 54 deaths 2,000,000 m3 Length of flow: 300 km

Year

Location

Name and type

1911

Tajikistan

Usoy MTDs

1914

Argentina

Rio Barrancas and Rio Colorado debris flow

1919

Indonesia (Java)

Kelut MTDs

1920

China, Gansu, Haiyuan

Loess flows, MTDs

1920

Mexico

Rio Huitzilapan debris flows

Earthquake magnitude 6.57.0

. 40 km (length) 600870 deaths

1921

Kazakh Republic

Alma-Ata debris flow

Snowmelt, subsequent rainfall

500 deaths

1929a Grand Banks of Newfoundland, Canada

Slumps and debris flows

Grand Banks earthquake

28 deaths

1933

China (Sichuan)

Deixi MTDs

1938

Japan (Hyogo)

Mount Rokko MTDs

Deixi earthquake Magnitude 7.5 Rainfall

. 150,000,000 m3 2500 deaths 505 deaths or missing, 130,000 homes were destroyed or badly damaged

1941

Peru

Huaraz debris flow

Failure of moraine dam

10,000,000 m3 40006000 deaths

1945

Peru

Cerro Condor-Sencca MTDs

Erosional undercutting

5,500,000 m3 13 bridges were destroyed

1949

Tajikistan (Tien Shan Mtns.)

Khait MTDs

Khait earthquake Magnitude 7.4

245,000,000 m3 7200 deaths

1953

Japan (Wakayama)

Arida River MTDs

1046 deaths

1953

Japan (City of Kyoto)

Arida River MTDs

Rainfall Major typhoon (cyclone) Rainfall

1958

Japan (Shizuoka)

Kanogawa MTDs

Rainfall

1960

Chile

Rupanco region MTDs

1962

Peru (Ancash)

Nevados Huascaran MTDs

1094 deaths 19,754 homes were destroyed Valdivia earthquake 40,000,000 m3 210 deaths Magnitude 7.5 Preceded by heavy rain Not known 13,000,000 m3 40005000 deaths

1963

Italy (FriuliVenezia Giulia)

Vaiont Reservoir MTDs

Not known

Eruption of Kelut Volcano

185 km (length) Lahars caused 5110 deaths, and destroyed or damaged 104 villages Haiyuan earthquake 50,000 km2 (area) 100,000 1 deaths Magnitude 8.5

336 deaths 5122 homes were destroyed

250,000,000 m3 2000 deaths (Continued)

Mass Transport, Gravity Flows, and Bottom Currents

25

2.1 Introduction

TABLE 2.1 (Continued) Triggering mechanism

Size, damage, and loss of human life

Alaska earthquake MTDs (also known as “Prince William Sound earthquake”)

Alaska earthquake Magnitude 9.0

China (Yunnan)

MTDs

Not known

1966

Brazil (Rio de Janeiro)

MTDs

Rainfall

211,000,000 m3 Submarine MTD at Seward Turnagain Heights MTD, 9,600,000 m3 Loss: $280,000,000 (1964 dollars); 122 deaths 450,000,000 m3 444 deaths 1,000 deaths

1970

Peru (Ancash)

Nevados Huascara´n MTDs

Earthquake Magnitude 7.7

30,000,00050,000,000 m3 18,000 deaths

1974

Peru

Mayunmarca MTDs

Rainfall

1976

Guatemala

Guatemala earthquake MTDs

1980

China (Yichang, Hubei)

Yanchihe MTDs

Guatemala earthquake Magnitude 7.5 Mining activity— occurred on manmade layered slopes

1,600,000,000 m3 450 deaths 10,000 MTDs over an area of 16,000 km2 200 deaths 150,000,000 m3 284 deaths

1980

United States (Washington)

Mount St. Helens MTDs

Eruption of Mount St. Helens volcano

1983

United States (Utah)

Thistle MTDs

Snowmelt and subsequent rainfall

1983

China (Gansu)

Saleshan MTDs

Rainfall

1983

Ecuador

Chunchi MTDs

Rain and/or snow (wettest year of century)

1985

Colombia (Tolima)

Nevado del Ruiz debris flows

Eruption of Nevado 23,000 deaths del Ruiz volcano

1985

Puerto Rico (Mameyes)

MTDs

Rainfall from tropical storm

129 deaths

1986

Papua, New Guinea (East New Britain)

Bairaman MTDs

Bairaman earthquake Magnitude 7.1

200,000,000 m3

Year

Location

Name and type

1964

United States (Alaska)

1965

This is the world’s largest historical MTD 3,700,000,000 m3 250 homes, 47 bridges, 24 km of rail, and 298 km of highway were destroyed 57 deaths 21,000,000 m3 This is the most expensive disaster to fix in the US history with a loss of $600,000,000 (1983 dollars) 35,000,000 m3 237 deaths 1,000,000 m3 150 deaths

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

26

2. Mass transport: slides, slumps, and debris flows

TABLE 2.1 (Continued) Triggering mechanism

Size, damage, and loss of human life

Reventador MTDs

Reventador earthquakes, magnitude 6.1 and 6.9 and rainfall

75,000,000110,000,000 m3 1000 deaths

Venezuela

Rio Limon debris flow

Rainfall

1987

Colombia

Villa Tina MTDs

Pond leakage

1988

Brazil

Rio de Janeiro and Petropolis MTDs

Rainfall

2,000,000 m3 210 deaths 20,000,000 m3 217 deaths Approximately 300 deaths

1989

China (Huaying, Sichuan)

Xikou MTDs

Rainfall

221 deaths

1991

China (Zhaotong, Yunnan)

Touzhai MTDs

Rainfall

18,000,000 m3 216 deaths

1991

Chile

Antofagasta debris flows

Rainfall

1993

Ecuador

La Josefina MTDs

Mine excavation and heavy rainfall

500,000,000700,000,000 m3 “Hundreds” of deaths were reported 20,000,00025,000,000 m3 13 bridges destroyed

1994

Colombia (Cauca) Paez MTDs

Paez earthquake, magnitude 6.0

250 km2 (area) 272 deaths

1998

Northern India (Malpa Himalaya Region)

Large MTDs

Rainfall

221 deaths

1998

Italy (Campania)

MTDs

Rainfall

More than 100 individual slope failures

1998

Honduras, Guatemala, Nicaragua, El Salvador

MTDs

Rainfall

Hurricane Mitch caused torrential rainfall. Approximately 10,000 deaths

1999

Venezuela (Vargas, northern coastal area)

MTDs

Rainfall

1999

Taiwan

MTDs

2000

Tibet

Yigong MTD

Year

Location

Name and type

1987

Ecuador (Napo)

1987

Nearly 1 m of heavy rain fell in a 3day period There were as many as 30,000 deaths Loss: $1,900,000,000 in 2001 US dollars Chi-Chi earthquake, 11,000 km2 (area) magnitude 7.3 158 deaths Meltwater from snow and glacier

100,000,000 m3 109 deaths (Continued)

Mass Transport, Gravity Flows, and Bottom Currents

27

2.1 Introduction

TABLE 2.1 (Continued) Triggering mechanism

Size, damage, and loss of human life

Year

Location

Name and type

2001

El Salvador

MTDs, lateral spreading, liquefaction

2 earthquakes 01/13/2001: Magnitude 7.7 02/13/2001: Magnitude 6.6

2002

Russia (North Ossetia)

2003

Sri Lanka (Ratnapura and Hambantota)

2003

United States (San Bernardino County, California)

Debris flows

Rainfall

2005

Pakistan and India

MTDs

Kashmir earthquake Thousands of MTDs Magnitude 7.6 25,500 deaths

2006

Philippines (Leyte)

MTDs

Rainfall

15,000,000 m3 1,100 deaths

2008

China (Sichuan)

MTDs

2008

Egypt (East Cairo)

Al-Duwayqa MTD

Wenchuan earthquake Magnitude 8.0 Destabilization due to man-made construction

2010

Uganda (Bududa) Debris flows

Heavy rainfall

2010

Brazil (Rio de Janeiro)

Debris flows

Heavy rainfall

15,000 MTDs 20,000 deaths Still being assessed Affected area was 6,500 m3 volume and rocks weighed about 18,000 tons 107 deaths 400 1 deaths Still being assessed 350 deaths Still being assessed

2013a India (Kedarnath, Debris flows Uttarakhand)

Heavy rainfall

5,700 deaths

2014a United States (Oso, Washington)

Debris flows

Heavy rainfall

43 deaths

2014a Afghanistan (Argo District, Badakhshan Province)

Mudslide

Heavy rainfall

3502,700 deaths

The January earthquake caused MTDs over a 25,000 km2 area, (including parts of Guatemala). The February earthquake caused MTDs over a 2500 km2 area B585 deaths Kolka Glacier debris flows Detachment of large Travel distance: 19.5 km glacier, causing a 110,000,000 m3 volume of glacial ice debris flow deposited 2,000,0005,000,000 m3 of ice debris at end of runout. 125 deaths MTDs Rainfall 24,000 homes and schools destroyed 260 deaths . 1,000,000 m3 (total volume) 16 deaths

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

28

2. Mass transport: slides, slumps, and debris flows

TABLE 2.1 (Continued) Name and type

Triggering mechanism

Size, damage, and loss of human life

2017a Colombia (Mocoa)

Debris flows

Heavy rainfall

329 1 deaths

2017a Bangladesh (Rangamati, Chittagong, and Bandarban)

Debris flows

Heavy rainfall

152 deaths

2018a Philippines (Naga, Cebu)

Debris flows

Heavy rainfall

78 deaths, 5 others missing

2019a Peru (Arequipa, Moquegua, Tacna, and Ica Regions)

Debris flows

Heavy rainfall and flooding

39 deaths

Year

Location

a Added in this book: https://en.wikipedia.org/wiki/List_of_landslides#21st_century_landslides. The term “landslide” was originally used to describe these examples. Modified after USGS (U.S. Geological Survey), 2010. Catastrophic landslids of the 20th century  worldwide. ,http://landslides.usgs.gov/ learning/majorls.php. (accessed 14.06.10).

TABLE 2.2 Comparison of large-volume ( . 100 km3) mass-transport deposits (MTDs) in submarine environments with four of the largest MTDs in subaerial environments. Volume (km3)

MTD (reference)

Environment (age)

Comments

1. Agulhas Slump SE African margin (Dingle, 1977)

20,331

Submarine (postPliocene)

The world’s largest submarine MTD triggered by earthquakes

2. Chamais Slump, SE African margin (Dingle, 1980) 3. Nuuanu Debris Avalanche, NE Oahu, Hawaii (Normark et al., 1993; Moore et al., 1994)

17,433

Submarine (Neogene)

Triggered by earthquakes

5,000

Submarine (2.7 Ma, Ward, 2001)

Triggered by volcanic activity; debris avalanche is a velocity-based term (see text)

4. Storegga Slide, Offshore Norway 2,4003,200 Submarine (8100 years (Bugge et al., 1987; Haflidason BP) et al., 2005) 5. WMTD, Amazon Fan, Equatorial 2,000 Submarine (Late Atlantic (Piper et al., 1997) Pleistocene)

Triggered by earthquakes

WMTD 5 western mass-transport deposits; possibly triggered during falling sea level (Damuth et al., 1988)

6. Insular Slope Slide, Puerto Rico (Schwab et al., 1993)

1500

Submarine (Quaternary?)

Triggered by earthquakes

7. Brunei Slide, NW Borneo (Gee et al., 2007)

1200

Submarine (Quaternary?)

Triggered by sediment loading, gas hydrates, and earthquakes (Continued)

Mass Transport, Gravity Flows, and Bottom Currents

29

2.1 Introduction

TABLE 2.2 (Continued) Volume (km3)

MTD (reference)

Environment (age)

Comments

8. Saharan Debris Flow, NW African Margin (Embley, 1976; Embley and Jacobi, 1977; Gee et al., 1999)

6001100

Submarine (60,000 years BP)

Long-runout volcaniclastic debris flows of over 400 km on gentle slopes that decrease to as little as 0.05 degrees

9. Orotava-Icod-Tino Debris Avalanche, NW African slope (Wynn et al., 2000) 10. Slump Complex, Israel (FreyMartinez et al., 2005)

1000

Submarine (Pleistocene)

Debris avalanche is a velocity-based term (see text)

1000

Submarine (PlioQuaternary)

Triggered by earthquakes

900

Submarine (Late Quaternary)

Triggered by earthquakes

200800

Submarine (300,000105,000 years BP)

Triggered by volcanic activity; debris avalanche is a velocity-based term (see text)

13. Nile MTC, Offshore Egypt (Newton et al., 2004)

670

Submarine (Quaternary)

MTC 5 mass-transport complex; triggered by rapid sedimentation

14. Copper River Slide, Kayak Trough, Northern Gulf of Alaska (Carlson and Molnia, 1977)

590

Submarine (Holocene)

Possibly triggered by earthquakes and rapid sedimentation

15. MTC 1, Trinidad (Moscardelli et al., 2006)

242

Submarine (PlioPleistocene)

MTC 1 5 mass-transport complex 1; triggered by tectonic activity and rapid sedimentation

16. Cape Fear MTD, The Carolina Trough, US Atlantic Margin (Popenoe et al., 1993; Lee, 2009) 17. The 1929 Grand Banks MTD, off the US Atlantic Coast and Canada (Heezen and Ewing, 1952; Piper and Aksu, 1987; Driscoll et al., 2000; Bornhold et al., 2003) 18. Currituck Slide, US Atlantic Margin (Locat et al., 2009)

200

Submarine (Pleistocene)

Triggered by salt tectonism and gashydrate decomposition

185200

Submarine (1929)

Triggered by earthquakes (magnitude 5 7.2)

165

Submarine (2450 ka)

Triggered by earthquakes and high pore pressure

Submarine (1520 ka)

Possibly triggered by salt tectonism

Submarine (Holocene)

Triggered by salt tectonism and rapid sedimentation

Submarine (Pliocene and Pleistocene)

Retrogressive movement

11. Bassein Slide, Sunda Arc, NE Indian Ocean (Moore et al., 1976) 12. Alika 1 and 2 Debris Avalanches, NE Oahu, Hawaii (Normark et al., 1993)

19. East Breaks Slide (western lobe) B160 NW Gulf of Mexico (McGregor et al., 1993) 20. MTD, Mississippi Canyon Area, 152 Gulf of Mexico (McAdoo et al., 2000) 21. Jan Mayen Ridge, Norwegian60 Greenland Sea (Laberg et al., 2014)

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

30

2. Mass transport: slides, slumps, and debris flows

TABLE 2.2 (Continued) Volume (km3)

Environment (age)

Comments

40

Submarine (Holocene)

Retrogressive slumps

17002000

Subaerial 2122 Ma

The world’s largest prehistoric subaerial volcanic MTD

20

Subaerial (10,370 6 120 years BP, Shoaei and Ghayoumian, 1998)

The world’s second largest prehistoric subaerial MTD triggered by earthquakes

25. Mount St. Helens, United States 2.8 (Schuster, 1983; Tilling et al., 1990)

Subaerial (May 18, 1980)

The world’s largest historic subaerial MTD triggered by volcanic eruption (USGS, 2004)

26. Usoy, Tadzhik Republic (Formerly USSR) (Bolt et al., 1975)

Subaerial (1911)

The world’s second largest historic subaerial MTD triggered by earthquakes (magnitude 5 7.4) (USGS, 2010)

MTD (reference) 22. Owen Ridge Oman Coast, Arabian Sea (Rodriguez et al., 2013) 23. Markagunt Gravity Slide, southwest Utah (United States) (Hacker et al., 2014) 24. Saidmarreh Slide, Kabir Kuh Anticline, SW Iran (Harrison and Falcon, 1938)

2.0

Note that the world’s largest submarine MTD (20,331 km3) is 10 times volumetrically larger than the world’s largest subaerial MTD (2000 km3). The term “landslide” was used to describe many of these examples by the original authors. Locations of selected examples are shown in Fig. 1. Compiled from several sources. After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166.

TABLE 2.3 Summary of 35 deep-water case studies, based on description of core and outcrop, carried out by G. Shanmugam on MTDs and SMTDs (19742011). Most case studies also contain intervals of bottom-current reworked sands (BCRS).

Location, symbol, and number in Fig. 2.31

Case studies

Thickness of core and outcrop describeda

B500 m DSDP core (selected intervals described) 1067 m Green Canyon, late Conventional Pliocene Garden Banks, middle core and piston core Pleistocene 25 wells Ewing Bank 826, Pliocene-Pleistocene South Marsh Island, late Pliocene

Comment (this paper)

1. Gulf of Mexico, United States (Shanmugam et al., 1988b)

1. Mississippi Fan, Quaternary, DSDP Leg 96

Mass-transport deposits, turbidites, bottom-current reworked sands

1. Gulf of Mexico, United States (Shanmugam et al., 1993a,b; Shanmugam and Zimbrick, 1996)

2.

Sandy mass-transport deposits and bottom-current reworked sands common

3. 4. 5.

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

31

2.1 Introduction

TABLE 2.3 (Continued)

Location, symbol, and number in Fig. 2.31

Case studies

2. California (Shanmugam and Clayton, 1989; Shanmugam, 2006a, 2012a)

6. South Timbalier, middle Pleistocene 7. High Island, late Pliocene 8. East Breaks, late Pliocene-Holocene 9. Midway-Sunset Field, upper Miocene, onshore

3. Ouachita Mountains, Arkansas and Oklahoma, United States (Shanmugam and Moiola, 1995)

10. Jackfork Group, Pennsylvanian

4. Southern Appalachians, 11. Sevier Basin, Middle Tennessee, United States Ordovician (Shanmugam, 1978; Shanmugam and Benedict, 1978) 5. Brazil (Shanmugam, 2006a, 2012a) 12. Lagoa Parda Field, lower Eocene, Espirito Santo Basin, onshore 13. Fazenda Alegre Field, upper Cretaceous, Espirito Santo Basin, onshore 14. Cangoa Field, upper Eocene, Espirito Santo Basin, offshore

Thickness of core and outcrop describeda

Comment (this paper)

650 m Conventional core 3 wells 369 m 2 outcrop sections

Sandy mass-transport deposits and bottom-current reworked sands Sandy mass-transport deposits and bottom-current reworked sands common

2,152 m 5 outcrop sections

Mass-transport deposits, turbidites, bottom-current reworked sands

200 m Conventional core 10 wells

Sandy mass-transport deposits and bottom-current reworked sands common

15. Peroa´ Field, lower Eocene to upper Oligocene, Espirito Santo Basin, offshore 16. Marlim Field, Oligocene, Campos Basin, offshore 17. Marimba Field, upper Cretaceous, Campos Basin, offshore 18. Roncador Field, upper Cretaceous, Campos Basin, offshore (Continued)

Mass Transport, Gravity Flows, and Bottom Currents

32

2. Mass transport: slides, slumps, and debris flows

TABLE 2.3 (Continued)

Location, symbol, and number in Fig. 2.31

Case studies

6. North Sea (Shanmugam et al., 1995a)

19. Frigg Field, lower Eocene, Norwegian North Sea 20. Harding Field (formerly Forth Field), lower Eocene, UK North Sea 21. Alba Field, Eocene, UK North Sea 22. Fyne Field, Eocene, UK North Sea 23. Gannet Field, Paleocene, UK North Sea 24. Andrew Field, Paleocene, UK North Sea 25. Gryphon Field, upper Paleocenelower Eocene, UK North Sea 7. UK Atlantic Margin (Shanmugam 26. Faeroe area, et al., 1995a) Paleocene, west of the Shetland Islands 27. Foinaven Field, Paleocene, West of the Shetland Islands

8. Norwegian Sea and vicinity (Shanmugam et al., 1994)

9. French Maritime Alps, Southeastern France (Shanmugam, 2002a, 2003a)

10. Nigeria (Shanmugam, 1997a; Shanmugam, 2006a, 2012a)

28. Mid-Norway region, Cretaceous, Norwegian Sea 29. Agat region, Cretaceous, Norwegian North Sea 30. Annot Sandstone, Eocene-Oligocene

31. Edop Field, Pliocene, offshore

Thickness of core and outcrop describeda

Comment (this paper)

3658 m Conventional core 50 wells

Sandy mass-transport deposits and bottom-current reworked sands common

Thickness included in the North Sea count 1 well Conventional core 1 well

Sandy mass-transport deposits and bottom-current reworked sands common

500 m Conventional core 14 wells

Sandy mass-transport deposits and bottom-current reworked sands common

610 mb 1 outcrop section (12 units described) 875 m Conventional core 6 wells

Sandy mass-transport deposits and bottom-current reworked sands common (deep tidal currents) Sandy mass-transport deposits and bottom-current reworked sands common (deep tidal currents) (Continued)

Mass Transport, Gravity Flows, and Bottom Currents

33

2.1 Introduction

TABLE 2.3 (Continued)

Location, symbol, and number in Fig. 2.31 11. Equatorial Guinea (Famakinwa et al., 1996; Shanmugam, 2006a, 2012b) 12. Gabon (Shanmugam, 2006a, 2012a)

13. Bay of Bengal, India (Shanmugam et al., 2009)

Kutei Basin, Makassar Strait (Saller et al., 2006)

Case studies 32. Zafiro Field, Pliocene, offshore 33. Opalo Field, Pliocene, offshore 34. Melania Formation, lower Cretaceous, offshore (includes four fields) 35. Krishna-Godavari Basin, Pliocene

Kutei Basin, Miocene

Total thickness of rocks described by the author

Thickness of core and outcrop describeda 294 m Conventional core 2 wells 275 m Conventional core 8 wells 313 m Conventional core 3 wells 2 wells? (Saller et al., 2006, 2008a,b)

Comment (this paper) Sandy mass-transport deposits and bottom-current reworked sands common Sandy mass-transport deposits and bottom-current reworked sands common Sandy debrites and tidalites common

Discussion of problematic turbidites (Shanmugam, 2008c, 2013b, 2014a)

11,463 m

a

The rock description of 35 case studies of deep-water systems comprises 32 petroleum-producing massive sands worldwide. Description of core and outcrop was carried out at a scale of 1:201:50, totaling 11,463 m, during 19742011, by G. Shanmugam as a Ph.D. student (197478), as an employee of Mobil Oil Corporation (19782000), and as a consultant (200011). Global studies of cores and outcrops include a total of 7832 m of conventional cores from 123 wells, representing 32 petroleum fields worldwide (Shanmugam, 2012a, 2013c). These modern and ancient deep-water systems include both marine and lacustrine settings. b The Peira Cava outcrop section was originally described by Bouma (1962), and later by Pickering and Hilton (1998, their Fig. 62), among others. In addition to the above 35 case studies on deep-water deposits, Shanmugam et al. (2000) have published a case study on tidedominated estuaries in Ecuador, which total 36 case studies by the author. Note that most SMTD examples are petroleumbearing deep-water reservoirs. DSDP, Deep Sea Drilling Project. Modified after Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166.

understanding of deep-marine processes is critical. Therefore, an objective of this chapter is to bring clarity to the classification of subaerial and submarine downslope processes by combining sound principles of fluid mechanics, soil mechanics, laboratory experiments, study of modern deep-marine systems, and detailed examination of core and outcrop. Specific objectives are: 1. to document modern and ancient examples of gravity-induced processes and their deposits worldwide, 2. to explain slope stability and sediment failure mechanics, 3. to review nomenclature and classification of downslope processes, 4. to establish criteria for recognizing depositional facies in the stratigraphic record, 5. to evaluate the potential mechanisms responsible for long-runout MTDs, and 6. to discuss reservoir characterization of MTDs.

Mass Transport, Gravity Flows, and Bottom Currents

34

2. Mass transport: slides, slumps, and debris flows

2.2 International projects and symposiums In understanding submarine mass-transport processes and their deposits, major international projects and symposiums have been organized during the past three decades. Selected examples are: 1. Arctic Delta Failure Experiment (ADFEX): 19891992 2. Geological Long-Range Inclined Asdic (GLORIA), a side-scan survey of the US Exclusive Economic Zone: 19841991 3. Sediment Transport on European Atlantic Margins (STEAM): 199396 4. European North Atlantic Margin (ENAM II): 199699 5. STRATA FORmation on the Margins (STRATAFORM): 19952001 6. Seabed Slope Process in Deep Water Continental Margin (Northwest Gulf of Mexico): 19962004 7. Continental Slope Stability (COSTA): 200004 8. IGCP-511 (IUGS-UNESCO’s International Geoscience Programme 511, which ended in 2009) and IGCP-585 (E-MARSHAL: Earth’s continental MARgins: asSessing the geoHAzard from submarine Landslides) projects (URL: http://www.igcp585.org/) (accessed 24.02.13): a. 2002: First International Symposium on “Submarine Mass Movements and Their Consequences”: Nice, France (Locat and Mienert, 2003) b. 2005: Second International Symposium on “Submarine Mass Movements and Their Consequences”: Oslo, Norway (Solheim, 2006) c. 2007: Third International Symposium on “Submarine Mass Movements and Their Consequences”: Santorini Is., Greece (Lykousis et al., 2007) d. 2009: Fourth International Symposium on “Submarine Mass Movements and Their Consequences”: Austin, Texas, United States (Mosher et al., 2010) e. 2011: Fourth International Symposium on “Submarine Mass Movements and Their Consequences”: Kyoto, Japan, 2011 (Yamada et al., 2012) f. 2013: Sixth International Symposium on “Submarine Mass Movements and Their Consequences”: Kiel, Germany (GEOMAR, 2012). g. 2015: Seventh International Symposium on “Submarine Mass Movements and Their Consequences”: Wellington, New Zealand, November 14 h. 2018: Eighth International Symposium on “Submarine Mass Movements and Their consequences”: Victoria, British Columbia, Canada, May 79, 2018 i. 2020: Ninth International Symposium on “Submarine Mass Movements and Their Consequences” (Summer, 2020): University College Dublin, Ireland, 2224 June, 2020. The symposium has been postponed until further notice due to COVID-19 (Coronavirus global pandemic). Continental margins provide an ideal setting for slope failure, which is the collapse of slope sediment from the shelf edge (Fig. 2.3). Following a failure, the failed sediment moves downslope under the pull of gravity when the shear stress exceeds the shear strength of the soil (see “Soil strength and slope stability” section below). Gravity-driven processes exhibit extreme variability in mechanics of sediment

Mass Transport, Gravity Flows, and Bottom Currents

2.4 Soil strength and slope stability

35

transport, ranging from mobility of kilometer-size solid blocks on the seafloor to transport of millimeter-size particles in suspension of dilute turbulent flows in deep-water environments (Shanmugam, 2009a). In communicating this variability without confusion, a review of mechanics of sediment failure, nomenclature, and classification of downslope processes is necessary.

2.3 Mechanics of sediment failure and sliding Sediment failures on continental margins are controlled by the pull of gravity, the source of the material (bedrock vs regolith), the strength of the soil (grain size, mineralogy, compaction, cementation, etc.), the weight of the material, the slope angle, the pore-water pressure, and the planes of weaknesses. In order to evaluate sediment failures in general, one needs to conduct a slope stability analysis for describing the sediment behavior and sediment strength during loading or deformation (Duncan and Wright, 2005; Shanmugam, 2009a, 2015a, 2018a).

2.4 Soil strength and slope stability The most fundamental requirement of slope stability is that the shear strength of the soil must be greater than the shear stress required for equilibrium (Duncan and Wright, 2005; Shanmugam, 2014a). The two conditions that result in slope instability are (1) a decrease in the shear strength of the soil and (2) an increase in the shear stress required for equilibrium. The decrease in the shear strength of the soil is caused by various in situ processes, such as an increase in pore-water pressure, cracking of the soil, swelling of clays, and leaching of salt. The increase in shear stress is induced by loads at the top of the slope, an increase in soil weight due to increased water content, seismic shaking, etc. A common method for calculating the slope stability is the “limit equilibrium analyses” in soil mechanics. A stable slope can be maintained only when the factor of safety for slope stability (F) is larger than or equal to 1 (Duncan and Wright, 2005, their Eqs. 6.1 and 13.2): F5

S Shear strength of the soil 5 $1 τ Shear stress required for equilibrium

where S 5 available shear strength, which depends on the soil weight, cohesion, friction angle, and pore-water pressure; τ 5 equilibrium shear stress, which is the shear stress required to maintain a just-stable slope. It depends on the soil weight, pore-water pressure, and slope angle.

Mass Transport, Gravity Flows, and Bottom Currents

36

2. Mass transport: slides, slumps, and debris flows

The shear strength is equal to the maximum shear stress, which can be absorbed by the slope without failure and can be defined by the MohrCoulomb failure criterion: S 5 c 1 σ tan φ where S 5 available shear strength (Fig. 2.26A); c 5 cohesion (nonfrictional) component of the soil strength; σ 5 total normal stress acting on the failure surface; ϕ 5 angle of internal friction of the soil. By combining the equations of shear strength and MohrCoulomb failure criterion, the factor of safety (F) can be expressed as: F5

c 1 σ tan ϕ τ

A sediment failure is initiated when the factor of safety for slope stability (F) is less than 1 (Fig. 2.26B). In other words, the sliding motion along the shear surface commences only when the driving gravitational force exceeds the sum of resisting frictional and cohesive forces. Initial porosity of the sediment plays a critical factor in controlling the behavior of the shear surface (Anderson and Riemer, 1995). Based on an experimental study on

FIGURE 2.26 (A) Plot showing that the shear strength of the soil (s) is composed of frictional (ϕ) and cohesive (c) components. (B) Conceptual diagram showing that a stable slope can be maintained only when the factor of safety for slope stability (F) is larger than or equal to 1 (Duncan and Wright, 2005). The sliding motion of failed soil mass commences along the shear surface when the factor of safety (F) is less than 1. Synonyms: Failure surface 5 slip surface 5 shear surface 5 primary glide plane. Source: Compiled from several sources (e.g., USACE, 2003; Duncan and Wright, 2005). From Shanmugam, G., 2014a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands: reply. AAPG Bull. 98, 858879, with permission from AAPG.

Mass Transport, Gravity Flows, and Bottom Currents

37

2.5 The role of excess pore-water pressure

FIGURE 2.27 Histogram showing frequency distribution of submarine slides with increasing slope angle, U.S. Atlantic Continental Slope. Note most slides occur on gentle slopes of less than 4 degrees. This compilation of empirical data from modern examples is helpful to the petroleum industry for understanding ancient slides and palaeogeography. Source: From Booth, J.S., O’Leary, D.W., Popenoe, P., Danforth, W.W., 1993. U.S. Atlantic continental slope landslides: their distribution, general attributes, and implications. In: Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone. U.S. Geological Survey Bulletin 2002, pp. 1422, with permission from USGS.

landslides initiated by rising pore-water pressures, Iverson (2000) reported that even small differences in initial porosity had caused major differences in mobility. For example, wet sandy soil with 50% porosity contracted during slope failure, partially liquefied, and accelerated to a speed of over 1 m s21, whereas the same soil with 40% porosity dilated during failure, slipped episodically, and traveled at a slow velocity of 0.2 cm s21. Finally, soil strength differs between drained and undrained conditions (Terzaghi et al., 1996; USACE, 2003; Duncan and Wright, 2005). Slides occur commonly on modern slopes of 14 degrees (Booth et al., 1993) (Fig. 2.27). Contrary to the popular belief, most submarine slides occur on gentle slopes of less than 4 degrees, sometimes even at 0.25 degrees. Submarine slides on slopes greater than 10 degrees are rare (Fig. 2.27).

2.5 The role of excess pore-water pressure Terzaghi (1936) first recognized that pore-water pressure controls the frictional resistance of slopes, which has remained the most important concept in understanding landslide behavior. A founding principle of slope stability is that a rise in pore-water pressure reduces the shear strength of the soil (Skempton, 1960). The shear strength of soil, in particular clays, is controlled by the frictional resistance and interlocking between particles (i.e., physical component), and interparticle forces (i.e., physicochemical component) (Karcz and Shanmugam, 1974; Parchure, 1980; Hayter et al., 2006). Furthermore, bed density and shear strength of soil increase with increasing consolidation (Hanzawa and Kishida, 1981; Dixit, 1982). A rise in pore-water pressure occurs when the saturated soil is stressed, and when the porosity cannot increase or the pore fluid cannot expand

Mass Transport, Gravity Flows, and Bottom Currents

38

2. Mass transport: slides, slumps, and debris flows

or escape through fractures. The excess pore-water pressure has been considered a vital factor in explaining the origin of subaerial mass-transport processes (Johnson, 1984; Anderson and Sitar, 1995; Iverson, 1997, 2000; Iverson et al., 1997; Jakob and Hungr, 2005). Iverson (1997), based on studies of coarse-grained subaerial debris flows, has developed a model in which excess pore-water pressure causes liquefaction of the sediment and thereby strongly reduces internal friction and increases sediment mobility or runout distances. Excess pore-water pressures have also been considered a characteristic property in explaining long-runout submarine MTDs (Suhayda and Prior, 1978; Hampton et al., 1996; Iverson et al., 1997; Gee et al., 1999; Major and Iverson, 1999). Submarine debris flows have lower yield strengths than subaerial debris flows due to entrainment of seawater (Pickering et al., 1989) and elevated pore-water pressure (Pierson, 1981). Laboratory measurements of pore-water pressure have shown that the front of the subaqueous clayey debris flow exhibits hydroplaning (Mohrig et al., 1998) on a thin layer of water, which causes low bed friction. Fronts of sandy debris flows show a fluidized head where bed friction is minimal (Ilstad et al., 2004). A 5-m thick submarine sandy debris flow, with a long-runout distance of over 400 km downslope of the Canary Islands, has been attributed to the development of excess pore-water pressure due to loading induced by a pelagic debrite package (Gee et al., 1999, their Fig. 2.12). Problems associated with long-runout MTDs are discussed below (Section 2.11).

2.6 Nomenclature and classification 2.6.1 Landslide versus mass transport Although the term “landslide” is deeply entrenched in the literature, there are inherent problems associated with the usage. For example: 1. For his first paper, Varnes (1958) used the title “Landslide types and processes” that included fall, topple, spread, translational slide, rotational slide, and flow. But for his second paper, Varnes (1978) changed the paper title to “Slope movement types and processes” to represent the same six processes, namely (1) fall, (2) topple, (3) spread, (4) translational slide, (5) rotational slide, and (6) flow (Fig. 2.28). In abandoning the term landslide, Varnes (1978, p. 11) eloquently explained that “One obvious change is the term slope movements, rather than landslides, in the title of this paper and in the classification chart. The term landslide is widely used, and no doubt, will continue to be used as an all-inclusive term for almost all varieties of slope movements, including some that involve little or no true sliding. Nevertheless, improvements in technical communication require a deliberate and sustained effort to increase the precision associated with the meaning of words, and therefore, the term slide will not be used to refer to movements that do not include sliding.” This cautionary note, which has been obviously ignored by other researchers, is the underpinning principle here (Shanmugam, 2015a).

Mass Transport, Gravity Flows, and Bottom Currents

2.6 Nomenclature and classification

39

FIGURE 2.28 Classification of subaerial slope movements into six types by Varnes (1978). Note that the term “landslide” is not used in his 1978 classification, but was used by Varnes (1958) in his previous classification. Also note that the processes fall, topple, and spread are not adopted in this article because deposits of these processes are difficult to distinguish from deposits of debris flows (i.e., debrites). See Cruden and Varnes (1996) and Wieczorek and Snyder (2009) for an expanded classification of Varnes (1978) with four additional types: (1) debris avalanche, (2) earthflow, (3) creep, and (4) lateral spread. See text for a critique of these additional terms. Source: Diagram modified after Highland, L.M., Bobrowsky, P., 2008. The Landslide Handbook—A Guide to Understanding Landslides. U.S. Geological Survey Circular 1325, Reston, Virginia, 129 p., with permission from USGS.

2. The term landslide literally implies sliding motion of a rigid body of earth or land along a shear surface. But debris flows, considered to be a part of the landslide family in some classifications (Cruden, 1991), are characterized by intergranular movements, not shear-surface movements (Shanmugam et al., 1994; Iverson et al., 1997). 3. The AGI Glossary of Geology (Bates and Jackson, 1980, p. 349) defined a landslide as “A general term covering a wide variety of mass movement landforms and processes involving the downslope transport, under gravitational influence, of soil and rock material en masse. Usually the displaced material moves over a relatively confined zone or surface of shear.” This definition, although implies that the shearsurface movement is a critical factor (see Fig. 2.26B), includes a variety of mass movements. 4. According to Cruden (1991), “A landslide is the movement of a mass of rock, earth or debris down a slope.” This broad definition includes not only slides, but also debris flows. There are at least five different definitions of the term landslide with conflicting meanings (Table 2.4).

Mass Transport, Gravity Flows, and Bottom Currents

40

2. Mass transport: slides, slumps, and debris flows

TABLE 2.4 Nomenclature of 79 different types of mass-transport processes and their deposits with overlapping and confusing meanings. Nomenclature

Characteristics

Reference

Comments

1. Landslide: Type 1 (First classification by J.D. Dana in 1862) (see Cruden, 2003) 2. Landslide: Type 2 (GSA Thematic Volume)

Refers to three processes: rock slides, earth spreads, and debris flows

Cruden (2003)

Impracticala (MTD or SMTD)

A general term used for various moderately rapid gravity-induced mass movements, which exclude creep and solifluction

Coates (1977)

Impractical (MTD or SMTD)

3. Landslide: Type 3 (AGI Glossary)

A general term for a variety of gravity-induced downslope mass movements, which include creep and solifluction

Bates and Jackson (1980)

Impractical (MTD or SMTD)

4. Landslide: Type 4 (NATO Workshop)

A sudden movement of earth and rocks down a steep slope

Saxov (1982) (see also Cruden, 1991)

Impractical (MTD or SMTD)

5. Landslide: Type 5 (USGS Handbook)

A downslope movement of rock or soil, or both, occurring on the surface of rupture in which much of the material often moves as a coherent or semicoherent mass with little internal deformation

Highland and Bobrowsky (2008) (see also Eckel, 1958)

Impractical (MTD or SMTD)

6. Fall or rockfall

Freefall of material from steep slopes

Varnes (1978)

Impractical (MTD or SMTD)

7. Sand Fall

Freefall of material at submarine canyon heads

Shepard and Dill (1966)

Impractical (SMTD)

8. Topple

Tilting without collapse

Varnes (1978)

Impractical (MTD or SMTD)

9. Slide

Coherent mass with translational movement

Dott (1963)

Slide

10. Slump

Coherent mass with rotational movement and internal deformation

Dott (1963)

Slump

11. Translational slump

Translational movement

Milia et al. (2006)

Translational movements are associated with slides (Dott, 1963), MTD

12. Drained slump

Slumping without excess pore pressure

Morgenstern (1967)

Impractical (MTD)

13. Undrained slump

Slumping with excess pore pressure

Morgenstern (1967)

Impractical (MTD) (Continued)

Mass Transport, Gravity Flows, and Bottom Currents

41

2.6 Nomenclature and classification

TABLE 2.4 (Continued) Nomenclature

Characteristics

Reference

Comments

14. Toreva block (named after the village of Toreva in Arizona, United States) 15. Spread

Backward rotational slip

Reiche (1937)

Impractical (MTD or SMTD)

Lateral extension accommodated by shear or tensile fractures

Varnes (1978)

Impractical (MTD or SMTD)

16. Debris flow

Plastic (en masse) flow with laminar state

Dott (1963), Hampton (1972)

Debrite

17. Debris avalanche

Extremely fast-moving debris flows

Varnes (1978)

Impractical (MTD or SMTD)

18. Cohesionless debris avalanche

Rolling, cascading, and collision of rock fragments on steep underwater slopes

Prior and Bornhold (1990)

Impractical (SMTD)

19. Rock-fragment flow

Large extremely rapid “rockfall-debris Varnes (1958, flows” 1978)

Impractical (MTD)

20. Debris slide

Slow-moving mass that breaks up into Varnes (1978) smaller blocks

Impractical (MTD or SMTD)

21. Flow slide (two words)

Disintegrating subaerial slide in coarse material where a temporary transfer of part of the normal stress onto the fluids of the void space, with a consequent sudden decrease in strength

Koppejan et al. (1948) (see also Rouse, 1984)

Impractical (MTD or SMTD)

22. Flow slide (two words)

High-velocity, transitional type between slumps and debris flows

Shreve (1968)

Impractical (MTD or SMTD)

23. Flow slide (one word)

Basal dense layer with viscoplastic behavior in stratified submarine sediment flows

Norem et al. (1990)

SMTD and associated turbidite

24. Marine flow slide

Liquefied marine sand with high Koning (1982) porosity and high pore-water pressure

Impractical (SMTD)

25. Retrogressive flow slide

Occurs along banks of noncohesive clean sand or silt and shows repeated fluctuations in pore-water pressure

Andresen and Bjerrum (1967)

Impractical (SMTD)

26. Deep creep

Slow-moving mass of bedrock (synonym: rock flow)

Varnes (1978, his Fig. 2.2)

Impractical (MTD or SMTD)

27. Soil creep

Slow-moving mass of fine soil

Varnes (1978, his Fig. 2.2)

Impractical (MTD or SMTD)

28. Seasonal creep

Slow-moving mass within the soil horizon affected by seasonal changes in soil moisture and temperature

Hansen (1984)

Impractical (MTD or SMTD) (Continued)

Mass Transport, Gravity Flows, and Bottom Currents

42

2. Mass transport: slides, slumps, and debris flows

TABLE 2.4 (Continued) Nomenclature

Characteristics

Reference

Comments

29. Continuous creep

Slow-moving mass where shear stress continuously exceeds the material strength

Hansen (1984) (USGS, 2004)

Impractical (MTD or SMTD)

30. Progressive creep

Slow-moving mass of associated with slopes reaching point of failure by other mass movements

Hansen (1984)

Impractical (MTD or SMTD)

31. Talus creep

Slow-moving large angular rock fragments on a gentle slope (synonym: scree creep)

Sharpe (1938)

Impractical (MTD or SMTD)

32. Slump creep

Slow-moving multiple processes

Carter and Lindqvist (1975)

Impractical (MTD or SMTD)

33. Mass creep

Slow-moving submarine slope sediments due to repeated loading effects by earthquakes

Almagor and Wiseman (1982)

Impractical (MTD or SMTD)

34. Rock-glacier creep

Slow-moving tongue of the rock glacier

Sharpe (1938)

Impractical (MTD or SMTD)

35. Solifluction (Soil flow of Varnes, 1978)

Slow-moving waterlogged soil over permafrost layers

Anderson (1906)

Impractical (MTD)

36. Earth flows

Slow- to fast-moving fine soil

Varnes (1978)

Impractical (MTD)

37. Sturzstrom (synonym: rock avalanche)

Fast-moving debris flows

Hsu¨ (1975, 2004)

Impractical (MTD or SMTD)

38. Inertia flow

Grain avalanching

Bagnold (1954)

SMTD

39. Grain flow

Sediment support by grain collision

Middleton and Hampton (1973)

SMTD

40. Fluidized flow

Full sediment support by upward intergranular flow

Middleton and Hampton (1973)

Impractical (SMTD)

41. Liquefied flow

Partial sediment support by upward intergranular flow

Lowe (1976a)

Impractical (SMTD)

42. Turbidity current

Sediment support by fluid turbulence

Middleton and Hampton (1973)

Turbidite, not MTD

43. Sand flow

A flow of wet sand that is subjected to Varnes (1958) fluctuations in pore-water pressure

Impractical (SMTD)

44. Loess flow

Intermediate stage between “liquefaction flow” and “sand flow” with increasing grain size

Impractical (MTD)

Coates (1977)

45. “High-density turbidity Stratified lower debris flow and upper Kuenen (1951) current” turbidity current

Debrite and turbidite (Shanmugam, 1996a,b)

46. Sandy debris flow

Sandy debrite

Sandy flow with plastic rheology and laminar state

Shanmugam (1996a)

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

43

2.6 Nomenclature and classification

TABLE 2.4 (Continued) Nomenclature

Characteristics

Reference

Comments

47. Cohesionless liquefied sand flow

Sliding-related sandy mass flows

Nemec (1990, his Fig. 32)

Impractical (SMTD)

48. Hyperconcentrated flow

Sediment concentration: 2060 by volume %

Pierson and Costa Impractical (SMTD) (1987)

49. Slurry flow

Cohesive debris flows

Carter (1975a)

Impractical (MTD)

50. Slurry flow

Synonym for “high-density turbidity current”

Lowe and Guy (2000)

Impractical (MTD or SMTD)

51. Lahar

Volcaniclastic debris flow

Bates and Jackson (1980)

MTD or SMTD

52. Nue´e ardente

Decoupling of pyroclastic flows (i.e., stratified flows)

Fisher (1995)

Impractical (MTD)

53. Cascading dense-water event

Analogous to “sand fall” of Shepard and Dill (1966)

Gaudin et al. (2006)

Impractical (SMTD)

54. Dense flow

Basal high-concentration layer in stratified sediment flows

Norem et al. (1990)

SMTD and associated turbidite

55. Fluidized cohesionless particle flow

Basal high-concentration layer in stratified sediment flows

Friedman et al. (1992)

SMTD and associated turbidite

56. Liquefied cohesionless coarse-particle flow

Basal high-concentration layer in stratified sediment flows

Sanders and Friedman (1997)

SMTD and associated turbidite

57. Slide

Basal high-concentration layer in stratified sediment flows

Kuenen (1951)

Impractical (MTD or SMTD)

58. Flowing-grain layer

Basal high-concentration layer in stratified sediment flows

Sanders (1965)

SMTD and associated turbidite

59. Laminar inertia flow

Basal high-concentration layer in stratified sediment flows

Postma et al. (1988)

SMTD and associated turbidite

60. Laminar sheared layer

Basal high-concentration layer in stratified sediment flows

Vrolijk and Southard (1997)

SMTD and associated turbidite

61. Traction carpet

Basal high-concentration layer in stratified sediment flows

Dzulynski and Sanders (1962)

SMTD and associated turbidite

62. Avalanching flow

Basal high-concentration layer in stratified sediment flows

Sanders (1965)

SMTD and associated turbidite

63. Mass flow

Basal high-concentration layer in stratified sediment flows

Friedman et al. (1992)

SMTD and associated turbidite

64. Mass flow

Plastic flow with shear stress distributed throughout the mass

Nardin et al. (1979)

MTD or SMTD

65. Laminar mass flow

Gradational processes involving sand flows, slumping, sliding, and spontaneous liquefaction

Carter (1975b)

SMTD

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

44

2. Mass transport: slides, slumps, and debris flows

TABLE 2.4 (Continued) Nomenclature

Characteristics

Reference

Comments

66. Granular mass flow

Concentrated grain ( . 0.06 mm)-fluid mixtures in rock avalanches, debris flows, and pyroclastic flows

Iverson and Vallance (2001)

Impractical (SMTD)

67. Hyperpycnal flow

Sinking river water that has higher density than basin water

Bates (1953), Mulder et al. (2003)

Impractical (MTD)

68. Dense flow

Multiple processes

Gani (2004)

SMTD and associated turbidite

69. Hybrid flow

Multiple processes

Houghton et al. (2009)

SMTD and associated turbidite

70. Tsunamite (deposit)

“Rope-ladder texture” and multiple processes

Michalik (1997)

Impractical Shanmugam (2006b)

71. Homogenite (deposit, Uniform texture, considered Kastens and Cita, 1981) synonymous with submarine landslide and megaturbidite

Camerlenghi et al. Turbidite, not MTD (2010) (Shanmugam, 2006b)

72. Olistostrome (deposit)

Submarine gravity sliding or slumping

Flores (1955); Hsu¨ Impractical (MTD or (1974) SMTD)

73. Gravitite (deposit)

Debris flows

Natland (1967)

Impractical (MTD or SMTD)

74. Gravite (deposit)

Slide, slump, debris flow, dense flow, and turbidity current

Gani (2004)

Impractical (MTD or SMTD)

Dzulynski et al. (1959)

Impractical (SMTD)

Mutti et al. (1984a)

Impractical (SMTD)

77. Megaturbidite (deposit) Large-scale debris flows

Labaume et al. (1987)

Impractical (SMTD)

78. Atypical turbidite (deposit)

Slumps, debris flows, and sand flows

Stanley et al. (1978)

Impractical (SMTD)

79. Duplex-like structures (deposit)

Slumps and debris flows

Shanmugam et al. (1988a)

Impractical (MTD)

75. Fluxoturbidite (deposit) Sand avalanche (see Hsu¨, 2004 for a critique of this term) 76. Seismoturbidite Large-scale mass flows (deposit)

a

In some cases, it is impractical to interpret a specific process from the rock record. In such cases, a nonspecific term of MTD or SMTD is preferred. MTD, Mass-transport deposits; SMTD, sandy mass-transport deposits. Compiled from several sources. Updated after Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 pp.

Mass Transport, Gravity Flows, and Bottom Currents

2.6 Nomenclature and classification

45

5. Geertsema et al. (2009, p. 59) state that “Landslides include debris flows and slides, earth flows and flowslides, rockfalls, slides, and avalanches, and complex landslides involving both rock and soil.” On the one extreme, the term landslide has been applied without any implication for a specific process (Gee et al., 2007; Camerlenghi et al., 2010), but on the other extreme, the term landslide represents only one category within a larger phenomenon called mass movements (Coates, 1977). 6. The U.S. Geological Survey uses the term landslide to include debris avalanche and creep with velocity connotations (Highland and Bobrowsky, 2008). However, velocities of transport processes cannot be interpreted from bathymetric images of modern seafloor or by examining the ancient rock record in core and outcrop (see Section 3.4 on “Processes based on transport velocity” below). 7. Even the International Geoscience Programme (IGCP-585), now called E-MARSHAL (Earth’s continental MARgins: asSessing the geoHAzard from submarine Landslides), uses the word landslide for all submarine mass movements (E-MARSHAL, 2013). 8. Similarly, the Springer journal “Landslides” (Editor-in-Chief: Kyoji Sassa) defines that “Landslides are gravitational mass movements of rock, debris or earth,” without a distinction between landslides and mass movement. Credit: http://www.springer.com/ earth 1 sciences 1 and 1 geography/natural 1 hazards/journal/10346 (accessed 27.12.14). The use of the term “landslide” is inappropriate as a general term to represent both the shear-surface “sliding” motion of a rigid body and the intergranular “flowing” motion of a plastic mass (Shanmugam et al., 1994). A more appropriate general term is “mass transport” or “mass movement,” which represents the failure, dislodgement, and downslope movement of either sediment or glacier under the influence of gravity. The advantage of the general term “mass transport” is that there is no built-in reference to a sliding motion. Nor is there any reference to sediment or glacier.

2.6.2 Subaerial processes based on the types of movement and material Varnes (1978, his Fig. 2.1) classified subaerial mass-transport processes into six movement-based types: (1) falls, (2) topples, (3) translational slides, (4) rotational kinds, (5) spreads, and (6) flows (Fig. 2.28). Further, Varnes (1978) added the prefix “rock” to the process names and established the material-based types: (1) rockfall, (2) rock topple, (3) rock slide, (4) rock slump, (5) rock spread, and (6) rock flow or deep creep. Although the spreads, topples, and falls could be observed in modern subaerial environments, the deposits of these three processes in the ancient rock record would not have any distinguishing attributes. This is because deposits of spreads, topples, and falls would resemble debrites (i.e., deposits of debris flows). Therefore, these three types are not adopted (Fig. 2.28).

2.6.3 Subaqueous processes based on mechanical behavior Dott (1963) proposed the most meaningful and practical classification of subaqueous mass-transport processes. It is somewhat analogous to the most widely accepted classification of subaerial mass-transport processes by Varnes (1958). In this scheme, subaqueous processes are broadly classified into (1) elastic, (2) elastic and plastic, (3) plastic, and (4) viscous fluid types based on mechanical behavior (Fig. 2.29). The elastic behavior

Mass Transport, Gravity Flows, and Bottom Currents

46

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.29 (A) Schematic diagram showing four common types of gravity-driven downslope processes that transport sediment into deep-marine environments. A slide represents a coherent translational mass transport of a block or strata on a planar glide plane (shear surface) without internal deformation. A slide may be transformed into a slump, which represents a coherent rotational mass transport of a block or strata on a concave-up glide plane (shear surface) with internal deformation. Upon addition of fluid during downslope movement, slumped material may transform into a debris flow, which transports sediment as an incoherent mass in which intergranular movements predominate over shear-surface movements. A debris flow behaves as a plastic laminar flow with strength. As fluid content increases in debris flow, the flow may evolve into Newtonian turbidity current. Not all turbidity currents, however, evolve from debris flows. Some turbidity currents may evolve directly from sediment failures. Turbidity currents can develop near the shelf edge, on the slope, or in distal basinal settings. (B) Sediment concentration (% by volume) in gravity-driven processes. Slides and slumps are composed entirely of sediment (100% by volume). Debris flows show a range of sediment concentration from 25% to 100% by volume. Note that turbidity currents are low in sediment concentration (1%23% by volume), implying low-density flows. These concentration values are based on published data by various authors (see Shanmugam, 2000, his Figure 4 for details). (C) Based on mechanical behavior of gravity-driven downslope processes, mass-transport processes include slide, slump, and debris flow, but not turbidity currents (Dott, 1963). (D) The prefix “sandy” is used for mass-transport deposits that have grain ( . 0.06 mm: sand and gravel) concentration value equal to or above 20% by volume. The 20% value is adopted from the original field classification of sedimentary rocks by Krynine (1948). See Shanmugam (1996a) for discussion on high-density turbidity currents. Source: From Shanmugam, G., Lehtonen, L.R., Straume, T., Syversten, S.E., Hodgkinson, R.J., Skibeli, M., 1994. Slump and debris flow dominated upper slope facies in the Cretaceous of the Norwegian and Northern North Seas (6167 N): implications for sand distribution. AAPG Bull. 78, 910937.

represents rockfall, the elastic and plastic behavior comprises slide and slump, the plastic behavior represents debris flow, and the viscous fluid represents Newtonian turbidity current. The importance of Dott’s (1963) classification is that mass-transport processes do not include turbidity currents (Fig. 2.29C). In this classification, a rockfall refers to sudden falling of rock fragments on steep slopes, such as submarine canyon heads. Because recognition of rockfall in the ancient record is impractical, it is not considered here as a separate

Mass Transport, Gravity Flows, and Bottom Currents

2.6 Nomenclature and classification

47

type. In short, mass-transport processes are composed of three basic types: (1) slide, (2) slump, and (3) debris flow (Fig. 2.29). I have adopted Dott’s (1963) classification in this review because theoretical analysis (Shanmugam, 1996a,b), experimental observations (Shanmugam, 2000a; Marr et al., 2001), and empirical data (Table 2.3) overwhelmingly show that turbidity currents are not mass-transport processes. The underpinning principle of Dott’s (1963) classification is the separation of solid from fluid mode of transport based on sediment concentration. In the solid (elastic and plastic) mode of transport, high sediment concentration is the norm (25%100% by volume, Fig. 2.29B). Mass-transport mechanisms are characterized by solid blocks or aggregate of particles (mass). In contrast, individual particles are held in suspension by fluid turbulence in turbidity currents (Dott, 1963; Sanders, 1965). Turbidity currents are characterized by low sediment concentration of 1%23% by volume (Fig. 2.29B). In other words, turbidity currents are innately low in flow density. A simple analogy to high-volume sediment transport by mass-transport processes is the human transport by a double-decker bus with a capacity to carry 73 passengers at a time (Fig. 2.30A). In contrast, low-volume sediment transport by

FIGURE 2.30

Comparison of human transport on land with gravity-driven sediment transport under water in illustrating the importance of sediment concentration. (A) Difference between a double-decker bus with a capacity to carry at least 73 passengers and a microcar with a capacity for only two passengers. (B) Difference between mass-transport processes with high sediment concentration (25%100% by volume) and turbidity currents with low sediment concentration (1%23% by volume). Sediment mass transport 5 bus transport. Turbidity current transport 5 microcar transport. Both bus and mass transport are extremely efficient systems for high-volume transport (long arrow). SC 5 sediment concentration. Source: Reproduced from Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166, with permission from Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

48

2. Mass transport: slides, slumps, and debris flows

turbidity currents is analogous to human transport by a microcar with a capacity to carry only two passengers at a time (Fig. 2.30B). Clearly, mass transport is a much more efficient mechanism for moving sediment downslope than a turbidity current. Mass transport can operate in both subaerial and subaqueous environments, whereas turbidity currents can operate only in subaqueous environments. The advantage of this classification is that physical features preserved in a deposit directly represent the physics of sediment movement that existed at the final moments of deposition.

2.6.4 Processes based on transport velocity The concept of velocity-based classification was first introduced by Sharpe (1938) and later adopted by Varnes (1958, 1978) for subaerial processes. There are at least 10 different factors that are commonly used in classifying landslides by various authors (Hansen, 1984, author’s Table 2.1). These factors are: (1) climate, (2) material moved, (3) coherence of material, (4) size of material, (5) geology, (6) type of movement, (7) speed of movement, (8) medium of movement: water/air/ice, (9) triggering mechanisms, and (10) morphological attributes. These 10 conflicting philosophies and related classifications have resulted in the current conceptual and nomenclatural crisis (Table 2.4). The velocity-based terms, such as avalanches, have also been adopted for downslope subaqueous processes when interpreting seismic and bathymetric data (Wynn et al., 2000; Lewis and Collot, 2001; Masson et al., 2006). Examples of velocitybased terms are as follows: 1. The term flow slide has been used for high-velocity subaerial processes that could be considered a transitional type between slumps and debris flows (Shreve, 1968; Rouse, 1984). 2. A slow-moving mass that breaks up into smaller blocks as it advances is called debris slide, whereas a fast-moving mass that breaks up into smaller blocks as it advances is called debris avalanche (Varnes, 1978). The velocity of debris avalanches is 5 m s21 (Cruden and Varnes, 1996; see also Hungr et al., 2001). 3. Catastrophic (fast-moving) debris flows are called sturzstrom (Hsu¨, 1975, 2004). 4. The term creep refers to a slow-moving mass movement (Bates and Jackson, 1980). There are nine kinds of creep depending on material and movement: (1) deep creep, (2) soil creep, (3) seasonal creep, (4) continuous creep, (5) progressive creep, (6) talus creep, (7) slump creep, (8) mass creep, and (9) rock-glacier creep (Shanmugam, 2012a). Although fast-moving and slow-moving mass-transport processes have been classified using absolute velocity values (Cruden and Varnes, 1996; Hungr et al., 2001), these velocity-based terms are not based on empirical data. For example, it is difficult to measure velocities of processes in modern deep-water environments because of common destruction of velocity meters by catastrophic mass-transport events (Inman et al., 1976; Shepard and Marshall, 1978). Cable breaks were used to estimate velocity of submarine mass-transport processes, triggered by the 1929 Grand Banks earthquake in offshore Newfoundland, Canada, which traveled at a speed of 67 km h21 (Piper et al., 1988). But such velocity values are not based on direct measurements. Therefore, we do not know

Mass Transport, Gravity Flows, and Bottom Currents

2.7 Recognition of the three basic types of mass-transport deposits

49

whether those cables were broken by slumps, debris flows, or turbidity currents. More importantly, there are no sedimentological criteria to determine the absolute velocities of sediment movement in the ancient rock record. This is because sedimentary features preserved in the deposits cannot and do not reflect absolute transport velocities. The prac˝ Diagram, meant tice of determining flow velocity from grain size using the Hjulstrom for fluvial processes (Sundborg, 1956), is inapplicable to MTD. This is because grain size is not proportional to flow velocity in mass-transport processes. In California, for example, it has been well documented that a slow-moving (at velocities of a few centimeters per day) debris flow with strength can detach a house from its foundation and transport it downslope. The other complication is that a debris flow can transform into turbidity current (Hampton, 1972), which is called surface transformation (Fisher, 1983). But there are no objective sedimentological criteria for interpreting flow transformation from the depositional record. In other words, the velocity-based terms are impractical and therefore, meaningless for interpreting the ancient geologic record.

2.6.5 Excessive synonyms Various classifications have given birth to a surplus of synonyms for mass-transport processes and their deposits (Reiche, 1937; Varnes, 1958, 1978; Hsu¨, 1974; Nardin et al., 1979; Bates and Jackson, 1980; Nemec, 1990; Palanques et al., 2006; Gaudin et al., 2006; Shanmugam, 1996a,b, 2006b; Camerlenghi et al., 2010; Tappin, 2010). Selected examples are: • • • • • • • •

landslide 5 slope movement 5 mass movement 5 mass transport 5 mass wasting, submarine landslide 5 megaturbidite 5 homogenite, mass-transport complex (MTC) 5 MTD 5 submarine mass failure, translational landslide 5 translational slip 5 block slide 5 glide 5 slide, rotational landslide 5 rotational slip 5 toreva block 5 slump, muddy debris flow 5 mud flow 5 cohesive debris flow 5 slurry flow 5 mass flow, debrite 5 olistostrome 5 sedimentary me´lange, sandy debris flow 5 granular flow 5 cohesionless debris flow 5 density-modified grain flow 5 cohesionless liquefied sand flow 5 grain flow 5 mass flow 5 high-density turbidity current 5 hyperconcentrated flow 5 slurry flow 5 hybrid flow, • flow slide 5 liquefaction slide, • rock avalanche 5 debris avalanche 5 sturzstrom, and • sand fall 5 cascading dense-water event 5 sand avalanche 5 grain flow 5 mass flow. This excessive use of synonyms is, obviously, not only not necessary but also even a reason for much confusion.

2.7 Recognition of the three basic types of mass-transport deposits Subaerial “landslides” were recognized as early as in 186 BCE in China (Li, 1989). Nevertheless, only during the past few decades, techniques of systematic recognition and

Mass Transport, Gravity Flows, and Bottom Currents

50

2. Mass transport: slides, slumps, and debris flows

mapping have been developed (Brabb, 1991; Lee, 2005). Also, geographical information systems have become an important part of databases on landslide research (e.g., Dikau et al., 1996). However, the ultimate recognition of individual types of MTD in the rock record must be based on principles of process sedimentology.

2.7.1 Process sedimentology Process sedimentology is the key to recognizing the basic MTD types in core and outcrop. Sanders (1963) published the pioneering paper on process sedimentology entitled “Concepts of fluid mechanics provided by primary sedimentary structures.” The discipline is concerned with the detailed bed-by-bed description of siliciclastic (and calciclastic) sedimentary rocks for establishing the link between the deposit and the physics and hydrodynamics of the depositional process. Basic requirements, principles, and methods of this discipline are: (1) a knowledge of physics, with emphasis on soil mechanics and fluid mechanics (Sanders, 1963; Brush, 1965); (2) the application of uniformitarianism principle; (3) the pragmatic, accurate, precise, and consistent description of the rock; (4) the preservation of absolute distinction between description and interpretation; (5) the documentation of excruciating details in sedimentological logs; (6) the interpretation of processes using exclusively primary sedimentary structures; (7) the mandatory consideration of alternative process interpretations; (8) the total exclusion of facies models; (9) the quantification of depositional facies; and (10) the routine use of common sense. A major problem in sedimentological studies is the failure to adopt the basic principles of process sedimentology, which has prompted lively debates on the deep-water petroleumproducing reservoirs of the Kutei Basin, Indonesia (Fig. 2.31) (Dunham and Saller, 2014; Saller et al., 2006; Shanmugam, 2008a, 2014a). Of the three basic types of mass-transport processes, namely slides, slumps, and debris flows (Fig. 2.29), the terms slide and slump are used for both a process and a deposit. The term debrite is used for deposit of a debris flow. The prefix “sandy” is used for lithofacies that have grain size values greater than 0.06 mm (sand and gravel) and have concentration value equal to or above 20% by volume (Fig. 2.29D). The three sandy SMTD types are emphasized here because of their reservoir potential. Criteria for recognizing MTD and SMTD types in core and outcrop have been developed by integrating my rock description (Table 2.3) with published information by other researchers (Dott, 1963; Helwig, 1970; Johnson, 1970; Fisher, 1971; Hampton, 1972; Middleton and Hampton, 1973; Enos, 1977; Dingle, 1977; Woodcock, 1976, 1979; Cook, 1979; Lowe, 1982; Maltman, 1987, 1994a,b,c; Pickering et al., 1989; Collinson, 1994). Numerous outcrop and core photographs of features associated with sandy slides, sandy slumps, and sandy debrites were published elsewhere (Shanmugam, 2012a).

2.8 Slides 2.8.1 Definition A slide is a coherent mass of sediment or a rigid body that moves along a planar glide plane and shows no internal deformation (Fig. 2.29A). Slides represent translational

Mass Transport, Gravity Flows, and Bottom Currents

2.8 Slides

51

FIGURE 2.31 Map showing 50 examples (locations) of submarine (black triangle) and subaerial (white triangle) mass-transport deposits (MTDs) that are often erroneously called “landslides” (see Tables 2.1, 2.2, and 2.5). Submarine and subaerial classification of each MTD denotes its depositional setting. Note locations of core studies (numbered yellow circles) and outcrop studies (numbered red circles) of deep-water successions carried out by the present author worldwide on MTDs and SMTDs (see Table 2.3 for details). 28 Submarine MTDs: Bering, Bering Sea (Karl et al., 1996; Nelson et al., 2011); Goleta, U.S. Pacific Margin (Greene et al., 2006); Monterey, U.S. Pacific Margin (Paull et al., 2005); Alika, Hawaii, Pacific (Normark et al., 1993); East Breaks, U.S. Gulf of Mexico (McGregor et al., 1993); Mississippi, U.S. Gulf of Mexico (Weimer, 1989, 1990; McAdoo et al., 2000; Nelson et al., 2011); Grand Banks, North Atlantic, Canada (Heezen and Ewing, 1952; Piper and Aksu, 1987; Bornhold et al., 2003); Currituck, U.S. Atlantic Margin (Locat et al., 2009); Hatteras, U.S. Atlantic Margin (Embley, 1980); Amazon, Equatorial Atlantic (Damuth et al., 1988; Piper et al., 1997); Alexander Island, Antarctica (Macdonald et al., 1993); Weddell Sea, Antarctica (Gales et al., 2014); Jan Mayen Ridge, NorwegianGreenland Sea (Laberg et al., 2014); Storegga, Norwegian Sea (Bugge et al., 1987; Haflidason et al., 2005); Nice, Mediterranean Sea (Dan et al., 2007); Nile, Mediterranean Sea (Newton et al., 2004); Canary, SW off Morocco, North Atlantic (Masson et al., 1997); Mauritania-Senegal, W Africa, North Atlantic (Jacobi, 1976); Zaire, W Africa, South Atlantic (formerly known as Congo) (Shepard and Emery, 1973); Owen Ridge, Oman coast, Indian Ocean (Rodriguez et al., 2013); Agulhas, SE Africa, Indian Ocean (Dingle, 1977); KG (Krishna-Godavari Basin), Bay of Bengal, NE Indian Ocean (Shanmugam et al., 2009); Bassein, NE Indian Ocean (Moore et al., 1976); Brunei, NW Borneo Margin (Gee et al., 2007); Kutei, Makassar Strait, Indonesia (Jackson, 2004); Unnamed, offshore New South Wales/Queensland, Australia (Clarke et al., 2012); Bass, SE Australia (Mitchell et al., 2007); Ruatoria, Hikurangi Margin, New Zealand (Collot et al., 2001). 22 Subaerial MTDs: Alaska, State of Alaska, United States (USGS, 2010); Frank, Canada (Cruden and Hungr, 1986); Mt. St. Helens, State of Washington, United States (Schuster, 1983; Tilling et al., 1990); Markagunt, State of Utah, United States (Hacker et al., 2014); Thistle, State of Utah, United States (USGS, 2010); Vargas, Venezuela (USGS, 2010); Nevada del Ruiz, Colombia (Pierson, 1990); Ancash, Peru (USGS, 2010); Santiago, Chile (Sepu´lveda et al., 2006); Rio de Janeiro, Brazil (USGS, 2010); Rio Colorado, Argentina (USGS, 2010); Elm, Swiss Alps (Heim, 1882); Aqaba, Gulf of Aqaba (Klinger et al., 1999); Bududa, Uganda (USGS, 2010); Kolka, Russia (North Ossetia) (USGS, 2010); Saidmerah, Iran (Harrison and Falcon, 1938); Usoy, Tajikistan (Bolt et al., 1975; USGS, 2010); Baikal, Olkhon Island (Lake Baikal, Siberia) (Tyszkowski et al. 2014); Gansu, China (USGS, 2010); Yigong, Tibet (USGS, 2010); Kyoto, Japan (USGS, 2010); Leyte, Philippines (USGS, 2010). Source: Blank world map credit: http://upload.wikimedia.org/wikipedia/commons/8/83/ Equirectangular_projection_SW.jpg (accessed 27.12.14). After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

52

2. Mass transport: slides, slumps, and debris flows

shear-surface movements. Such sliding movements are also common in glaciers (Easterbrook, 1999). Submarine slides can travel hundreds of kilometers on continental slopes. Long-runout distances of up to 810 km for slides have been documented for submarine MTDs (Table 2.5).

TABLE 2.5 Comparison of long-runout MTDs on Earth (submarine and subaerial) with Venus, Iapetus, and Mars (extraterrestrial). Runout distance (km)

Environment

Data

Process

1. Storegga Slide, Norwegian Continental Margin (Bugge et al., 1987; Jansen et al., 1987; Haflidason et al., 2005)

810

Submarine

Seismic and GLORIA side-scan sonar images and core

Slide, slump, and debris flow

2. Agulhas, S. Africa (Dingle, 1977)

750

Submarine

Seismic

Slide and slump

3. Saharan Debris Flow, NW African Margin (Embley, 1982) 4. Canary Debris Flow, NW African Margin (Masson et al., 1997) 5. Hatteras, US Atlantic Margin (Embley, 1980)

700

Submarine

Seismic

Debris flow

600

Submarine

Seismic and core

Debris flow

B500

Submarine

Seismic and core

Slump and debris flow

6. Mauritania-Senegal, NW African Margin (Jacobi, 1976)

B300

Submarine

Seismic and core

Slump and debris flow

7. Nuuanu, NE Oahu (Hawaii) (Normark et al., 1993; Moore et al., 1994) 8. Wailau, N. Molakai (Hawaii) (Normark et al., 1993)

235

Submarine

GLORIA side-scan sonar images

Mass transport

, 195

Submarine

GLORIA side-scan sonar images

Mass transport

9. Rockall, NE Atlantic (Prior and Coleman, 1979, 1984) 10. Clark, SW Maui, Hawaii (Normark et al., 1993)

160

Submarine

Seismic

Mass transport

150

Submarine

GLORIA side-scan sonar images

Mass transport

11. N. Kauai, N. Kauai, Hawaii (Normark et al., 1993)

140

Submarine

GLORIA side-scan sonar images

Mass transport

12. East Breaks (West), Gulf of Mexico (McGregor et al., 1993)

110

Submarine

Seismic and core

Slump and debris flow

13. Grand Banks, Newfoundland (Heezen and Ewing, 1952; Driscoll et al., 2000; Bornhold et al., 2003)

. 100

Submarine

Seismic and core

Mass transport and turbidity currenta

Name and location

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

53

2.8 Slides

TABLE 2.5 (Continued) Runout distance (km)

Environment

Data

Process

14. Ruatoria, New Zealand (Collot et al., 2001) 15. Alika-2, W Hawaii (Hawaii) (Normark et al., 1993)

100

Submarine

Seismic

Mass transport

95

Submarine

GLORIA side-scan sonar images

Mass transport

16. Kaena, NE Oahu (Hawaii) (Normark et al., 1993)

80

Submarine

GLORIA side-scan sonar images

Mass transport

17. El Golfo, Western Canary Islands (Masson et al., 2002) 18. Bassein, Bay of Bengal (Moore et al., 1976) 19. Kidnappers, New Zealand (Lewis, 1971) 20. Munson-Nygen, New England (O’Leary, 1993)

65

Submarine

Seismic

Mass transport

55

Submarine

Seismic

Slide and debris flow

45

Submarine

Seismic

Slump and slide

45

Submarine

Seismic

Slump and debris flow

21. Ranger, Baja California (Prior and Coleman, 1984) 1. Osceola mudflow, Mount Rainier (Vallance and Scott, 1997) 2. Nevado del Ruiz, Colombia (Pierson, 1990) 3. Pine Creek and Muddy River lahars, Mount St. Helens (Pierson, 1985)

35

Submarine

Seismic

Mass transport

120

Subaerial

Outcrop

Mass transport

103

Subaerial

Outcrop

Mass transport

31

Subaerial

Outcrop

Mass transport (the world’s largest historical subaerial MTD)

4. Saidmarreh Slide, Zagros FoldThrust Belt, SW Iran (Roberts and Evans, 2009)

19

Subaerial

Outcrop

Mass transport (the world’s largest prehistoric subaerial MTD)

1. Venus (Malin, 1992)

550

Extraterrestrial (Venus)

Radar images by the Magellan spacecraft

Mass transport

2. Iapetus, a satellite of Saturn (Singer et al., 2012, their Fig. 5)

780

Extraterrestrial (Iapetus)

Cassini mission images

Mass transport

3. Thaumasia Plateau (Montgomery et al., 2009, their Fig. 9)

2500

Extraterrestrial (Mars)

Thermal Emission Imaging System infrared (THEMIS IR)

Mass transport

Name and location

a

See Shanmugam (2012a) for discussion on the evidence for turbidity currents. The term “landslide” was used to describe many of these examples by the original authors. Locations of selected examples are shown in Fig. 2.31. Long-runout MTDs provide empirical data for developing depositional models for deep-water sandstone petroleum reservoirs. The change in numbering is to reflect the change in type of environment (subaerial, submarine, and extraterrestrial). Compiled from several sources. After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166.

Mass Transport, Gravity Flows, and Bottom Currents

54

2. Mass transport: slides, slumps, and debris flows

2.8.2 Origin In submarine environments, slides tend to occur on continental margins commonly near the shelf-slope breaks, in submarine canyons, and in fjords. However, long-runout slides may occur in basinal settings as well. Slides are commonly associated with triggering events such as earthquakes, meteorite impacts, volcanic activities, glacial loading, sediment loading, cyclones, and tsunamis (Chapter 7: Triggering Mechanisms of Downslope Processes). Some of the best-studied seismic examples of submarine MTDs are in the area of the Storegga Slide on the mid-Norwegian continental margin (Solheim et al., 2005b). Even in these cases, the authors acknowledged the practical difficulties in distinguishing slides from debrites on seismic profiles. This is because both slides and debrites exhibit homogeneous (i.e., transparent) to chaotic reflections (Fig. 2.32). In distinguishing slides from debrites, Solheim et al. (2005b) used additional criteria such as the existence of a

FIGURE 2.32 Seismic profile showing transparent (homogeneous) to chaotic internal reflections of slide deposits (SD). Note continuous and parallel internal reflections of contourite deposits (CD). The Storegga Slide on the mid-Norwegian continental margin. Source: Profile courtesy of A. Solheim. Modified after Solheim, A., Berg, K., Forsberg, C.F., Bryn, P., 2005b. The Storegga Slide complex: repetitive large scale sliding with similar cause and development. Mar. Pet. Geol. 22 (12), 97107, with permission from Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

2.8 Slides

55

FIGURE 2.33 Geological setting of the Nankai accretionary wedge. (A) Shaded relief map of the Nankai Trough showing the regional setting of the IODP NanTroSEIZE drilling transect. The red line through IODP drill sites C0004 and C0008 shows the location of the seismic line cross section in (B). The inset is a tectonic map of the northern Philippine Sea plate. (B) Interpreted composite seismic line of the NanTroSEIZE transect showing the predominant morphotectonic zones. Source: Modified from Moore et al. (2009). VE, 2X (twofold vertical exaggeration). From Strasser, M., Moore, G., Kimura, G., Kopf, A., Underwood, M., Guo, J., et al., 2011. Slumping and mass-transport deposition in the Nankai forearc: evidence from IODP drilling and 3-D reflection seismic data. Geochem. Geophys. Geosyst. 12, 124. AGU.

headwall as well as sidewalls. Similar problems of recognizing individual depositional facies (e.g., slides vs debrites) on seismic profiles have been acknowledged by Tripsanas et al. (2008) and Twichell et al. (2009). In recognizing slides, McAdoo et al. (2000) used bathymetry and GLORIA side-scan sonar data. But such large-scale images are unreliable for distinguishing the sliding motion from flowing motion. These real-world examples reveal the limitations of relying on seismic data for distinguishing specific types of deep-water depositional facies. The solution is to examine the rocks directly by using core or outcrop. A detailed seismic-core study of the Nankai Subduction Zone, Japan (Fig. 2.33) by Strasser et al. (2011) has documented slumping and MTD using both 3D reflection seismic data (Fig. 2.34A and B) and IODP drilling core data (Fig. 2.34C and D). A depositional model illustrates the origin of MTD by sediment failure in this tectonically active setting (Fig. 2.35).

2.8.3 Identification Slides are capable of transporting gravel and coarse-grained sand because of their inherent strength. General characteristics of slides are: • slab geometry (outcrop) (Fig. 2.10); • blocky log motif (Fig. 2.36). Cored interval of this log motif (Fig. 2.36) reveals complex internal features in core (Fig. 2.37); • sand injection beneath the slide block (Fig. 2.38); • primary basal glide plane or decollement (core and outcrop) (Fig. 2.39);

Mass Transport, Gravity Flows, and Bottom Currents

56

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.34 (A) Seismic profile with core sites. (B) Interpretation showing MTD packages. Nankai Basin. (C and D) Core showing MTD features. Source: From Strasser, M., Moore, G., Kimura, G., Kopf, A., Underwood, M., Guo, J., et al., 2011. Slumping and mass-transport deposition in the Nankai forearc: evidence from IODP drilling and 3-D reflection seismic data. Geochem. Geophys. Geosyst. 12, 124. AGU.

• • • • • • • • • •

basal shear zone (core and outcrop) (Fig. 2.39); secondary internal glide planes (core and outcrop) (Fig. 2.40); unusual steep fabric (Fig. 2.41); preservation of original strata from the provenance region (Fig. 2.42); multiple internal layers within a single slide unit (Fig. 2.42); subaerial to shallow-water facies encased in deep-water host muddy facies (Fig. 2.43) due to long-distance transport of MTD (Fig. 2.44); associated slumps (Fig. 2.42); sheet-like geometry (Fig. 2.42); in cores and outcrops, unusually steep, near-vertical, razor-sharp, basal contacts of thick sand units that truncate bedding layers of underlying mudstone units (Fig. 2.45); MTDs are characterized by chaotic reflections on seismic profiles.

Mass Transport, Gravity Flows, and Bottom Currents

2.8 Slides

57

FIGURE 2.35 Proposed origin of MTD in three stages explaining observed seismic geometries. Source: From Strasser, M., Moore, G., Kimura, G., Kopf, A., Underwood, M., Guo, J., et al., 2011. Slumping and mass-transport deposition in the Nankai forearc: evidence from IODP drilling and 3-D reflection seismic data. Geochem. Geophys. Geosyst. 12, 124. AGU.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 2.36 An actual example of a Middle Eocene deep-marine sand package in the U.K. North Sea showing blocky Gamma-ray log motif. Details of this sand are shown in Figures 2.372.41. Source: Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166. Elsevier.

FIGURE 2.37 (A) Gamma-ray log showing blocky motif of a real-world subsurface cored Eocene interval, North Sea. (B) Cored intervals (cores 14) showing distribution of muddy and sandy lithologies. (C.) Based on detailed core description, sedimentological features of the cored interval (cores 14) are illustrated in Figures 2.382.41. Source: After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166. Elsevier.

59

2.8 Slides

FIGURE 2.38 Core photograph showing sand injection. See Fig. 2.37C for photo location. After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166. Elsevier.

FIGURE 2.39 Core photograph showing primary glide plane and basal shear zone. See Fig. 2.37C for photo location. After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

60

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.40 Core photograph showing secondary glide plane. See Fig. 2.37C for photo location. After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166. Elsevier.

FIGURE 2.41 Core photograph showing steep fabric. See Fig. 2.37C for photo location. After Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109166. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 2.42 Outcrop photograph showing sheet-like geometry of an ancient sandy submarine slide (1000 m long and 50 m thick) encased in deep-water mudstone facies. Note the large sandstone sheet with rotated/slumped edge (left). Person (arrow): 1.8 m tall. Blocky wireline log motif is inserted for comparison with subsurface slides. Note slumped left edge. Ablation Point Formation, Kimmeridgian (Jurassic), Alexander Island, Antarctica. Source: Photo courtesy of D.J.M. Macdonald. From Macdonald, D.I.M., Moncrieff, A.C.M., Butterworth, P.J., 1993. Giant slide deposits from a Mesozoic fore-arc basin, Alexander Island, Antarctica. Geology 21, 10471050. Labels by G. Shanmugam. GSA.

FIGURE 2.43 Outcrop photograph showing a sandy slide encased in deep water muddy units, Lower Jurassic, Longobucco Basin, Calabria, Southern Italy. This slide block was called “olistolith” by Teale and Young (1987). Arrow points to the late T. Teale. Source: After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524. Elsevier.

62

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.44 Model showing long runout blocks in the Longobucco Basin. Source: After Teale, T., Young, J.R., 1987. Isolated olistoliths from the Longobucco Basin Calabria, S. Italy. In: Leggett, J.K., Zuffa Shaver, G.G. (Eds), Advances in Marine Clastic Sedimentology. Graham & Trotman, pp. 7588. FIGURE 2.45 Core photograph showing sheared contact between overlying sand (left) and underlying mudstone. The steeply dipping (80 degrees) contacts represent primary glide plane (arrow) of a major slump/slide sheet. Note the effects of shearing in the form of drag lamination in the underlying mudstone. 73890 (2253 m) core depth. Pliocene, Edop Field, Offshore Nigeria.

Mass Transport, Gravity Flows, and Bottom Currents

2.9 Slumps

63

2.8.4 Case studies The above discussion of numerous field examples of MTD is based on individual case studies, such as the Nankai fore-arc Basin, Japan (e.g., Strasser et al., 2011) and the Longobucco Basin, Calabria, Italy (Teale and Young, 1987).

2.8.5 Facies models There are no facies models for slides as a distinct depositional facies.

2.8.6 Problems In nature, MTDs occur commonly as a combined product of mixed or transitional processes than as an end-member type, such as slide or slump. Therefore, distinguishing end-member facies is a challenge. Because MTDs generate large-scale fan geometries on amplitude maps or other images (Figs. 2.14, 2.19, and 2.20), caution must be exercised in applying facies models for turbidite-dominated submarine fans (Mutti, 1992), without sediment core from the package.

2.9 Slumps 2.9.1 Definition A slump is a coherent mass of sediment that moves on a concave-up glide plane and undergoes rotational movements causing internal deformation (Fig. 2.29A). Slumps represent rotational shear-surface movements. Slumps are capable of transporting gravel and coarse-grained sand because of their inherent strength. SSDS, discussed in detail below in Chapter 9, Soft-Sediment Deformation Structures, are considered as part of the slump family because slump folds are commonly classified as SSDS.

2.9.2 Origin A distinguishing attribute of synsedimentary slumps, as opposed to tectonic slumps, is their sandwiched occurrence between undeformed layers in sedimentary lithofacies, such as limestones (Fig. 2.46), anhydrites (Fig. 2.47), among others. Slumps are commonly associated with triggering events such as earthquakes, meteorite impacts, volcanic activities, glacial loading, sediment loading, cyclones, and tsunamis (Chapter 7: Triggering Mechanisms of Downslope Processes). I have selected two case studies in explaining the two different origins of slump structures. 2.9.2.1 Submarine channel deposition In the Pennsylvanian Jackfork Group in the Ouachita Mountains (Fig. 2.48) (Shanmugam et al., 1988a). Duplex-like structure is a sigmoidal deformation feature that has been observed in deep-water sandy lithofacies that are sandwiched between

Mass Transport, Gravity Flows, and Bottom Currents

64

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.46 Detailed sketches by Sir William Edmond Logan of localized deformed beds within otherwise undeformed Devonian limestones, Gaspe´ Peninsula, Quebec, Canada (Logan, 1863). Such deformed beds are commonly called “soft-sediment deformation structures” (SSDS). Source: Diagram reproduced from Maltman, A., 1994a. Deformation structures preserved in the rocks. In: Maltman, A. (Ed.), The Geological Deformation of Sediments. Chapman & Hall, London, pp. 261307.

FIGURE 2.47 Core photographs showing microfolds in anhydrite (white) layers with intervening undeformed anhydrite layers. Dark layers represent calcite with organic matter. (A) Core slab dominated by undeformed layers with rare layers of microfolds; (B) Core slab dominated by layers of microfolds with rare unreformed layers. These examples are classic SSDS because deformed layers are sandwiched between undeformed layers. SSDS 5 soft-sediment deformation structures. Castile Formation, Permian, Delaware Basin, New Mexico. See Kirkland and Anderson (1970) for a detailed study of the Castile microfolds. Source: Samples courtesy of D.W. Kirkland.

Mass Transport, Gravity Flows, and Bottom Currents

2.9 Slumps

65

FIGURE 2.48 Location map showing study locality used in outcrop study of duplex-like structures in the Ouachita flysch in Arkansas.

FIGURE 2.49 Outcrop photograph showing duplex-like structures (i.e., sigmoidal deformation structures) and laterally extensive nature (arrow). Pennsylvanian, Jackfork Group, DeGray Spillway East Wall Section, Ouachita Mountains, Arkansas. Source: After Shanmugam, G., Moiola, R.J., Sales, J.K., 1988a. Duplex-like structures in submarine fan channels, Ouachita Mountains, Arkansas. Geology 16, 229232, with permission from Geological Society of America (GSA).

undeformed layers (Fig. 2.49). In particular, two adjacent layers exhibit opposing dip directions (Fig. 2.50). Conventionally, duplex features have been attributed to tectonic deformation of lithified units (Boyer and Elliott, 1982) (Fig. 2.51). However, a tectonic origin for the sigmoidal slices is considered unlikely in the Jackfork Group because of observed opposing

Mass Transport, Gravity Flows, and Bottom Currents

66

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.50 Outcrop photograph showing two adjacent sigmoidal deformation structures (i.e., duplexlike structures) with opposing dips of imbricate slices. Pennsylvanian, Jackfork Group, DeGray Spillway East Wall Section, Ouachita Mountains, Arkansas. Source: After Shanmugam, G., Moiola, R.J., Sales, J.K., 1988a. Duplexlike structures in submarine fan channels, Ouachita Mountains, Arkansas. Geology 16, 229232, with permission from Geological Society of America (GSA).

FIGURE 2.51 Theoretical stages of development of duplex-like structures in thrust tectonics. Simplified after Boyer and Elliott (1982). Source: Diagram from Shanmugam, G., Moiola, R.J., Sales, J.K., 1988a. Duplex-like structures in submarine fan channels, Ouachita Mountains, Arkansas. Geology 16, 229232, with permission from Geological Society of America (GSA). Mass Transport, Gravity Flows, and Bottom Currents

67

2.9 Slumps

FIGURE 2.52 A side view of flume tank showing sandy debris flows with imbricate slices (inclined arrow). Such imbrications develop in sandy debris flows when the front of a flow freezes, the body of the flow breaks and thrusts over the slice in the front due to compression. Similar features (duplex-like structures) in the rock record have been ascribed to synsedimentary duplex-like structures (Shanmugam et al., 1988a). Flow direction is from right to left. Source: Photo from Shanmugam (2000a), with permission from Elsevier.

FIGURE 2.53 Subaerial slurry flows (i.e., plastic debris flow with movement from right to let, arrow) showing development of synsedimentary folding in the frontal zone of Matanuska Glacier, Alaska (Lawson, 1981). This folding is analogous to the origin of imbricate slices in experiments on sandy debris flows (see Fig. 2.52). Source: Photo courtesy of G.D. Klein.

directions of imbrication in stratigraphically adjacent units (Fig. 2.50). Such opposing orientations would require an unrealistic tectonic history. Therefore, the imbricate slices (i.e., duplexes) have been attributed to sedimentary slumping (Shanmugam et al., 1988a,b). This conclusion was based, in part, on an experimental model of a small-scale duplex structure generated in soft plaster in the laboratory (Shanmugam et al., 1988a,b, their Fig. 2.3). Sigmoidal deformation structures with imbricate slices have also been generated in flume experiments on sandy debris flows (Marr et al., 2001; Shanmugam, 2000a) (Fig. 2.52). Glacial debris flows also are known to generate imbricate bedding in Alaska (Fig. 2.53).

Mass Transport, Gravity Flows, and Bottom Currents

68

2. Mass transport: slides, slumps, and debris flows

Ni et al. (2015) attributed the origin of duplex-like structures in the Jurassic strata of western Qaidam Basin, China, to synsedimentary slumping, but related to earthquakes. In light of knowledge gained from experiments and field observations (Fig. 2.54A), a depositional model or the origin of duplex-like structures in submarine channels has been proposed using three stages (Fig. 2.54B). Stage 1: Deposition of sediments by axial sediment-gravity flows in a submarine channel. Stage 2: Mass flows, triggered by sediment failure along the right-hand channel wall, glided over the sediment from stage 1 in a perpendicular direction, causing duplex 1. Note that the dip direction of duplex 1 is opposite to the flow direction of mass flows above.

FIGURE 2.54 (A) An ideal stratigraphic column with duplex-like structures of different composition; (B) three stages of development of duplex-like structures with opposing dips in a submarine channel. Stage 1: deposition from axial sediment gravity flows. Stage 2: deformation of unlithified sediment into duplex 1 by mass flows triggered from right-hand channel wall. Stage 3: deformation of unlithified sediment into duplex 2 by mass flows triggered from left-hand channel wall. Source: From Shanmugam, G., Moiola, R.J., Sales, J.K., 1988a. Duplex-like structures in submarine fan channels, Ouachita Mountains, Arkansas. Geology 16, 229232, with permission from Geological Society of America (GSA).

Mass Transport, Gravity Flows, and Bottom Currents

2.9 Slumps

69 FIGURE 2.55 Outcrop photograph showing two layers of seismicityinduced soft-sediment deformation structures (SSDS), in this case slump folds, with an intervening interval of undeformed layers. Perazim Wadi in the Quaternary Lisan Formation, a dry wash in the Ami’az Plain SW of Ein Boquet in Israel. Although this formation is not of deep-water origin, it illustrates the seismicity-induced sediment deformation in tectonically active settings. Source: Photo courtesy of Professor Emeritus R.D. Hatcher, Jr., Department of Earth and Planetary Sciences, The University of Tennessee, Knoxville.

FIGURE 2.56 (A and B) Photographs showing slump folded layers with undeformed layers above and below. (C) Sketch. Dead Sea Basin. Source: Compiled from Alsop, G.I., Marco, S., 2013. Seismogenic slump folds formed by gravity-driven tectonics down a negligible subaqueous slope. Tectonophysics 605, 4869.

Mass Transport, Gravity Flows, and Bottom Currents

70

2. Mass transport: slides, slumps, and debris flows

Stage 3: Mass flows, triggered by sediment failure along the left-hand channel wall, glided over the sediment in a perpendicular direction, causing duplex. This synsedimentary origin of duplexes is a more realistic explanation than the conventional tectonic origin for explaining field disposition of beds. From an overall kinematic style viewpoint, representing either rooted or gravity-driven (Waldron and Gagnon, 2011), the proposed duplex origin is strictly a gravity-driven deformation. 2.9.2.2 Earthquake-induced gravity tectonics Slump folds, sandwiched between undeformed layers (Fig. 2.55), have been documented to be triggered by earthquakes. In the Dead Sea Basin, seismogenic slump folds (Fig. 2.56) have been explained by gravity-driven tectonics down negligible subaqueous slope (Fig. 2.57) (Alsop and Marco, 2013). FIGURE 2.57 Model explaining the origin of slump folds in four stages in the Dead Sea Basin. Source: From Alsop, G.I., Marco, S., 2013. Seismogenic slump folds formed by gravity-driven tectonics down a negligible subaqueous slope. Tectonophysics 605, 4869.

Mass Transport, Gravity Flows, and Bottom Currents

2.9 Slumps

71

2.9.3 Identification General characteristics of slumps are: • large-scale slump sheets with upslope tensional glide planes and downslope compressional thrusts (Fig. 2.58), • basal zone of shearing (core and outcrop), • slump folds (Helwig, 1970) interbedded with undeformed layers (core and outcrop) (Figs. 2.592.62), • irregular upper contact (core and outcrop) (Fig. 2.63), • chaotic bedding in heterolithic facies (core and outcrop) (Fig. 2.64), • rotated elongate grains (Maltman, 1987) (microscopic), • steeply dipping and truncated layers (core and outcrop) (Fig. 2.65), • associated ptygmatic folding (Fig. 2.66), • associated slides (core and outcrop) (Fig. 2.42), • associated brecciated clasts (core and outcrop) (Fig. 2.67), and • chaotic facies in high-resolution seismic profiles. In submarine environments, slumps tend to occur commonly on slope settings.

FIGURE 2.58 An ideal slump sheet with tensional and compressional features. Source: From Lewis, K.B., 1971.

Slumping on a continental slope inclined at 1 4 . Sedimentology 16, 97110. Figure from Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524, with permission from Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

72

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.59 Outcrop photograph showing slump-folded heterolithic facies overlain by undeformed deep-water sandstone, Eocene, La Jolla, California. Source: After Shanmugam, G., 2006a. DeepWater Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476, Elsevier.

2.9.4 Case studies The two case studies discussed earlier are the Ouachita flysh and Dead Sea Basin.

2.9.5 Facies models There are no facies models for slumps.

2.9.6 Problems Like slides, slumps are often too complicated because of associated facies.

Mass Transport, Gravity Flows, and Bottom Currents

2.10 Debris flows: a prelude

73 FIGURE 2.60 Core photograph showing interbedded occurrence of deformed (convolute bedding) sandstone and siltstone (light gray) layers with undeformed claystone (dark gray) layers. Paleocene, U.K. North Sea. Source: From Shanmugam, G., 2012a. New perspectives on deepwater sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524, with permission from Elsevier.

2.10 Debris flows: a prelude A debris flow is a sediment flow with plastic rheology and laminar state from which deposition occurs through freezing en masse. The terms debris flow and mass flow are used interchangeably because each exhibits plastic flow behavior with shear stress distributed throughout the mass (Nardin et al., 1979). In debris flows, intergranular movements predominate over shear-surface movements. Although most debris flows move as incoherent mass, some plastic flows may be transitional in behavior between coherent mass movements and incoherent sediment flows (Marr et al., 2001). Debris flows may be mud-rich (i.e., muddy debris flows), sand-rich (i.e., sandy debris flows), or mixed types. Debris flow is considered both as MTD and as a gravity low. Because there is so much confusion in distinguishing debris flow from turbidity currents (Mutti et al., 1999; Zavala, 2019), debris flows are considered in great detail in Chapter 3, Gravity Flows: Debris

Mass Transport, Gravity Flows, and Bottom Currents

74

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.61

(A) Core photograph showing slump-fold axis (arrow) of a heterolithic facies unit in sandstone, Cretaceous, West Africa. (B) Core photograph showing slump-folded heterolithic (sand and mud) facies and associated sand injection, Paleocene, Faeroe Basin, U.K. Continental Margin. Source: From Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524, with permission from Elsevier. (B) After Shanmugam, G., Bloch, R.B., Mitchell, S.M., Beamish, G.W.J., Hodgkinson, R.J., Damuth, J.E., et al., 1995a. Basin-floor fans in the North Sea: sequence stratigraphic models vs. sedimentary facies. AAPG Bull. 79, 477512, with permission from American Association of Petroleum Geologists (AAPG).

Flows, Grain Flows, Liquefied/Fluidized Flows, Turbidity Currents, Hyperpycnal Flows, and Contour Currents.

2.11 Long-runout mechanisms An understanding of long-runout mechanisms is important not only for academic reasons, but also for economic reasons. For example, long-runout MTDs (Table 2.5) provide empirical data for developing predictive depositional models for deep-water sandstone petroleum reservoirs in the subsurface.

2.11.1 Basic concept The basic premise of long-runout MTDs is that they travel further than the distance predicted by simple frictional models. Heim’s (1932) study of the subaerial “Elm Slide” in the

Mass Transport, Gravity Flows, and Bottom Currents

75

2.11 Long-runout mechanisms

FIGURE 2.62 (A) Core photograph showing slump-folded heterolithic (sand and mud) facies) and associated sand injection, Paleogene, Faeroe Basin, U.K. continental margin. (B) Close-up photograph. Arrow shows stratigraphic position. Source: Partly after Shanmugam, G., Bloch, R.B., Mitchell, S.M., Beamish, G.W.J., Hodgkinson, R.J., Damuth, J. E., et al., 1995a. Basin-floor fans in the North Sea: sequence stratigraphic models vs. sedimentary facies. AAPG Bull. 79, 477512, reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

FIGURE 2.63 Core photograph showing sharp and irregular contact between underlying argillaceous sandstone (dark gray) and overlying cleaner sandstone (light gray). Note deformed and brecciated shale clasts in the argillaceous sandstone. Cretaceous, Nise Formation, offshore Norway. Source: After Shanmugam, G., Lehtonen, L.R., Straume, T., Syversten, S.E., Hodgkinson, R. J., Skibeli, M., 1994. Slump and debris flow dominated upper slope facies in the Cretaceous of the Norwegian and Northern North Seas (6167 N): implications for sand distribution. AAPG Bull. 78, 910937, reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

Mass Transport, Gravity Flows, and Bottom Currents

76

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.64

Outcrop photograph showing lateral pinchout geometry of a heterolithic slump facies with contorted bedding (light color unit). These interbedded fine sandstone and shale are of deepwater origin. Aberystwyth Grit Formation, Silurian, Wales. Source: Photo courtesy of K.F. Keller.

FIGURE 2.65 Core photograph showing steeply dipping clay-rich layers (white) adjacent to horizontal layers in sandstone. Cretaceous, Agat Formation, offshore Norway. Source: After Shanmugam, G., Lehtonen, L.R., Straume, T., Syversten, S.E., Hodgkinson, R.J., Skibeli, M., 1994. Slump and debris flow dominated upper slope facies in the Cretaceous of the Norwegian and Northern North Seas (6167 N): implications for sand distribution. AAPG Bull. 78, 910937, reprinted by permission from American Association of Petroleum Geologists (AAPG).

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 2.66 photograph showing ptygmatically folded sandstone dike (arrow) in mudstone associated with sandy slumps, Cretaceous, Agat Formation, offshore Norway.

Slump folds

5 cm

10 m

Secondary glide plane

Brecciated zones Slump folds

Sand layer mud 10 cm

10 cm

Contorted layers

Steeply dipping layers

Glide plane (fault)

Brecciated mud clasts

Abrupt change in fabric

10 cm

10 cm

Mud layer

Internal shear surface (Secondary glide plane) Basal shear zone (Primary glide plane) Sand

Clastic injections

10 cm

5m

Basal shear zone

Inclined dish structures

Mud

Injected sand

FIGURE 2.67 Summary of features associated with slump deposits observed in core and outcrop. Slump fold, an intraformational fold produced by deformation of soft sediment; contorted layer, deformed sediment layer; basal shear zone, the basal part of a rock unit that has been crushed and brecciated by many subparallel fractures due to shear strain; glide plane, slip surface along which major displacement occurs; brecciated zone, an interval that contains angular fragments caused by crushing of the rock; dish structures, concave-up (like a dish) structures caused by upwardescaping fluids in the sediment; clastic injections, natural injection of clastic (transported) sedimentary material (usually sand) into a host rock (usually mud). Source: From Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476.

78

2. Mass transport: slides, slumps, and debris flows

Swiss Alps has been the source of the following basic equations for understanding the mobility of MTDs: 1. H/L 5 tanφ, where H represents the vertical fall height, L represents the runout distance, and ϕ is the Coulomb angle of sliding friction (e.g., Griswold and Iverson, 2008). 2. H/Lα 1/V, where V is the initial volume of the moving mass (e.g., McEwen, 1989). 3. H/L 5 1, where L is the normal-runout distance (Fig. 2.68A) (e.g., Collins and Melosh, 2003). 4. H/L # 1, where L is the long-runout distance (Fig. 2.68B) (e.g., Hampton et al., 1996).

FIGURE 2.68

Conceptual models showing sliding movement of a rigid body in subaerial environments. (A) An ideal model in which the predicted runout length (L) is equal to vertical fall height (H) (e.g. Collins and Melosh, 2003). (B) Long-runout model in which the runout length (L) exceeds the vertical fall height (e.g., Hampton et al., 1996). (C) Basic equations derived from the work of Heim (1932) on the “Elm Slide” in the Swiss Alps.

Mass Transport, Gravity Flows, and Bottom Currents

2.11 Long-runout mechanisms

79

Although there are many documented cases of long-runout MTDs in both subaerial (Table 2.5) and submarine (e.g., submarine slides in Hawaii with more than 200 km of runout distances, Moore et al., 1989) environments with runout distances measuring up to 100 times their vertical fall height and high speeds of up to 500 km h 21 (Martinsen, 1994), the geologic community was reluctant to accept mechanisms that attempted to explain MTDs that travel farther and faster than expected. A major turning point on the skepticism over long-runout MTDs occurred on May 18, 1980 when the eruption of Mount St. Helens in the United States generated impressive longrunout subaerial MTDs that were captured on videotapes (see The Learning Channel, 1997).

2.11.2 Subaerial environments There are at least 20 potential mechanisms that could explain the mechanical paradox of long-runout MTDs (Terzaghi, 1950; Brunsden, 1979; Schaller, 1991, among others). Selected examples of subaerial mechanisms are: 1. lubrication by liquefied saturated soil entrained during transport (Heim, 1882; Hungr and Evans, 2004), 2. dispersive pressure in grain flows (Bagnold, 1954), 3. fluidization by entrapped air (Kent, 1966), 4. cushion of compressed air beneath the slide (Shreve, 1968), 5. fluidization by dust dispersions (Hsu¨, 1975), akin to grain flows (Bagnold, 1954), 6. spontaneous reduction of friction angle at high rates of shearing (Scheidegger, 1975; Campbell, 1989), 7. vaporization of water at the base and related excess pore-water pressure (Goguel, 1978), 8. Frictional heating along a basal fluid-saturated shear zone and related rise in porewater pressure (Voight and Faust, 1982; Goren and Aharonov, 2007), 9. self-lubrication by frictionally generated basal melt layers (Erismann, 1979; De Blasio and Elverhøi, 2008; Weidinger and Korup, 2009), 10. acoustic fluidization (Melosh, 1979), 11. mechanical fluidization or inertial grain flow (Davies, 1982), 12. fluidization by volcanic gases (Voight et al., 1983), 13. excess pore-water pressure (Cruden and Hungr, 1986; Iverson, 1997), 14. self-lubrication by granular flows acting as basal shear zone (Cleary and Campbell, 1993), and 15. seismic energy released during meteorite impacts, proposed for Mars (Akers et al., 2012), is also applicable to Earth.

2.11.3 Submarine environments Submarine environments with long-runout MTDs have been broadly grouped into five types: (1) fjords, (2) active river deltas on the continental margin, (3) submarine

Mass Transport, Gravity Flows, and Bottom Currents

80

2. Mass transport: slides, slumps, and debris flows

canyon-fan systems, (4) open continental slopes, and (5) oceanic volcanic islands and ridges by Hampton et al. (1996). To this list, a sixth type “glacially influenced continental margins” (Elverhøi et al., 1997) needs to be added. Submarine MTDs with long-runout distances of over 100 km commonly occur on slopes of less than 2 degrees on the US Atlantic Continental Slope (Fig. 2.4A). Several potential mechanisms are available for explaining long-runout submarine MTDs over low-angle slopes: 1. 2. 3. 4. 5.

hydroplaning (Mohrig et al., 1998), excess pore-water pressure (Pierson, 1981; Gee et al., 1999), elevated gas pressure (Coleman and Prior, 1988), dispersive pressure in grain flows (Bagnold, 1954; Norem et al., 1990), self-lubrication by granular flows acting as basal shear zone (Cleary and Campbell, 1993), 6. self-lubrication at the base of gas-hydrate stability window that coincides with the base of MTDs (Bugge et al., 1987; Cochonat et al., 2002), 7. flow transformation (Talling et al., 2007), and 8. seismic energy released during meteorite impacts, proposed for Mars (Akers et al., 2012), is also applicable to Earth. Of various mechanisms listed above, the hydroplaning concept (Mohrig et al., 1998) has gained acceptance (McAdoo et al., 2000; Shanmugam, 2000a; Marr et al., 2001; Elverhøi et al., 2002; Ilstad et al., 2004; De Blasio et al., 2006). Nevertheless, the hydroplaning mechanism is inapplicable to explaining long-runout debris flows in subaerial and extraterrestrial environments.

2.11.4 Extraterrestrial environments Analogous to subaerial and submarine environments on Earth, there are numerous published examples of long-runout MTDs on other planets of the Solar System (Table 2.5). Although submarine MTDs show much longer runout distances than those of subaerial MTDs on Earth, the longest runout distance of 2500 km has been documented for an extraterrestrial MTD on Mars (Montgomery et al., 2009, their Fig. 2.9). The following mechanisms have been offered for explaining long-runout MTDs on extraterrestrial environments: 1. self-lubrication by released groundwater, wet debris, or mud (Lucchitta, 1979, 1987); 2. aqueous pore-pressure support (Harrison and Grimm, 2003); 3. continental-scale salt tectonics coupled with overpressured fluids (Montgomery et al., 2009); 4. movement on ice (De Blasio, 2011); 5. movement on evaporitic salt (De Blasio, 2011); 6. friction reduction during flash heating (Singer et al., 2012); and 7. seismic energy released during meteorite impacts (Akers et al., 2012). Similar explanations were offered previously for landslides on the Moon (Guest, 1971; Howard, 1973).

Mass Transport, Gravity Flows, and Bottom Currents

2.11 Long-runout mechanisms

81

2.11.5 H/L ratio problems Although the H/L model has been influential for nearly a century, many problems still remain. 1. Because the original work by Heim (1932) was written in German, there have been differences of opinion among later workers as to the meaning of the German nomenclature used by Heim to describe the type of motion, ranging from sliding (Shreve, 1968) to flowing (Hsu¨, 1975). 2. The H/L ratios for submarine MTDs are lower (0.0010.3, Hampton et al., 1996, their Table 2.5) than those for subaerial counterparts (1.621, Ritter et al., 1995). The model underestimates the extent of runout distance (L) for water-saturated debris flows (Iverson, 1997; Griswold and Iverson, 2008) and if the volumes of moving mass exceed about 106 m3 (Heim, 1932; Hsu¨, 1975; Scheidegger, 1973). Also, the model does not take into account the effect of runout-path topography on the distal or lateral limits of inundation (Griswold and Iverson, 2008). 3. At a given value of H/L, the Martian MTDs are typically about 50100 times more voluminous than the terrestrial counterparts (McEwen, 1989). However, there is no universally accepted physical basis for explaining the equation H/Lα 1/V (Dade and Huppert, 1998). 4. Dade and Huppert (1998) have used L/H as a measure of the efficiency of MTD movement, which is the inverse of the friction coefficient (H/L). 5. On Earth, submarine MTDs are much larger in size than subaerial MTDs (Hampton et al., 1996), and submarine MTDs travel longer distances than subaerial MTDs (Fig. 2.69) (Hampton et al., 1996, their Table 2.1; and Elverhøi et al., 2002, their Table 2.1). 6. Venusian MTDs (Malin, 1992, his Fig. 2.11) and Martian MTDs (Collins and Melosh, 2003, their Fig. 2.1) travel longer distances than those on Earth’s subaerial environments (Fig. 2.70). 7. The H/L model has been applied to both “landslides” and “debris flows” without acknowledging the basic differences in sediment movement between the two processes (McEwen, 1989; Malin, 1992; Hampton et al., 1996; Ablay and Hu¨rlimann, 2000; McAdoo et al., 2000; Elverhøi et al., 2002; Legros, 2002; Collins and Melosh, 2003; Geertsema et al., 2009; Singer et al., 2012). The problem is that slides represent a rigidbody sliding motion over a shear surface (Varnes, 1978; Dott, 1963), whereas debris flows represent an intergranular flowing motion (Shanmugam et al., 1994; Iverson et al., 1997). 8. In subaerial environments on Earth, H/L ratios were measured from outcrops (Heim, 1932), but in submarine environments H/L ratios were measured using bathymetric and/or side-scan sonar images (McAdoo et al., 2000). On Mars, H/L ratios were measured from the Viking Orbiter images (McEwen, 1989). Clearly, there is no consistency among these methodologies. 9. Unlike on Earth, field measurements of motion type and direct examination of the rock in situ are impractical in distinguishing slides from debris flows on other planets. Nevertheless, Malin (1992) interpreted slides and debris flows on Venus based on types

Mass Transport, Gravity Flows, and Bottom Currents

82

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.69 Plot of H/L (vertical fall height/length of runout) ratio versus volume of submarine MTDs by Hampton et al. (1996). For comparison, the average value for subaerial MTDs (upper curve) proposed by Scheidegger (1973) is shown. Note the upper-bound values from Edgers and Karlsrud (1982) for submarine (upper curve) and subaerial (lower curve) MTDs. Source: Redrawn from Hampton, M.A., Lee, H.J., Locat, J., 1996. Submarine landslides. Rev. Geophys. 34, 3359, with permission from American Geophysical Union.

of landforms seen on radar images acquired from the Magellan spacecraft. Costard et al. (2002) interpreted debris flows based on the observation of small gullies on Mars, seen on images obtained from the Mars Observer Camera (MOC) aboard the Mars Global Surveyor spacecraft, and using the similarities of Martian gullies with gullies in East Greenland. Miyamoto et al. (2004) interpreted debris flows on Mars using MOC images and numerical simulation. The problem is that debrite depositional facies should be interpreted using cm-scale primary sedimentary features in core or outcrop for establishing plastic rheology and laminar state of the debris flow (see “Recognition of Depositional Facies” section above). Such detailed observations cannot be made using seismic data and radar images. After over 130 years of research, since the work of Heim (1882), there is still no agreement on a unified scientific theory on long-runout MTDs. The reason is that each case is unique. More importantly, there are no consistencies in concepts, nomenclatures, data sources, and methodologies when investigating MTDs on different planets.

Mass Transport, Gravity Flows, and Bottom Currents

2.12 Reservoir characterization

83

FIGURE 2.70 Plot of H/L (vertical fall height/length of runout) ratio versus volume of MTDs on Mars and Earth. Filled circles 5 data points from Valles Marineris on Mars (McEwen, 1989, his Table 1). Filled squares 5 data points for dry-rock avalanches of nonvolcanic origin on Earth (Scheidegger, 1973; Hsii, 1975). Lines are linear least-squares fits. Source: Redrawn from McEwen, A.S., 1989. Mobility of rock avalanches: evidence from Valles Marineris, Mars. Geology 17, 11111114, with permission from Geological Society of America.

2.12 Reservoir characterization An accurate depiction of depositional facies is crucial in reservoir characterization of deep-water MTDs. However, there are cases in which the use of the term landslide has created unnecessary confusion. For example, Welbon et al. (2007, p. 49) state, “Landslides can consist of rotational slips, translational slide blocks, topples, talus slopes, debris flows, mudslides and compressional toes which combine in different proportions to form complex landslides . . . Processes of landslide deformation include slip-on discrete surfaces, distribution of shear within the landslide, vertical thinning and lateral spreading through shear, fluidization, porosity collapse and loss of material from the top or toe of the complex. These processes control the quality of the resultant reservoirs.” This reservoir characterization raises the following fundamental questions:

Mass Transport, Gravity Flows, and Bottom Currents

84

2. Mass transport: slides, slumps, and debris flows

• What are the criteria for distinguishing deposits of topples with no sliding motion from those of debris flows with flowing motion in core or on seismic profiles? • Does the porosity collapse occur in deposits of topples? • If so, what are the criteria to recognize porosity collapse in deposits of topples in the subsurface? • What is the point in including a landform (talus slope) along with a process (debris flow) under the term landslide? For clarity, reservoir characterization of deep-water sands must identify the process-specific depositional facies such as slides, slumps, and debrites. In reservoir characterization, wireline (e.g., gamma-ray) log motifs are the basic subsurface data that are routinely used by the petroleum industry. Interpreting a process-specific depositional facies (e.g., slide vs debrite) from a log motif, without corresponding sediment core, is impossible. For example, analogous to sandy slide blocks that are sandwiched between deepwater mudstones in outcrops (Fig. 2.42), long-runout sandy debrite bodies (Fig. 2.71) are likely to generate blocky motifs on wireline logs in the subsurface (Fig. 2.72A). In distinguishing sandy slides (Fig. 2.72B) from sandy debrites (Fig. 2.72C) in the ancient stratigraphic record, direct examination of the rocks is crucial. The other issue is the differences in reservoir quality between slides and debrites. Large sandy slides commonly contain multiple original strata (Fig. 2.42). In cases where lithified strata are transported as sandy slides almost intact, the slided bodies are likely to represent original porosity and permeability (i.e., pretransport reservoir quality) from the provenance

FIGURE 2.71 Conceptual model showing long-runout sandy debrite blocks away from the shelf edge. Based on studies of sandy debris flows and their deposits in flume experiments (Shanmugam, 2000a; Marr et al., 2001), documentation of long-runout sandy debris flows in modern oceans (Gee et al., 1999) and interpretation of longrunout ancient “olistolith” (Teale and Young, 1987). This model is useful in developing deep-water depositional models for sandstone reservoirs of debrite origin. Source: After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524, with permission from Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

2.12 Reservoir characterization

85

FIGURE 2.72 (A) Hypothetical wireline log showing blocky motif for both sandy slide and sandy debrite units. Blocky gamma-ray (wireline) log motifs, among other motifs, are basic subsurface data that are routinely used by the petroleum industry (e.g. Shanmugam et al., 1995a). The primary control of log motif is sediment texture (i.e., sand vs mud), not individual primary sedimentary structures. Without direct examination of the rocks for sedimentary structures, distinguishing between a slide and a debrite facies is impossible from wireline log motifs alone. (B) Hypothetical sedimentological log of a sandy slide unit, composed of three original layers representing pretransport strata and texture from the provenance region, with basal shear surface and sand injection. 1. Sandstone. 2. Shale. 3. Sandstone. Note that layer 2 (shale) may act as a permeability barrier and that layer 3 (upper sandstone) and layer 1 (lower sandstone) may behave as two separate flow units during production. (C) Hypothetical sedimentological log of a sandy debrite unit with floating mudstone clasts and quartz granules (red circles). This debrite sandstone without permeability barrier would behave as a single flow unit. VF 5 very fine sand; F 5 fine sand; M 5 medium sand. After Shanmugam (2015a). With permission from Elsevier.

(Fig. 2.72B). On the other hand, debrites are likely to represent posttransport depositional texture and reservoir quality (Fig. 2.72C). Furthermore, if a sandy slide unit contains two sandstone reservoirs with an intervening shale layer, the shale layer could act as a permeability barrier (Fig. 2.72B). In such cases, a single slide unit would have to be characterized as two separate petrophysical flow units. By contrast, a single debrite unit, without a permeability layer, would be characterized as a single petrophysical flow unit (Fig. 2.72C).

Mass Transport, Gravity Flows, and Bottom Currents

86

2. Mass transport: slides, slumps, and debris flows

FIGURE 2.73 Bathymetric map of the study area showing Seismic Profile IT-EG08B, which is highlighted in yellow. Off NW Barents Sea. Source: From Madrussani, G., Rossi, G., Rebesco, M., Picotti, S., Urgeles, R., Llopart, J., 2018. Sediment properties in submarine mass-transport deposits using seismic and rock-physics off NW Barents Sea. Mar. Geol. 402, 264278, with permission from Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

2.13 Synopsis

87

2.13 Synopsis • The general term “mass transport” (i.e., slides, slumps, and debris flows) represents the failure, dislodgement, and downslope movement of either sediment or glacier under the influence of gravity. As a collective sedimentologic phenomenon, MTDs reveal remarkable cosmic congruity in geometry on Mars, Jupiter, and Earth (i.e., subaerial, sublacustrine, and submarine environments). In understanding the process-product link, this chapter is devoted to defining and distinguishing individual process types and their deposits.

FIGURE 2.74 Profile IT-EG08B (see location in Figure 2.73). (A) Prestack-depth migrated seismic section showing two packets of chaotic reflections (PLS-1 and PLS-2). (B) The same seismic section with the picked horizons showing two MTD-1 and MTD-2 packets. Source: From Madrussani, G., Rossi, G., Rebesco, M., Picotti, S., Urgeles, R., Llopart, J., 2018. Sediment properties in submarine mass-transport deposits using seismic and rock-physics off NW Barents Sea. Mar. Geol. 402, 264278, with permission from Elsevier. Additional labels by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

88

2. Mass transport: slides, slumps, and debris flows

• The term “landslide” has been in use in a variety of scientific domains since 1838 without conceptual clarity. During the past 181 years, our failure to adopt a sound process-specific terminology has resulted in 79 superfluous MTD types in the geologic and engineering literature. This profligate period of “kicking the can down the road” must end now. Only slides, slumps, and debrites can be meaningfully interpreted in the sedimentary record. Therefore, the term “landslide” should be restricted solely to MTDs in which a sliding motion can be empirically determined. A precise interpretation of a depositional facies (e.g., sandy slide vs sandy debrite) is vital not only for maintaining conceptual clarity but also for characterizing petroleum reservoirs. • However, in the real world of deep-marine settings (Fig. 2.73), one has to deal with packets of deposits that exhibit chaotic reflections on seismic profiles (Fig. 2.74). An example is the one reported by Madrussani et al. (2018) off NW Barents Sea (Fig. 2.73). In cases like this, the most practical way is to apply the collective term MTD for the entire packet (Fig. 2.74). • Available data show that debris flows alone can generate tongue-shaped geometries in both experiments (Fig. 2.12) and natural environments (Figs. 2.7 and 2.9). • At present, there are no reliable criteria for distinguishing SSDS associated with masstransport processes from those SSDS that are associated with early deformation of sediments during tectonic deformation.

Mass Transport, Gravity Flows, and Bottom Currents

C H A P T E R

3 Gravity flows: debris flows, grain flows, liquefied/fluidized flows, turbidity currents, hyperpycnal flows, and contour currents O U T L I N E 3.1 Introduction

90

3.2 Gravity (i.e., density) flows

92

3.3 Gravity-driven downslope processes 3.3.1 Mass transport versus turbidity currents 3.3.2 Sediment-gravity flows 3.3.3 Newtonian versus plastic fluid rheology 3.3.4 Turbulent versus laminar flow state 3.3.5 Sediment concentration

92 92 93 95 96 97

3.4 Debris flows 3.4.1 Definition 3.4.2 Origin 3.4.3 Identification 3.4.4 Case studies 3.4.5 Facies models 3.4.6 Problem

98 98 99 99 101 120 120

3.5 Liquefied/fluidized flows

126

Mass Transport, Gravity Flows, and Bottom Currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00003-5

3.5.1 3.5.2 3.5.3 3.5.4 3.5.5 3.5.6

89

Definition Origin Identification Case studies Facies models Problem

126 127 130 131 131 131

3.6 Grain flows 3.6.1 Definition 3.6.2 Origin 3.6.3 Identification 3.6.4 Facies models 3.6.5 Problem

131 131 132 132 133 133

3.7 Turbidity currents 3.7.1 Definition 3.7.2 Origin 3.7.3 Identification 3.7.4 Case studies 3.7.5 Facies models 3.7.6 Problem

133 133 135 136 136 136 138

© 2021 Elsevier Inc. All rights reserved.

90

3. Gravity flows

3.8 Hyperpycnal flows: a prelude 3.8.1 Definition

140 140

3.9 Thermohaline contour currents: a prelude 145

3.9.1 Definition 3.9.2 Downslope initiation 3.10 Synopsis

145 145 146

3.1 Introduction Gravity flows are the most consequential sedimentary phenomena in the geologic record. From a sedimentological perspective, gravity flows are ubiquitous in both subaerial and subaqueous environments. Importantly, gravity flows dominate in shelf, slope, and basin environments. They are caused not only by sediment density, but also by changes in temperature and salinity. Furthermore, density-driven flows travel not only downslope, but also alongslope. Therefore, the key objectives of this chapter are (1) to identify and discuss basic types of gravity flows, (2) to provide a clear definition of each flow type, (3) to identify their origins or triggering mechanisms, (4) to suggest identification markers of their deposits, and (5) to identify the remaining unresolved problems in aiding future research. I have attempted to accomplish these objectives by integrating: 1. 2. 3. 4. 5. 6.

theoretical considerations, experimental verifications, modern submarine observations, modern subaerial observations, ancient outcrop examples, and modern and ancient subsurface (sediment core) examples.

The following six basic types of gravity (density) flows are selected for discussion in this review (Table 3.1). The density value cited for each example is to provide a relative sense, and they should not be considered typical of the given example. 1. 2. 3. 4. 5. 6.

hyperpycnal flows (ρ): 0.025 g cm23 (Wright and Nittrouer, 1995), turbidity currents (ρ): 1.1 g cm23 (Kuenen, 1966), debris flows (ρ): 2.0 g cm23 (Hampton, 1972), liquefied/fluidized flows (ρ): 1.8 g cm23 (Breien et al., 2010), grain flows (ρ): 2.1 2.3 g cm23 (Parsons et al., 2001a), and Antarctic bottom water (thermohaline contour currents, THCC) (ρ): 0.03 g cm23 (Purkey et al., 2018).

According to Marr et al. (2001, their Table 3.1), sandy debris flows with densities ranging from 1.62 to 1.86 g cm 3 were used in flume experiments. This topic should be helpful from both an academic and an applied point of view. For example, identification markers of deposits of gravity flows are of practical value because sandy debrites and associated mass-transport deposits are important petroleum reservoirs in the North Sea (Shanmugam et al., 1995a); Nigeria (Shanmugam, 1997b; Shanmugam

Mass Transport, Gravity Flows, and Bottom Currents

91

3.1 Introduction

TABLE 3.1 Six types of gravity flows and their characteristics in downslope and alongslope settings. Flow type (Fig. 3.61)

Reliability of identification markers

Flow attributes

Environment

Origins (triggers)

Density of river water . density of basin water SSC (suspended sediment concentration) (ρ): 0.025 g cm23 (Wright and Nittrouer, 1995) Newtonian rheology Turbulent state C , 9% by volume (Bagnold, 1962) Flow density (ρ): 1.1 g cm23 (Kuenen, 1966) Deposition by settling Plastic rheology Laminar state C: 25% 100% Flow density (ρ): 2 g cm23 (Hampton, 1972)

Subaqueous only, near shoreline

River floods

Unreliable facies model because of absence of modern sediment core and experimental observations (Shanmugam, 2018a)

Subaqueous only, shelf, slope, and basin

Earthquake, slope instability, oversupply of sediment, volcanism, meteorite impact, tsunamis, cyclones

Reliable normal grading

Subaerial and subaqueous

Earthquake, slope instability, oversupply of sediment, volcanism, meteorite impact, tsunamis, cyclones

Reliable markers because of modern examples and experimental observations (Shanmugam, 2000a; Marr et al., 2001)

4. Liquefied/ Upward-moving fluid fluidized flow Flow density (ρ): 1.8 g cm23 (Breien et al., 2010)

Subaerial and subaqueous

Earthquakes, volcanism, Reliable markers meteorite impacts, because of modern tsunamis, cyclones examples in earthquakeinduced SSDS (Shanmugam, 2017a)

5. Grain flow

Frictional strength Grain collision (dispersive pressure) Flow (ρ): 2.1 2.3 g cm23 (Parsons et al., 2001a,b)

Subaerial and Climate, wind, steep subaqueous, gradients Aeolian dunes, and submarine canyons

Reliable markers because of modern examples in Aeolian dunes and in submarine canyons

6. Thermohaline contour current (THCC)

Current reworking Antarctic bottom water (AABW) density (ρ): 0.03 g cm23 (Purkey et al., 2018) Bottom water density in Ross Sea, Antarctica at 4000 m water depth (ρ): 0.03 g cm23 (Henze, 2015, her Fig. 2.14)

Subaqueous only, shelf edge, slope, basin

Reliable markers because of modern sediment cores (Hollister, 1967)

1. Hyperpycnal flow

2. Turbidity current

3. Debris flow

En masse freezing

Shelf freezing (temperature and salinity) in Antarctica. Note that THCC began as downslope gravity flows (Fig. 3.60), but became a contour current

Mass Transport, Gravity Flows, and Bottom Currents

92

3. Gravity flows

et al., 1995b); Bay of Bengal (Shanmugam et al., 2009); Gulf of Mexico, Russia, and Australia (Meckel, 2009, 2010); and China (Zou et al., 2012; Fudol et al., 2019). Global economic significance of sandy contourites has been discussed by Viana (2008), Stow et al. (2011), and Shanmugam (2016b). In terms of regional importance, Mullins et al. (1980), Mulder et al. (2019), and Eberli and Betzler (2019) discussed carbonate sandy contourites in the Straits of Florida, and Shanmugam et al. (1993a,b, 1995c) documented measured porosity and permeability values of petroleum-producing sandy contourites in the Ewing Bank area of the Gulf of Mexico.

3.2 Gravity (i.e., density) flows The term “flow” is used here for a continuous, irreversible deformation of sedimentwater mixture that occurs in response to applied shear stress, which is gravity in most cases (Pierson and Costa 1987, p. 2). In this book, density flows and gravity flows are considered to be one and the same, although density represents mass per unit volume and gravity represents a force. Gravity flows have been of great interest to sedimentologists and engineers for over 100 years, since the first discussion of theory of turbulence in fluid mechanics (Prandtl, 1925, 1926). Selected publications on this domain are Kuenen (1951 and 1953), Bates (1953), Bagnold (1954 and 1962), Dott (1963), Sanders (1965), Middleton (1965, 1966, 1967, 1970, 1993), Klein (1966 and 1975), Middleton and Bouma (1973), Hampton (1972), Middleton and Hampton (1973), Lowe (1976a, 1976b, 1982), Kneller (1995), Shanmugam, 1996a, 2000a, 2002a, 2006a, 2012a, 2015a,b, 2018a, 2019a), Iverson (1997), Rebesco et al. (2008), and Zenk (2008).

3.3 Gravity-driven downslope processes In Chapter 2, Mass Transport: Slides, Slumps, and Debris Flows (Fig. 2.29), I discussed four types of gravity-driven downslope processes. Because all six types discussed here begin their journey as downslope gravity flows, some basic principles are briefly explained.

3.3.1 Mass transport versus turbidity currents As discussed in Chapter 2, Mass Transport: Slides, Slumps, and Debris Flows, Dott (1963) proposed the most meaningful and practical classification of subaqueous gravitydriven processes. In this scheme, subaqueous processes are broadly classified into (1) elastic, (2) elastic and plastic, (3) plastic, and (4) viscous fluid types based on mechanical behavior (Fig. 2.29). The elastic behavior represents rockfall; the elastic and plastic behavior comprises slide and slump; the plastic behavior represents debris flow, and the viscous fluid represents Newtonian turbidity current. The importance of Dott’s (1963) classification is that mass-transport processes do not include turbidity currents. In this article, although mass-transport processes are composed of three basic types: (1) slide, (2) slump, and (3) debris flow, only debris flow is considered as a “flow.” The reason is that slides and slumps are coherent masses and they are not composed of sediment-water mixtures, a

Mass Transport, Gravity Flows, and Bottom Currents

3.3 Gravity-driven downslope processes

93

condition that is a prerequisite in defining a “flow” (Pierson and Costa, 1987). For this reason, debris flows are discussed under both “Mass transport” (Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows) and “Gravity flows” this chapter. The underpinning principle of Dott’s (1963) classification is the separation of solid from fluid mode of transport based on sediment concentration. In the solid (elastic and plastic) mode of transport, high sediment concentration is the norm (25% 100% by volume, Fig. 2.29B). Mass-transport mechanisms are characterized by solid blocks or aggregate of particles (mass). In contrast, individual particles are held in suspension by fluid turbulence in turbidity currents (Dott, 1963; Sanders, 1965). Turbidity currents are characterized by low sediment concentration, commonly below 9% sediment concentration by volume (Bagnold, 1962). Mass transport can operate in both subaerial and subaqueous environments, but turbidity current can operate only in subaqueous environments. The advantage of this classification is that physical features preserved in a deposit directly represent the physics of sediment movement that existed at the final moments of deposition. The link between the deposit and the physics of the depositional process can be established by practicing the principle of process sedimentology (Sanders, 1963), which is detailed bed-by-bed description of sedimentary rocks and their process interpretation (Sanders, 1963; Shanmugam, 2006a).

3.3.2 Sediment-gravity flows Middleton and Hampton (1973) used the term “sediment-gravity flow” (synonyms: gravity flows, density flows, mass flows, sediment flows) as a general term for flow of sediment-fluid mixtures under the action of gravity. Middleton and Hampton (1973) also distinguished sediment-gravity flows from fluid-gravity flows. Sediment-gravity flows are mixtures of water and sediment particles in which the gravity acting on the sediment particles moves the fluid downslope. In contrast, as in rivers, where fluid is directly driven by gravity that moves the particles. Furthermore, Middleton and Hampton (1973) classified sediment-gravity flows into four types based on sediment-support mechanisms (Fig. 3.1). They are: 1. 2. 3. 4.

turbidity current with turbulence; fluidized sediment flow with upward-moving intergranular flow; grain flow with grain interaction (i.e., dispersive pressure); and debris flow with matrix strength. Sandy debris flows occupy an intermediate region between debris flows and grain flows (Fig. 3.1).

Despite the simple nature of the classification, the sediment-support concept has been subjected to a major critique (Dasgupta, 2003). Furthermore, the complex variability in character of sediment-gravity flows, with emphasis on turbidity currents, has been the subject of numerous publications since the 1950s (Kuenen, 1950a; Bagnold, 1962; Dott, 1963; Sanders, 1965; Simpson, 1972; Lowe, 1979, 1982; Allen, 1985; Parker et al., 1986; McCave and Jones, 1988; Postma et al., 1988; Ghibaudo, 1992; Mutti, 1992; Kneller, 1995; Shanmugam, 1996a, 2000a; Kneller and Buckee, 2000; Marr et al., 2001; Mulder and

Mass Transport, Gravity Flows, and Bottom Currents

94

3. Gravity flows

FIGURE 3.1 Classification of sediment-gravity flows based on sediment-support mechanisms by Middleton and Hampton (1973). The position of sandy debris flow is shown for comparison.

Alexander, 2001; Baas et al., 2004; Parsons et al., 2007; Manica, 2009, 2012; Khripounoff et al., 2012, among many others). Because vital properties of deep-water turbidity currents have never been measured directly, some publications have proposed unrealistic nomenclature and attributes of turbidity currents and their deposits, which resulted in confusion. For example: • Middleton and Hampton (1973) used the term “mass flow” as a synonym for “sediment-gravity flow.” This is incongruous because turbidity currents are not mass flows. However, debris flows are indeed mass flows because they exhibit plastic flow behavior with shear stress distributed throughout the mass (Nardin et al., 1979). • McCave and Jones (1988, p. 250) advocated, “. . .deposition of ungraded muds from high-density nonturbulent turbidity currents.” By definition, turbidity currents are turbulent (Bagnold, 1962; Dott, 1963; Sanders, 1965; Middleton and Hampton, 1973). • Postma et al. (1988) considered the bottom laminar (nonturbulent) layer as part of highdensity turbidity currents. • Kneller (1995) proposed inverse grading as a product of turbidity currents. • Kneller and Buckee (2000) claimed that turbidity currents can be nonturbulent (i.e., laminar) in state. Furthermore, they claimed that turbidity currents are natural phenomena whose exact hydrodynamic properties are unclear. If so, what is the meaning of turbidite facies models, such as those proposed by Bouma (1962) and by Lowe (1982)?

Mass Transport, Gravity Flows, and Bottom Currents

3.3 Gravity-driven downslope processes

95

• Mulder and Alexander (2001) introduced a new classification of density flows in which “grain flows” are classified as “hyperconcentrated density flows.” According to Pierson and Costa (1987), hyperconcentrated flow, which is intermediate between stream flow and debris flow (Beverage and Culbertson, 1964), has a sediment concentration of 20% 60% by volume. However, grain flows have a concentration value of 100% by volume (Fig. 3.4). The redundant use of an existing nomenclature “hyperconcentrated flow” with a specific concentration for grain flows is unnecessary and confusing. • Mutti et al. (2003, p. 745) justified Kuenen’s (1951) flawed definition of “high-density turbidity currents” for the sake of maintaining a stable terminology. • Manica (2012) states that “Gravity (or density currents) currents are a general class of flows (also known as stratified flows) in which flow takes place because of relatively small differences in density between two flows (Middleton, 1993).” Not all gravity flows are stratified. Stratified flows refer specifically to gravity flows with two distinct layers of different rheology, state, and sediment concentration (see Fig. 3.38). Nonstratified gravity flows (i.e., single layer) are common in nature. • In an observational study of the Var submarine canyon (northwestern Mediterranean Sea), Khripounoff et al. (2012) recognized three kinds of sediment-gravity flows. The first kind, triggered during a flood of the Var River, was determined to be a hyperpycnal current with a large vertical extent ( . 100 m high) and relatively low velocity (40 cm s21). The second kind, observed after a Var River flood, was more energetic with a maximum horizontal current peak of 60 cm s21 but with a low vertical extent (30 m high). This event was considered to be a turbidity landslide. The third was the result of a local canyon wall failure. It was characterized by a speed of .85 cm s21. As discussed in Chapter 2, Mass Transport: Slides, Slumps, and Debris Flows, landslides are not sediment-gravity flows. • In a recent study of turbidity currents in the Monterey Canyon entitled “Powerful turbidity currents driven by dense basal layers,” Paull et al. (2018) acknowledged the following points: • We lack detailed in situ seabed measurements of how dense remobilized layers originate. • The fundamental structure of turbidity currents has remained unresolved despite being essential input for modeling and predicting turbidity current dynamics, their impact on seafloor infrastructure, and the architecture of their deposits. • Additional problems have been identified and discussed in Section 3.7.6 under turbidity currents.

3.3.3 Newtonian versus plastic fluid rheology In this chapter, the focus is on debris flows and turbidity currents because of their importance. These two processes are distinguished from one another on the basis of fluid rheology and flow state. The rheology of fluids can be expressed as a relationship between applied shear stress and rate of shear strain (Fig. 3.2). Newtonian fluids (i.e., fluids with no inherent strength), such as water, will begin to deform the moment shear stress is applied, and the deformation is linear. In contrast, some naturally occurring materials (i.e.,

Mass Transport, Gravity Flows, and Bottom Currents

96

3. Gravity flows

FIGURE 3.2 Rheology (stress strain relationships) of Newtonian fluids and Bingham plastics. Graph shows that the fundamental rheological difference between debris flows (Bingham plastics) and turbidity currents (Newtonian fluids) is that debris flows exhibit strength, whereas turbidity currents do not. Reynolds number is used for determining whether a flow is turbulent (turbidity current) or laminar (debris flow) in state (after Shanmugam, 1997a). Source: Compiled from several sources: Dott Jr., R.H., 1963. Dynamics of subaqueous gravity depositional processes. AAPG Bull. 47, 104 128; Enos, P., 1977. Flow regimes in debris flow. Sedimentology 24, 133 142; Pierson, T.C., Costa, J.E., 1987. A rheologic classification of subaerial sediment-water flows. In: Costa, J.E., Wieczorek, G.F. (Eds.), Debris Flows/Avalanches: Process, Recognition, and Mitigation, Geological Society of America Reviews in Engineering Geology, VII, pp. 1 12; Phillips, C.J., Davies, T.R.H., 1991. Determining rheological parameters of debris flow material. Geomorphology 4, 101 110; Middleton, G.V., Wilcock, P.R., 1994. Mechanics in the Earth and Environmental Sciences. Cambridge University Press, Cambridge, p. 459. Elsevier.

fluids with strength) will not deform until their yield stress has been exceeded (Fig. 3.2); once their yield stress is exceeded, deformation is linear. Such materials (e.g., wet concrete) with strength are considered to be Bingham plastics (Fig. 3.2). For flows that exhibit plastic rheology, the term plastic flow is appropriate. Using rheology as the basis, deep-water sediment flows are divided into two broad groups, namely, (1) Newtonian flows that represent turbidity currents and (2) plastic flows that represent debris flows.

3.3.4 Turbulent versus laminar flow state In addition to fluid rheology, flow state is used in distinguishing laminar debris flows from turbulent turbidity currents. The difference between laminar and turbulent flows was demonstrated in 1883 by Osborne Reynolds, an Irish engineer, by injecting a thin stream of dye into the flow of water through a glass tube. At low rates of flow, the dye stream traveled in a straight path. This regular motion of fluid in parallel layers, without

Mass Transport, Gravity Flows, and Bottom Currents

3.3 Gravity-driven downslope processes

97

FIGURE 3.3 Depth-velocity diagram showing laminar and turbulent fields of fluids (partly after Allen, 1984 and Enos, 1977; after Shanmugam, 2012a). Source: Elsevier.

macroscopic mixing across the layers, is called a laminar flow. At higher flow rates, the dye stream broke up into chaotic eddies. Such an irregular fluid motion, with macroscopic mixing across the layers, is called a turbulent flow. The change from laminar to turbulent flow occurs at a critical Reynolds number (the ratio between inertia and viscous forces) of about 2000 (Fig. 3.3).

3.3.5 Sediment concentration Sediment concentration is the most important property in controlling fluid rheology (Fig. 3.4). Classification of gravity-driven sediment flows into Newtonian and plastic types is based on fluid rheology. Turbidity currents are Newtonian flows, whereas all mass flows (muddy debris flows, sandy debris flows, and grain flows) are plastic flows. Turbidity currents occur only as subaqueous flows, whereas debris flows and grain flows can occur both as subaerial and as subaqueous flows. High-density turbidity currents are not meaningful in this rheological classification because their sediment concentration values represent both Newtonian and plastic flows (see Shanmugam, 1996a). It is important to note that a typical turbidity current can exist only in sediment concentration less than 9% by volume (Bagnold, 1962). In the following discussion, each gravity flow is evaluated with the above principles in mind.

Mass Transport, Gravity Flows, and Bottom Currents

98

3. Gravity flows

Rheology

Newtonian

Plastic

Sediment support

Dispersive pressure Cohesive strength, frictional strength and buoyancy

Transport mode

Turbulence Mass transport Suspended load

Grain flows

Flow type

Sandy debris flows Muddy debris flows Hyperconcentrated flows “High density turbidity currents” Low density turbidity currents

0

10

20

30

40

50

60

70

80

90

100

Sediment concentration (% by volume) FIGURE 3.4 Classification of gravity-driven sediment flows, based on sediment concentration, into Newtonian and plastic types. Turbidity currents are Newtonian flows, whereas all mass flows (muddy debris flows, sandy debris flows, and grain flows) are plastic flows. Turbidity currents occur only as subaqueous flows, whereas debris flows and grain flows can occur both as subaerial and as subaqueous flows. Sediment concentration is the most important property in controlling fluid rheology. High-density turbidity currents are not meaningful in this rheological classification because their sediment concentration values represent both Newtonian and plastic flows. High-density turbidity currents are included here solely for purposes of discussion. Also, for purposes of comparison, subaerial flows (river currents and hyperconcentrated flows) are considered. Published values of sediment concentration by volume percent are: (1) river currents (1% 5%; e.g., Galay, 1987), (2) lowdensity turbidity currents (1% 23%; e.g., Middleton, 1967, 1993), (3) high-density turbidity currents (6% 44%; Kuenen, 1966; Middleton, 1967), (4) hyperconcentrated flows (20% 60%; Pierson and Costa, 1987), (5) muddy debris flows (50% 90%; Coussot and Muenier, 1996), (6) sandy debris flows (25% 95%; Shanmugam, 1997a; which was partly based on reinterpretations of various processes that exhibit plastic rheology in papers by Middleton, 1966, 1967; Wallis, 1969; Lowe, 1982; Shultz, 1984), and (7) grain flows (50% 100%; partly based on Rodine and Johnson, 1976; Shultz, 1984; Pierson and Costa, 1987) (after Shanmugam, 2000a). Source: Reproduced with permission from Elsevier.

3.4 Debris flows 3.4.1 Definition A debris flow is a sediment flow with plastic rheology and laminar state from which deposition occurs through freezing en masse. The terms debris flow and mass flow are used interchangeably because each exhibits plastic flow behavior with shear stress

Mass Transport, Gravity Flows, and Bottom Currents

3.4 Debris flows

99

distributed throughout the mass (Nardin et al., 1979). In debris flows, intergranular movements predominate over shear-surface movements. Although most debris flows move as incoherent mass, some plastic flows may be transitional in behavior between coherent mass movements and incoherent sediment flows (Marr et al., 2001). Debris flows may be mud-rich (i.e., muddy debris flows), sand-rich (i.e., sandy debris flows), or mixed types. In multibeam bathymetric data, recognition of debrites is possible. Sandy debris flows are considered because of their importance in petroleum geology (Shanmugam et al., 2009). Sandy debris flow represents an intermediate stage between grain flow and cohesive debris flow (Fig. 3.1). The concept of sandy debris flows was first introduced by Hampton (1972). Sandy debris flows are defined here on the basis of (1) plastic rheology, (2) multiple sediment-support mechanisms (cohesive strength, frictional strength, hindered settling, and buoyancy), (3) mass-transport mode, (4) more than 25% 30% sand and gravel; (5) 25% 95% sediment (gravel, sand, and mud) concentration by volume (Fig. 3.4), and (6) variable clay content (as low as 0.5% by weight) (Shanmugam, 2000a). Sandy debris flows could develop in slurries of any grain size (very fine sand to gravel), any sorting (poor to well), any clay content (low to high), and any modality (unimodal and bimodal). Sandy debris flow was misclassified as “high-density turbidity currents” (see review by Shanmugam, 1996a).

3.4.2 Origin • • • • • • •

Earthquake Slope instability on alluvial fans Oversupply of sediment Volcanism Meteorite impact Tsunamis Cyclones

3.4.3 Identification The reliability of identification of ancient debrites in the rock record is high. This is because reliable field criteria have been developed on the basis of modern analogs of gravelly debrites (Fig. 3.5) and on the basis of large laboratory flume experiments in reproducing excellent examples of sandy debris flows with various diagnostic features, such as snout and other identification markers (Section 3.4.4.1: Experimental sandy debris flows). Debris flows are capable of transporting gravel and coarse-grained sand because of their inherent strength. General characteristics of muddy and sandy debrites are: • gravel to mud lithofacies; • floating or rafted mudstone clasts, which occur at some distance above the base of bed or near the tops of sandy or muddy beds (core and outcrop) (Fig. 3.6); • planar clast fabric, which is alignment of long axis of clasts parallel to bedding (i.e., horizontal) (core and outcrop) (Figs. 3.6 and 3.7);

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.5 Underwater photograph showing a pocket of rounded cobbles up to 15 cm in diameter in massive sandy matrix at a depth of 130 m (427 ft) in Los Frailes Canyon, Baja California. Note projected nature of clasts from the upper sediment surface. Source: Photo by R.F. Dill. (after Shepard and Dill, 1966), Rand McNally & Company. Published in Shanmugam, G. 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9.: Elsevier, Amsterdam, 524 p.

FIGURE 3.6 Core photograph showing a floating mudstone clast near the top of a sand unit, Paleocene, North Sea. Planar fabric, indicative of laminar flow, and sharp upper contact, indicative of flow freezing, are considered evidence for deposition from debris flows. Source: Published in Shanmugam, G. 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 p.

101

3.4 Debris flows

FIGURE 3.7 Polished slab showing projected clasts, interpreted as freezing by debris flows. Source: From Shanmugam, G. Benedict, G. L., 1978. Fine-grained carbonate debris flow, Ordovician basin margin, Southern Appalachians. J. Sediment. Petrol. 48, 1233 1240. SEPM.

• • • • • • • • • • • • • • • • • •

floating armored mudstone balls in sandy or muddy matrix (core and outcrop); projected clasts (core and outcrop) (Fig. 3.7); imbricate clasts (experiment); brecciated mudstone clasts in sandy matrix (core and outcrop) (Fig. 3.8); concentration of larger clasts (pumice blocks) near the front of volcanic debris flows or lahars, which would result in inverse grading of clasts in the rock record; inverse grading of clasts and rock fragments with random fabric (core and outcrop) (Fig. 3.9); inverse grading of quartz granules in sandy matrix (core and outcrop); inverse grading, normal grading, inverse to normal grading, and absence of any grading of matrix (core and outcrop); huge tabular carbonate clasts (3 3 15 m) in channel-ill settings (Fig. 3.10); unusually large blobs of sandy clasts in muddy matrix (Fig. 3.11); floating quartz granules in sandy matrix (core and outcrop); pockets of gravels in sandy matrix (core and outcrop) (Fig. 3.12); preservation of delicate mud fragments with planar fabric in sandy matrix (core and outcrop); irregular, sharp upper contacts (core and outcrop); side-by-side occurrence of garnet granules (density: 3.5 4.3) and quartz granules (density: 2.65) (core and outcrop); lenticular (Fig. 3.13) to sheet-like in geometry; lobe-like geometry (map view) in the Gulf of Mexico; tongue-like geometry (map view) in the North Atlantic (Fig. 2.4A).

3.4.4 Case studies 3.4.4.1 Experimental sandy debris flows As noted earlier, the concept of sandy debris flows was first introduced by Hampton (1972). Sandy debris flows are defined on the basis of (1) plastic rheology (Fig. 3.2), (2)

Mass Transport, Gravity Flows, and Bottom Currents

102

3. Gravity flows

FIGURE 3.8 Core photograph showing brecciated mudstone clasts (light gray) in fine-grained sand, Eocene, North Sea. In some intervals, mudstone clasts are dark gray in color suggesting possible differences in provenance. Source: Published in Shanmugam, G. 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 p.

multiple sediment-support mechanisms (cohesive strength, frictional strength, hindered settling, and buoyancy), (3) mass-transport mode (Fig. 3.3), (4) more than 25% 30% sand and gravel, (5) 25% 95% sediment (gravel, sand, and mud) concentration by volume (Fig. 3.4), and (6) variable clay content (as low as 0.5% by weight). Rheology is more important than grain-size distribution in controlling sandy debris flows. Sandy debris flows could develop in slurries of any grain size (very fine sand to gravel), any sorting (poor to well), any clay content (low to high), and any modality (unimodal and bimodal) (Shanmugam, 1996a, 1997a, 2000a). Theoretically, grain flows (i.e., cohesionless debris flows) and muddy debris flows (i.e., cohesive debris flows) may be considered to be two end-members of plastic flows (Fig. 3.14). Sandy debris flows are considered to represent an intermediate position

Mass Transport, Gravity Flows, and Bottom Currents

103

3.4 Debris flows

FIGURE 3.9 Outcrop photograph showing inverse grading with floating boulder-size clasts near the top of sandstone unit (arrow). Note random fabric of clasts. Middle Miocene, San Onofre Breccia, Dana Point, California. This lithofacies has been interpreted to be sandy debrite associated with alluvial fan and fan delta. Source: Reproduced from Shanmugam, G. (2012a). New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9, Elsevier, Amsterdam, 524 p.

FIGURE 3.10 Photograph showing a large tabular carbonate clast (3 3 15 m), interpreted as carbonate debris flow deposit in a submarine channel setting. Note base of channel marked by a black line. Hales Limestone, Lower Ordovician, Nevada. Source: Photo Courtesy of Harry. E. Cook. After Cook (1979), SEPM.

Mass Transport, Gravity Flows, and Bottom Currents

104

3. Gravity flows

FIGURE 3.11 Photograph showing a large intrabasinal sandstone clast in mudstone matrix. Pennsylvanian Jackfork Group, Arkansas. Note 15-cm scale near bottom. Source: From Shanmugam, G., Moiola, R.J., 1995. Reinterpretation of depositional processes in a classic flysch sequence (Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma. AAPG Bull. 79, 672 695. AAPG.

FIGURE

3.12 (A) Core photograph showing floating quartz pebbles in fine-grained massive sand implying flow strength. Eocene, North Sea. (B) Core photograph showing pockets of gravel in medium-grained sand implying flow strength. Pliocene, offshore Equatorial Guinea. Source: From Shanmugam, G. (2012a). New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9, Elsevier, Amsterdam, 524 p. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

3.4 Debris flows

105

FIGURE 3.13 Outcrop photograph showing lenticular geometry of debrite with floating clasts (arrowheads), Cretaceous, Tourmaline Beach, California (after Shanmugam, 2006a). Source: Reproduced with permission from Elsevier.

between grain flows (with frictional strength) and muddy debris flows (with cohesive strength) (Shanmugam, 1997a). An advantage of this concept is that it requires neither the steep slopes necessary for grain flows nor the high matrix content necessary for cohesive debris flows. One of the main criticisms leveled against the concept of sandy debris flows was the flawed notion that all debris flows must have high clay content in order to provide the necessary strength (e.g., D’Agostino and Jordan, 1997). Although Hampton (1975) noted that as little as 2% clay is sufficient to provide the strength for sandy debris flows, Costa and Williams (1984) described a number of mud-poor debris flows in which mud constituted less than 2% of the debris flow. It is worth noting that for the first time, to understand mechanics of sandy debris flows and their deposits, a Mobil-funded experimental flume study was carried out at St. Anthony Falls Laboratory (SAFL), University of Minnesota (1996 98) under the direction of Professor G. Parker. Results were published in two major articles (Shanmugam, 2000a; Marr et al., 2001). The major advantage of experiments is that it allows researchers to measure fluid dynamical properties and observe sedimentary characteristics of deposits. In our flume experiments, the distinction between debris flows and turbidity currents in terms of fluid rheology and flow state is inescapable. Importantly, in order to verify the concept of sandy debris flows with low clay content, experiments were conducted on subaqueous sandy debris flows (Marr et al., 2001; Shanmugam, 2000a). The experimental flume used was 10 m in length, 30 cm in width, and 80 cm in depth (Figs. 3.15 and 3.16). The flume was fitted with three different slopes: 4.6, 1.1, and 0 degrees to observe changes in deposition at points of slope change. These slope angles are analogous to those of modern continental slope, rise, and abyssal plain. Sediment slurries were composed of silica sand (120 µm size), clay (bentonite or kaolinite), coal slag (same bulk density as silica sand: 2.6 g cm23), and water (Fig. 3.17). Coal slag of 500 µm size (coarse sand) was used as a tracer material to establish flow behavior

Mass Transport, Gravity Flows, and Bottom Currents

106

3. Gravity flows

FIGURE 3.14 Theoretical versus natural debris flows. Theoretically, grain flows (i.e., cohesionless debris flows) and muddy debris flows (i.e., cohesive debris flows) are considered to be two end-members of rheological debris flows (Lowe, 1979). Following Lowe (1979), the rheological term “plastic” is used for both grain flows (frictional strength) and muddy debris flows (cohesive strength). Sandy debris flows, not studied before in experiments, are considered to represent an intermediate position between the end-member types. Multiple sediment support mechanisms are proposed for sandy debris flows. An advantage of this concept is that it requires neither the steep slopes required for grain flows nor the high matrix content necessary for cohesive debris flows (after Shanmugam, 1997a). Source: Reproduced with permission from Elsevier.

FIGURE 3.15

Dimensions of the flume used in sandy debris flow experiments (after Shanmugam, 2000a) Source: Reproduced with permission from Elsevier. Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.16 Flume used in the experiments of sandy debris flows. See Shanmugam (2000a) and Marr et al. (2001) for details on experiments. Source: Photo by G. Shanmugam.

FIGURE 3.17 (A) Materials used in experiments. (B) Sand-clay-water mix.

108

3. Gravity flows

and depositional pattern of coarse-grained grains in comparison to very fine-grained sandy matrix. Sandy debris flows were generated with bentonite clay content as low as 0.5% by weight or with kaolinite clay as low as 5% by weight. Sandy debris flows were also generated using medium-grained sand (300 µm size) with bentonite clay content as low as 1.5% by weight or kaolinite clay as low as 5% by weight. On the basis of experimental sandy debris flows, the following general observations and inferences have been made (Shanmugam, 2000a): 1. Sandy debris flows are a viable mechanism for transporting and depositing sand in subaqueous environments. 2. Sandy debris flows can travel long distances on gentle slopes of less than 1 degree. 3. Contrary to popular belief, sandy debris flows do not require high clay content. A clay content as low as 0.5% is sufficient to generate sandy debris flows. However, without at least 0.5% of clay, debris flows will not develop. In the complete absence of clay, the sand water slurry either becomes a short-lived grain flow or a short-lived turbidity current. 4. Sandy debris flows are developed from slurries of both bimodal and unimodal grainsize distribution. 5. The ratio of water to clay and the types of clay determined the flow behavior. For example, by maintaining a constant amount of kaolinite at 15% by weight, and by increasing the water content to 25%, 30%, and 40% by weight, three different types of sandy debris flows (i.e., strong, moderate, and weak) were generated (Fig. 3.18). The increase in water reflects a decrease in fluid strength. The significance of this observation is that changes in water content alone can make a difference in the flow behavior without changing clay content. Thus, the amount of clay in the deposit is not always a useful criterion for interpreting the nature of flow. Primary sedimentary features are more reliable for interpreting flow behavior than clay content. Deposits of sandy debris flows with low clay content (e.g., 0.5%) are potential candidates for misinterpretation as deposits of high-density turbidity currents or grain flows. 6. Strong debris flows developed thick fronts with well-defined body, whereas weak debris flows developed poorly defined body (Fig. 3.18). Strong flows also showed well-developed snouts (Fig. 3.19). 7. Weak flows developed thick turbulent suspension (i.e., turbidity current) on top (Fig. 3.20). Such density-stratified flows could be mistakenly classified as “highdensity turbidity currents” (e.g., Postma et al., 1988). 8. Subaqueous debris flows developed hydroplaning (Fig. 3.21), whereas subaerial debris flows did not (Mohrig et al., 1998). Experimental studies of subaqueous debris flows have shown that hydroplaning can dramatically reduce the bed drag and thus, increase head velocity. This explains why subaqueous debris flows can travel fast and afar on gentle slopes. 9. Water-escape structures (dishes and pillars) have been observed in experimental sandy debris flows. Dish structures in experimental sandy debris flows formed in three stages, namely: (1) hydroplaning, (2) water entrapment (Fig. 3.22), and (3) water escape (Fig. 3.23). During the hydroplaning stage, water penetrates underneath the plastic flow layer (Fig. 3.23, Stage 1). When the deposit begins to settle, water gets

Mass Transport, Gravity Flows, and Bottom Currents

109

3.4 Debris flows

FIGURE 3.18 Weak, moderate, and strong types of sandy debris flows and their properties (after Shanmugam, 2000a). Source: Reproduced with permission from Elsevier.

FIGURE 3.19 Side view of flume tank showing strong debris flows with welldeveloped snout. Note the absence of turbulent suspension on top. Also note irregular upper surface caused by sudden freezing of the flow. Deformation in the front suggests strongly coherent character of flow, which may be called a slump. Source: Reproduced from Shanmugam, G. (2006a). Deep-water processes and facies models: implications for sandstone petroleum reservoirs. In: Handbook of Petroleum Exploration and Production, vol. 5. Elsevier, Amsterdam, 476 p.

trapped in cavities underneath the bed (Fig. 3.23, Stage 2). Finally, further settling of sediment causes the trapped water to escape by bursting open the top of the cavity, resulting in a sand volcano. A fully developed volcano would form a dish-shaped basal surface (Fig. 3.23, Stage 3), which would eventually mimic dish structures in the rock record. Water escape also results in vertical pipes or pillars. Water-escape structures in deep-water sands were previously used as evidence for deposition from

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE

3.20 Side view of flume tank showing weak debris flows with well-developed turbulent clouds (TC) on top of sandy debris flows (SDF). Dashed line marks the boundary between laminar debris flow and turbulent turbidity current. Such density-stratified flows may be erroneously classified as high-density turbidity currents (see Figs. 3.38 and 3.39B).

FIGURE 3.21 Three profiles of experimental debris flows. (A) Slow-moving subaerial debris flow without hydroplaning. (B) Fast-moving subaqueous debris flow with hydroplaning (arrow) beneath the head of debris flow. (C) Fast-moving subaqueous debris flow with detached head. Note turbulent flows above subaqueous debris flows. Based on experiments of Mohrig et al. (1998) (after Shanmugam, 2002a). Source: Reproduced with permission from Elsevier.

111

3.4 Debris flows

FIGURE

3.22 Side view of flume tank showing sandy debris flows with water entrapment (arrow) beneath a debris-flow layer (i.e., Stage 2 in Fig. 3.23). The trapped water would escape when the sandy debris flow layer begins to settle toward the flume floor, causing sand volcanoes. Horizontal distance between the 4.0 and 4.1 markers is 10 cm. Flow direction is from right to left (after Shanmugam (2000a). Source: Elsevier.

FIGURE 3.23 Diagram showing stages of development of water-escape structures in sandy debris flows: (1) hydroplaning, (2) water entrapment in cavities (see Fig. 3.22), and (3) water escape. Deposition of debris-flow layer squeezes the water in the cavities causing it to escape upward, resulting in sand volcanoes, dish structures, and vertical pipes. Diagram is based on direct observations of experiments as well as observations of videotapes of experiments (after Shanmugam, 2000a). Source: Reproduced with permission from Elsevier.

liquefied flows and high-density turbidity currents (Lowe, 1975, 1982). Our experiments suggest that water-escape structures are common in subaqueous sandy debris flows with hydroplaning. Perhaps, the presence of water-escape structures in

Mass Transport, Gravity Flows, and Bottom Currents

112

3. Gravity flows

FIGURE 3.24 Normal grading developed by sandy debris flows in experiments. Note a zone of clean sand at bottom.

sandy debris flows may be used to infer hydroplaning in the rock record. The origin of dish structures and pillars is commonly ascribed to liquefaction and fluidization. 10. Both normal grading (Fig. 3.24) and inverse grading developed in sandy debris flows. Coarse-tail normal grading was observed only in weak and moderate debris flows. Settling of coarser grains occurs from suspension through hindered settling after the flow had stopped. This settling of grains from a nonturbulent (i.e., laminar) flow after a flow had halted is different from settling of grains that may occur from a turbulent turbidity current even during transport. 11. In the rock record, sandy debrites with normal grading may be misinterpreted as turbidites. However, normally graded sandy debrites can be distinguished from normally graded sandy turbidites by associated features. For example, floating clasts and granules are common in normally graded debrites, whereas floating clasts and granules are unlikely to be present in true turbidites. 12. In experiments using 300 µm size silica sand and 5 wt.% kaolinite, sandy debrites developed not only a normal grading, but also a relatively clean basal sand layer. This clean basal sand layer is attributed to sudden settling of sand grains coupled with upward migration of mud due to elutriation (Fig. 3.24). In experiments, debris flows commonly undergo flow transformation and generate turbidity currents on their top (Hampton, 1972). Surface and elutriation flow transformations are commonly responsible for transferring mud from underlying debris flows into overlying turbidity currents. Such a mechanism has important implications for developing alternative deep-water depositional models. This is because the conventional wisdom dictates that

Mass Transport, Gravity Flows, and Bottom Currents

113

3.4 Debris flows

13.

14.

15.

16. 17. 18.

only turbidites form mud-poor reservoir sands. Our experiments, however, have shown that sandy debris flows are capable of forming clean, mud-poor sands. Massive sands emplaced by sandy debris flows often exhibit random distribution of coal slag (coarse sand) throughout the bed composed of very fine-grained sand. This is analogous to floating granules in sandstone. Such massive sands in the rock record are potential candidates for misinterpretation as deposits of high-density turbidity currents. Internal layers in sandy debrites were developed by postdepositional movement along failure planes (or secondary glide planes) during remobilization (Fig. 3.25). In the rock record, such layers could be misidentified as parallel lamination, resulting in an erroneous process interpretation (i.e., traction). Imbricate slices developed when the front of a flow froze, and the body of the same flow broke away from the front end and overrode the front as successive thrust slices (Fig. 3.26). Imbricate slices suggest compression. Large-scale compressional ridges have been reported from a modern submarine “flow slide” in a fjord, British Columbia (Prior et al., 1982). Such ridges have been reported from modern glacial deposits as well (Fig. 2.53). Imbricate slices (duplex-like structures) have also been reported from the subaerial Blackhawk landslide (Shreve, 1968). The origin of duplex-like structures (i.e., imbricate slices) in the Pennsylvanian Jackfork Group, Ouachita Mountains, has been attributed to synsedimentary slumping (Shanmugam et al., 1988a). Such debris flows may be classified both as mass transport and as sediment flows because of their transitional behavior. Strong flows commonly generated sharp and irregular upper surfaces by freezing (Fig. 3.27). As noted earlier, freezing of a strong flow developed an irregular snout in the front of the flow (Fig. 3.19). During remobilization, frontal parts of sandy debris flows detached themselves from the main body and started to move ahead (outrun) of the main body as isolated blocks FIGURE 3.25 Side view of flume tank showing sandy debris flows with nearly horizontal or gently dipping internal layers (horizontal arrow). These layers are caused by postdepositional movement along failure planes (or secondary glide planes) during remobilization of flows. In the rock record, these horizontal layers could be misidentified as parallel lamination, which could result in erroneous interpretations of traction processes. Horizontal distance between the 4.3 and 4.4 markers is 10 cm. Flow direction is from right to left (after Shanmugam, 2000a). Source: Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

114

3. Gravity flows

FIGURE 3.26 Side view of flume tank showing sandy debris flows with imbricate slices (inclined arrow). Such imbrications develop in sandy debris flows when the front of a flow freezes, the body of the flow breaks and thrusts over the slice in the front due to compression. Similar features (duplex-like structures) in the rock record have been ascribed to synsedimentary slumping (Shanmugam et al., 1988a) (after Shanmugam, 2000a). Source: Elsevier.

FIGURE 3.27 Side view of flume tank showing sandy debris flow with a sharp upper contact (arrow), caused by freezing of the flow. Random distribution of coal slags is due to freezing of the flow with strength. Horizontal distance between the 5.8 and 5.9 markers is 10 cm. Flow direction is from right to left.

(Fig. 3.28). Such isolated (outrunner) blocks are evidence of tensional movement. Large bodies of isolated muddy debris flows and slumps have been reported from modern oceans (Embley, 1976, 1980; Embley and Jacobi, 1977). In summary, sandy debris flows have developed a variety of sedimentological features (Fig. 3.29). Some of these features may be misinterpreted as deposits of turbidity currents (e.g., normal grading), or even as tectonic features (e.g., duplex structures). Because of the complexity of the features of sandy debris flows, there are no simple vertical facies models for deposits of sandy debris flows. For the same reason, interpretation of sandy debris flows in the rock record would require excruciatingly detailed observations of intricate sedimentary features. The importance of basin-filling by “cohesionless debris flows” was emphasized by Syvitski and Farrow (1989, p. 30). These researchers described deposits of these sandy

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.28 Map view of experimental sandy debris flows showing isolated blocks of sand bodies (arrow). These sandy debrite bodies slowly got detached from the main body by tension. Detachments may be explained by hydroplaning and related faster-moving head with respect to the body. Width of photo is approximately 10 cm. Flow direction is from right to left (after Shanmugam, 2000a). Source: Reproduced with permission from Elsevier.

Features

Observation Sharp upper contact

Interpretation Freezing of flow and plastic rheology

Irregular upper contact and Freezing of primary relief and plastic rheology lateral pinch-out geometry

Flow direction

Irregular front (snout)

Freezing of primary relief and plastic rheology

Non-erosive base and water entrapment ( )

Laminar flow and hydroplaning

Dish structures and water ) entrapment (

Hydroplaning and water escape

Vertical pipes

Hydroplaning and water escape

Grain segregation and normal grading

Grain settling from weak flow

Planar fabric and inverse grading

Laminar flow and flow strength

Random fabric

Flow strength and freezing of flow

Internal layers

Mass movement and secondary glide planes

Imbricate slices

Mass movement and compression

Isolated blocks

Mass movement and tension

120 µm silica sand 500 µm coal slag (bulk density: 2.6 g cm–3)

FIGURE 3.29 Summary of identification markers associated with debris flows based on experiments. Source: From Shanmugam, G., 2000a. 50 years of the turbidite paradigm (1950s 1990s): deep-water processes and facies models a critical perspective. Mar. Pet. Geol. 17, 285 342.

116

3. Gravity flows

flows as: “They tend to contain structureless clean well-sorted sand with mixed grading, that is, both reverse and normal grading. They may contain synsedimentary clasts and there appears to be little difference between the characteristics of a pebbly grain flow and a sandy debris flow deposit.” Sandy debrites have been interpreted to be the dominant depositional facies in several hydrocarbon-producing reservoirs composed of massive sands in the North Sea and Norwegian Sea (Shanmugam et al., 1994 and 1995a). 3.4.4.2 Krishna-Godavari Basin, Bay of Bengal, India The Krishna-Godavari (KG) Basin is located on the eastern continental margin of India (Fig. 3.30). Operator Reliance Industries Limited and Niko Resources discovered gas in

FIGURE 3.30 (A) Index map showing locations of the Krishna-Godavari (KG) Basin and the KG-D6 block (offshore, State of Andhra Pradesh) on the eastern continental margin of India. (B) Map showing location of our study area in the Block KG-D6. (C) Root mean square (RMS) seismic amplitude map of our study area showing locations of cored wells 1, 2, and 3. RMS map represents the entire reservoir (400 ms time window). Amplitude color code: bright red, high amplitude (gas-charged sandy lithologies); yellow, intermediate amplitude (mixed lithologies); blue-to-dull green, low amplitude (nonsandy or muddy lithologies). Sinuous and lobate planform geometries are present. Note position of well 2 in a sinuous form. The seismic profile, which passes through well 2, represents an oblique strike section across a sinuous form (submarine canyon) Source: (A C) From Shanmugam, G., Shrivastava, S.K., Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736 756, with permission from SEPM.

Mass Transport, Gravity Flows, and Bottom Currents

3.4 Debris flows

117

Pliocene deep-water siliciclastic reservoirs of the KG Basin in 2002. A comprehensive study, based on integration of modern seafloor bathymetry, conventional cores, and seismic attributes, was carried out (Shanmugam et al., 2009). The cored Pliocene intervals in three wells represent the deep offshore component of KG Basin. The modern seafloor bathymetry of our study area reveals that the upper-slope setting is characterized by widespread mass-transport deposits and submarine canyons (Fig. 2.21). Most canyons are in their incipient stages of development. The western shelf edge is characterized by headwall scarps. Slide blocks have detached from these headwall scarps (i.e., slide scars) and moved downslope developing chutes. As a result, slide blocks occur at the mouths of chutes (Fig. 2.21). Such genetic links between chutes and slides on other slope settings have been documented by Prior et al. (1981). Shelf-indenting submarine canyons characteristically originate at the shelf edge (Fig. 2.21). Straight and slightly sinuous canyons are present. These canyons are at least 20 25 km long. The modern canyon walls have steep slopes of more than 8 degrees, and the canyon floors exhibit slopes of 2 4 degrees. Such steep gradients are conducive for triggering mass movements along canyon walls (e.g., Lastras et al., 2007). Based on examination of 313 m of conventional cores from three wells, five depositional facies have been interpreted (Table 3.2): 1. Facies 1: sandy debrite (Fig. 3.31) with planar fabric (Fig. 3.32A), sandy slump (Fig. 3.32B) and brecciated clasts (Fig. 3.33), sandy slide, and sandy cascading flow; 2. Facies 2: muddy slump and debrite (Fig. 3.34); 3. Facies 3: sandy tidalite (Fig. 3.35); 4. Facies 4: muddy tidalite; 5. Facies 5: hemipelagite. Debrites and slumps constitute up to 99% (Table 3.3). Sand injectites are common (Fig. 3.34). Pliocene environments are interpreted to be comparable to the modern upper continental slope with widespread mass-transport deposits and submarine canyons in the KG Basin. Both mass movements and faulting are important mechanisms in explaining the origin of submarine canyons in the KG Basin area. Deformational features observed in the mudstone unit that underlies the Pliocene sinuous canyon in the KG Basin have been explained by major mass movements (Shanmugam et al., 2009). The initiation of submarine canyons by mass movements has been discussed and documented worldwide by other researchers (Shepard, 1981; Moore et al., 1989; Ridente et al., 2007). Pliocene canyons are sinuous, exhibit 90-degree deflections, at least 22 km long, relatively narrow (500 1000 m wide), deeply incised (250 m), and asymmetrically walled in the KG Basin (Fig. 3.36). Sandy debrites occur as sinuous canyon-fill massive sands (Fig. 3.37), intercanyon sheet sands (1750 m long or wide and 32 m thick), and canyonmouth slope-confined lobate sands (3 km long, 2.5 km wide, and up to 28 m thick). Continuous bright-red amplitude in sinuous forms suggests continuous distribution of sand along the entire length of the sinuous canyon. Because upper-slope sandy debrites mimic base-of-slope turbidite channels and lobes in planform geometries, the conventional use of submarine fan models as a template to predict the distribution of deep-water sand is dubious.

Mass Transport, Gravity Flows, and Bottom Currents

118

3. Gravity flows

TABLE 3.2 Descriptions of lithofacies and interpretations of depositional facies in the Krishna-Godavari Basin. Lithofacies (description)

Depositional facies (interpretation)

Lithofacies 1 • Unindurated fine- to coarse-grained Sandy debrite, sandy slump, sandy slide, and sandy cascading flow (sandy mass-transport deposits composed mostly of sandy debrite) massive sand • Massive sand intervals with amalgamation surfaces • Floating quartz granules common • Floating mudstone clasts abundant • Mudstone clasts with random fabric • Mudstone clasts with planar fabric • Clast-rich intervals • Inverse grading of floating mudstone clasts • Inverse grading of sand matrix • Poorly sorted matrix • Sharp, irregular, and embayed upper contacts • Sharp and irregular basal contact • Shear plane immediately beneath the basal contact • Sand pillars near upper contacts • Slumped mudstone clasts • Contorted layers • Steep layers with dip up to 60 degrees • Dispersed carbonaceous debris in some intervals • Low mud matrix (,1% by volume) Lithofacies 2 • • • • • • • • • • • • • • • • • •

Indurated mudstone and claystone Internal shear planes Drag folds Slump folds common Contorted layers common Chaotic fragments Steeply dipping fabric Stretched clasts, rock fragments, and boudins Sandy injectites Sand offshoots Floating sandstone rock fragments common Floating mudstone clasts common Planar clast fabric Random clast fabric Brecciated clasts Fractures filled with quartz granules Fossil (calcareous) fragments (bivalves) Nodules (calcareous)

Muddy slump and debrite

(Continued) Mass Transport, Gravity Flows, and Bottom Currents

3.4 Debris flows

119

TABLE 3.2 (Continued) Lithofacies (description)

Depositional facies (interpretation)

Lithofacies 3 • • • • • • • •

Unindurated fine-grained sand Amalgamated units Rhythmic bedding Double mud layers common Lenticular and wavy bedding Parallel laminae Ripple laminae with mud drapes Deformed double mud layers and ripples • Flame structures • Concentration of carbonaceous fragments along mud layers • Thick-thin bundles • Calcareous nodules Lithofacies 4

Sandy tidalite

• • • •

Indurated silty mudstone Rhythmic bedding (rhythmites) Double mud layers common Mud offshoots in mud-draped ripples • Lenticular and wavy bedding • Flame structures • Calcareous nodules • Burrows Lithofacies 5

Muddy tidalite

• Indurated mudstone and claystone • Parallel silt laminae • Closely interbedded with mudstone with double mud layers • Burrows • Trace fossil (Zoophycos) • Calcite-cemented zones

Hemipelagite

Notes: Processes and their products are distinguished using the following nomenclature: debris flow (process): debrite (product); turbidity current: turbidite; tidal current: tidalite; hemipelagic settling: hemipelagite; injection: injectite; cascading flow: cascading flow; slump: slump; slide: slide. From Shanmugam, G., Shrivastava, S.K., Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736 756.

Canyon-fill facies are characterized by the close association of sandy debrites and tidalites in the KG Basin (Fig. 3.37). Reservoir sands, composed mostly of amalgamated units of sandy debrites, are thick (up to 32 m), and low in mud matrix (less than 1% by volume). The best reservoir facies is composed of sandy debrites. This facies exhibits high values of measured porosities (35% 40%) and permeabilities (850 18,700 mD) (Table 3.4). Sandy tidalites and related bottom-current reworked facies exhibit moderate porosity (31% 40%) and permeability (525 6930 mD). Muddy tidalites are poor reservoirs (Table 3.3). Muddy

Mass Transport, Gravity Flows, and Bottom Currents

120

3. Gravity flows

FIGURE 3.31 (A) Sedimentological log of core 14 m in well 2 showing floating mudstone clasts in amalgamated massive sand. (B) Core photograph showing horizontal (planar fabric) and vertical (random fabric) positions of floating mudstone clasts (arrows) in massive sand (after Shanmugam et al., 2009). Source: With permission from SEPM.

slumps and debrites and hemipelagites are considered to be nonreservoirs. Postdepositional sandy injectites, closely associated with sandy debrites, also exhibit high values of porosity (34% 35%) and permeability (8680 11,760 mD). Sandy debrites on the upper-slope settings develop not only sheet-like geometries in the intercanyon areas but also exhibit high porosities and permeabilities because of low mud matrix.

3.4.5 Facies models Debrites are perhaps the best studied and best documented facies. Ironically, there are no facies models for debrites.

3.4.6 Problem There are three major problem areas regarding interpreting coarse-grained deposits either as high-density turbidites or as sandy debrites in deep-water strata.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.32 Core photos showing (A) planar clast fabric and (B) associated slump facies. Source: From Shanmugam, G., Shrivastava, S.K., Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736 756.

FIGURE 3.33 (A) Sedimentological log showing massive sand with floating brecciated mudstone clasts, deformed double mud layers, and truncated ripples in massive sand (lithofacies 1 and 3). (B) Lithofacies 1 core photograph showing brecciated mudstone clasts. Arrow shows stratigraphic position of photograph (after Shanmugam et al., 2009). Source: With permission from SEPM.

122

3. Gravity flows

FIGURE 3.34 (A) Sedimentological log showing sand injection in a mudstone unit (lithofacies 2) that is sandwiched between massive sand units (lithofacies 1). (B) Lithofacies 2 core photograph showing injection of sand into host mudstone. Truncation of horizontal laminae in mudstone by injected sand is evident. Arrow shows stratigraphic position of photograph (after Shanmugam et al., 2009). Source: With permission from SEPM.

First, are high-density turbidity currents sandy debris flows? Conventionally, stratified flows have been classified as high-density turbidity currents. I (Shanmugam, 1996a) argued against such classifications. The prevailing differences of opinion on nomenclature can be explained by our flume experiments (Shanmugam, 2000a; Marr et al., 2001). For example, the stratified flow with lower laminar layer and an upper turbulent layer in our experiment (Fig. 3.38) would be classified differently by different researchers as follows: 1. Group 1 of researchers would recognize the importance of bottom layer with different rheology and flow state (Bagnold, 1956; Sanders, 1965; Shanmugam, 1996a). 2. Group 2 would not (Kuenen, 1951; Postma et al., 1988; Mutti et al., 1999; Zavala, 2019). Postma et al. (1988) would combine both layers and classify them together as “highdensity turbidity currents” (Fig. 3.38). Second, are floating clasts in deep-water sandstones represent sandy debrites? Experiments have shown (Fig. 3.39) that clasts indeed form along rheological boundaries on top of sandy debris flows (Fig. 3.40).

Mass Transport, Gravity Flows, and Bottom Currents

3.4 Debris flows

123

FIGURE 3.35 (A) Sedimentological log of core 8 for the interval 2072 2077.5 m in well 2 showing alternation of sand (lithofacies 3) and mudstone (lithofacies 4) intervals with continuous presence of double mud layers (DML). Note floating sandstone rock fragments and mudstone clasts in a basal mudstone interval (lithofacies 2). The cored interval represents core 8 of canyon-fill deposits in seismic profile (Fig. 3.36). (B) Lithofacies 3 core photograph showing rhythmic bedding (rhythmites) and double mud layers (DML, arrows) in sand. N 5 Neap (thin) bundle; S 5 Spring (thick) bundle (after Shanmugam et al., 2009). Source: With permission from SEPM.

In explaining the origin of floating clasts (Fig. 3.40), Postma et al. (1988, p. 55) state that “On the basis of our laboratory experiments, we suggest that outsized clasts in highdensity turbidity currents can glide and be subsequently be deposited at some height within the flow, at the boundary between a highly concentrated, nonturbulent “carpet” (inertia-flow layer) and the overlying, faster-moving turbulent layer. . .” Clearly, the origin of clasts is controlled by the basal nonturbulent layer, which I consider as “sandy debris flow.” Although Postma et al. (1988) attributed the origin of clasts to high-density turbidity currents, I would classify the genesis of clasts as a product of sandy debris flows (Fig. 3.38). The reader can decide as to which one of the two hypotheses (i.e., turbulent turbidity currents vs nonturbulent debris flows) is rational in explaining the origin of floating clasts. Third, the major unresolved issue is flow transformation in sediment-gravity flows. Fisher (1983) proposed four types of transformations for sediment-gravity flows (Fig. 3.41):

Mass Transport, Gravity Flows, and Bottom Currents

124

3. Gravity flows

TABLE 3.3 Distribution of lithofacies and depositional facies in the Krishna-Godavari Basin. Well 1

Well 2

Depositional facies

m

Lithofacies 1 (sandy debrite, slump, slide, and cascading flow)

18.2

65

48

56

1.7

1

Lithofacies 2 (muddy slump and debrite)

9.5

34

28.5

33

23.5

12

Lithofacies 3 (sandy tidalite)

0.3

1

8

9

0

0

Mixed lithofacies 3 and 4 (sandy and muddy tidalite)

0

0

0

0

29.3

15

Mixed lithofacies 5 and 4 (hemipelagite and muddy tidalite)

0

0

0

0

140

70

Lithofacies 5 (hemipelagite)

0

0

1.5

2

0

0

4.9

2

a

%

m

Well 3

%

m

Sandy injectite

Present

Present

Present

Present

Total

28

100

86

100

%

199

100

a

Sandy injectite is a postdepositional feature, but it is difficult to quantify independently of sandy debrite, slump, and slide (lithofacies 1) in wells 1 and 2. From Shanmugam, G., Shrivastava, S.K., Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736 756.

FIGURE 3.36 Seismic profile showing boundaries of a major erosional feature of Pliocene age, which we have interpreted as a submarine canyon on the upper-slope environment. Cored intervals are shown by yellow bars on the wireline log of well 2. The southeast canyon wall, which corresponds to the contact between cores 10 and 11 (rectangle box, see Fig. 3.37 for details), is characterized by slump folds, sand injections, and other sediment deformation in core. Both walls of the canyon are aligned in trend with underlying normal faults. Immediately beneath the canyon, a seismic unit (with cores 12, 13, and 14) exhibits continuous and parallel reflections. This seismic unit, which is 1750 m long or wide, is composed primarily of sandy debrites in core in the intercanyon environments. This NW-SE seismic profile represents an oblique strike section across a sinuous canyon with well 2 (after Shanmugam et al., 2009). Source: With permission from SEPM.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.37 An enlarged version of boxed area in Fig. 3.36 showing the position of southeast canyon wall and canyon-fill facies. The canyon-fill facies is composed of sandy debrites, sandy tidalites, and muddy slumps. The intercanyon facies is composed of muddy slumps and debrites with sand injectites in core. Severe sediment deformation is evident both below and above the canyon wall. The lack of core recovery at the canyon wall may be due to extreme sediment deformation (after Shanmugam et al., 2009). Source: With permission from SEPM.

TABLE 3.4 Representative porosity (%) and permeability (mD) values from various facies. Well 1 Facies

Depth (m)

Porosity (%)

Well 2

Permeability (mD)

Depth (m)

Porosity (%)

Well 3

Permeability (mD)

Depth (m)

Porosity (%)

Permeability (mD)

Sandy debrite

2129.68 35.00

2703

2224.76 30.50

7555

2166.60 39.0

10,018

2114.50 31.00

3072

2046.42 34.10

1788

2109.90 34.2

18,691

Sandy tidalite

1887.25 40.39

6930

2072.22 40.90

586

2114.37 38.7

5477

2047.03 34.00

525

2106.33 37.9

5977

Muddy tidalite

2074.11 39.30

3375

2150.20 29.2

Sandy injectitea

2108.76 35.50

11,760

2161.76 34.0

8681

1958.18 34.90

11,205

2126.47 34.7

11,212

a

2.76

Postdepositional facies. All measurements were made at 300 psi (Krishna-Godavari Basin). From Shanmugam, G., Shrivastava, S.K., Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736 756.

126

3. Gravity flows

FIGURE 3.38 Diagram illustrating the importance of distinguishing bottom layer based on fluid rheology and flow state in density-stratified gravity flows, which is based on a photograph of experimental density-stratified gravity flows showing the rheological difference between plastic debris flow (bottom layer) in massive sand and Newtonian turbidity current (top layer). Note that only Group 1 of researchers would recognize the importance of bottom layer with different rheology and flow state. Note that Postma et al. (1988) would classify both layers together as “high-density turbidity current” (see Fig. 3.39B). This Mobil-funded experimental flume study was carried out at St. Anthony Falls Laboratory (SAFL), University of Minnesota (1996 98) under the direction of Professor G. Parker to evaluate the fluid dynamical properties of sandy debris flows. Results were published in two major articles (Shanmugam, 2000a,b; Marr et al., 2001) (after Shanmugam, 2019c). Source: Springer. Open Access.

(1) body transformation, (2) gravity transformation, (3) surface transformation, and (4) elutriation transformation. Flow transformations cannot be established without knowing: (1) initial flow behavior, (2) transport mechanisms, and (3) final flow behavior. There are, however, no established criteria for recognizing initial flow behavior and transport mechanisms in the depositional record (Dott, 1963; Middleton and Hampton, 1973). In discussing the physics of debris flows, Iverson (1997) states, “When mass movement occurs, the sediment-water mixtures transform to a flowing, liquid-like state, but eventually they transform back to nearly rigid deposits.” Although such transformations can occur during transport, evidence for flow transformations cannot be inferred from the final deposit. We may never resolve this issue of flow transformation.

3.5 Liquefied/fluidized flows 3.5.1 Definition In contrast to the classification of Middleton and Hampton (1973), Lowe (1976a) made a clear distinction between liquefied and fluidized systems. In liquefied beds and flows, the Mass Transport, Gravity Flows, and Bottom Currents

3.5 Liquefied/fluidized flows

127

FIGURE 3.39

(A) Plot of sediment concentration for different flow types. Note that a typical turbidity current can exist only in sediment concentration less than 9% by volume (Bagnold, 1962). Note overlap in sediment concentration among low-density, turbidity currents, high-density turbidity currents, and hyperconcentrated flows. (B) Experimental stratified flows with a basal laminar-inertia flow and an upper (turbulent) turbidity current that have been termed as “high-density turbidity currents.” Note clasts near the top of sandy debris flows along the rheological interface. Compare with Fig. 3.38 and related text. Source: (A) Modified after Shanmugam, G., 1996a. High-density turbidity currents: are they sandy debris flows? J. Sediment. Res. 66, 2 10. Reproduced with permission from SEPM. (B) Postma, G., Nemec, W., Kleinspehn, K.L., 1988. Large floating clasts in turbidites: a mechanism for their emplacement. Sediment. Geol. 58, 47 61. Publication: Sedimentary Geology.

solids settle downward through the fluid, displacing it upward, whereas, in fluidized beds, the fluid moves upward through the solids, which are temporarily suspended without net downward movement.

3.5.2 Origin • • • • • •

Earthquake Sediment loading Volcanism Meteorite impacts Tsunamis Cyclones Mass Transport, Gravity Flows, and Bottom Currents

128

3. Gravity flows

FIGURE 3.40 Inferred flow velocity and sediment concentration in explaining the origin of mudstone clasts along rheological interface between underlying laminar-inertia flow and upper turbulent flow. Note that “highdensity turbidity currents” are considered as sandy debris flows (Shanmugam, 1996a). Source: From Postma, G., Nemec, W., Kleinspehn, K.L., 1988. Large floating clasts in turbidites: a mechanism for their emplacement. Sediment. Geol. 58, 47 61. Elsevier.

In understanding this type of phenomenon, one needs to discuss liquefaction. Allen (1984, Volume II, p. 343) provided an accurate account of soft-sediment deformation in terms of physics. Important points are: 1. Stratigraphical and sedimentological studies over many years have shown that soft sediments often become deformed nontectonically. The structures induced take myriad forms and are increasingly called “soft-sediment deformations.” 2. Soft-sediment deformation is associated in time with the earliest stages of sediment consolidation, when the deposit is weakest and pore fluid is being expelled most rapidly. This process is popularly known as “prelithification deformation.” Lowe (1975) classified such soft-sediment deformations as “water-escape structures.” 3. Liquefaction is significant in the production of many kinds of soft-sediment deformations. He and Qiao (2015, their Fig. 3.1) classified deformations of seismites, based on structural styles, preserved positions, activity times, formation mechanisms, and dynamics of soft-sediment deformation structures (SSDS) triggered by seismic activity, into five primary types: 1. liquefied deformation, 2. thixotropic deformation, 3. hydroplastic deformation,

Mass Transport, Gravity Flows, and Bottom Currents

3.5 Liquefied/fluidized flows

129 FIGURE 3.41 Types of low transformations. Source: Modified after Fisher RV, 1983. Flow transformations in sediment gravity flows. Geology 11, 273 274. GSA.

4. superimposed gravity driving deformation, and 5. brittle deformation. Further, based on the main genetic types, composition of sediments and deformation styles, the authors proposed 35 secondary types (e.g., liquefied breccia, liquefied droplet, homogenite, tepee structure, fault grading, shatter breccia, etc.).

Mass Transport, Gravity Flows, and Bottom Currents

130

3. Gravity flows

3.5.3 Identification • • • •

Fluid escape structures Dish and pillar structures (Fig. 3.42A) Dewatering pipes (Fig. 3.42B) Over 100 types of SSDS (Shanmugam, 2017a)

Su and Sun (2012) proposed that the following SSDS are common identification markers associated with earthquake-induced liquefaction: • • • •

diapirs, clastic dikes, convolute bedding, compressional deformation features (accordion folds, plate-spine breccias, mound-andsag structures), and • extensional plastic features (loop-bedding).

FIGURE 3.42 (A) Core photograph showing water-escape dish structures by liquidization in fine-grained, well-sorted sand. The arrow shows a concave-up (dish structure) color couplet with left-wing dipping at 45 degrees from the core horizontal due to deformation. Note cross-cutting relationship between two dish structures in which an earlier formed dish structure (1) has been terminated by a later one (2). Eocene, U.K. North Sea. (B) Core photograph showing pipes (water-escape structures). Paleocene, U.K. Atlantic Margin. Source: (A) From Shanmugam, G., 2006a. Deep-water processes and facies models: implications for sandstone petroleum reservoirs. Elsevier, Amsterdam, p. 476, with permission from Elsevier. (B) From Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 p., with permission from Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

3.6 Grain flows

131

Li et al. (2008) suggested the following criteria for recognizing features induced by seismicity: • • • • • • • • • • •

seismic micro-fractures microcorrugated laminations liquefied veins “vibrated liquefied layers” deformed cross-laminations convolute laminations load structures flame structures breccias slump structures seismo-disconformity

3.5.4 Case studies Seismogenic slump folds and their origin in the Dead Sea area were discussed by Alsop and Marco (2013) (see Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows).

3.5.5 Facies models There are no facies models.

3.5.6 Problem Fluidized flows are transitional and transient in nature (Lowe, 1982). They are also not important sediment-transport processes. For these reasons, I combined the two processes and call it “liquefied/fluidized flow.” In the rock record, it is a challenge to distinguish liquefaction features induced by earthquakes from those generated by rapid sedimentation. The other problem is that there are no objective criteria to recognize earthquakes as a unique triggering mechanism (among 20 others) of SSDS (Shanmugam, 2017a). Major problems in recognizing seismicity-induced SSDS are discussed by Shanmugam (2016a, 2017c).

3.6 Grain flows Bagnold (1954) first introduced the concept of dispersive pressure by colliding grains in high-concentration dispersions, which would later be termed “grain flows.”

3.6.1 Definition According to Lowe (1976b), the term grain flow is restricted to sediment-gravity flows in which a dispersion of cohesionless grains is maintained against gravity by grain

Mass Transport, Gravity Flows, and Bottom Currents

132

3. Gravity flows

dispersive pressure and in which the fluid interstitial to the grains is the same as the ambient fluid above the flow. Modified flows include those in which a dense interstitial fluid, current, or escaping pure fluid aids in maintaining the dispersion.

3.6.2 Origin • Arid climate, wind, Aeolian dunes (Fig. 3.43); • steep gradients associated with submarine canyon-heads where sand fall occurs (Fig. 3.44). Submarine sand falls are considered somewhat analogous to grain flows.

3.6.3 Identification • • • •

Massive sand layers thin layers (,5 cm) well sorted inverse grading

FIGURE 3.43 Grain flows. (A) Photograph showing grain avalanches (i.e., grain flows) on the “slip face” or lee side of an Aeolian dune. (B) Photograph showing grain avalanches (i.e., grain flows) on the “slip face” or lee side of an Aeolian dune. Saudi Arabia. Red scale 5 5 cm. Source: (A) Photo was taken at Kelso in the Mojave Desert, California by Mark. A. Wilson. Wikipedia. Public domain. (B) Photo by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

133

3.7 Turbidity currents

FIGURE 3.44 Underwater photograph showing a cascading sand fall at a depth of 40 m (130 ft) in gully leading down into San Lucas Canyon, Baja California. Such pure sand falls would develop massive sand intervals in the rock record (after Shanmugam, 2012a). Source: Photo by R.F. Dill. (after Shepard and Dill (1966), Rand McNally & Company). Elsevier.

3.6.4 Facies models There are no facies models because convincing examples are seldom recognized (Stauffer, 1967).

3.6.5 Problem Deposits of grain flows are volumetrically insignificant in submarine environments, but included here for completeness.

3.7 Turbidity currents 3.7.1 Definition A turbidity current is a sediment flow (Fig. 3.45) with Newtonian rheology (Fig. 3.3) and turbulent state (Fig. 3.46) in which sediment is supported by turbulence and from which deposition occurs through suspension settling (Dott, 1963; Sanders, 1965; Middleton and Hampton, 1973; Shanmugam, 1996a). They tend to spread out and develop fan geometry (Fig. 3.47; compare with tongue geometry of debrites in Fig. 2.12). Turbidity currents exhibit unsteady and nonuniform flow behavior (Fig. 3.48), and they are surge-type waning flows. As they flow downslope, turbidity currents (Fig. 3.48) invariably entrain ambient fluid (seawater) in their frontal head portion due to turbulent mixing (Allen, 1985). With increasing fluid content, plastic debris flows may tend to become Newtonian

Mass Transport, Gravity Flows, and Bottom Currents

134

3. Gravity flows

FIGURE 3.45 Photograph of side view of experimental turbidity current with well-developed head. Source: Photo from experiments conducted by M.L. Natland and Courtesy of G.C. Brown.

FIGURE 3.46

Photograph of front view of experimental turbidity current showing flow turbulence. Source: Photo from experiments conducted by M.L. Natland and Courtesy of G.C. Brown. Published in Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 p.

FIGURE 3.47 Photograph of map view of a spreading turbidity current developing fan geometry. This flow is purely mud in texture. Source: Published in Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9.Elsevier, Amsterdam, 524 p.

Mass Transport, Gravity Flows, and Bottom Currents

3.7 Turbidity currents

135

FIGURE 3.48

Schematic illustration showing the leading head portion of an unsteady, nonuniform, and turbulent turbidity current. Due to turbulent mixing, turbidity currents invariably entrain ambient fluid (seawater) at their head regions. Source: Modified from Allen, J.R.L., 1985. Loose-boundary hydraulics and fluid mechanics: selected advances since 1961. In: Brenchley, P.J., Williams, P.J. (Eds.), Sedimentology: Recent Developments and Applied Aspects, published for the Geological Society by Blackwell Scientific Publications, Oxford, pp. 7 28. Published in Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 p.

turbidity currents. However, not all turbidity currents evolve from debris flows. Some turbidity currents may evolve directly from sediment failures. Although turbidity currents may constitute a distal end-member in basinal areas, they can occur in any part of the system (i.e., shelf edge, slope, and basin).

3.7.2 Origin The origins of four sediment-gravity flows are closely related to sediment failures and slope instability. There are 21 triggering mechanisms in causing sediment failures (Shanmugam, 2015a), but only important mechanisms are listed in each case. • • • •

earthquake oversupply of sediment volcanism meteorite impact

Mass Transport, Gravity Flows, and Bottom Currents

136

3. Gravity flows

• tsunamis • cyclones

3.7.3 Identification Turbidity currents cannot transport gravel and coarse-grained sand in suspension because they do not possess the strength like debris flows. General characteristics of turbidites are: • fine-grained sand to mud; • flute casts as sole marks (Fig. 3.49). However, bottom currents could also generate such sole marks (Klein, 1966); • normal grading (core and outcrop) (Fig. 3.50); • sharp or erosional basal contact (core and outcrop) (Fig. 3.50); • gradational upper contact (core and outcrop) (Fig. 3.50); • thin layers, commonly centimeters in thickness (core and outcrop) (Fig. 3.51); • sheet-like geometry in basinal settings (outcrop) (Fig. 3.51); • lenticular geometry may develop in channel-fill settings.

3.7.4 Case studies The classic case study of turbidites is based on the examination of the Annot Sandstone in SE France (Bouma, 1962).

3.7.5 Facies models In his seminal publication on turbidite fans, Bouma (1962, p. 98) used the term “cone” for describing submarine fans. Bouma (1962) proposed the most convincing link between the turbidite facies model with five divisions, namely Ta, Tb, Tc, Td, and Te (Fig. 3.52A), FIGURE 3.49 Flute casts, common sole marks, often used as evidence for erosion by turbidity currents. Flow direction is from left to right. However, alternative interpretations by bottom currents are suggested (Klein, 1966). Jackfork Group, Pennsylvanian, Oklahoma. Source: Photo by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

137

3.7 Turbidity currents

FIGURE 3.50 Two examples of classic turbidites. (A) Core photograph showing turbidite units with sharp basal contact, normal grading, and gradational upper contact. Arrow marks a normally graded unit with finegrained sand at bottom (light gray) grading into clay (dark gray) near top. Note that these thin-bedded units cannot be resolved on seismic data. Zafiro Field, Pliocene, Equatorial Guinea. (B) Core photograph showing turbidite units with sharp basal contact, normal grading, and gradational upper contact (yellow arrow). Cretaceous, West Africa. Source: From Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476. Elsevier. FIGURE 3.51 Outcrop photograph showing tilted thin-bedded turbidite sandstone beds with sheet-like geometry, Lower Eocene, Zumaya, northern Spain. Source: Reproduced from Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476. Elsevier.

and their areal distribution on a submarine fan (Fig. 3.52B). In illustrating the lateral distribution of the five divisions (i.e., the Bouma Sequence) on a submarine fan (Fig. 3.52B), Bouma (1962, p. 98) stated that “It can be understood that at the outset that the deposit of

Mass Transport, Gravity Flows, and Bottom Currents

138

3. Gravity flows

FIGURE 3.52 The first turbidite-fan link proposed by Bouma (1962). (A) The turbidite facies model with five internal divisions (Ta, Tb, Tc, Td, and Te). This vertical facies model is commonly known as “the Bouma Sequence.” (B) Areal distribution of turbidite facies in a submarine fan (after Bouma, 1962; after Shanmugam, 2016a). Source: Elsevier.

a large turbidity current contains all the components belonging to the complete sequence type Ta e. The coarseness of the material and the velocity decrease in the current direction. Also, the thickness of the graded interval decreases in the downstream direction. . .”

3.7.6 Problem 1. There are no theoretical (Sanders, 1965; Van der Lingen, 1969; Hsu¨, 2004; Shanmugam, 1997a), experimental (Leclair and Arnott, 2005), or observational (Shanmugam, 2002a, 2006a) basis for validating the complete Bouma Sequence. Leclair and Arnott (2005, p. 4) state that “. . .the debate on the upward change from massive (Ta) to parallel laminated (Tb) sand in a Bouma-type turbidite remains unresolved.” 2. The complete “Bouma” Sequence with all five divisions (Fig. 3.53A) has never been reproduced in experimental studies and has never been documented in sediments on the modern ocean floor. Importantly, these divisions can be interpreted as products of alternative processes (Fig. 3.53A). 3. Overzealous turbidite research has resulted in too many facies models in explaining downslope changes in turbidite deposition with 18 divisions (Fig. 3.53B). 4. The ideal turbidite bed with 16 divisions (Fig. 3.54A) is untenable from a fluid dynamic point of view. No one has ever documented the vertical facies model showing the R1, R2, R3, S1, S2, and S3 divisions of the Lowe (1982) sequence and the

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.53 (A) The turbidite facies model (i.e., the Bouma Sequence) showing Ta, Tb, Tc, Td, and Te divisions. Conventional interpretation is that the entire sequence is a product of a turbidity current (Bouma, 1962; Walker, 1965; Middleton and Hampton, 1973). According to Lowe (1982), the Ta division is a product of a highdensity turbidity current and Tb, Tc, and Td divisions are deposits of low-density turbidity currents. In this article, the Ta division is considered to be a product of a turbidity current only if it is normally graded, otherwise it is a product of a sandy debris flow; the Tb, Tc, and Td divisions are considered to be deposits of bottom-current reworking (after Shanmugam, 1997a). (B) Schematic diagram showing downslope changes in turbidite divisions from coarse-grained turbidites (Lowe, 1982), through classic turbidites (Bouma, 1962), to fine-grained turbidites (Stow and Shanmugam, 1980). If existing turbidite facies models were realistic, then an ideal turbidite bed should develop 16 divisions. However, no one has ever documented such a turbidite bed with 16 divisions in the field or in flume experiments (after Shanmugam, 2000a). Source: (A and B) Reproduced with permission from Elsevier.

140

5. 6.

7. 8. 9.

10.

3. Gravity flows

Ta, Tb, Tc, Td, and Te divisions of the Bouma (1962) sequence in ascending order in modern deep-sea sediments. No one has ever replicated in flume experiments of turbulent turbidity currents that could carry coarse sand and gravel in suspension in laboratory flume experiments that could produce the R1, R2, R3, S1, S2, and S3 divisions in ascending order. The ultimate objective of facies models is to interpret ancient strata (i.e., the unknown). However, the turbidite facies models, developed exclusively from the ancient strata without validation from the modern environment (i.e., the known), promote circular reasoning (Fig. 3.54B). Misapplication of the term “turbidite” for deposits of all four types of sedimentgravity flows, including debris flows (Fig. 3.55) by Mutti et al. (1999) and by Zavala (2019). There is no agreement on the defining criteria of turbidity currents (Fig. 3.56). There is no agreement on the density value that separates “low-density” from “highdensity” turbidity currents (Fig. 3.39A). Turbidity currents are inherently low in sediment concentration or low in flow density (Fig. 3.39A). According to Bagnold (1962), typical turbidity currents can function as truly turbulent suspensions only when their sediment concentration by volume is below 9% or C , 9% (Fig. 3.39A). Therefore, high-density turbidity currents (Fig. 3.39B) cannot exist in nature. There is no agreement on the importance of normal grading in turbidites (Fig. 3.57).

Because of the above problems, turbidite research is still in a state of flux.

3.8 Hyperpycnal flows: a prelude Because of the importance of hyperpycnal flows and their deposits (i.e., hyperpycnite) to the scope of this book, Chapter 6, Hyperpycnal Flows, is devoted to this process.

3.8.1 Definition In advocating a rational theory for delta formation, Bates (1953) suggested three types (Fig. 3.58): (1) hypopycnal plume for floating river water that has lower density than basin water (Fig. 3.58A), (2) homopycnal plume for mixing river water that has equal density as basin water (Fig. 3.58B), and (3) hyperpycnal plume for sinking river water that has higher density than basin water (Fig. 3.58C). It is worth noting that Middleton and Hampton (1973) did not consider hyperpycnal flows in their original classification, although hyperpycnal flows are indeed driven by sediment gravity (Chapter 6: Hyperpycnal Flows). For the following reasons, hyperpycnal flows are considered as sediment-gravity flows in this chapter as well as in Chapter 6, Hyperpycnal Flows. 1. River-mouth hyperpycnal flows are caused by higher density of the entering river flows in comparison to density of seawater (Bates, 1953). Sediment particles in the flow are the cause of higher flow density.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.54

(A) Schematic diagram showing an ideal turbidite bed with nine turbidite divisions by combining the five divisions of the “Bouma Sequence” (Bouma, 1962) and the five divisions of the “Lowe Sequence” of high-density turbidites (Lowe, 1982). Note R1 is not shown. According to Lowe (1982), S3 1/4 Ta. On the righthand column, I have included my interpretation of these divisions. (B) Summary diagram revealing the total lack of empirical data for high-density turbidity currents (see Shanmugam, 2012a for details) Source: From Shanmugam, G., 2016a. Submarine fans: a critical retrospective (1950 2015). J. Palaeogeogr. 5 (2), 110 184. Elsevier.

142

3. Gravity flows

Sediment-gravity flows (Middleton and Hampton, 1973)

Turbidity currents

Fluidized flows

Debris flows

Turbidites (Sanders, 1965)

Grain flows

FIGURE 3.55 Original classification of sediment-gravity flows by Middleton and Hampton (1973). Confusing application of the term “turbidites” to deposits of all four types by Mutti et al. (1999) without regard for fluid mechanics, which Zavala (2019) has adopted in his comment. I have adopted Sanders’ (1965) classification in which only deposits of turbidity currents are considered as turbidites. Source: From Shanmugam, G., 2002a. Ten turbidite myths. Earth-Sci. Rev. 58, 311 341. Elsevier.

Turbidites (Mutti et al., 1999)

FIGURE 3.56 Four different definitions of turbidity currents based on (1) rheology of fluids, (2) sedimentsupport mechanism, (3) driving force, and (4) velocity-time factors (of these four types, only the Newtonian and turbulent types are useful in interpreting the behavior of flow because evidence for fluid rheology and sedimentsupport mechanism is preserved in the deposit. However, evidence for driving force and waxing velocity is not always preserved). Source: From Shanmugam, G., 2000a. 50 years of the turbidite paradigm (1950s 1990s): deep-water processes and facies models a critical perspective. Mar. Pet. Geol. 17, 285 342. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 3.57 Three publications showing how opinions on nature of grading in turbidites have changed through time. Top: Bouma (1962) suggested normal grading for turbidites. Middle: Harms and Fahnestock (1965) proposed normal grading and massive (i.e., no grading) types for turbidites. Bottom: Kneller (1995) advocated normal grading, massive (i.e., no grading), and inverse grading for turbidites (after Shanmugam, 2000a). Source: Reproduced with permission from Elsevier.

FIGURE 3.58 Concepts and examples of density plumes. (A C) Schematic diagrams showing three types of density variations in riverine plumes in deltaic environments based on concepts of Bates (1953). (A) Hypopycnal plume in which density of river water is less than density of basin water; (B) homopycnal plume in which density of river water is equal to density of basin water; (C) hyperpycnal plume in which density of river water is greater than density of basin water. Source: From Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 p., with permission from Elsevier.

144

3. Gravity flows

FIGURE 3.59 (A) Conceptual diagram of a continental margin showing relative positions of plunge point (red filled circle) at river mouth and submarine fan at base-of-slope. (B) Note that fluid-gravity flows can transform into sediment-gravity flows at plunge points and deposit sediments as hyperpycnites near the shoreline in shallow-water environments. Compare with Fig. 6.1.

2. The other option for higher density of entering flow is by changes in salinity and/or temperature, such as thermohaline ocean-bottom contour currents (Gordon, 2019), which is unlikely to occur at river mouths. 3. By applying the concept of Middleton and Hampton (1973), where the river waters enter the ocean, density of ambient fluid changes from air (1.225 kg m23) to seawater (1030 kg m23) (Beicher, 2000). In other words, at river-mouth plunge points, fluid-gravity flows could transform into sediment-gravity flows (Fig. 3.59). However, fluid mechanics of hyperpycnal flows is mired in controversies (Shanmugam, 2018b, 2019c). Importantly, this flow transformation does not imply that all river flows routinely become turbidity currents (see Chapter 6: Hyperpycnal Flows).

Mass Transport, Gravity Flows, and Bottom Currents

3.9 Thermohaline contour currents: a prelude

145

3.9 Thermohaline contour currents: a prelude 3.9.1 Definition Thermohaline-induced bottom currents that follow regional bathymetric contours in deep-water (200 m bathymetry) environments. They are called thermohaline contour currents (THCC). A detailed account of this process is given in Chapter 8, Bottom Currents.

3.9.2 Downslope initiation The origin of thermohaline water masses is best studied using the Antarctic bottom water (AABW) (Gordon, 2001, 2019; Gordon, 2013; Purkey et al., 2018, among others). The AABW is initiated as downslope gravity flows on the continental slope (Fig. 3.60).

FIGURE 3.60 Schematic of the origin of the Antarctic bottom water as downslope gravity flows on the continental slope. Cold shelf water forms through brine rejection in coastal polynyas during ice formation and export. The shelf water flows down the slope in dense plumes, mixing with ambient Warm Deep Water (also referred to as modified Circumpolar Deep Water). Potential temperatures pertinent to Weddell Sea Bottom Water formation are also shown (after Gordon, 2019). Source: Modified after Gordon, A.L., 2001. Bottom water formation. In: Encyclopedia of Ocean Sciences, second ed. 334 340. Elsevier; and Purkey, S.G., Smethie, W.M., Gebbie, G., Gordon, A.L., Sonnerup, R.E., Warner, M.J., Bullister, J.L., 2018. A synoptic view of the ventilation and circulation of Antarctic bottom water from chlorofluorocarbons, Ann. Rev. Mar. Sci. 10, 503 527, with additional labels by G. Shanmugam. Elsevier. Mass Transport, Gravity Flows, and Bottom Currents

146

3. Gravity flows

The AABW has a density of 0.03 g cm23 with temperatures ranging from 20.8 C to 2 C (35 F), salinities from 34.6 to 34.7 psu. Being the densest water mass of the oceans (Purkey et al., 2018), AABW is found to occupy the depth range below 4000 m. The AABW is formed in the Weddell and Ross Seas, off the Ade´lie Coast and by Cape Darnley from surface water cooling in polynyas and below the ice shelf. A unique feature of AABW is the cold surface wind blowing off the Antarctic continent (Fig. 3.60). The surface wind creates the polynyas (i.e., an area of open water surrounded by sea ice), which opens up the water surface to more wind. This Antarctic wind is stronger during the winter months and thus, the AABW formation is more pronounced during the Antarctic winter season.

3.10 Synopsis Six basic types of gravity flows (i.e., density flows) have been identified in subaerial and subaqueous environments. They are: (1) hyperpycnal flows (subaqueous), (2) turbidity currents (subaqueous), (3) debris flows (subaerial and subaqueous), (4) liquefied/fluidized flows (subaerial and subaqueous), (5) grain flows (subaerial and subaqueous), and (6) THCC (subaqueous). The first five types are flows in which the density is caused by sediment in the flow, whereas in the sixth type (THCC: Thermohaline Contour Currents), the density is caused by variations in temperature and salinity. Although all six types originate initially as downslope gravity flows, only the first five types are truly downslope processes, whereas the sixth type eventually becomes an alongslope process. • Hyperpycnal flows are triggered by river floods in which density of incoming river water is greater than the basin water. These flows are confined to proximity of the shoreline. They transport mud, and they do not transport sand into the deep sea. There are no sedimentological criteria yet to identify hyperpycnites in the ancient sedimentary record. • A turbidity current is a sediment-gravity flow with Newtonian rheology and turbulent state in which sediment is supported by flow turbulence and from which deposition occurs through suspension settling. Typical turbidity currents can function as truly turbulent suspensions only when their sediment concentration by volume is below 9% or C , 9%. This requirement firmly excludes the existence of “high-density turbidity currents.” Turbidites are recognized by their distinct normal grading in deep-water deposits. Turbidite facies models are obsolete. • A debris flow (C: 25% 95%) is a sediment-gravity flow with plastic rheology and laminar state from which deposition occurs through freezing en masse. The terms debris flow and mass flow are used interchangeably. General characteristics of muddy and sandy debrites are floating clasts, planar clast fabric, inverse grading, etc. Most sandy deep-water deposits are sandy debrites and they comprise important petroleum reservoirs worldwide. • A liquefied/fluidized low ( . 25%) is a sediment-gravity flow in which sediment is supported by upward-moving intergranular fluid. They are commonly triggered by seismicity. Water-escape structures, dish and pillar structures, and SSDS are common.

Mass Transport, Gravity Flows, and Bottom Currents

3.10 Synopsis

147

• A grain flow (C: 50% 100%) is a sediment-gravity flow in which grains are supported by dispersive pressure caused by grain collision. These flows are common on the slip face of Aeolian dunes. Massive sand and inverse grading are potential identification markers. • THCC originate in the Antarctic region due to shelf freezing and the related increase in the density of cold saline (i.e., thermohaline) water. Although they begin their journey as downslope gravity flows, they eventually flow alongslope as contour currents. Contourites are characterized by traction structures, such as cross-laminae, and current ripples. Facies models of hyperpycnites, turbidites, and contourites are obsolete. Of the six types of gravity flows, hyperpycnal flows and their deposits are the least understood. • Gravity flows constitute the single most important sediment-transport mechanism on land, shelf, slope, and basin environments. They play important roles not only in downslope, but also in alongslope directions (Fig. 3.61). In terms of transporting large

FIGURE 3.61 Summary diagram of six types of density flows and their characteristics in downslope and alongslope settings. Red dot suggests common position of flow.

Mass Transport, Gravity Flows, and Bottom Currents

148

3. Gravity flows

volumes of coarse-grained sediment into the deep sea, debris flows and related mass movements are the most important of all other processes. Also, identification markers of debrites discussed in this review are of value for recognizing them in the ancient sedimentary record because sandy debrites are important petroleum reservoirs worldwide.

Mass Transport, Gravity Flows, and Bottom Currents

C H A P T E R

4 A paradigm shift O U T L I N E 4.1 Introduction

4.6 The Annot Sandstone (Eocene Oligocene), Peira-Cava Area, Maritime Alps, SE France

149

4.2 Amazon Fan, Equatorial Atlantic 150 4.2.1 Sinuous channels 150 4.2.2 High-amplitude reflection packet (HARP) units 153 4.2.3 Lower-fan lobes 153 4.3 Mississippi Fan, Gulf of Mexico 4.3.1 Channelized lobes

155 155

4.4 Monterey Fan, North Pacific 4.4.1 Monterey Canyon 4.4.2 Depositional lobe

156 156 159

4.5 Krishna-Godavari (KG) Basin, Bay of Bengal, India

160

161

4.7 The Jackfork Group, Pennsylvanian, Ouachita Mountains 169 4.8 Basin-floor fan model, Tertiary, North Sea

171

4.9 Mass-flow lobes, Ulleung Basin, East Sea, Korea

176

4.10 Upper Triassic Yanchang Formation, Ordos Basin, central China 176 4.11 Supercritical and subcritical fans

177

4.12 Synopsis

179

4.1 Introduction The turbidite exuberance that dominated the deep-water domain for nearly a century is waning as a consequence of emerging emphasis on sandy mass-transport deposits (SMTDs) and bottom-current reworked sands (BCRS). In fact, numerous published classic “turbidites” have been reinterpreted as SMTDS and BCRS in the ancient stratigraphic record (Shanmugam, 2012a). The turbidite paradigm is fundamentally defective because it is built on facies models, such as the “Bouma sequence” (Fig. 3.52) for classic turbidites deposited

Mass Transport, Gravity Flows, and Bottom Currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00004-7

149

© 2021 Elsevier Inc. All rights reserved.

150

4. A paradigm shift

by low-density turbidity currents (Bouma, 1962) and the “Lowe sequence” (Fig. 3.53B) for coarse-grained turbidites deposited by high-density turbidity currents (HDTCs; Lowe, 1982). These vertical facies models were derived solely from the study of ancient rock record using outcrops, without any empirical data of “sandy turbidity currents” from modern oceans. The primary attraction to the vertical facies model of HDTCs and their deposits, composed of R1, R2, R3, S1, S2, and S3 divisions in ascending order (Lowe, 1982), is that it allows one to interpret ancient deep-water coarse sandstone and conglomerate deposits as turbidites (Mulder, 2011; Mutti, 1992). But turbidite facies models are fatally flawed because of the empirical data from the following case studies show a clear dominance of MTD and BCRS. Core and outcrop examination of modern and ancient deep-water systems shows a dominance of MTD (Fig. 4.1). Selected examples are discussed here.

4.2 Amazon Fan, Equatorial Atlantic 4.2.1 Sinuous channels Submarine channels on mature passive-margin settings tend to be relatively long, bifurcating, of low gradient, and largely sinuous. The relatively fine-grained (mud-rich) character of the transported sediment associated with channels on mature passive-margin fans, such as the modern Amazon Fan (Damuth et al., 1988), gives rise to excellent bank stability and favors development of a single, largely sinuous, commonly meandering channel. According to Pirmez and Flood (1995), sinuosity of the Amazon Fan channel is smaller than 1.5 over most of the upper half of the channel, but increases downdip to about 2.3 becoming a meandering channel (Fig. 4.2A). In cross-sectional view, these channel-levee complexes generate gull-wing geometry because of vertical levee build up. The Amazon Fan with 4 km of sediment core, taken from 17 drill sites during Ocean Drilling Program Leg 155, has provided a great opportunity to understand the depositional origin of various fan elements (Normark, Damuth et al., 1997). The purpose of studying modern deep-water systems with abundant cores, such as the Amazon Fan, is to: (1) describe cores from known elements [i.e., channel, levee, high-amplitude reflection packet (HARP), etc.], (2) interpret processes in each element, and, (3) most importantly, quantify the relative importance of depositional facies in each element. Such a quantitative data would help to establish a genetic link between a given element (e.g., channel-fill) and a process (e.g., turbidity currents). Then, sedimentologists could use this element-process linkage, derived from modern systems, for reconstructing ancient deep-water environments. Unfortunately, Normark, Damuth et al. (1997) did not quantify depositional facies in various elements of the Amazon Fan. To establish a genetic link between processes and elements in the Amazon Fan, I have examined published graphic sedimentological columns (Shipboard Scientific Party, 1995a, their Fig. 4), grain-size variations, core description, and photographs (Shipboard Scientific Party, 1995a). Using this information, I have interpreted depositional processes and quantified the amount of various lithofacies from the Site 934. This quantification suggests that the principal sand unit is composed of fine-to-medium sand (i.e., Unit IV, Site 934, Shipboard Scientific Party, 1995a) in the Brown Channel. The cored interval is composed

Mass Transport, Gravity Flows, and Bottom Currents

4.2 Amazon Fan, Equatorial Atlantic

151

FIGURE 4.1 Map showing locations of modern and ancient deep-water systems, commonly known as submarine fans, discussed during the COMFAN Meeting in 1982 (Bouma et al., 1985a). Note that plotting of each example at a precise location is not practical because the length varies from 16 km for the Crati Fan to 3000 km for the Bengal Fan. This size variability is also the reason why that this map does not have a scale. Data from Barnes and Normark (1985). After Shanmugam (2016a). JOP, Elsevier. Creative Common.

of sediments deposited by sandy debris flows (30%), muddy slumps (20%), and mixed sandy debris flows and bottom currents (11%). The remaining interval is composed of pelagic/hemipelagic mud (34%) and turbidites (5%) (Shanmugam, 2006a, his Table 6.6).

Mass Transport, Gravity Flows, and Bottom Currents

152

4. A paradigm shift

FIGURE 4.2 A. Ocean Drilling Program (ODP) Leg 155 Sites, Amazon Fan. B. Sedimentological log of channel-fill element at Site 934. Note rafted clasts and deformed clasts in sandy matrix, suggesting deposition from sandy debris flow. Location map after Normark, Damuth et al. (1997) and Site Log after Shipboard Scientific Party (1995a). C. Core photograph showing floating pebble in fine sandy matrix suggesting deposition from sandy debris flow. ODP Leg 155, Site 934B, channel fill, Amazon Fan. Photo courtesy of J. E. Damuth. D. Core photograph showing sharp-based and normally graded bed (arrow) with medium sand at the base (39.5 cm) and silt at top from channel-fill deposits. Normal grading is interpreted as evidence for deposition from turbidity currents. ODP Leg 155, Site 934A, 10H 6, 26 44 cm, Amazon Fan. (After Normark and Damuth et al., (1997). Photo courtesy of Ocean Drilling Program (ODP), College Station, Texas. After Shanmugam (2016a). JOP, Elsevier. Creative Common.

My interpretations of Unit IV at Site 934 as deposits of sandy debris flows and Unit III as deposits of muddy slumps are consistent with the interpretation reached by the Shipboard Scientific Party (1995a). Also, Normark, Damuth et al. (1997, p. 632) described that “The most prevalent facies is thick-bedded, disorganized structureless to chaotic sand.” Their description suggests a dominance of deposition from debris flows and slumps in the Amazon channel-fill elements. General characteristics of deposits of sinuous channels of the Amazon Fan are: • • • • • • • • • •

medium- to fine-grained sand, interbedded mudstone, erosional base, rip-up clasts, contorted mud layers (muddy slump) (Fig. 4.2B), thick units of sand with rafted and deformed clasts (sandy debris flow) (Fig. 4.2B), floating pebbles in sandy matrix (sandy debris flow) (Fig. 4.2C), thin units of normally graded sand and silt layers (turbidity currents) (Fig. 4.2D), several meters thick to thin units, and lenticular sand-body geometry.

Graphic sedimentological columns for cores from the Leg 155 Site 944 from a middle fan levee element of the Amazon Fan show a striking muddy facies in which contorted

Mass Transport, Gravity Flows, and Bottom Currents

4.2 Amazon Fan, Equatorial Atlantic

153

layers and mud clasts are common (Shipboard Scientific Party, 1995b, their Fig. 2). This muddy facies is interpreted to be deposits of muddy slumps and debris flows in a levee element. Core evidence is compelling that the Amazon channels and levees were deposited by debris flows and slumps. This core-based interpretation, however, is not embraced by all researchers. Pirmez and Imran (2003), for example, concluded that the Amazon channels were originated by turbidity currents based mainly on seismic information.

4.2.2 High-amplitude reflection packet (HARP) units In the Amazon Fan, channel bifurcation (Fig. 4.3A) through avulsion is thought to lead initially to deposition of nonchannelized sandy flows in the interchannel area (Flood et al., 1991). Subsequent progradation of channels and levees over these sandy deposits has produced a sheet-like geometry at the base of the new channel-levee system. This sheet-like geometry returns HARPs on seismic data (Fig. 4.3B). A HARP unit at Site 931 was cored beneath Channel 5 (Fig. 4.3B). Core shows floating mud clasts in muddy matrix at Site 931B (Fig. 4.3C). These clasts are evidence for deposition from muddy debris flow. Normark, Damuth et al. (1997) have summarized facies distribution in HARP units. In a sequence-stratigraphic model, these sheet-like HARP units overlain by a channellevee system (gull-wing) are identical in appearance to a basin-floor fan overlain by a slope fan (Vail et al., 1991). However, there is a major difference between a basin-floor fan and a HARP unit. A basin-floor fan is formed by progradation primarily during lowstands of sea level (allocyclic process), whereas HARPs are formed by channel bifurcation (autocyclic process). More importantly, a basin-floor fan and a slope fan are not contemporaneous features (Vail et al., 1991), whereas a HARP unit and its overlying channel-levee system are contemporaneous elements. Thus, caution must be exercised in interpreting seismic geometries in terms of depositional elements. HARP units are also considered to be analogous to depositional lobes in the conventional submarine fan model (Normark, Damuth et al., 1997, their Fig. 4B, p. 617). They described deposits of depositional lobes as deposits with irregularly shaped mud clasts in poorly sorted, fine-to-coarse sand (their Facies 2). This lobe facies is several meters in thickness. I have interpreted this lobe facies as debrites. A typical depositional lobe, however, should be composed of turbidites (Mutti, 1977). Normark, Damuth et al. (1997, p. 613) wrote “it is difficult to explain why thick intervals with numerous mud clasts of variable size are randomly scattered into the sand matrix and are not also graded.” The reason is that these sandy intervals with mud clasts represent debrites, not turbidites.

4.2.3 Lower-fan lobes Site 946 is located on a lower-fan lobe (Flood, Piper, Klaus et al., 1995, their Plate 1, inside pocket of back cover). Cores taken from the Leg 155 Site 946, located in the lowerfan setting (Fig. 4.4A), show floating siltstone pebbles in a sandy matrix (Fig. 4.4B). These pebbles suggest deposition from sandy debris flows. At Site 946, normally graded

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 4.3 A. Conceptual model showing that channel bifurcation through avulsion on a deep-sea fan can result in unchannelized sandy flows (top diagram) by breaching their confining levee through a crevasse and spreading out initially as unchannelized flows into a lower interchannel areas. New channel reestablishment over these sandy deposits (bottom diagram) can result in sheet-like geometry (Flood, Piper et al., 1995) that return high-amplitude reflections (HARPs) on seismic data (Flood et al., 1991). Sheet-like HARPs overlain by channel-levee complex (gull-wing geometry) are identical to basin-floor fan overlain by slope fan in a sequence-stratigraphic framework (see Fig. 4.21). However, there is a major difference between a basin-floor fan and HARP. For example, a basin-floor fan is formed by progradation during lowstands of sea level (allocyclic process), whereas HARPs are formed by channel bifurcation (autocyclic process). Therefore, caution must be exercised in interpreting seismic geometries in terms of processes. B. Seismic profile showing HARP units (horizontal dashed lines) and overlying Channel 5 with levee units. Note position of Site 931B, Amazon Fan. Modified after Pirmez et al. (1997). C. Core photograph showing floating mud clasts in silty matrix. suggesting deposition from muddy debris flow. Site 931B, HARP unit, Leg 155, Site 931, Amazon Fan. See also Shipboard Scientific Party (1995c, their Figure 7B). Photo courtesy of J. E. Damuth. After Shanmugam (2016a). JOP, Elsevier. Creative Common.

4.3 Mississippi Fan, Gulf of Mexico

155

FIGURE 4.4 A. Left: Ocean Drilling Program (ODP) Leg 155 Sites, Amazon Fan. Right: Sedimentological log of lower-fan lobe and levee facies at Site 946. Note rafted clasts in sandy matrix, suggesting deposition from sandy debris flow. Location map is after Normark, Damuth et al. (1997) and Site Log is after Shipboard Scientific Party (1995d). B. Core photograph showing rafted siltstone pebbles in sandy (fine- to medium grained) matrix, suggesting deposition from sandy debris flow in lower-fan lobe environments, Site 946A, Amazon Fan. Photo courtesy of J. E. Damuth. After Shanmugam (2016a). JOP, Elsevier. Creative Common.

turbidite sand layers have been recognized (Shipboard Scientific Party, 1995d, their Fig. 3), but they are rare in comparison to deposits of sandy slumps and debris flows. Legou et al. (2008) reported channel-mouth lobes, based on bathymetric, acoustic imagery and 3.5 kHz subbottom profiler data. Although these authors assumed the origin of lobes by turbidity currents, they did not provide any empirical data using sediment cores on the depositional origin of these lobes by turbidity currents.

4.3 Mississippi Fan, Gulf of Mexico 4.3.1 Channelized lobes The Mississippi Fan in the Gulf of Mexico was the first modern deep-sea fan from which long Deep Sea Drilling Project cores were retrieved. These cores were taken during the Leg 96 (Bouma et al., 1985b; Bouma, Coleman, et al., 1985). The cores from the Leg 96 suggest a variety of depositional facies. Interpreted depositional processes of these facies are slumps, debris flows, turbidity currents, and bottom currents (Stow et al., 1985; Shanmugam et al., 1988b). Conventionally, sheet-like geometries are associated with turbidites deposited at the terminus of a submarine fan. These sheet sands are also known as depositional lobes (Mutti,

Mass Transport, Gravity Flows, and Bottom Currents

156

4. A paradigm shift

1977). The outer fan areas of the Mississippi Fan were used as the modern analog for turbidite fans with sheet-like geometries (Shanmugam et al., 1988b). Such an analogy was based strictly on the parallel and continuous reflection patterns observed on seismic profiles. However, subsequent SeaMARC 1A sidescan-sonar data (Twichell et al., 1992), as well as piston and gravity cores (Nelson et al., 1992; Schwab et al., 1996) taken from channels in the outer Mississippi Fan, have revealed some important information: 1. The terminus of the Mississippi Fan is not sheet-like as previously thought. 2. Piston and gravity cores taken from the terminus of the Mississippi Fan shows channels in the terminus of the Mississippi Fan (Fig. 4.5A) are filled with debrites for the most part (Fig. 4.5B). 3. Contrary to popular belief, the terminus of the Mississippi Fan is channelized and dendritic in nature (Twichell et al., 1995). SeaMARC 1A sidescan-sonar image mosaic of the distal Mississippi Fan shows dendritic pattern with abrupt edges (Fig. 4.5C). Twichell et al. (1995, their Fig. 41.2, p. 283) explained this pattern as “The association of these high-backscatter deposits with channels and their abrupt edges suggest channelized transport of the sediments that compose these deposits and a sudden freezing of the flows at the site of deposition.” Sudden freezing of flows is characteristic of plastic debris flows. Core 44 taken from these dendritic distal edges of the Mississippi Fan, composed of chaotic silt beds with floating clay clasts, has been interpreted to be composed of 100% debrites (Fig. 4.5B). These dendritic features are called “channelized lobes” in this article for distinguishing them from nonchannelized depositional lobes. These “channelized lobes” do not imply slope/upper fan environments as advocated by Nelson et al. (1985). The application of the term “depositional lobe,” meant for turbidite-dominated fans (Mutti, 1977), to debrite-dominated terminus of the Mississippi Fan (Twichell et al., 1992; Nelson et al., 1992; Schwab et al., 1996) is inappropriate. Such a misapplication perpetuates the notion that turbidites are more common in modern fans than they actually are.

4.4 Monterey Fan, North Pacific 4.4.1 Monterey Canyon The Monterey Canyon in offshore northern California (Fig. 4.6) provides an opportunity to understand the sand distribution in sinuous submarine canyons (Paull et al., 2005). The tight meander of the Monterey Canyon was explained by structural (fault) patterns (Martin and Emery, 1967). Paull et al. (2005) have documented the occurrence of clean massive sand on the canyon floor of the Monterey Canyon. They have described a total of 92 cores from the Monterey Canyon axis. Core lengths varied between 19 and 252 cm, with an average length of 134 cm. In 21 cores, gravel content varied between 30% and 100%. In 35 cores, sand and silt content varied between 50% and 100%. Floating rounded cobble grains are up to 7.5 cm in length. Floating clay clasts commonly occur within the coarse sand and gravel lithofacies. These floating cobbles and clasts in sandy matrix (Fig. 4.7) are considered evidence for deposition from sandy debris flows (Shanmugam, 2012a). Individual sand layers are greater than

Mass Transport, Gravity Flows, and Bottom Currents

4.4 Monterey Fan, North Pacific

157

FIGURE 4.5 A. Map showing location of piston and gravity cores taken from ‘channelized lobes’ in the outer Mississippi Fan, Gulf of Mexico. Compiled from Twichell et al. (1992) and Schwab et al.(1996). B. Histograms showing dominance of debris-flow facies in cores from ‘channelized lobes’ in the outer Mississippi Fan (see Fig. 4.5A for location of cores). Percentages of facies were calculated by the author using published data from Schwab et al. (1996). Note that all nine cores contain debris flows, whereas only three cores comprise turbidites. In seven out of nine cores, the amount of debris-flow facies far exceeds the amount of turbidite facies. In core GC 44, debris-flow facies comprises 100%. This facies distribution has important implications for submarine fan models. C. SeaMARC 1A side-scan-sonar image mosaic of ‘depositional lobes’ of the distal Mississippi Fan showing dendritic pattern with abrupt edges. Strong acoustic returns (high backscatter) are white and light grey; weak acoustic returns (light backscatter) are black and dark grey. Note position of core 44, which contains chaotic silt beds and floating clay clasts (see Twichell et al.,1995, their Figure 41.4, p.286),suggesting deposition from slumps and debris flows. Core 44 is composed of 100% debris flow (Figure 4.5B). (Modified after Lee et al. (1996). Image courtesy of D. C. Twichell. After Shanmugam (2016a). JOP, Elsevier. Creative Common.

50 cm in thickness in 21 cores, and they are more than a meter in thickness in 11 cores. The density of core sampling and related grain-size analysis (Paull et al., 2005, their Fig. 2 and their Table 1) suggests the possibility of rather continuous distribution of sand-sized material along the Monterey Canyon floor. Importantly, these massive sands are clean, with implications for high porosity and permeability. Similar continuous distribution of clean massive

Mass Transport, Gravity Flows, and Bottom Currents

158

4. A paradigm shift

FIGURE 4.6 High-resolution shaded relief image of the Monterey Canyon showing sinuous planform geometry (dashed line), offshore northern California. MTD, mass-transport deposit. Credit: NGDC (National Geophysical Data Center) Coastal Relief Model Vol. 06 Shaded Relief Images, NOAA Satellites and Information. Uniform Resource Locator (URL), http://www.ngdc.noaa.gov/mgg/coastal/grddas06/html/gna37123.htm. Accessed June 22, 2004. Credi: NOAA. Public Domain. Labels by G. Shanmugam.

sands in sinuous canyons has been documented using conventional cores and root mean square seismic amplitude maps in the Krishna-Godavari (KG) Basin (Pliocene), Bay of Bengal (discussed in Section 4.5).

Mass Transport, Gravity Flows, and Bottom Currents

4.4 Monterey Fan, North Pacific

159

FIGURE 4.7 Core photographs showing sand-rich lithofacies of split sections of vibracores collected from the floor and flanks of modern Monterey Canyon (northern California). (A) Clean, massive, well-sorted, medium to coarse sand. (B) Clean gravelly sand. (C) Chaotic mixture of coarse sand, gravel, and clay clasts. (D) Floating cobble in clean sandy matrix. (E) Floating clay clasts (arrows) in coarse sandy matrix. Photos courtesy of C. K. Paull. Similar lithofacies were originally published by Paull et al. (2005). GSA. Fair usage. Labels by G. Shanmugam.

4.4.2 Depositional lobe Based on surface morphology, interpreted from geophysical data, the term “depositional lobe” has been applied to this modern Monterey Fan (Normark et al., 1985; Gardner et al., 1996). There are large boulders, which are 20 m or more in diameter, on the most recent depositional lobe of Monetery Fan (Fig. 4.8B). They occur both as isolated individual boulders and as fields of boulders (Gardner et al., 1996). A total of 492 boulders were counted. The fields of boulders show a rough alignment with lineations in backscatter intensity that was interpreted as sediment-transport flow patterns (Fig. 4.8C). Gardner et al. (1996) interpreted these boulders as deposits of large mass-transport flows. Box cores taken from the Monterey Fan also show floating silt clasts in sand (Lee et al., 1996; their Fig. 13.11). This lithofacies may be interpreted as sandy debrites. In studying channel-mouth lobes of the Monterey Fan, Klaucke et al. (2004, p. 181) noted, “Sand is particularly concentrated in finger-like areas of low backscatter intensity and is interpreted as the result of nonturbulent sediment-gravity flows depositing meters thick massive, fine sand.” This interpretation clearly implies that finger-like deposits are products of laminar sandy debris flows. Studies of the Monterey Fan, the Amazon Fan, and the Mississippi Fan clearly suggest that lobe-like features on modern deep-water systems are indeed dominated by debrite

Mass Transport, Gravity Flows, and Bottom Currents

160

4. A paradigm shift

FIGURE 4.8 A. Location map of Monterey Fan showing channel-mouth lobe, offshore California. Note locations of the "Monterey meander" (Greene and Ward, 2003) and the "Shepard meander" (Fildani and Normark, 2004) associated with the Monterey Canyon. Figure from Klaucke et al. (2004). Publication: Marine Geology. With permission from Elsevier. B. TOBI (Towed Ocean Bottom Instrument) Sidescan-sonar image of boulders on surface of the most recent "depositional lobe" of the Monterey Fan. C. Interpretation showing distribution of boulders. Lines represent sediment-transport flow patterns. Arrow on sonograph shows direction of insonification (i.e., sonar illumination). (After Gardner et al. (1996; their Figure 12 17). Figure published by Cambridge University Press, is not subject to U.S. copyright law. Reproduced with permission from J.V.Gardner.

deposition. The routine application of the popular term “depositional lobe,” implying a dominance of turbidites, to debrite-dominated modern fans is erroneous.

4.5 Krishna-Godavari (KG) Basin, Bay of Bengal, India As noted in Chapter 3, Gravity Flows: Debris Flows, Grain Flows, Liquefied/Fluidized Flows, Turbidity Currents, Hyperpycnal Flows, and Contour Currents, the KG Basin is located on the eastern continental margin of India (Fig. 3.30A). Operator Reliance Industries Limited and Niko Resources discovered gas in Pliocene deep-water siliciclastic reservoirs of the KG Basin in 2002. A comprehensive study, based on integration of modern seafloor bathymetry, long sediment cores, and seismic attributes, was carried out (Shanmugam et al., 2009). The cored Pliocene intervals in three wells represent the deep offshore component of the KG Basin. The modern seafloor bathymetry of our study area reveals that the upperslope setting is characterized by widespread MTDs and submarine canyons (Fig. 2.21).

Mass Transport, Gravity Flows, and Bottom Currents

4.6 The Annot Sandstone (Eocene Oligocene), Peira-Cava Area, Maritime Alps, SE France

161

Most canyons are in their incipient stages of development. The western shelf edge is characterized by headwall scarps. Slide blocks have detached from these headwall scarps (i.e., slide scars) and moved downslope developing chutes. As a result, slide blocks occur at the mouths of chutes (Fig. 2.21). Pliocene canyons are sinuous, exhibit 90-degree deflections, at least 22 km long, relatively narrow (500 1000 m wide), deeply incised (250 m), and asymmetrically walled. Based on examination of 313 m of conventional cores from three wells (Fig. 3.30B), five depositional facies have been interpreted: (1) sandy debrite (Figs. 3.31, 3.32, and 3.34), sandy slump, sandy slide, and sandy cascading flow; (2) muddy slump and debrite; (3) sandy tidalite; (4) muddy tidalite; and (5) hemipelagite. Debrites and slumps constitute up to 99% in one well. Sand injectites are common. Pliocene environments are interpreted to be comparable to the modern upper continental slope with widespread MTDs and submarine canyons in the KG Basin. Frequent tropical cyclones, tsunamis, earthquakes, shelf-edge canyons with steep-gradient walls of more than 30 degrees, and seafloor fault scarps are considered to be favorable factors for triggering mass movements. Sandy debrites occur as sinuous canyon-fill massive sands, intercanyon sheet sands (1750 m long or wide and 32 m thick), and canyon-mouth slope-confined lobate sands (3 km long, 2.5 km wide, and up to 28 m thick). Canyon-fill facies are characterized by the close association of sandy debrites and tidalites (see Fig. 3.35) (Table 3.3). Reservoir sands, composed mostly of amalgamated units of sandy debrites, are thick (up to 32 m), low in mud matrix (less than 1% by volume), and high in measured porosity (35% 40%) and permeability (850 18,700 mD). Because upper-slope sandy debrites mimic base-of-slope turbidite channels and lobes in planform seismic geometries (Fig. 2.20B), the use of conventional turbidite-fan models as a template to predict the distribution of deep-water sand is tenuous, without the validation from process sedimentological studies of sediment cores.

4.6 The Annot Sandstone (Eocene Oligocene), Peira-Cava Area, Maritime Alps, SE France Bouma (1962) used the Annot Sandstone [Gre`s d’ Annot Formation (Eocene Oligocene)], exposed in the Peira-Cava Area and vicinity of the French Maritime Alps, for developing the first turbidite facies model. Later, he extended his study to Switzerland and other areas in Europe. Bouma (1962) documented his field observations in eight photographic plates, seven of which contain outcrop photographs from the Peira Cava type locality (Plates A, B, C, D, E, F, and H) and the eighth one contains outcrop photographs from Switzerland (Plate G). Although the Annot Sandstone appears to show normal grading, detailed description offers a different story. To understand the nature of normal grading and related complexities in the Annot Sandstone, Mobil Oil Company initiated a field research project with three geologists (R.J. Moiola, R.B. Bloch, and G. Shanmugam) in 1998. We reexamined in detail 12 sandstone units of the Annot Sandstone, exposed along a road section in the Peira-Cava Area of French Maritime Alps (Fig. 4.9). This is the same road section that Bouma (1962, p. 93, his Fig. 23) and Lanteaume et al. (1967) first used in their studies of the Annot Sandstone. Results of our reexamination of the Annot Sandstone were

Mass Transport, Gravity Flows, and Bottom Currents

162

4. A paradigm shift

FIGURE 4.9 Location map showing sites of 12 measured units of the Annot Sandstone(Eocene Oligocene) along a road section, nearPeira Cava area, French Maritime Alps, during a field study by Mobil Oil Company. Bouma(1962, p.93, his Figure 23) used this road section, among others, in his Ph.D. study of the Annot Sandstone and other formations in developing the turbidite facies model. Unit 9 in this study corresponds to Bouma’s Layer No.1 in his measured section K (seeBouma,1962, Enclosures I in inside pocket of back bookcover). An outcrop photograph of Layer No.1 is published by Bouma (1962,see his Plate H1). Each unit in this study represents a major sandstone body and closely related minor sandstone and mudstone beds. Map is simplified after Lanteaume et al. (1967). Figure from Shanmugam (2002a). Publication: Earth-Science Reviews. With permission from Elsevier.

published (Shanmugam, 2002a, 2003a, 2006a), The key features that contradict the conventional turbidite interpretation of the Annot Sandstone are: • Basal contorted layers (Fig. 4.10): The contorted intervals beneath the sandstone are interpreted to be a product of shearing and slumping produced by stress exerted by overriding mass flows. Large-scale shear structures have been reported in the Annot Sandstone in other areas as well (Clark and Stanbrook, 2001). • Basal inverse grading (Fig. 4.11A and C): A combined mechanism of dispersive pressure, matrix strength, hindered settling, and buoyant lift would explain the development of inverse grading. The inverse grading is attributed to a plastic debris-flow origin.

Mass Transport, Gravity Flows, and Bottom Currents

4.6 The Annot Sandstone (Eocene Oligocene), Peira-Cava Area, Maritime Alps, SE France

163

FIGURE 4.10 A. Sedimentological log of amalgamated sandstone Unit 2. B. Outcrop photograph of Unit 2 showing basal contorted layrs. Annot Sandstone (Eocene Oligocene), Peira Cava area, French Maritime Alps. SE France. Figures from Shanmugam (2002a). Publication: Earth-Science Reviews. With permission from Elsevier.

• Basal normal grading: Because these sandstone intervals are not only amalgamated but composed of complex internal features, no simple origin is meaningful. • Lenticular layers (Fig. 4.12A): Lenticular layers with quartzose granules in sandstone may be interpreted as deposits of non-Newtonian flows with strength. The presence of planar fabric supports the laminar state of flow (Fisher, 1971). By simply describing these sedimentary features without using the “Bouma” divisions, lenticular layers would be interpreted to be deposits of plastic laminar flows. • Pockets of gravel (Fig. 4.12B): Unit 8 with pockets of gravel cannot be explained by a single waning turbidity current. The depositing flow must have had enough flow strength to support granules near its upper part. The pockets of gravels near the top of the bed reflect freezing of a plastic flow (Fig. 4.12B). • Floating armored mudstone balls (Fig. 4.13A): Stanley et al. (1978) interpreted armored mudstone balls in the Annot Sandstone to be associated with the filling of canyons by mass flows. • Floating mudstone clasts (Fig. 4.11E): In Unit 7, intervals of floating mudstone clasts are interpreted as deposits of plastic debris flows. A combination of dispersive pressure, matrix strength, hindered settling, and buoyant lift is proposed as the cause of floating clasts. • Floating quartzose granules (Fig. 4.11A): Even a single floating quartzose granule in a quartz-rich sandy matrix is of rheologic and hydrodynamic significance. In the Annot

Mass Transport, Gravity Flows, and Bottom Currents

164

4. A paradigm shift

FIGURE 4.11 A. Sedimentological log of amalgamated sandstone Unit 7 showing basal inverse grading overlain by an interval of complex normal grading with floating granules and mudstone clasts, parallel laminae, and lenticular layers. Note sudden increase in grain size at 5 m. Note conventional description using Bouma notations (Ta, Tb, and Tc). B. Outcrop photograph of Unit 7 showing sheet-like geometry. C. Outcrop photograph of Unit 7 showing basal inversely graded interval in coarse- to granule-grade sandstone. D. Outcrop photograph of a pocket of clasts and matrix in the middle of the unit. Arrow shows stratigraphic position of photo. E. Outcrop photograph of Unit 7 showing a floating mudstone clast in the middle of the unit. Annot Sandstone (Eocene Oligocene), Peira Cava area, French Maritime Alps. SE France. Figures from Shanmugam (2002a). Publication: Earth-Science Reviews. With permission from Elsevier.

Sandstone, quartzose granules floating in a sandy matrix are evidence that the flow had strength and that settling of the grains is hindered. Because Unit 7 contains floating quartz granules (Fig. 4.11A) and amalgamation surfaces, it has been interpreted to be deposits of multiple episodes of sandy debris flows and bottom currents. • Double mud layers (DMLs, Fig. 4.13A and B): There are no analogous divisions for DMLs in the Bouma Sequence. DMLs are unique to shallow-water tidal environments and have been ascribed to alternating ebb and flood tidal currents with extreme time-velocity asymmetry in subtidal settings (Visser, 1980). In the Annot Sandstone (Fig. 4.13A), DMLs have been interpreted to be deposits of deep-marine tidal currents (Shanmugam, 2003a). Hydrodynamically, turbidity currents are unsuitable to explain the DML. • Sigmoidal cross-bedding (Figs. 4.14 and 4.15) (Shanmugam, 2000a, 2003a): There are no analogous divisions for sigmoidal cross-bedding in the Bouma Sequence. These bedforms, typical of tidal currents in estuarine environments (Fig. 4.16) (Terwindt, 1981;

Mass Transport, Gravity Flows, and Bottom Currents

4.6 The Annot Sandstone (Eocene Oligocene), Peira-Cava Area, Maritime Alps, SE France

165

FIGURE 4.12

A. Sedimentological log of amalgamated sandstone Unit 8 showing basal inverse grading, floating armored mudstone balls, lenticular layers, and pockets of gravel. Note conventional description using Bouma notations (Ta). B. Outcrop photograph of Unit 8 showing a pocket of gravel (bounded by dashed line) with quartz, feldspar, rock fragments, and mudstone clasts in fine-grained sandstone. Arrow shows stratigraphic position of photo. Annot Sandstone (Eocene Oligocene), Peira Cava area, French Maritime Alps. SE France. Figures from Shanmugam (2002a). Publication: Earth-Science Reviews. With permission from Elsevier.

Banerjee, 1989; Shanmugam et al., 2000), can also be attributed to deep-marine tidal currents in submarine canyons (Shanmugam, 2003a). Although the turbidite facies model advocates a simple origin by turbidity currents, details of the Annot Sandstone clearly reveal a complex origin by processes involving slumping, sandy debris flows, and tidal bottom currents. Deposits of true turbidity currents are extremely rare. Our observations are nothing new. Stanley (1963, 1975) was one of the early researchers to recognize the importance of slumps, debris flows, grain flows, and liquefied flows in the origin of the Annot Sandstone, SE France. In a recent comprehensive study, Etienne et al. (2012, p. 3) have acknowledged the complex origin of the Annot Sandstone by the following statement: The Annot Sandstones are up to 1200 m thick (Inglis et al., 1981) and composed of siliciclastic deposits resulting from various types of gravity flows such as slumps, debris flows (sensus Hampton, 1972; Middleton and Hampton, 1973), slurry flows (sensus Lowe and Guy, 2000), high-density turbidity currents (sensus Kuenen, 1966; Middleton, 1967; Lowe, 1982; Postma, 1986) also defined as sandy debris flows (sensus Shanmugam, 1996a, 2000a) and low-density turbidites (sensus Bouma, 1962; Middleton and Hampton, 1973) i.e. classical turbidites. —Cited references Shanmugam, 1996 represents Shanmugam, 1996a in this book.

The problem remains as to how one can explain deep-water units that show a partial Bouma Sequence composed of a basal massive division and an upper parallel-laminated

Mass Transport, Gravity Flows, and Bottom Currents

166

4. A paradigm shift

FIGURE 4.13 A. Sedimentological log of Annot Unit 2. Note conventional description using Bouma notations (Ta, Tb, and Tc). B. Outcrop photo showing double mud layers (DML). Arrow shows stratigraphic position of photo. Annot Sandstone (Eocene Oligocene), Peira Cava area, French Maritime Alps. SE France. Figures A and B from Shanmugam (2003a). Publication: Marine and Petroleum Geology. With permission from Elsevier.

division. In areas in which both downslope sandy debris flows and alongslope-bottom currents operate concurrently, the reworking of the tops of sandy debris flows by bottom currents may be expected. Such a scenario could generate a basal massive sand division and an upper reworked division, mimicking a partial Bouma Sequence (Shanmugam, 2006a, 2012a). The reworking of deep-water sands by bottom currents has been suggested by other researchers as well (e.g., Stanley, 1993; Ito, 2002). In summary, the most influential turbidite facies model (i.e., the Bouma Sequence) was acceptable in 1962 when our understanding of deep-water processes was limited. However, given the cumulative knowledge that we have acquired during the past 58 years, such as double mud layers (Fig. 4.13B) and sigmoidal cross-bedding (Figs. 4.14B and 4.15B) in the Annot Sandstone, the Bouma is obsolete in 2020. Furthermore, several authors have proposed alternative origins for the basal massive division (Ta): 1. antidune (Harms and Fahnestock, 1965); 2. bed load (Sanders, 1965);

Mass Transport, Gravity Flows, and Bottom Currents

4.6 The Annot Sandstone (Eocene Oligocene), Peira-Cava Area, Maritime Alps, SE France

167

FIGURE 4.14

(A) Sedimentological log of an amalgamated sandstone unit showing sigmoidal cross-bedding with mud (mica) drapes. Note a ‘‘normally graded’’ bed at 1 2 m interval with floating clasts. (B) Outcrop photograph showing sigmoidal cross-bedding in medium- to coarse-grade sandstone. Note mud/mica-draped (dark colored) stratification. Arrow shows stratigraphic position of photo. Pickering and Hilton (1998, their Fig. 4K) published a reverse view (i.e., foreset is dipping to the right in outcrop as shown in (B), but the authors published a view in which the foreset is dipping to the left, apparently a printing error!) of this cross-stratified sandstone unit and interpreted it as deposits of ‘‘high-density turbidity currents.’’ Unit 10 (see Fig. 4.9 for location), Annot Sandstone (Eocene Oligocene), Peira Cava area, French Maritime Alps. From Shanmugam (2002a).

3. 4. 5. 6. 7. 8.

grain flows (Stauffer, 1967); pseudoplastic quick bed (Middleton, 1967); density- modified grain flows (Lowe, 1976b); high-density turbidity currents (Lowe, 1982); upper-plane-bed (Arnott and Hand, 1989); sandy debris flow (Shanmugam, 1996a).

Under this umbrella of new knowledge, the Bouma Sequence can no longer function as a genetic facies model for turbidity currents and their deposits Although, it was reasonable in 1962, at a time of limited knowledge on deep-water processes, to introduce a simplistic turbidite facies model, it is unreasonableto apply this model to the rock record in 2020.

Mass Transport, Gravity Flows, and Bottom Currents

168

4. A paradigm shift

FIGURE 4.15. (A) Sedimentological log of an amalgamated sandstone unit showing sigmoidal cross-bedding with tangential toeset. Note inverse grading below and lenticular layers above. (B) Outcrop photograph showing sigmoidal cross-bedding (top arrow) with tangential toeset in coarse- to granule-grade sandstone. Note mud/ mica-draped (dark colored) stratification. Note inversely graded gravel layer below (bottom arrow). Arrows show stratigraphic position of photo. See Bouma and Coleman (1985, their Fig. 5) for an overall view of this unit. Unit 11 (see Fig. 7 for location), Annot Sandstone (Eocene Oligocene), Peira Cava area, French Maritime Alps. From Shanmugam (2002a).

FIGURE 4.16 A model for tidal bundles in shallow-marine environments. This model has been adopted to explain sigmoidal cross-bedding in deep-water Annot Sandstone in SE France in this book. Original diagram from Terwindt (1981). Modified by Banerjee (1989), and further simplified by Shanmugam et al. (2000).

Mass Transport, Gravity Flows, and Bottom Currents

4.7 The Jackfork Group, Pennsylvanian, Ouachita Mountains

169

FIGURE 4.17 A. Location map showing the DeGray Spillway section in Arkansas and the Kiamichi Mountain section in Oklahoma that were used in the study of the Jackfork Group. (After Shanmugam and Moiola (1995). Reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.). B. Turbidite-fan model for the Ouachita flysch by Shanmugam and Moiola (1988) showing a longitudinal fan system in a regional tectonic framework. Note that the fan is not subdivided into inner, middle, and outer segments. C. Debrite-model of the Jackfork Group showing the importance of debris flows and slumps on a slope setting. Tectonic framework is after Thomas (1985). Paleocurrent data of the Jackfork suggest a westward flow in southeastern Oklahoma (Briggs and Cline, 1967). (After Shanmugam and Moiola (1995). Figures A, B, and C from AAPG. With permission from AAPG.

4.7 The Jackfork Group, Pennsylvanian, Ouachita Mountains The Pennsylvanian Jackfork Group in the Ouachita Mountains of Arkansas and Oklahoma (Fig. 4.17A) has conventionally been interpreted by many workers (e.g., Briggs and Cline, 1967), including us (Shanmugam and Moiola, 1988), as a classic flysch sequence dominated by turbidites in a submarine fan setting; however, normal size grading and Bouma sequences, indicative of turbidite deposition, are essentially absent in these sandstone beds. They appear massive (i.e., structureless) in outcrop, but when slabbed reveal diagnostic internal features. These beds exhibit sharp and irregular upper bedding contacts, inverse size grading, floating mudstone clasts, a planar clast fabric, lateral pinch-out geometries, moderate-to-high detrital matrix (up to 25%), sigmoidal deformation (duplex)

Mass Transport, Gravity Flows, and Bottom Currents

170

4. A paradigm shift

structures (Shanmugam et al., 1988a), and contorted layers. All these features indicate sand emplacement by debris flows (mass flows) and slumps. Mud matrix in these sandstones was sufficient to provide cohesive strength to the flow. Discrete units of current ripples and horizontal laminae have been interpreted to represent traction processes associated with bottom-current reworking. As a consequence, we admitted our earlier misinterpretations as a turbidite fan and proposed a totally different debrite model (Shanmugam and Moiola, 1995). This was an epiphany. The dominance of sandy debris-flow and slump deposits (nearly 70% at DeGray Spillway section) and bottom-current reworked deposits (40% at Kiamichi Mountain section) and the lack of turbidites in the Jackfork Group have led us to propose a slope setting. Our rejection of a submarine fan setting has important implications for predicting sandbody geometry and continuity because deposits of fluidal turbidity currents in fans are laterally more continuous than those of plastic debris flows and slumps on slopes. A turbiditedominated fan model (Fig. 4.17B) would predict an outer fan environment with laterally continuous, sheet-like sandstones for the Jackfork Group in southern Oklahoma and western Arkansas, whereas a debrite/slump model (Fig. 4.17C) would predict predominantly a slope environment with disconnected sandstone bodies for the same area. Our field study of the Jackfork Group (Fig. 4.18) revealed abundant evidence for MTD in the form of: • floating quartzite pebbles (Fig. 4.19), • duplex-like structures (Fig. 4.20A), • floating mudstone clasts (Fig. 4.20B), among other features.

FIGURE 4.18

Stratigraphic column of the Pennsylvanian Jackfork Group. Ouachita Mountains.

Mass Transport, Gravity Flows, and Bottom Currents

4.8 Basin-floor fan model, Tertiary, North Sea

171

FIGURE 4.19 Outcrop photograph showing floating quartzite pebbles in sandstone. Pennsylvanian Jackfork Group. Ouachita Mountains.

Our (Shanmugam and Moiola, 1995) controversial reinterpretation had resulted in 42 printed pages of discussions and replies by some of the leading authorities in the field, which included the following: • • • • •

A.H. Bouma, M.B. DeVries, and C.G. Stone, (1997) J.L. Coleman, (1997) A.E. D’Agostino and D.W. Jordan (1997) D.R. Lowe (1997) R.M. Slatt, P. Weimer, and C.G. Stone (1997)

We promptly responded (Shanmugam and Moiola, 1997). These academic discussions had resulted in 42 printed pages in the AAPG Bulletin. It is worth noting that no other paper in the AAPG Bulletin history (1917 present) has generated this much controversy.

4.8 Basin-floor fan model, Tertiary, North Sea The concept of basin-floor fan (Vail et al., 1991) in a sequence-stratigraphic template (Fig. 4.21) has been popular both in the academia (Catuneanu, 2006) and in the petroleum industry (Saller et al., 2004).

Mass Transport, Gravity Flows, and Bottom Currents

172

4. A paradigm shift

FIGURE 4.20. A. Outcrop photograph showing duplex-like structures. B. Photograph of polished slab showing floating clasts. Pennsylvanian Jackfork Group. Ouachita Mountains.

The conventional basin-floor fans are believed to be composed of sandy turbidites (Vail et al., 1991). Mobil Oil Company initiated a major field study in the early 1990s to understand the depositional origin of deep-water strata in the North Sea and adjacent regions. Our examination of nearly 12,000 ft. (3658 m) of conventional core from Paleogene and Cretaceous deep-water sandstone reservoirs cored in 50 wells in 10 different areas or fields reveals that these reservoirs are predominantly composed of MTDs, mainly sandy slumps and sandy debris flows (Shanmugam et al., 1995a). Classic turbidites are extremely rare and comprise less than 1% of all cores. Sedimentary features indicating slump and debris-flow origin include sand units with sharp upper contacts; slump folds; discordant, steeply dipping layers (up to 60 degrees); glide planes; shear zones; brecciated clasts; clastic injections; floating mudstone clasts; planar clast fabric; inverse grading of clasts; and moderate-to-high matrix content (5% 30%). Many of the cored reservoirs either have been previously interpreted as basin-floor fans or exhibit seismic (e.g., mounded forms) and wireline-log signatures (e.g., blocky motif, Fig. 4.21B) and stratal relationships (e.g., downlap onto sequence boundary) indicating basinfloor fans within a sequence-stratigraphic framework. This model predicts that basin-floor

Mass Transport, Gravity Flows, and Bottom Currents

4.8 Basin-floor fan model, Tertiary, North Sea

173

FIGURE 4.21 Conceptual sequence-stratigraphic fan models. A. Seismic geometries showing basin-floor fans (sheet mound) and slope fans (complex or gull-wing mound). B. Wireline-log motifs showing blocky motif of basin-floor fans. C. Slope-fan model. D. Basin-floor fan model. Simplified from Vail et al. (1991). After Shanmugam et al. (1995a). Figures A, B, C and D from AAPG. With permission from AAPG.

Mass Transport, Gravity Flows, and Bottom Currents

174

4. A paradigm shift

FIGURE 4.22 Depositional model of Gryphon Field area, UK North Sea. A. Well-developed blocky log motif. Lower Eocene, Gryphon Field, Kerr-McGee 9/18b-7. B. Depth-tied sedimentological log showing facies distribution. Stratigraphic positions of core features C and D are shown by arrows. C. Facies 2: Core photograph showing a large mudstone clast (arrow) in fine-grained massive sand. Note irregular upper surface of mudstone clast. 5725

Mass Transport, Gravity Flows, and Bottom Currents

4.8 Basin-floor fan model, Tertiary, North Sea

175

fans are predominantly composed of sand-rich turbidites with laterally extensive, sheet-like geometries. However, calibration of sedimentary facies in our long (400 700 ft.) cores with seismic and wireline-log signatures through several of these basin-floor fans (including the Gryphon-Forth, Frigg, and Faeroe areas) shows that these features are actually composed almost exclusively of MTDs consisting mainly of slumps and debris flows (Figs. 4.22 and 4.23). Our core studies thus, underscore the complexities of deep-water depositional systems and indicate that model-driven interpretation of remotely sensed data (i.e., seismic and wireline logs) to predict specific sedimentary facies and depositional features should proceed with caution. Process sedimentological interpretation, using long sediment cores, is commonly critical for determining the true origin and distribution of reservoir sands. Our reinterpretation of massive sands in the North Sea had also resulted in a major discussion by Hiscott et al. (1997) and in a reply by Shanmugam et al. (1997a). This debate was mostly about HDTCs. Like the turbidite facies model (i.e., the Bouma Sequence), sequencestratigraphic models are obsolete for interpreting deep-water sediments (Shanmugam, 2007)

FIGURE 4.23

Plot showing dominance of MTD and BCRS in the deep-water systems of North Sea.rom Shanmugam et al. (1995a), AAPG.

L

ft (1746 m). (D) Facies 2: Core photograph showing steeply dipping layer (arrow),interpreted as internal shear plane at 5726.7 ft (1746.6 m). (E) Schematic depositional model, based on integration of core, log and seismic data, showing mounded seismic facies are a result of slumps and debris flows. Note that the 9/18b-7 well is located at the intersection of seismic lines KMG-881-35 and KMG-881-8 (shown by sketches) and that cored interval is shown by a solid black bar. Areal distribution of slump masses is speculative. (Shanmugam et al. (1995a). Figures A, B, C, D and E from AAPG. With permission from AAPG.

Mass Transport, Gravity Flows, and Bottom Currents

176

4. A paradigm shift

and they promote groupthink. In summary, each model is unique. At present, there is no single fan model with universal applicability because their depositional origin is complex.

4.9 Mass-flow lobes, Ulleung Basin, East Sea, Korea Lee et al. (2011) conducted a detailed analysis of MR1 sonar and chirp (2 7 kHz) seismic data with 5 (2.9 9.4 m long) piston cores and 17 AMS 14C ages. On MR1 sonar image, eight mass-flow lobes are identified in the western basin plain ( . 2100 m). They are covered by c. 2-m thick Holocene pelagic sediments and show the northward flow direction. Lobes 1 4 have large dimensions ( . 27 km long and 15 25 km wide) and occupy in the lower stratigraphic position. In contrast, lobes 5 8, deposited in the more proximal area, have small dimensions (8.8 31.5 km long and 1.2 12 km wide) and occur in the upper stratigraphic position. Lobes 1 4 deposited retrogressively. Lobes 1 and 2 are characterized by relatively strong back-scattering intensity with smooth surfaces on MR1 image and show flat, sharp bottom echo and several distinct to diffuse internal reflectors in chirp profiles. Sediments near their edges consist of fine-grained turbidites (laminated sand/ mud and homogeneous mud) with minor massive clay-rich sand. However, they change to mud-matrix disorganized gravel and massive sand with the overlying fine-grained turbidites toward the proximal part. In this study, the term “mass flow” represents “turbidity currents,” following the classification of Middleton and Hampton (1973). Four other examples from SE Asia have been considered. They are: (1) distal turbidite lobe, Late Paleozoic Taean Formation, Western Korea (So et al., 2013); (2) Miocene submarine fan, Pearl River Mouth Basin, South China Sea (Wang et al., 2012); (3) lacustrine turbidites in the Eocene Shahejie Formation, Dongying Sag, Bohai Bay Basin, North China Craton (Wang et al., 2013); and (4) deep-water sublacustrine fan system from the syn-rift Lower Cretaceous Nantun Formation of the Tanan Depression (Tamtsag Basin), Mongolia (Jia et al., 2014). However, process sedimentological details on the origin of these sands are insufficient.

4.10 Upper Triassic Yanchang Formation, Ordos Basin, central China In China, the concept of sandy debris flows has been applied to petroleum reservoirs in the Ordos Basin (Li et al., 2011; Zou et al., 2012). In a detailed sedimentological study, based on conventional cores from 30 wells representing the Upper Triassic Yanchang Formation of the Ordos Basin in central China, Zou et al. (2012) had recognized four lithofacies: (1) fine-grained massive sandstone with floating mudstone clasts and planar clast fabric (sandy debrite); (2) fine-grained sandstone and siltstone showing contorted bedding, sand injection, and ptygmatic folding (sandy slump); (3) fine-grained sandstone with thin layers of normal grading and flute casts (turbidite); and (4) mudstone with faint laminae (suspension fallout). The proposed model shows an upslope increase in sandy debrites toward the lake margin and a downslope increase in turbidites toward the lake center (Fig. 4.24). Zou et al. (2012) documented facies quantitatively. For example, the Chang 6 sublayer is characterized by sandy debrites in 95% of the cored wells. Sandy debrites are reservoir facies, whereas turbidites are

Mass Transport, Gravity Flows, and Bottom Currents

4.11 Supercritical and subcritical fans

177

FIGURE 4.24. Deep-lacustrine depositional model showing a non-channelized distribution of sandy debrites. Four depositional facies have been recognized: (1) sandy debrite, (2) sandy slump, (3) turbidite, and (4) suspension fallout. Note an upslope increase in sandy debrites and a downslope increase in turbidites in the Upper Triassic Yanchang Formation, Ordos Basin, central China. These trends have been used to predict the distribution of reservoir facies composed of sandy debrites (see text for details). Modified after Zou et al. (2012).

nonreservoir facies. Zou et al. (2012) had used these trends as a template for predicting the distribution of sandstone reservoir facies, composed of sandy debrites, in the Huaqing oil field with 100 million tons of oil reserves. The implication is that this predictive model may be applicable to analogous deep-lacustrine basins worldwide.

4.11 Supercritical and subcritical fans Based on the basic assumption that submarine fans are deposited by stratified HDTCs and nonstratified low-density turbidity currents, Postma and Cartigny (2014, their Fig. 3) classified various elements of submarine fans as deposits of supercritical (Fr . 1) and subcritical (Fr , 1) turbidity currents. In this classification, the underpinning principle is the Froude Number (Fr). Froude Number U 5 flow velocity U Fr 5 pffiffiffiffiffi g 5 gravitational acceleration gh h 5 flow depth On steep slopes, with gradients .0.5 degrees, supercritical flows are likely to develop (Komar, 1971). Supercritical flows generate distinct bedforms in experiments (Cartigny

Mass Transport, Gravity Flows, and Bottom Currents

178

4. A paradigm shift

et al., 2014). In explaining supercritical fans, Hamilton et al. (2015) conducted experiments with steep slopes and saline density currents with crushed plastic to emulate sustained turbidity currents and bedload transport. This bedload transport is analogous to experimental stratified flows with a basal laminar-inertia flow and an upper (turbulent) turbidity current that has been termed as “high-density turbidity currents” by Postma et al. (1988) (see Fig. 3.39B). Hoyal et al. (2014) applied the concept of “Froude supercritical submarine fans” to examples of “steep” fans that include the Brushy Canyon (Texas), Ainsa (Spain), Tabernas/Sorbas (Spain), Karoo (South Africa), Golo (Corsica), and East Breaks (Gulf of Mexico) using outcrop and high-resolution seismic data. They have concluded that sandy submarine fans from steep and tectonically active margins are typically small in radius (,10 km), and could be dominated by deposits emplaced at hydraulic jumps. Despite their popularity in both academia (Postma and Cartigny, 2014) and in the petroleum industry (Hoyal et al., 2014), the Froude fan models suffer for the following reasons: • No one has documented supercritical turbidity currents, which Postma and Cartigny (2014, their Fig. 3) equate with high-density (stratified) turbidity currents (see Fig. 13B), in the modern oceans. • There is no experimental basis for Froude fans. For example, Kostic and Parker (2007) observe that “The transition from supercritical to subcritical flow is accomplished through an internal hydraulic jump. Consider a steady turbidity current flowing from a steep canyon onto a milder fan, and then exiting the fan down another steep canyon. The flow might be expected to undergo a hydraulic jump to subcritical flow near the canyon fan break and then accelerate again to critical flow at the fan canyon break downstream. The problem of locating the hydraulic jump is here termed the ‘jump problem.’ Experiments with fine-grained sediment have confirmed the expected behavior outlined above. Similar experiments with coarse-grained sediment suggest that if the deposition rate is sufficiently high, this ‘jump problem’ may have no solution with the expected behavior, and in particular no solution with a hydraulic jump. In such cases, the flow either transits the length of the low-slope fan as a supercritical flow and shoots off the fan canyon break without responding to it or dissipates as a supercritical flow before exiting the fan. The analysis presented below confirms the existence of a range associated with rapid sediment deposition where no solution to the ‘jump problem’ can be found.” No further explanation is necessary. • In the Froude-supercritical low (Fr . 1), where the water surface is in phase with the bed, upstream-migrating antidune bedforms commonly develop (Simons et al., 1965). Although antidunes have been recognized in outcrops (Skipper, 1971; Andreetta, 2009), such bedforms (i.e., “backset bedding” or cross-stratification that dips against the direction of flow of the depositing currents) are impossible to recognize in the subsurface sediment cores of ancient deep-water systems. • In the Brushy Canyon Formation, Delaware Basin, the presence of hummocky crossstratification in outcrops has been interpreted as storm-modified hyperpycnites in shallow-water environments (Higgs, 2009), which contradicts deposition by supercritical flows.

Mass Transport, Gravity Flows, and Bottom Currents

4.12 Synopsis

179

• In the Golo (Corsica) example, there is uncertainty about the origin of sand beds (Shanmugam, 2016a). • In the East Breaks (Gulf of Mexico) example, the dominance of sandy debrites and slumps in cores and on seismic data (Woodbury et al., 1978; Rothwell et al., 1991; Shanmugam, 2006a, his Section 6.3.2) contradicts deposition by supercritical flows. Given these uncertainties, one should proceed with caution in classifying submarine fans using Froude regimes.

4.12 Synopsis • This chapter documents observational evidence from case studies, based on Amazon Fan (Equatorial Atlantic), Mississippi Fan (Gulf of Mexico), Monterey Fan (North Pacific), KG Basin (Bay of Bengal), the Annot Sandstone (Eocene Oligocene, Peira-Cava Area, Maritime Alps, SE France), the Jackfork Group (Pennsylvanian, Ouachita Mountains, United States), basin-floor fans in the North Sea, Upper Triassic Yanchang Formation (Ordos Basin, central China), among others, that compel a paradigm shift from turbidites to MTD and BCRS. • The three principal fan elements (i.e., canyons, channels, and lobes) are historically considered to be dominated by turbidites. In reality, case studies demonstrate that modern canyons, channels, and lobes are dominated by MTDs, not turbidites. A supporting empirical evidence for a paradigm shift is the study of tidal currents in modern submarine canyons worldwide. In their seminal study, Shepard et al. (1979) documented 25,000 hours of velocity measurements of mostly tidal currents from over 150 stations located in 25 submarine canyons around the world. This study has never been matched by any other studies on velocity measurements of modern turbidity currents in submarine canyons, including by the most recent study on the Monterey Canyon (Maier et al., 2019). The previous 12 case studies of modern turbidity currents that were published during a period of 56 years (1952-2008) (e.g., Heezen and Ewing, 1952; Inman et al., 1976; Hay et al., 1982; Normark, 1989; Parsons et al., 2003; Xu et al., 2004, among others) were also unconvincing (see explanations by Shanmugam, 2012a). This extreme rarity of velocity measurements of modern turbidity currents is truly mystifying given the fact that thousands of ancient strata were interpreted routinely as turbidites in the sedimentary record. In my view, the driving force behind most turbidite interpretations was turbidite facies models and groupthink. • Empirical data suggest that MTDs, mostly debrites, can and do occur in a number of settings (Fig. 4.25), usually known for turbidites. More importantly, these examples show that debrite sands are distributed in a variety of manner in deep-water settings, not just as channels and lobes. This new information should help petroleum geoscientists to develop realistic depositional models in future deep-water exploration. • Finally, the turbidite paradigm has failed on all fronts, namely theoretical, experimental, and empirical grounds. In understanding this conclusion, all one has to

Mass Transport, Gravity Flows, and Bottom Currents

180

4. A paradigm shift

A. Coalesced debris fans N

B. Isolated debris tongue

N

C. Multiple debris tongues N

5 km Lake Tahoe

Lake Tahoe

D. Detached blocks

5 km

E. Sinuous & braided canyons N

Bear Island Fan

100 km

F. Sinuous channel

N

N

NW Africa

100 km

G. Channelized lobe N

LA Margin

10 km

H. Nonchannelized lobe

Amazon Fan

50 km

i. Intraslope basin

N

Mississippi Fan

FIGURE 4.25

10 km

Monterey Fan

50 km Beaumont Basin

10 km

Depositional settings of MTD composed mostly of debrites. From Shanmugam (2006a).

FIGURE 4.26 Comparison of number of direct observations on deep-water tidal and turbidity currents in modern settings with number of interpreted deep-water tidalites and turbidites in the ancient sedimentary record. Importantly, Shepard et al. (1979) documented 25,000 hours of velocity measurements of mostly tidal currents from 25 submarine canyons around the world. But no such robust dataset on modern turbidity currents exists. In explaining the rarity of turbidity currents in modern oceans, sequence stratigraphers could use the notion that turbidity currents occur preferentially during periods of sea-level lowstands (Damuth and Fairbridge, 1970; Shanmugam and Moiola, 1982, 1988; Vail et al., 1991). However, such notions are obsolete (Shanmugam, 2007).

do is to rigorously examine the obvious disconnect between direct observations on deep-water tidal currents and turbidity currents in modern settings and interpreted deep-water deposits in the ancient sedimentary record (Fig. 4.26).

Mass Transport, Gravity Flows, and Bottom Currents

C H A P T E R

5 Density plumes: types, deflections, and external controls O U T L I N E 5.1 Introduction

182

5.2 Dataset

187

5.3 General types of density plumes

188

5.4 Deflected sediment plumes and their control 189 5.4.1 Elwha sediment plume, Strait of Juan de Fuca, United States 189 5.4.2 Connecticut River, New England region, United States, Long Island Sound, United States 189 5.4.3 Eel River, California, Pacific Ocean, United States 191 5.4.4 Mississippi River, Gulf of Mexico, United States 191 5.4.5 Rı´o de la Plata Estuary, South Atlantic Ocean, Argentina, and Uruguay 192 5.4.6 Rhone River, Gulf of Lions, Mediterranean Sea, France 194 5.4.7 Ebro Delta, Mediterranean Sea, Iberian Peninsula 195 5.4.8 Guadalquivir River, Gulf of Ca´diz, Southern Spain 196

Mass Transport, Gravity Flows, and Bottom Currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00005-9

5.4.9 Tiber River, Tyrrhenian Sea, Italy 197 5.4.10 Mornos and Fonissa Rivers, Gulf of Corinth, Greece 198 5.4.11 Congo (Zaire) River, Atlantic Ocean, West Africa 199 5.4.12 Yellow River, Bohai Bay, China 199 5.4.13 Yangtze River, East China Sea, China 200 5.4.14 Pearl River, South China Sea, China 200 5.4.15 Krishna-Godavari Rivers, Bay of Bengal, India 202 5.4.16 Brisbane River, Moreton Bay, Australia 203 5.4.17 Dart River, Lake Wakatipu, South Island, New Zealand 203 5.5 Global significance of wind forcing on sediment plumes 205 5.6 Implications for sediment transport 207 5.7 Implications for provenance

210

5.8 Synopsis

211

181

© 2021 Elsevier Inc. All rights reserved.

182

5. Density plumes: types, deflections, and external controls

5.1 Introduction Density plumes and their various configurations seen on satellite images have been a source of curiosity to the geologic community as well as to the general public. The U.S. National Aeronautics and Space Administration (NASA, 2019) has archived satellite images on density plumes in its online publishing outlet “Earth Observatory” since 1999. NASA has used a variety of satellites, such as Aqua, Terra, and Topex/Poseidon. However, there has not been a systematic attempt to compile variations in natural configurations of density plumes in the world’s oceans and lakes (Shanmugam, 2018b, 2018c). Following Shanmugam (2019d), this chapter is an attempt to document common occurrence of density plumes that are deflected at river mouths worldwide due to various external controlling factors (Fig. 5.1; Table 5.1). Bates (1953) suggested three types of density plumes at river-mouth deltaic environments: (1) hypopycnal plume for floating river water that has lower density than basin water (Fig. 5.2A), (2) homopycnal plume for mixing river water that has equal density as basin water (Fig. 5.2B), and (3) hyperpycnal plume for sinking river water that has higher density than basin water (Fig. 5.2C). Although rivermouth hyperpycnal plumes have received much attention (Bates, 1953; Mulder et al., 2003), plumes in other environments (e.g., lakes) are equally important (Shanmugam, 2018b,c). River mouths constitute an important intersectional setting between terrestrial and marine or lacustrine environments. In terms of processes that influence river-mouth sedimentation are waves (Komar, 1976), tides (Klein, 1970), gravity-driven downslope processes (Middleton and Hampton, 1973), cyclones (Shanmugam, 2008a), tsunamis (Shanmugam, 2008a), and shelf edge related currents (Southard and Stanley, 1976).

FIGURE 5.1 Location map of 29 rivers used in this study. See Table 5.1 for details.

Mass Transport, Gravity Flows, and Bottom Currents

TABLE 5.1 Case studies of 29 rivers, their sediment plumes, and external controls. Serial number (Fig. 5.1) Case study and location

Type

Environment

External control

Comments

1

Amazon River, Brazil, Equatorial Atlantic

Deflecting

Open marine

Ocean currents and phytoplankton

Natural sediment plume

2

Betsiboka River, NW Madagascar

Massive

Bay

Tides

Natural sediment plume

3

Brisbane River, Australia, Moreton Bay

Deflecting (Fig. 5.18)

Bay

Tides

Anthropogenic, due to Port of Brisbane

4

Chignik River, Alaska, Pacific Ocean

Linear

Braid delta in a lagoon, Pacific Ocean

Coarse-grained braid delta (McPherson et al., 1987)

Natural sediment plume

5

Congo (Zaire) River, West Africa

Deflecting (Fig. 5.13)

Marine

Tidal currents (Shanmugam, 2003a)

Natural sediment plume

6

Connecticut River, New England region, United States

Deflecting (Fig. 5.5A)

Long Island Sound

Wind forcing (Hurricane Irene, August 21 30, 2011)

Natural sediment plume

7

Copper River, United States, Coalescing Gulf of Alaska

Braid delta, Marine

Eolian

Natural sediment plume

8

Dart River, South Island, New Zealand, Lake Wakatipu

Deflecting (Fig. 5.19C)

Braid delta, lacustrine

Tidal? (Heath, 1975)

Natural sediment plume

9

Ebro Delta, Iberian Peninsula, Mediterranean Sea

Deflecting (Fig. 5.9B)

River-dominated delta

Wind forcing, cyclonic events, and ocean currents (Arnau et al., 2004)

Natural sediment plume

10

Eel River, California, United States

Deflecting (Fig. 5.6B)

Shelf

Wind forcing and shelf currents (Geyer et al., 2000; Imran and Syvitski, 2000)

Natural sediment plume

11

Elwha River, Washington, United States, Strait of Juan de Fuca

Deflecting (Fig. 5.4C)

Strait

Tidal currents (Cannon, 1978; Thomson Anthropogenic plume et al., 2007) and upwelling wind currents caused by dam demolition (Foreman et al., 2008)

12

Fonissa River, Greece, Gulf of Corinth

Deflecting (Fig. 5.12C)

Delta

Bottom currents (Beckers et al., 2016)

Natural sediment plume (Continued)

TABLE 5.1 (Continued) Serial number (Fig. 5.1) Case study and location

Type

Environment

External control

Comments

U-Turn (deflecting) (Fig. 5.10C)

River-dominated delta

Surface and slope currents (Peliz et al., 2009)

Natural sediment plume

Tide-dominated estuary (Balasubramanian and Ajmal Khan, 2002)

Tidal currents

Natural sediment plume

13

Guadalquivir River, Southern Spain, Gulf of Ca´diz

14

Hugli River, a tributary of Anastomosing the Ganges River, India, Bay of Bengal

15

Krishna-Godavari Rivers, India, Bay of Bengal

Deflecting (Figs. 5.16 and 5.17)

Estuary

Tidal currents (Shanmugam et al., 2009), monsoonal currents (Jagadeesan et al., 2013), and geostrophic currents (Sridhar et al., 2008)

Natural sediment plume

16

Mackenzie River Delta, Canada, Beaufort Sea

Swirly

River-dominated delta

Arctic ocean currents

Natural sediment plume

17

Mississippi River, United States, Gulf of Mexico

Deflating (Figs. 5.2D and 5.7B)

River-dominated delta

Wind forcing, shelf currents (Walker and Rouse, 1993)

Natural sediment plume

18

Mornos River, Greece, Gulf of Corinth

Deflecting (Fig. 5.12B)

Delta

Bottom currents (Beckers et al., 2016)

Natural sediment plume

19

Niger River, West Africa

Linear

Wave-dominated delta

Wind forcing, wave currents

Natural sediment plume

20

Nile Delta, Egypt, Mediterranean Sea

Lobate

River-dominated delta

Wind forcing

Natural sediment plume

21

Onibe River, Eastern Madagascar

Dissipating with sharp front

Marine

Wind forcing, cyclone (Giovanna, Natural sediment plume February 7 24, 2012), and ocean currents

22

Pearl River, South China Sea Deflecting (Fig. 5.15C)

Marine

Upwelling jets (Chen et al., 2017)

Natural sediment plume

23

Rhone Delta, France, Gulf of Deflecting (Fig. 5.9A) Lions, Mediterranean Sea

River-dominated delta

Ocean currents (Arnau et al., 2004)

Natural sediment plume

24

Rı´o de la Plata Estuary, Argentina and Uruguay, South Atlantic Ocean

Dissipating and deflecting (Fig. 5.8C)

Marine

Ocean currents (Gonzalez-Silvera et al., 2006; Matano et al., 2010)

Natural sediment plume

25

Rupert Bay, Quebec

Swirly

Bay

Mixing of river and seawater combined with churn of tides

Natural sediment plume

26

Tiber River, Italy, Tyrrhenian Sea

Deflecting (Fig. 5.11B)

Marine

Wind forcing, longshore currents (Mikhailova et al., 1998)

Natural sediment plume

27

Yangtze River, China, East China Sea

Deflecting (Fig. 5.14A)

Tide-dominated estuary

Shelf currents (Liu et al., 2006), vertical mixing by tides in winter months (Luo et al., 2017)

Natural sediment plume

28

Yellow River, China, Bohai Bay

Horse’s tail (deflecting) (Fig. 5.2E), lobate (Fig. 5.2F)

River-dominated delta

Tidal shear front (Wang et al., 2010)

Natural sediment plume

29

Zambezi River, Central Mozambique, India Ocean

Coalescing lobate, Wave-dominated delta associated with multiple river mouths

Wind forcing, longshore currents (Mikhailov et al., 2015)

Natural sediment plume

186

5. Density plumes: types, deflections, and external controls

FIGURE 5.2 (A C) Schematic diagrams showing three types of density plumes at river mouths in deltaic environments based on concepts of Bates (1953). (D) Image of the Mississippi River showing well-developed deflecting plume (yellow arrow). Circle shows river mouth. (E) Satellite image of the Yellow River showing welldeveloped lobate plume at the old river mouth. (F) Satellite image of the Yellow River showing horse’s tail (deflecting) plume at the modern river mouth that was initiated in 1996. Two circles show old and modern river mouths. Source: (A C) From Shanmugam, G., 2012a. New perspectives on deep-water sandstones: Origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration And Production, vol. 9, Elsevier, Amsterdam, p. 524, with permission from Elsevier. (D F) From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 238.

In addition to longshore currents (Komar, 1976), there are cross-shelf currents (Brink, 2016) and upwelling currents (Milliff et al., 2004; Foreman et al., 2008) that are active at the river-mouth and shelf environments. In evaluating river-mouth phenomena, the primary purpose of this chapter is to document the common occurrence of deflected sediment plumes at river mouths. The second objective is to document the types of external controls involved in deflecting sediment plumes. The third objective is to discuss implications of sediment deflections in understanding paleocurrents, paleogeography, provenance, and reservoir distribution in the ancient sedimentary record. Although the primary dataset for this study is from NASA’s satellite images, other published photographic and other images are also used.

Mass Transport, Gravity Flows, and Bottom Currents

5.2 Dataset

187

5.2 Dataset The term “sediment plume” is used here for plumes in which the primary cause of density is sediment, although salinity and temperature are important in some cases. A plume is defined as a fluid enriched in sediment, ash, biological or chemical matter that enters another fluid. NOAA Fisheries Glossary (2006, p. 42) defines a River Plume as “Turbid freshwater flowing from land and generally in the distal part of a river (mouth) outside the bounds of an estuary or river channel.” Because the original concept of hyperpycnal flows is closely tied to river floods, density plumes at river mouths are considered using the following 29 rivers in this study (Fig. 5.1): 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29.

Amazon River, Equatorial Atlantic, Brazil; Betsiboka River, Bombetoka Bay, NW Madagascar; Brisbane River, Moreton Bay, Australia; Chignik River, Alaska, Pacific Ocean, United States; Congo (Zaire) River, South Atlantic Ocean, West Africa; Connecticut River, New England region, United States, Long Island Sound, United States; Copper River, Gulf of Alaska, United States; Dart River, South Island, Lake Wakatipu, New Zealand; Ebro River, Mediterranean Sea, Iberian Peninsula; Eel River, California, Pacific Ocean, United States; Elwha River, Strait of Juan de Fuca, United States; Fonissa River, Gulf of Corinth, Greece; Guadalquivir River, Gulf of Ca´diz, Southern Spain; Hugli River (a distributary of the Ganges River), Bay of Bengal, India; Krishna-Godavari Rivers, Bay of Bengal, India; Mackenzie River, Beaufort Sea, Canada; Mississippi River, Gulf of Mexico, United States; Mornos River, Gulf of Corinth, Greece; Niger River, North Atlantic Ocean, West Africa; Nile River Delta, Mediterranean Sea, Egypt; Onibe River, Indian Ocean, Eastern Madagascar; Pearl River, South China Sea, China; Rhone River, Gulf of Lions, Mediterranean Sea, France; Rı´o de la Plata Estuary, South Atlantic Ocean, Argentina and Uruguay; Rupert River, Quebec, Rupert Bay, Canada; Tiber River, Tyrrhenian Sea, Italy; Yangtze River, East China Sea, China; Yellow River, Bohai Bay, Bohai Sea, China; Zambezi River, Indian Ocean, Central Mozambique.

Mass Transport, Gravity Flows, and Bottom Currents

188

5. Density plumes: types, deflections, and external controls

FIGURE 5.3 Summary diagram showing 14 general types of plumes that include 12 marine examples and two lacustrine examples. (A) Lobate plume developed by a single river channel. (B) Coalescing lobate plume developed by multiple river channels. (C) Dissipating plume with an irregular front developed within a major estuary. (D) Linear plume developed in a braid delta (McPherson et al., 1987). (E) U-turn plume developed in response to influence by ocean currents. (F) Meltwater plume developed from glacier. (G) Dust plume from eolian processes that can transport dust beyond the shelf edge. (H) Cascading plume developed during cyclones that tend to transport sediment (gravel, sand, and mud) beyond the shelf break (Shanmugam, 2008a). (I) Backwash plume developed during tsunamis that tend to transport sediment (gravel, sand, and mud) beyond the shelf break (Shanmugam, 2006a). (J) Whitings plume and ring plume developed in carbonate environments. (K) Ash plume developed during volcanic eruptions. (L) Tendril and swirly plumes developed in lakes. Note that with the exception of cyclones and tsunamis, none of the other plumes can transport sand and gravel to the deep sea by bedload mode. Although these different types can be recognized on modern systems using satellite or other photographic images, individual types cannot be distinguished in the ancient sedimentary record yet. Source: Updated after Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 238.

5.3 General types of density plumes A review of 45 case studies of density plumes worldwide at various settings (Shanmugam, 2018a,b), represents not only conventional river-mouth plumes, but also estuarine, meltwater, eolian, volcanic, cyclonic, upwelling, and other environments (Fig. 5.3). Although I tried to describe plume types based on shapes, it is not always possible because shapes are too complex in some cases. For this reason, I used their origin in classifying plumes in selected cases (e.g., volcanic plume, meltwater plume, dust plume, etc.). In this chapter, I have selected the following rivers in understanding plume deflections at river mouths by external controlling factors.

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

189

5.4 Deflected sediment plumes and their control External controls are allogenic in nature, which are external to the depositional system, such as uplift, subsidence, climate, eustasy, etc. However, external controls of density plumes are much more variable and include some common depositional processes (e.g., tidal currents, wind forcing, cyclones, etc.). In addition, local physiographic elements, such as seafloor ridges and channels could influence the path of plumes. Anthropogenic structures are also known to control deflection of plumes. In this study, the deflection of sediment transport away from the normal downslope transport, by mechanisms, such as longshore currents, is emphasized.

5.4.1 Elwha sediment plume, Strait of Juan de Fuca, United States The Elwha River is 72 km long in the Olympic Peninsula in the U.S. state of Washington. From its source at Elwha Snowfinger in the Olympic Mountains, it flows generally north to the Strait of Juan de Fuca at the United States Canada border (Fig. 5.4A). A spectacular example of an anthropogenic Elwha sediment plume was triggered by the demolition of Elwha Dam in the Olympic Peninsula, State of Washington (Fig. 5.4A). This sediment plume (Fig. 5.4C) was the result of sediment released from the world’s largest dam demolition (Ritchie et al., 2018). The University of Washington (Seattle, WA) first reported this phenomenal event and its oceanographic and sedimentologic implications in the UW News (Hickey, 2013). According to USGS (2018), this demolition event flushed out 20 million tons of sediment into the Strait of Juan de Fuca. One could classify these Elwha sediment plumes (Fig. 5.4C) as modern hyperpycnal flows based on visual observation alone. However, without measurements of fluid theology, flow state, and flow density, any classification of these Elwha plumes either as hyperpycnal flows, or as turbidity currents, or as sandy debris flows is problematic. Despite the uncertain nature of flow types, an important lesson learned from the Elwha sediment plume is that external factors are critical in redirecting sediment transport. The deflection of Elwha plume to the east (Fig. 5.5) could be attributed to tidal currents in this estuarine environment (Cannon, 1978; Thomson et al., 2007; Warrick et al., 2011). Also, summer upwelling winds move easterly into the Strait of Juan de Fuca (Fig. 5.5B). Such summer winds could also explain deflecting sediment plume to the east of the Elwha River mouth (Fig. 5.4C). The reason is that winds reach a maximum speed of 8 m s21 off Vancouver Island with increasing magnitudes eastward in the Strait of Juan de Fuca (Foreman et al., 2008).

5.4.2 Connecticut River, New England region, United States, Long Island Sound, United States The Connecticut River is the longest river in the New England region of the United States. It flows roughly southward for 653 km through four states. It originates at the U.S. border with Quebec, Canada, and empties into Long Island Sound. After Hurricane Irene drenched New England with rainfall in late August 21 30, 2011, the Connecticut River was spewing muddy sediment into Long Island Sound.

Mass Transport, Gravity Flows, and Bottom Currents

190

5. Density plumes: types, deflections, and external controls

FIGURE 5.4 Sediment plume triggered by Elwha Dam demolition in the State of Washington (United States). (A) Index map showing Elwha Dam (arrow). The 108-foot dam, built in 1910 and demolished in 2012, is located approximately 7.9 km upstream from the river mouth. (B) Aerial photograph of the Olympic Peninsula and the Strait of Juan de Fuca. Filled yellow circle 5 Elwha River mouth. (C) Elwha sediment plume triggered by the demolition of Elwha Dam in 2012. Red arrow shows easterly deflecting plume, away from the Pacific Ocean. This deflection could be attributed to tidal currents in this estuarine environment. Also, the Strait of Juan de Fuca is subjected to easterly upwelling winds (see Fig. 5.5). Aerial photo was taken on March 30, 2012. (D) Aerial photo of Elwha River mouth showing absence of sediment plume in 2019 (compare with C). Aerial photo was taken on February 28, 2019. Source: (A) Credit: U.S. Geological Survey Public Domain map. (B) From Duda, J.J., Warrick, J.A., Magirl, C.S., 2011. Coastal and lower Elwha River, Washington, prior to dam removal—history, status, and defining characteristics. In: Duda, J.J., Warrick, J.A., Magirl, C.S. (Eds.), Coastal Habitats of the Elwha River: Biological and Physical Patterns and Processes Prior to Dam Removal. U.S. Geological Sand Survey Scientific Investigations Report 2011-5120. Chapter 1, with additional labels by G. Shanmugam. (C) Photo credit: Tom Roorda. (D) Photo courtesy of Tom Roorda, Roorda Aerial, Port Angeles, WA.

A satellite image showed that the Connecticut River entering the Long Island Sound generates deflected lobate plumes (Fig. 5.6A). The image was acquired on September 2, 2011, 2 days after a storm dissipated in this area. However, the storm became extratropical cyclone on August 28 and lingered on for a few days. Therefore, the cause of plume deflection could be the posthurricane winds associated with the extratropical cyclone.

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

191

FIGURE 5.5 External control of sediment plumes in the Strait of Juan de Fuca. (A) MERIS (MEdium Resolution Imaging Spectrometer) satellite image showing oceanographic setting of the Strait of Juan de Fuca. Filled yellow circle 5 Elwha River mouth (Satellite image courtesy of the European Space Agency). (B) Average summer upwelling winds, which move easterly in the strait could explain deflecting plumes observed at the Elwha River mouth (see Fig. 5.4C). According to Foreman et al. (2008), winds reach a maximum speed of 8 m s 1 off Vancouver Island with increasing magnitudes eastward in the Strait of Juan de Fuca. Filled yellow circle 5 Elwha River mouth. Source: From Foreman, M. G. G., Callendar, W., MacFadyen, A., Hickey, B.M., Thomson, R.E., di Lorenzo, E., 2008. Modeling the generation of the Juan de Fuca Eddy. J. Geophys. Res. Oceans 113 (C3), CiteID C03006, with additional labels by G. Shanmugam.

5.4.3 Eel River, California, Pacific Ocean, United States The Eel River is about 315 km long in northern California where it empties into the Pacific Ocean. A satellite image of the Eel River shows a southerly deflection of sediment plume (Fig. 5.6B). Imran and Syvitski (2000) studied the Northern California Margin near the mouth of the Eel River and suggested that hyperpycnal flows may be influenced by the along-shelf currents and be deflected northward. Geyer et al. (2000) reported both southerly and northerly winds in the area, and thus, the southerly deflection of plume shown in the NASA image (Fig. 5.6B) can be attributed to wind forcing.

5.4.4 Mississippi River, Gulf of Mexico, United States The Mississippi River is the second-largest drainage system on the North American continent, second only to the Hudson Bay drainage system. Its source is Lake Itasca in northern Minnesota and it flows generally south for 2320 mi (3730 km) to the Mississippi

Mass Transport, Gravity Flows, and Bottom Currents

192

5. Density plumes: types, deflections, and external controls

FIGURE

5.6 (A) Satellite image showing the Connecticut River entering the Long Island Sound with a deflected lobate plume. Image acquired on September 2, 2011. (B) Satellite image showing the Eel River in California with a deflected lobate plume. Image acquired on December 9, 2012. Source: (A) NASA Earth Observatory image by Robert Simmon. (B) NASA image courtesy Jeff Schmaltz, LANCE MODIS Rapid Response. Caption by Adam Voiland. Additional labels and symbols by G. Shanmugam.

River Delta in the Gulf of Mexico. A convincing example of deflecting sediment plume is revealed by the Mississippi River Delta in a recent NASA satellite image (Fig. 5.7). These deflecting (Fig. 5.7, yellow arrow) sediment plumes (i.e., hyperpycnal plumes) away from the shelf edge due to external factors, such as wind forcing and shelf currents (Walker and Rouse, 1993), are clear evidence that hyperpycnal flows do not transport sediment across the shelf into deep-water environments.

5.4.5 Rı´o de la Plata Estuary, South Atlantic Ocean, Argentina, and Uruguay The Rı´o de la Plata Estuary is located on the east coast of South America, bordering Argentina and Uruguay. It is 280 km long and 220 km wide at its mouth, and its water depth does not exceed 10 m (Fig. 5.8B). It receives water and sediment from both the Parana´

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

193

FIGURE 5.7 Deflecting sediment plumes associated with the Mississippi River Delta in the U.S. Gulf of Mexico. (A) Index map showing position of the Mississippi River Delta (Box). (B) NASA satellite image of the Mississippi River (United States) showing deflecting (yellow arrow) sediment plumes (i.e., hyperpycnal plumes) away from the shelf edge due to external factors such as wind forcing and shelf currents (Walker and Rouse, 1993). Image acquired on March 4, 2018. Source: (A) Image credit: NASA. (B) Image credit: NASA Earth Observatory image by Joshua Stevens, using MODIS data from LANCE/EOSDIS Rapid Response. Additional labels and interpretation by G. Shanmugam.

and Uruguay rivers with an annual mean discharge of 22,000 m3 s21. Satellite images show dissipating in the north and deflecting in the south plume with an irregular front (Fig. 5.8C). The dissipating and deflecting functions of the plume can be attributed to ocean currents that are active at the mouth of the estuary in the South Atlantic (Gonzalez-Silvera et al., 2006; Matano et al., 2010; Shanmugam, 2018a).

Mass Transport, Gravity Flows, and Bottom Currents

194

5. Density plumes: types, deflections, and external controls

FIGURE 5.8 (A) Index map of the Rı´o de la Plata Estuary, a location of the Rı´o de la Plata Estuary (white circle). (B) Satellite image showing the Rı´o de la Plata Estuary. (C) Satellite image showing the Rı´o de la Plata ˇ Estuary with hyperpycnal plumes that tend to deflect toward the Argentinian shelf to the south. Framinan and Brown (1996) used the term “turbidity front” for this hyperpycnal plume. Note that the entire, 220-km wide, plume gets diluted and dissipated with an irregular front, which fails to advance into the South Atlantic. This dilution of plume is attributed to external controls, such as ocean currents operating on the shelf. The Parana´ River, the second longest river in South America after the Amazon, supplies three-quarters of the freshwater that enters the estuary, with the remainder arriving from the Uruguay River. See Fossati and Piedra-Cueva (2013). Source: (A) Image credit: ETOPO1 Global Relief Model, C. Amante and B.W. Eakins, ETOPO1 Arc-Minute Global Relief Model: Procedures, Data Sources and Analysis, NOAA Technical Memorandum NESDIS NGDC-24, March 2009. (B and C) Figure from Shanmugam (2018b).

5.4.6 Rhone River, Gulf of Lions, Mediterranean Sea, France The Rhone River is one of the major rivers of Europe. It originates in the Rhone Glacier in the Swiss Alps and empties into the Mediterranean Sea. In understanding river mouth plume-dispersion patterns in the Mediterranean Sea, Arnau et al. (2004) have utilized satellite imagery products, including various types of thermal and visible images [advanced very high 2 resolution radiometer (AVHRR), sea-viewing wide field-of-view sensor (SeaWiFS), and moderate resolution imaging spectroradiometer (MODIS)]. These images were used to describe plume-formation events, their association with coastal oceanography, and their dispersal in the northwestern Mediterranean Sea. At this location, two of the largest Mediterranean rivers (Rhone and Ebro) open into this virtually land-locked sea. Arnau et al. (2004) discussed whether flood events in the study area, as conditioned by riverine, oceanographic, and climatic factors. In this study, we are concerned with the

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

195 FIGURE 5.9 Rivers flowing into the northern Mediterranean Sea. (A) Satellite image showing deflected plume of the Rhone River. (B) Satellite image showing deflected plume of the Ebro River. Source: From Arnau, P., Liquete, C. Canals, M., 2004. River mouth plume events and their dispersal in the northwestern Mediterranean Sea. Oceanography 17 (3), 23 31, with additional labels by G. Shanmugam.

spectacular images of the Rhone River and its deflecting plume (Fig. 5.9A). Ocean currents are considered an important factor in deflecting these plumes.

5.4.7 Ebro Delta, Mediterranean Sea, Iberian Peninsula The Ebro River is the second longest river in the Iberian Peninsula after the Tagus. Arnau et al. (2004) discussed various aspects of Ebro River. A deflecting plume at the mouth of the Ebro River is striking (Fig. 5.9B). Cyclonic events and ocean currents are considered an important factor in deflecting the Ebro River plume.

Mass Transport, Gravity Flows, and Bottom Currents

196

5. Density plumes: types, deflections, and external controls

FIGURE 5.10 (A) Location map of the Gulf of Ca´diz (red filled circle). (B) Circulation patterns of ocean currents in the Gulf of Ca´diz (Peliz et al. 2009). MO 5 Mediterranean outflow; GCC 5 Gulf of Ca´diz slope current. (C) Satellite image showing sediment plumes with a U-turn pattern (white arrow). Note that the U-turn pattern is mimicking the circulation of ocean currents (B). White open circle 5 Guadalquivir River mouth. Source: NASA Additional symbols and labels all by G. Shanmugam.

5.4.8 Guadalquivir River, Gulf of Ca´diz, Southern Spain The Guadalquivir River is a major river in the Iberian Peninsula with its entire length of 657 km in Spain. It empties into the Gulf of Cadiz to the south. The Gulf of Ca´diz is located in the northeastern Atlantic Ocean (Fig. 5.10A). It is enclosed by the southern Iberian and northern Moroccan margins, west of the Gibraltar Strait. Two major rivers, the Guadalquivir and the Guadiana, as well as smaller rivers, such as the Odiel, the Tinto, and the Guadalete, reach the ocean here. In terms of ocean currents (Peliz et al., 2009), it is one of the most complex oceanographic settings (Fig. 5.10B). Mimicking the current patterns, sediments that are emptied into the gulf by the Guadalquivir River exhibit a U-Turn shape for the plume (Fig. 5.10C). In cases like this, one must consider the influence of ocean currents on the dispersal of hyperpycnite sediments. The problem is that how these hyperpycnite sediments would differ from those hyperpycnites unaffected by ocean currents. In other words, do plume configurations (i.e., U-Turn vs. lobate) matter in the depositional record? No one has addressed this issue. However, the fact that sediment plumes are deflected by external factors, such as ocean currents, is not in dispute.

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

197

FIGURE 5.11 Deflecting sediment plumes associated with the Tiber River, Tyrrhenian Sea. (A) Index map of Tiber River Delta near Rome, Italy. (B) The Copernicus Sentinel-2B satellite true-color image showing deflecting sediment plume at the mouth of the Tiber River. Note northwest-trending plume (arrow) controlled by northwesterly wind (Manca et al., 2014) and by northwesterly flowing longshore currents at the river mouth (Mikhailova et al., 1998). (C) Rose diagrams showing the velocity and the direction of prevailing wind (left panel) and the maximum velocity and direction of gusts (right panel) in the Giglio Island, which is located northwest of the Tiber delta. Measurements were made every 10 minutes at the weather station of Giglio Porto during the study period 2012 13. Giglio Porto is located 185 km northwest of Rome. Note that wind is trending from SSE to NNW direction (compare with northwest-trending plume direction in B). Source: (A) Map modified after Wikipedia. (B) Copernicus Sentinel data (2019), processed by ESA, CC BY-SA 3.0 IGO. Image captured on February 5, 2019. (C) From Cutroneo, L., Ferretti, G., Scafidi, D., Ardizzone, G. D., Vagge, G., Capello, M., 2017. Current observations from a looking down vertical V-ADCP: interaction with winds and tide? The case of Giglio Island (Tyrrhenian Sea, Italy). Oceanologia 59 (2), 139 152, with permission from Elsevier.

5.4.9 Tiber River, Tyrrhenian Sea, Italy The Tiber River originates in the Apennine Mountains in Emilia-Romagna and flowing 406 km (252 mi) through Tuscany, Umbria, and Lazio, and empties into the Tyrrhenian Sea, near the city of Rome, Italy. A satellite image of sediment plumes associated with the Tiber River in Italy also shows deflected plumes due to northwesterly wind and longshore currents along the Tyrrhenian coast (Fig. 5.11). The northwest-trending plume (arrow in Fig. 5.11B) controlled by northwesterly wind (Manca et al., 2014) and by northwesterly

Mass Transport, Gravity Flows, and Bottom Currents

198

5. Density plumes: types, deflections, and external controls

FIGURE 5.12 (A) Index map showing Gulf of Corinth, Greece. (B) Deflected plume associated with Mornos River. (C) Deflected plume associated with Fonissa River. Source: From Beckers, A, Beck, C., Hubert-Ferrari, A., Tripsanas, E., Crouzet, C., Sakellariou, D., Papatheodorou. G., de Batist, M. 2016. Influence of bottom currents on the sedimentary processes at the western tip of the Gulf of Corinth, Greece. Mar. Geol. 378, 312 332, with additional labels by G. Shanmugam.

flowing longshore currents at the river mouth (Mikhailova et al., 1998). Such trends are normal, considering that this setting is a wave-dominated delta (Milliff et al., 2013).

5.4.10 Mornos and Fonissa Rivers, Gulf of Corinth, Greece The Gulf of Corinth is a 120 km long, up to 30 km wide, and 867 m deep-water body connected to the Ionian Sea, in Greece (Fig. 5.12A). The Gulf is connected at its western tip to the Mediterranean Sea through three shallow sills. Beckers et al. (2016) discussed the influence of bottom currents on deflecting sediment plumes in two rivers, namely the south-flowing Mornos River (Fig. 5.12B) and the north-flowing Fonissa River (Fig. 5.12C) into the Gulf of Corinth. Twelve cores from 0.4 to 2.2 m long were retrieved in 2011 and 2014 with UWITEC and BENTOS gravity corers. The cores are located at various depths and various distances from the Rion straits. These cores indicate that drifts are composed of homogenous bioturbated mud in their upper part. This core study is one of the rare cases in which the authors attempt to relate deflection of sediment plumes at river mouths to bottom currents and to link the shallow-water drift deposit to bottom currents. Types of bottom currents in deep-water environments, including contour currents, are discussed by Shanmugam (2016b).

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

199

5.4.11 Congo (Zaire) River, Atlantic Ocean, West Africa The Congo (Zaire) River is the second longest (4700 km in length) river in Africa, next to Nile. A Landsat 8 image collected on March 2, 2015, by NASA shows a distinct northerly deflection of sediment plume at the river mouth (Fig. 5.13). This setting is highly influenced by tidal currents (Shanmugam, 2003a). In a numerical modeling of sediment plumes at Congo River mouth, Denamiel et al. (2013) considered tides, wind stress, surface heat flux, and ocean boundary conditions. Hopkins et al. (2013) traced Congo plumes for hundreds of kilometers and attributed plumes’ deflection to winds and the Angola Current.

5.4.12 Yellow River, Bohai Bay, China The Yellow River is the second longest river in China, after the Yangtze River, and at the estimated length of 5464 km. It originates in the Bayan Har Mountains in Qinghai

FIGURE 5.13 Deflected plume associated with Congo (Zaire) River. West Africa. Source: Landsat 8 image was collected on March 2, 2015 by NASA. Additional labels and symbols by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

200

5. Density plumes: types, deflections, and external controls

province of Western China and flows through nine provinces before emptying into the Bohai Bay to the east. It is regarded as the world’s largest contributor of fluvial sediment load to the ocean (Yu et al., 2011). This river has developed both lobate and horse’s tail plumes (Fig. 5.2D and E). Wang et al. (2010) documented the position of the tidal front about 5 km seaward off the Yellow River mouth and explained the tide-induced density flows on the shelf (Wang et al., 2010). The importance of these numerical experiments is that the topography with a strong slope off the Yellow River mouth was a determining factor on the generation of a shear front. The sedimentologic implication of the shear front is that it limits seaward transport of sediments (Li et al., 2001; Wang et al., 2010). If so, the extent of sediment transport into the deep sea by hyperpycnal flows comes into question. In other words, the entire concept of hyperpycnal flows transporting sediment into the deep sea (Mulder et al., 2003; Steel et al., 2016; Warrick et al., 2013; Zavala and Arcuri, 2016) is unsupported by the Yellow River, which is considered to be a classic river for hyperpycnal flows.

5.4.13 Yangtze River, East China Sea, China The Yangtze River is the longest river (about 6300 km) in Asia. Satellite images show that the Yangtze River generates both hyperpycnal and deflected hypopycnal plumes (Fig. 5.14A). The Yangtze River mouth is a complex setting in which both ocean currents and tidal currents are affecting sediment dispersal. Unlike the Yellow River that enters a protected Bohai Bay from major ocean currents, the Yangtze River enters the East China Sea affected by the warm, north-flowing Kuroshio Current (Fig. 5.14B). As a consequence, muddy sediments brought by the Yangtze River are redistributed and deposited as a mud belt on the inner shelf (Wu et al., 2016). This mud belt is evident on the satellite images (Fig. 5.14A).

5.4.14 Pearl River, South China Sea, China The Pearl River system is China’s third-longest river, 2400 km (1500 mi) in length, after the Yangtze River and the Yellow River. Satellite sea surface temperature ( C) in the northern South China Sea on July 9, 2009 (Fig. 5.15A) and on July 24, 2011 (Fig. 5.15B) shows the trend of sediment plumes associated with the Pearl River. Chen et al. (2017) proposed a model for upwelling water and pathway of the Pearl River plume in the northern South China Sea (Fig. 5.15C). There is clear evidence for deflecting flume and that the current direction is parallel to plume direction. Satellite images show that there was a belt of turbid water appearing along an upwelling front near the Chinese coast of Guangdong and indicate that the turbid water of the Pearl River plume water could be transported to a far-reaching area east of the Taiwan Bank. Numerical modeling results are consistent with the satellite observations and reveal that a strong jet exists at the upwelling front with a speed as high as 0.8 m s21, which acts as a pathway for transporting the high-turbidity plume water. The dynamical analysis of Chen et al. (2017) suggests that geostrophic equilibrium dominates in the upwelling front and plume areas, and the baroclinicity of the upwelling front resulting from the horizontal density gradient is

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

201

FIGURE 5.14 Data from the Yangtze River, China. (A) Satellite image showing the Yangtze River plunging into the East China Sea. Note development of both hyperpycnal plume (yellow color due to high sediment concentration) near the river mouth and hypopycnal plume (blue color due to low sediment concentration) on the seaward side. Note deflected hypopycnal flows that move southward (white arrow), possibly due to modulation by south-flowing shelf currents. In a recent study, Luo et al. (2017) recognized that extended and deflected density plumes (white arrow) tend to develop during winter months, which are absent during the summer months. Note sheet-like mud belt developed along the inner shelf due to contour-following shelf currents. White dashed circle 5 Yangtze River mouth. (B) Map showing warm Kuroshio Current (KC) in the East China Sea and Yellow Sea. TWC 5 Taiwan Warm Current; YSWC 5 Yellow Sea Warm Current; ZFCC 5 Zhejiang 2 Fujian Coastal Current; JCC 5 Jiangsu Coastal Current. Blue circles: Yangtze and Yellow River mouths. (C) Conceptual model of sedimentary and oceanographic processes affecting the sediment dispersal at both subaqueous river mouth and alongshore deposits associated with the Yangtze River. Blue circle 5 Yangtze River mouth. Source: (A) NASA. (B and C) From Liu, J.P., Li, A.C. Xu, K.H. Velozzi, D.M. Yang, Z.S. Milliman, J.D. DeMaster. D.J., 2006. Sedimentary features of the Yangtze River-derived along-shelf clinoform deposit in the East China Sea. Cont. Shelf Res. 26, 2141 2156, with additional labels by G. Shanmugam. With permission from Elsevier.

responsible for the generation of the strong jet, which enhances the far-reaching transport of the terrigenous nutrient-rich water of the Pearl River plume. Model sensitivity analyses also confirm that this jet persists as long as the upwelling front exists, even when the wind subsides and becomes insignificant.

Mass Transport, Gravity Flows, and Bottom Currents

202

5. Density plumes: types, deflections, and external controls

FIGURE 5.15 Data from the Pearl River, China. Satellite sea surface temperature (SST) ( C) in the northern South China Sea on July 9, 2009 (A) and on July 24, 2011 (B) showing the Pearl River mouth and associated plumes. (C) Schematics of upwelling water and pathway of the Pearl River plume in the northern South China Sea. Note that the deflecting flume (red arrow) direction is parallel to the South China Sea Warm Current (blue arrow) direction. Contours show the bathymetry in meters. Source: Compiled from Chen, Z., Pan, J. Jiang, Y. Lin, H., 2017. Far-reaching transport of Pearl River plume water by upwelling jet in the northeastern South China Sea. J. Mar. Syst. 173, 60 69, with additional color labels by G. Shanmugam.

5.4.15 Krishna-Godavari Rivers, Bay of Bengal, India Both Krishna and Godavari Rivers originate in the Western Ghats and flow across the Deccan Plateau and empty their sediments into the Bay of Bengal. Well-developed deflecting plumes are evident in NASA images of the Godavari River mouth (Fig. 5.16). The Krishna-Godavari Basin is a tide-influenced setting (Shanmugam et al. (2009). In this setting, monsoonal currents play an important role in southerly deflection of sediment plume (Fig. 5.16). Sridhar et al. (2008) studied the influence of seasonal geostrophic currents on sediment plumes in the Krishna-Godavari Basin. Based on the data from Indian remote sensing satellite, Oceansat-1, carries ocean color monitor (OCM) sensor, Sridhar et al. (2008) documented the suspended sediment concentrations during 1999 2006 and illustrated a unique plume off Krishna-Godavari River Basin (Fig. 5.17). Though high sediment concentration is present all along the east coast of India, the offshoot of the plume is present only at Krishna-Godavari Basin. The presence, extent, orientation, and intensity of this plume have both seasonal and interannual variations. Sridhar et al. (2008) superimposed the geostrophic currents over the OCM observations and documented a convincing influence of geostrophic currents on the deflection of sediment plumes (Fig. 5.17).

Mass Transport, Gravity Flows, and Bottom Currents

5.4 Deflected sediment plumes and their control

203

FIGURE 5.16 Data from the Krishna-Godavari Rivers, India. (A) Index map of India showing location of the Krishna-Godavari (KG) Basin and northeast monsoonal currents. (B) Image showing southerly deflecting plume at the mouth of the Godavari River. Source: (A) Modified after Jagadeesan, L., Jyothibabu, R., Anjusha, A., Mohan, A.P., Madhu, N.V., Muraleedharan, K.R., Sudheesh, K., 2013. Ocean currents structuring the mesozooplankton in the Gulf of Mannar and the Palk Bay, southeast coast of India. Prog. Oceanogr. 110, 27 48. (B) Image credit: NASA. Additional labels and symbols by G. Shanmugam.

5.4.16 Brisbane River, Moreton Bay, Australia The Brisbane River is located on the east coast of Australia (Fig. 5.18A). A storm on May 1, 2015, dropped more than 360 mm (14 inches) of rain within about 3 hours in southeast Queensland. As a result of the rainfall, flash flooding caused distinct river plumes to form along the coastline. On May 3, after the storm had passed, the operational land imager on Landsat 8 of NASA acquired a good view of a deflected plume from the Brisbane River entering Moreton Bay (Fig. 5.18B). In this case, the deflection was caused by the anthropogenic structure of the Port of Brisbane (Fig. 5.18B).

5.4.17 Dart River, Lake Wakatipu, South Island, New Zealand The Dart River originates from the Dart Glacier in the heart of the Southern Alps in the South Island of New Zealand (Fig. 5.19A). A photograph taken from a helicopter showing a well-developed braid delta with deflecting density plumes (i.e., hyperpycnal plumes) at the mouths of the Dart and Rees Rivers flowing into Lake Wakatipu at Glenorchy near

Mass Transport, Gravity Flows, and Bottom Currents

204

5. Density plumes: types, deflections, and external controls

FIGURE 5.17 Sediment plumes and geostrophic currents in the Krishna-Godavari Basin. Source: From Sridhar, P. N., Ali, M. M., Vethamony, P., Babu, M. T., Ramana, V., Jayakumar, S., 2008. Seasonal occurrence of unique sediment plume in the Bay of Bengal. Eos Trans. 89 (3), 22 23, with additional labels by G. Shanmugam.

FIGURE 5.18 Data from the Brisbane River, Australia. (A) Index map of Australia showing location of city of Brisbane. (B) Satellite image showing deflection of the Brisbane River Plume. Note the location of Port of Brisbane and its influence on the plume direction. Source: NASA. Additional labels and symbols by G. Shanmugam.

Queenstown, South Island, New Zealand (Fig. 5.19B). High gradients of this setting are typical of coarse-grained braid deltas (McPherson et al., 1987). The cause of plume deflection is unclear, although the area is subjected to tidal influence (Heath, 1975).

Mass Transport, Gravity Flows, and Bottom Currents

5.5 Global significance of wind forcing on sediment plumes

205

FIGURE 5.19 Data from the Dart River, New Zealand. (A) Index map showing New Zealand and position of Dart braid delta in the South Island of New Zealand. (B) Aerial photograph showing Southern Alps and related fluvial setting. Note steep gradient that is typical of braid deltas (McPherson, Shanmugam, and Moiola, 1987). (C) Aerial photograph taken from a helicopter showing a well-developed braid delta with deflecting density plumes (i.e., hyperpycnal plumes) at the mouths of the Dart and Rees Rivers flowing into Lake Wakatipu at Glenorchy near Queenstown, South Island, New Zealand. The Dart River originates from the Dart Glacier in the heart of the Southern Alps to the north (i.e., left of image). Approximate width of braid delta in the image is 1.5 km. Source: Photo by John G. McPherson, Melbourne, Australia. Additional labels and symbols by G. Shanmugam.

5.5 Global significance of wind forcing on sediment plumes A review of sediment plumes suggests that there are 22 external controls (Fig. 5.20). Although there are 22 external factors, wind forcing is the most significant worldwide (Table 5.1). Wind forcing refers to wind tress exerted by the wind on bodies of water. Wind forcing is the umbrella term for various wind-related phenomena, such as wind waves, longshore currents, cyclonic currents, monsoonal currents, upwelling currents, and seiche. Milliff et al. (2004, their Fig. 5) discussed aspects of wind stress curl and wind stress divergence in illustrating the storm tracks of the Northern Hemisphere (e.g., 35 50 N, emanating from western boundaries in both ocean basins), storm tracks in the Southern Hemisphere [e.g., 35 50 S, the intertropical convergence zone in the eastern tropical Pacific (0 10 N)], and tropical cyclone regions of the Indian and western Pacific Oceans. Although tropical cyclones occur worldwide, they tend to concentrate on certain

Mass Transport, Gravity Flows, and Bottom Currents

206

5. Density plumes: types, deflections, and external controls

FIGURE 5.20

Summary diagram showing 22 external controls. Source: Updated after Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 238.

key locations, namely the Gulf of Mexico and the Bay of Bengal (Fig. 5.21). Kao et al. (2010) discussed injection of freshwater hyperpycnal flows into the deep sea by cyclones in the subtropics. In the context of this chapter, the Hurricane Floyd (Fig. 5.22) generated 100-km wide sediment plumes on the U.S. Atlantic Margin (Fig. 5.23). An understanding of wind forcing is critical on deflecting sediment plumes in the world’s oceans at various depths. This issue can be demonstrated using empirical data from the Gulf of Mexico, which is an ideal location to study wind forcing. For example, the Loop Current in the Gulf of Mexico is a wind-driven current system (Mullins et al., 1980). Velocities in eddies that have detached from the Loop Current have been recorded as high as 200 cm s21 at a depth of 100 m (Cooper et al., 1990). The Loop Current and related eddies pose significant problems for deep-water drilling (Koch et al., 1991). For example, drilling operations in the Green Canyon 166 area were temporarily suspended in August of 1989 because of high-current velocities that reached nearly 150 cm s21 at a depth of 45 m and 50 cm s21 at a depth of 250 m. These intense bottom currents affect the ability of a drilling rig to hold station over a wellhead (Koch et al., 1991). Currentvelocity measurements, bottom photographs, high-resolution seismic records, and GLORIA side-scan sonar records indicate that the Loop Current influences the seafloor at least periodically in the Gulf of Mexico (Pequegnat, 1972). Computed flow velocities of the Loop Current vary from nearly 100 cm s21 at the sea surface to more than 25 s21 at 500 m water depth (Nowlin and Hubert, 1972). This high surface velocity suggests a

Mass Transport, Gravity Flows, and Bottom Currents

5.6 Implications for sediment transport

207

FIGURE 5.21 Map showing the tracks of all tropical cyclones, which formed worldwide during the period 1985 2005. The points show the locations of the cyclones at six-hourly intervals. The color scheme represents tropical depression, tropical storm, and the Saffir Simpson Hurricane Scale of 1 5. Note high concentration of cyclones in the Gulf of Mexico and Bay of Bengal. The Saffir 2 Simpson Hurricane Scale: Category 1: 120 153 km h21; Category 2: 154 177 km h21; Category 3: 178 209 km h21; Category 4: 210 249 km h21; Category 5: . 249 km h21. Source: From Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration And Production, vol. 9. Elsevier, Amsterdam, p. 524, based on data from NASA.

wind-driven origin for these currents. Flow velocities measured using a current meter reach up to 19 cm s21 at a depth of 3286 m (Pequegnat, 1972). Kenyon et al. (2002b) reported 25 cm s21 current velocity measured 25 m above the seafloor. Such currents are capable of reworking fine-grained sand on the seafloor. Current ripples, composed of sand at a depth of 3091 m on the seafloor (Pequegnat, 1972), are the clear evidence of deep bottom-current activity in the Gulf of Mexico today (Pequegnat, 1972). Therefore, wind forcing is a powerful agent in deflecting sediment plumes at various depths varying from 10s of meters to 1000s of meters.

5.6 Implications for sediment transport A summary diagram illustrates the differences between two depositional systems, namely the normal sediment transport versus the deflected sediment transport (Fig. 5.24). In the normal mode, the following conventional concepts are applicable:

Mass Transport, Gravity Flows, and Bottom Currents

208

5. Density plumes: types, deflections, and external controls

FIGURE 5.22 The NOAA Geostationary Operational Environmental Satellite (GOES) image showing Hurricane Floyd off the Florida coast on September 14, 1999. Floyd was nearly 933 km (580 mi) across with hurricane-force winds extending 200 km (125 mi) from the hurricane’s eye and sustained winds up to 224 km hr21 (140 mi hr21, category 4). Scale bar is approximate because the distance between degrees of longitude ranges between 111.32 km (69.17 mi) at 0_ (the equator) to 1.95 km (1.21 mi) at 89_ latitude (DOP, 2007). FL5Florida. Image credit: NASA (2007). See text for explanation of geography. Source: After Shanmugam, G., 2008a. The constructive functions of tropical cyclones and tsunamis on deepwater sand deposition during sea level highstand: implications for petroleum exploration. AAPG Bull. 92, 443 471. Reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

• • • • • • • •

Normal downslope transport from source to sink is common. Paleocurrent directions are reliable. Depositional settings cover shelf, slope, and basin. Wind forcing and tidal currents are common in shelf environment. Mass transport and bottom currents are common in slope and basin. There is a general increase in grain size from sink to source. There is a reliable compositional trend from sink to source. There is a reliable inference to provenance.

Mass Transport, Gravity Flows, and Bottom Currents

5.6 Implications for sediment transport

209

FIGURE 5.23 Shelf-wide sediment plumes generated by Hurricane Floyd of 1999 on the U.S. Atlantic margin. (A) Satellite image showing calm shelf waters (dark blue) on a fair-weather day (April 5, 2000) along the FloridaGeorgia-South Carolina-North Carolina coast. Note the influx of suspended sediments and organic matter (yellowish brown) from four rivers into the Atlantic Ocean along the coast. Dashed line indicates approximate position of the shelf edge. (B) Satellite image showing shelf-wide sediment plume (cyan color) as Hurricane Floyd (storm weather) passed over these waters on September 16, 1999. Note that the turbid zone (i.e., sediment plume) is occupying the entire shelf width, which is approximately 100 km (62 mi). Three bent arrows show trends of sediment transport into the deep Atlantic Ocean. FL 5 Florida. Source: After Shanmugam, G., 2008a. The constructive functions of tropical cyclones and tsunamis on deepwater sand deposition during sea level highstand: implications for petroleum exploration. AAPG Bull. 92, 443 471. Reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

• At river mouths, lobate plumes (deltas, Fig. 5.2E) develop with predictable sand distribution. In basinal environments, submarine fans tend to develop at canyon mouths with predictable sand distribution. In the deflected mode, conventional concepts do not apply. For example (Fig. 5.24): • A major shift occurs in sediment-transport direction occurs at river mouths. • Deflected sediment plumes at river mouths are potential candidates for being classified as “hyperpycnal flows.” • Sediment accumulation tends to occur in shelf environments, close to the shoreline. • Sediment distribution occurs on only one side of the river mouth. • These sediments would be conventionally classified as “hyperpycnites.”

Mass Transport, Gravity Flows, and Bottom Currents

210

5. Density plumes: types, deflections, and external controls

FIGURE 5.24 Summary diagram showing the difference in sediment transport between normal mode and deflected mode with corresponding implications for paleocurrents and provenance. Compare with case study of the Elwha plume in the Strait of Juan de Fuca (Fig. 5.4C). After Shanmugam, G., (2019d). Global significance of wind forcing on deflecting sediment plumes at river mouths: Implications for hyperpycnal flows, sediment transport, and provenance. Journal of the Indian Association of Sedimentologists, Vol. 36, No. 2, 1 37.

• • • •

Wind, tidal, and longshore currents are dominant. Sediment-transport directions could vary with processes (e.g., bidirectional tidal currents). Transport processes may vary with external control, such as cyclones and upwelling. Current reworking is common and therefore, traction structures are also common in sediment. • Current reworking may increase reservoir quality. • Because current directions are complex, it is unreliable to infer provenance accurately. • Unlike normal sediment transport systems, sand distribution is quite different in deflected systems. For example, deflected sediment may develop tongue-like geometry (Fig. 5.4C), whereas normal mode develops lobate geometry at river mouths (Fig. 5.2E). Such a difference is important in petroleum exploration. In the case of the deflected mode, the sand abruptly ends at river mouths, and there is a total absence of sand on the western side of the river mouth in the Elwha River (Fig. 5.4C).

5.7 Implications for provenance Aspects of sediment provenance have been documented in thematic edited volumes (e.g., Zuffa, 1985; Mazumder, 2016). Principal data used in provenance studies are (1) pale current directions, (2) grain size, (3) sediment composition, (4) diagenetic alteration, (5) stratigraphic framework, (6) depositional setting, and (7) tectonic deformation. Commonly, primary sedimentary structures and related current directions are used in deciphering sediment provenance (Pettijohn, 1975; Potter and Pettijohn, 1977; Zuffa, 1985). However, complex current directions associated with deep-water bottom currents pose

Mass Transport, Gravity Flows, and Bottom Currents

5.8 Synopsis

211

immense challenges in inferring the primary sediment source (Shanmugam, 2016b). This challenge is equally acute in shelf environments where sediment transport is diverted from the normal downslope mode due to external factors, such as wind forcing (Fig. 5.4C).

5.8 Synopsis • Empirical data based on satellite images and aerial photographs show that over 50% of the cases studied (i.e., 18 out of 29 cases) have been subjected to varying degrees of deflection of sediment transport, away from the normal course in a downslope direction, by external controls. These controls include a plethora of oceanographic, meteorological, and anthropogenic phenomena. However, wind forcing is the most dominant external control. Failure to establish external controls on sediment plumes could result in erroneous depositional models in terms of paleocurrents, paleogeography, and provenance. • There are many unresolved issues in understanding the deposits of sediment plumes in the ancient record (Shanmugam, 2018b). For example, no one has established sedimentological properties of deflected plumes in the sedimentary record, modern or ancient. Perhaps, a starting point is to recover sediment cores from modern examples, such as the Elwha sediment plume (Fig. 5.4C), when they develop and examine sediment cores in great detail, including, texture, structure, mineralogy, current direction related features, etc., as a modern analog of deflected plumes. • Most importantly, these empirical data defy the popular wisdom that hyperpycnal flows transport sand into the deep sea from river mouths (Chapter 6: Hyperpycnal Flows).

Mass Transport, Gravity Flows, and Bottom Currents

C H A P T E R

6 Hyperpycnal flows O U T L I N E 6.1 Definition

214

6.2 Origin

214

6.3 Identification

214

6.4 Hyperpycnites and related issues 6.4.1 The incentive 6.4.2 The history 6.4.3 The hyperpycnite problem 6.4.4 The objective

215 216 216 217 218

6.5 Basic concepts 6.5.1 Hyperpycnite 6.5.2 Continental margin 6.5.3 Plunge point 6.5.4 Plume versus flow 6.5.5 Types of river-mouth flows 6.5.6 River currents versus turbidity currents 6.5.7 Transformation of river currents into turbidity currents 6.5.8 Fine-grained deltas versus coarse-grained deltas

218 224 224 224 224 225

6.6 The Yellow River, China: a case study 6.6.1 Delta versus estuary 6.6.2 Bathymetry 6.6.3 River-mouth processes 6.6.4 Bottom-turbid layers 6.6.5 Multilayer hyperpycnal flows Mass Transport, Gravity Flows, and Bottom Currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00006-0

226 228 228 229 231 233 233 233 234

213

6.6.6 Tide-modulated hyperpycnal flows 234 6.6.7 Internal waves 234 6.6.8 Velocity measurements 234 6.6.9 Tidal shear front 235 6.6.10 M2 tidal dynamics in Bohai and Yellow Seas 236 6.7 The Yangtze River, China: a case study 237 6.7.1 Hyperpycnal and hypopycnal plumes 237 6.7.2 Ocean currents 237 6.7.3 Tidal river dynamics 237 6.8 External controls

239

6.9 Recognition of ancient hyperpycnites 6.9.1 The hyperpycnite facies model 6.9.2 Inverse to normal grading 6.9.3 Internal erosional surface 6.9.4 Traction structures 6.9.5 Massive sandstones 6.9.6 Lofting rhythmites 6.9.7 Plant remains 6.9.8 Hyperpycnite fan models 6.9.9 Flawed principles 6.9.10 Grain size 6.9.11 Modern analogs 6.9.12 Submarine canyons

241 242 243 244 244 245 245 246 246 248 249 249 250

© 2021 Elsevier Inc. All rights reserved.

214

6. Hyperpycnal flows

6.10 Cyclone-induced hyperpycnal turbidity currents in canyons 253 6.11 Configurations of density plumes

6.12.6 Ring plume: South Pacific Ocean 261 6.12.7 Tendril plume: Lake Michigan, United States 261 6.12.8 Swirly plume: Lake Erie, United States 262

254

6.12 Global case studies 255 6.12.1 Dissipating plume with irregular front: the Rı´o de la Plata Estuary, Argentina and Uruguay, South Atlantic Ocean 256 6.12.2 Coalescing lobate plume: Zambezi River, Indian Ocean 260 6.12.3 Tidal lobate plume: San Francisco Bay, Pacific Ocean 260 6.12.4 Swirly cyclonic plume: Tropical Storm Ida, Northern Gulf of Mexico 261 6.12.5 Whitings plume: the Great Bahama Bank, North Atlantic Ocean 261

6.13 Challenges

262

6.14 Future research directions

262

6.15 Academic discussions 263 6.15.1 Stages of scientific development 263 6.15.2 Hyperpycnites and the remaining unresolved issues 268 6.16 Synopsis

270

6.1 Definition In advocating a rational theory for delta formation, Bates (1953) suggested three types of river-mouth sediment plumes: (1) hypopycnal plume for floating river water that has lower density than basin water (Fig. 5.2A), (2) homopycnal plume for mixing river water that has equal density as basin water (Fig. 5.2B), and (3) hyperpycnal plume for sinking river water that has higher density than basin water (Fig. 5.2C). Although Middleton and Hampton (1973) did not classify hyperpycnal flows as a sediment-gravity flow in their classification, these flows are indeed sediment-gravity flows. Sediment density plays a vital role in initiating hyperpycnal flows. An example of density of hyperpycnal flow (ρ) is 0.025 g cm23 (Wright and Nittrouer, 1995). Hyperpycnites represent deposits of hyperpycnal flows.

6.2 Origin Hyperpycnal flows originate from river floods at the plunge point near the shoreline (Fig. 6.1).

6.3 Identification • Facies model (Mulder et al., 2003) • inverse to normal grading • internal erosion surface Mass Transport, Gravity Flows, and Bottom Currents

6.4 Hyperpycnites and related issues

215

FIGURE 6.1 Continental margin and flume experiments. (A) Conceptual diagram of a continental margin showing relative positions of plunge point (red filled circle) at river mouth and submarine fan at base-of-slope. Average shelf width 5 80 km. Maximum shelf width 5 1500 km. (B) Schematic diagram, based on flume experiments conducted using freshwater as standing body, showing transformation of river current into turbidity current at plunge point (red filled circle). Note that this experiment using freshwater is applicable to freshwater lakes, but not to marine settings (sea or ocean). (C) Schematic diagram with backwater zone showing transformation of river plume into turbidity currents at plunge point (red filled circle). Note the close similarity between (B) and (C) on the initiation of turbidity currents at plunge point. In this study, the term "hyperpycnal flow" is used for flows seaward of the plunge point, instead of turbidity current (see text). Source: (B) From Kostic, S., Parker, G., Marr, J.G., 2002. Role of turbidity currents in setting the foreset slope of clinoforms prograding into standing fresh water. J. Sediment. Res. 72 (3), 353 362, with additional labels by G. Shanmugam. (C) From Lamb, M.P., McElroy, B., Kopriva, B., Shaw, J., Mohrig, D., 2010. Linking river-flood dynamics to hyperpycnal-plume deposits: experiments, theory, and geological implications. GSA Bull. 122, 1389 1400, with additional symbols by G. Shanmugam. From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 2 238. Springer, Open Access.

Various problems associated with these criteria are discussed by Shanmugam (2018b) (Section 6.9 below).

6.4 Hyperpycnites and related issues Shanmugam (2018b) has identified inherent problems associated with hyperpycnites in deepwater environments. This paper has resulted in two comments (Zavala, 2019; Van Loon et al., 2019), a “Words of the Editor-in-Chief—some ideas about the comments and discussions of

Mass Transport, Gravity Flows, and Bottom Currents

216

6. Hyperpycnal flows

hyperpycnal flows and hyperpycnites” (Feng, 2019), and my combined reply (Shanmugam, 2019c). My paper also led Prof. Z.-Z. Feng to organize a thematic session titled “Definitions and names of over density flow (hyperpycnal flow) and identification markers of deposits of ancient over density flows (hyperpycnites)” during the Fourth International Conference of Palaeogeography held in Beijing, China (September 20 22, 2019). In addressing the hyperpycnite problem, the following sections are organized along the lines discussed in two recent papers (Shanmugam, 2018b,c).

6.4.1 The incentive The term “hyperpycnite” (i.e., deposits of hyperpycnal flows) was first introduced by Mulder et al. (2002) in an academic debate with me (Shanmugam, 2002b) on the origin of inverse grading by hyperpycnal flows. The following year, Mulder et al. (2003) published their review paper with the introduction of the genetic facies model of hyperpycnites. I have been an ardent critic of all genetic facies models since the 1980s. Examples of my article titles are: 1. “Is the turbidite facies association scheme valid for interpreting ancient submarine fan environment?” (Shanmugam et al., 1985a), 2. “High-density turbidity currents: Are they sandy debris flows?” (Shanmugam, 1996a), 3. “The Bouma Sequence and the turbidite mind set” (Shanmugam, 1997a), 4. “The tsunamite problem” (Shanmugam, 2006b), 5. “The obsolescence of deep-water sequence stratigraphy in petroleum geology” (Shanmugam, 2007), 6. “The landslide problem” (Shanmugam, 2015a), 7. “Submarine fans: A critical retrospective (1950 2015)” (Shanmugam, 2016a), 8. “The contourite problem” (Shanmugam, 2016b), 9. “The seismite problem” (Shanmugam, 2016c), and 10. “The hyperpycnite problem” (Shanmugam, 2018b). Additional incentives for publishing multiple articles on hyperpycnal flows are explained in Chapter 10.

6.4.2 The history Forel (1885, 1892) first reported the phenomenon of density plumes in the Lake Geneva (Loc Le´man), Switzerland. Later, Bates (1953) suggested three types, namely (1) hypopycnal plume, (2) homopycnal plume, and (3) hyperpycnal plume. Mulder et al. (2003) expanded the applicability of the concept of hyperpycnal plumes from shallow water (deltaic) to deep-water (continental slope and abyssal plain) environments. In this new development, hyperpycnal flows are considered analogous to turbidity currents in many respects (Mulder et al., 2003; Steel et al., 2016; Zavala and Arcuri, 2016). During the past four decades, there has been an accelerated effort to understand these density plumes through (1) observational and/or interpretational (Arnau et al., 2004; Bhattacharya and MacEachern, 2009; Collins et al., 2017; Gihm and Hwang, 2016; Johnson et al., 2001; Lewis et al., 2018; Luo et al., 2017; Milliman et al., 2007; Mulder

Mass Transport, Gravity Flows, and Bottom Currents

6.4 Hyperpycnites and related issues

217

et al., 2003; Mutti et al., 1996; Ogston et al., 2000; Pan et al., 2017; Petter and Steel, 2006; Pierce, 2012; Puig et al., 2014; Schillereff et al., 2014; Shanmugam, 2018b,c; Soyinka and Slatt, 2008; Steel et al., 2016, 2018; Sun et al., 2016; Talling, 2014; Warrick et al., 2013; Wilson and Schieber, 2014, 2017; Wright et al., 1986, 1988; Yang et al., 2017a; Zavala and Arcuri, 2016; Zavala and Pan, 2018; Zavala et al., 2006; among others), (2) experimental (Kostic and Parker, 2003; Kostic et al., 2002; Lamb and Mohrig, 2009; Lamb et al., 2010; Parsons et al., 2001b), and (3) numerical (Chen et al., 2013; Kassem and Imran, 2001; Khan et al., 2005; Kostic and Parker, 2003; Morales de Luna et al., 2017; Qiao et al., 2008; Wang and Wang, 2010; Wang et al., 2017; among others) studies.

6.4.3 The hyperpycnite problem Despite popular claims that (1) river flows transform into turbidity currents at plunge points near the shoreline (Kostic et al., 2002; Lamb et al., 2010), (2) hyperpycnal flows have the power to erode the seafloor and cause submarine canyons (Lamb et al., 2010), (3) hyperpycnal flows develop a unique vertical sequence (i.e., facies model) (Mulder et al., 2003), and (4) hyperpycnal flows are efficient in transporting sand across the shelf and can deliver sediments into the deep sea for developing submarine fans (Zavala and Arcuri, 2016), our understanding of hyperpycnal flows and their deposits, in particular, in deepwater settings (i.e., seaward of the shelf-slope break at about 200 m water depth, Fig. 6.1), is highly speculative. Specific issues are: 1. There is not a single documented case of hyperpycnal flow, which is transporting sand across the continental shelf, and supplying sand beyond the modern shelf break (Fig. 6.1). 2. Thus far, the emphasis has been solely on river-mouth hyperpycnal flows (Mulder et al., 2003), thus ignoring density plumes in other environments, such as open marine settings, far away from the shoreline. 3. Despite their common occurrence, density plumes triggered by tidal currents, glacial meltwater, eolian dust, volcanic explosion, cyclones, tsunamis, upwelling, etc., are largely ignored from sedimentological investigations. 4. Specifically, there are fundamental problems associated with the concept of hyperpycnal flows in terms of fluid dynamics, depositional mechanisms, sedimentary structures, etc., which generated a lively debate (Mulder et al., 2002; Shanmugam, 2002b). 5. Finally, hyperpycnite facies models have implications for the petroleum industry for predicting sandy reservoirs in deep-water petroleum exploration and exploitation. For example, Yang et al. (2017a, p. 115) in their article published in the AAPG Bulletin stated that “The lacustrine hyperpycnites of the Yanchang Formation have important implications for unconventional petroleum exploitation.” Shanmugam (2019b) discussed this study in terms of inherent problems with data, documentation, and facies model.

Mass Transport, Gravity Flows, and Bottom Currents

218

6. Hyperpycnal flows

6.4.4 The objective In addressing the problems, the primary purpose of this chapter is to rigorously evaluate the merits of various claims on hyperpycnal flows and related facies models. This evaluation, which began nearly 30 years ago (Shanmugam, 2002b), is based on 45 case studies (Fig. 6.2; Table 6.1). Each case study is used in identifying problem areas. In particular, the Yellow River in China is used as the prime example because of its historical significance (Milliman and Meade, 1983) and its data-rich environments (Wright et al., 1986). This discussion broadly covers the following topics: (1) basic concepts, (2) the Yellow River, (3) the Yangtze River, (4) external controls, (5) recognition of ancient hyperpycnites, (6) submarine fans, (7) submarine canyons, and (8) configurations of density plumes. The ultimate goal here is to identify problem areas and to alert students of challenges in their future research and to identify opportunities for future research.

6.5 Basic concepts In this review, which covers multiple disciplines (e.g., process sedimentology, physical oceanography, meteorology, hydraulic engineering, etc.), it is necessary to establish at the outset some basic concepts and related nomenclatures.

FIGURE 6.2 Location map of 45 case studies of marine and lacustrine environments (Table 6.1). Forel (1885, 1892) first reported the phenomenon of density plumes in the Lake Geneva, which is considered as the birthplace of concepts related to density plumes. Source: Modified after Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 2 238. Springer, Open Access.

Mass Transport, Gravity Flows, and Bottom Currents

TABLE 6.1 Case studies showing various configurations of density plumes on satellite images in modern marine and lacustrine environments. Serial number Case study and (Fig. 6.2) location

Configuration of density plumes on satellite images

Environment

External control

Comments

1

Mississippi River, Gulf Deflecting (Fig. 5.2D) of Mexico (Walker and Rouse, 1993)

River-dominated delta

Shelf currents

Interpretation of a specific type of plume in the ancient record is impractical at present

2

Yellow River, China, Bohai Bay

Simple lobate, associated with a single river mouth (Old river mouth, 1995) (Fig. 5.2E)

River-dominated delta

Tidal shear front (Wang et al., 2010)

Interpretation of a specific type of plume in the ancient record is impractical at present

3

Yellow River, China, Bohai Bay

Horse’s tail caused by River-dominated deflecting plume (Modern delta river mouth, 1999) (Fig. 5.2F)

Tidal shear front (Wang et al., 2010)

Interpretation of a specific type of plume in the ancient record is impractical at present

4

Yangtze River, China, East China Sea

Deflecting (Fig. 5.14A)

Shelf currents (Liu et al., 2006)

Interpretation of a specific type of plume in the ancient record is impractical at present

Tide-dominated estuary

Vertical mixing by tides in winter months (Luo et al., 2017) 5

Chignik River, Alaska, Linear (Fig. 6.5) Pacific Ocean

Braid delta in a lagoon, Pacific Ocean

Coarse-grained braid delta (McPherson et al., 1987)

Interpretation of a specific type of plume in the ancient record is impractical at present

6

Copper River, Gulf of Alaska

Coalescing irregular, associated with multiple river mouths

Braid delta, marine

Coarse-grained braid delta

Interpretation of a specific type of plume in the ancient record is impractical at present

7

Copper River, Gulf of Alaska

Blanketing eolian dust plume

Braid delta, marine

Eolian

Interpretation of a specific type of plume in the ancient record is impractical at present

8

Mackenzie River Delta, Canada, Beaufort Sea

Swirly

River-dominated delta

Arctic ocean currents

Interpretation of a specific type of plume in the ancient record is impractical at present (Continued)

TABLE 6.1 (Continued) Serial number Case study and (Fig. 6.2) location

Configuration of density plumes on satellite images

Environment

External control

Comments

Rupert Bay, Quebec

Swirly

Bay

Mixing of river and seawater combined with churn of tides

Interpretation of a specific type of plume in the ancient record is impractical at present

10

Connecticut River, New England region, United States

Deflecting and lobate (Fig. 5.6A)

Long Island Sound

Cyclone (Hurricane Irene, August 21 30, 2011)

Interpretation of a specific type of plume in the ancient record is impractical at present

11

Eel River, California, United States

Deflecting (Fig. 5.6B)

Shelf

Shelf currents (Imran and Syvitski, 2000)

Interpretation of a specific type of plume in the ancient record is impractical at present

12

Golden Gate Bridge, San Francisco Bay, Pacific Ocean

Tidal lobate (Fig. 6.21C)

Bay mouth

Tidal currents (Barnard et al., 2006)

Interpretation of a specific type of plume in the ancient record is impractical at present

13

Amazon River, Brazil, Equatorial Atlantic

Deflecting

Open marine

Ocean currents and phytoplankton

Interpretation of a specific type of plume in the ancient record is impractical at present

14

Rio de la Plata Estuary, Argentina and Uruguay, South Atlantic Ocean

Dissipating (Fig. 5.8C)

Marine

Ocean currents (GonzalezSilvera et al., 2006; Matano et al., 2010)

Interpretation of a specific type of plume in the ancient record is impractical at present

15

Rhone Delta, France, Gulf of Lions, Mediterranean Sea (Arnau et al., 2004)

Deflecting (Fig. 5.9A)

River-dominated delta

Ocean currents

Interpretation of a specific type of plume in the ancient record is impractical at present

16

Ebro Delta, Iberian Peninsula, Mediterranean Sea (Arnau et al., 2004)

Deflecting (Fig. 5.9B)

River-dominated delta

Cyclonic events and ocean currents

Interpretation of a specific type of plume in the ancient record is impractical at present

17

Guadalquivir River, Southern Spain, Gulf of Ca´diz

U-Turn (Fig. 5.10C)

River-dominated delta

Surface and slope currents (Peliz et al., 2009)

Interpretation of a specific type of plume in the ancient record is impractical at present

9

18A

Strait of Gibraltar, Mediterranean Sea

Swirly plume

Strait mouth

Ocean water moving through Interpretation of a specific type of the strait and forming internal plume in the ancient record is waves (Shanmugam, 2013a) impractical at present

18B

Strait of Gibraltar, Mediterranean Sea

Internal waves

Strait mouth

Same as above

19

Otsuchi Bay, located on the Sanriku Coast, Iwate Prefecture, Japan

Complex sediment mixing Bay and plumes

Internal bore (Masunaga et al., 2015)

20

Hugli River (a distributary of the Ganges River), India, Bay of Bengal

Anastomosing

Tide-dominated estuary (Balasubramanian and Ajmal Khan, 2002)

Tidal currents

21

Gulf of Mannar, India and Sri Lanka, Indian Ocean

Massive and swirly

Marine

Monsoonal currents Interpretation of a specific type of (Jagadeesan et al., 2013), wave plume in the ancient record is actions (Sridhar et al., 2008) impractical at present

22

Zambezi River, Central Mozambique, Indian Ocean

Coalescing lobate, associated with multiple river mouths (Fig. 6.22B)

Wave-dominated delta

Longshore currents (Mikhailov et al., 2015)

Interpretation of a specific type of plume in the ancient record is impractical at present

23

Betsiboka River, NW Madagascar

Massive

Bay

Tides

Interpretation of a specific type of plume in the ancient record is impractical at present

24

Onibe River, Eastern Madagascar

Dissipating with sharp front

Marine

Cyclone (Giovanna, February 7 24, 2012) and ocean currents

Interpretation of a specific type of plume in the ancient record is impractical at present

25

Niger River, West Africa

Linear

Wave-dominated delta

Wave currents

Interpretation of a specific type of plume in the ancient record is impractical at present

26

Congo (Zaire) River, West Africa

Lobate

Marine

Tidal currents (Shanmugam, 2003a)

Interpretation of a specific type of plume in the ancient record is impractical at present

Conceptual model

Interpretation of a specific type of plume in the ancient record is impractical at present

(Continued)

TABLE 6.1 (Continued) Serial number Case study and (Fig. 6.2) location

Configuration of density plumes on satellite images

27

Nile Delta, Egypt, Mediterranean Sea

28

Environment

External control

Comments

Lobate

River-dominated delta

Currents

Interpretation of a specific type of plume in the ancient record is impractical at present

Off Namibia, South Atlantic

Cloudy

Marine

Upwelling (plankton) (Shillington et al., 1992)

Interpretation of a specific type of plume in the ancient record is impractical at present

29

Off Namibia, South Atlantic

Swirly

Marine

Upwelling (hydrogen sulfide)

Interpretation of a specific type of plume in the ancient record is impractical at present

30

U.S. Atlantic shelf

Cascading (Shanmugam, 2008a)

Shelf (Marine)

1999 Hurricane Floyda

Interpretation of a specific type of plume in the ancient record is impractical at present

31

Northern Gulf of Mexico

Swirly (Fig. 6.20B)

Shelf (Marine)

2009 Tropical Storm Idaa

Interpretation of a specific type of plume in the ancient record is impractical at present

32

Kalutara Beach, Sri Lanka, Arabian Sea

Backwash (Shanmugam, 2006b)

Marine

2004 Indian Ocean Tsunamia

Interpretation of a specific type of plume in the ancient record is impractical at present

33

Greenland, Labrador Sea

Meltwater

Marine

Subglacial, meltwater (Chu, 2014). See also Cuffey and Paterson (2010)

Interpretation of a specific type of plume in the ancient record is impractical at present

34

Thwaites Glacier,

Meltwater

Marine

Subglacial, meltwater

Interpretation of a specific type of plume in the ancient record is impractical at present

Whitings (Fig. 6.34B)

Marine

Fish activities (Broecker et al., Interpretation of a specific type of 2000), wind (Dierssen et al., plume in the ancient record is 2009), Florida Current (Purkis impractical at present et al., 2017)

Western Antarctica 35

The Great Bahama Bank, Atlantic Ocean

36

Tagula Island, South Pacific Ocean

Ring (Fig. 6.34C)

Marine

Coral reef (Khanna and Yadav, 2008)

Interpretation of a specific type of plume in the ancient record is impractical at present

37

Egypt, Red Sea

Dust

Marine

Eolian

Interpretation of a specific type of plume in the ancient record is impractical at present

38

Yucatan Peninsula, Southern Gulf of Mexico

Feathery

Marine

Complex mix of sediment and plankton

Interpretation of a specific type of plume in the ancient record is impractical at present

39

Paluweh Volcano, Indonesia, Indian Ocean

Volcanic ash

Marine

Volcanic

Interpretation of volcanic ash plume is possible

40

Bogoslof Island, Bering Sea

Volcanic ash

Marine

Volcanic (Shipley and SarnaWojcicki, 1982)

Interpretation of volcanic ash plume is possible

41

Pribilof Islands Bering Sea

Swirly

Marine

Phytoplankton

Interpretation of a specific type of plume in the ancient record is impractical at present

42

Entire Globe

Numerous

Marine

Chlorophyll

Interpretation of a specific type of plume in the ancient record is impractical at present

43

Carolina Continental Rise, North Atlantic

Gas hydrate (Paull et al., 1995). See also Ruppel and Kessler (2017)

Marine

Pockmarked seafloor associated with active chemosynthetic biological communities

Interpretation of Gas hydrate plume is possible

44

Lake Michigan, United Tendril (Fig. 6.24B) States

Lacustrine

Eolian

Interpretation of a specific type of plume in the ancient record is impractical at present

45

Lake Erie, United States

Lacustrine

Seiche (de Jong and Battjes, 2004)

Interpretation of a specific type of plume in the ancient record is impractical at present

a

Swirly (Fig. 6.24B)

Transport of gravel, sand, and mud to deep-water environments by cyclone- and tsunami-related flows is possible. See also Shanmugam (2018b,c).

224

6. Hyperpycnal flows

6.5.1 Hyperpycnite As mentioned at the outset, the term “hyperpycnite” was introduced by Mulder et al. (2002) in an academic debate with me (Shanmugam, 2002b) on the origin of inverse grading by hyperpycnal flows. Mulder et al. (2002) attempted to differentiate “hyperpycnites” deposited by hyperpycnal turbidity currents from “classic turbidites” deposited from failure-related turbidity currents. The problem is that triggering mechanisms of turbidity currents (or any other process) cannot be determined from the depositional record (Shanmugam, 2015a, 2016a,b,c).

6.5.2 Continental margin A basic conceptual framework is used in which a river mouth is located near the shoreline, whereas a submarine fan is located at the base of the continental slope, separated by a wide continental shelf (Fig. 6.1A). In order for river plumes to act as hyperpycnal flows and deliver sediment to the deep sea for developing submarine fans (Zavala and Arcuri, 2016), hyperpycnal flows must travel 10 100s of kilometers across the shelf from their point of origin.

6.5.3 Plunge point The term “plunge point” is used for both “plunging waves” and “plunging rivers.” According to the Glossary of Coastal Terminology (1998), a plunging wave as is defined as the point at which the wave curls over and falls. According to Assireu et al. (2011), for plunging rivers, the plunge point is the main mixing point between river and epilimnetic reservoir. In other words, it is the point at which sediment-laden river flow plunges down into a standing body of water, be it a lake, a reservoir, or a sea. Plunging occurs very close to the shoreline in shallow water (Fig. 6.1B). In the Yellow River in China, for example, the plunge point occurs at 5 m of depth in the Bohai Bay (Wright et al., 1986). When a river flow crosses the plunge point at the river mouth, it transforms into a river plume of various densities, which include hyperpycnal plumes (Fig. 5.2C). At the plunge point, the river flow moves from a momentum-dominated type to a buoyancy-dominated type and marks the transition of an inflow to an underflow (Dallimore et al., 2004). At the plunge point, the river has already dropped its coarse fractions (gravel and sand) upstream as delta-plain facies. The remaining fine fractions in muddy suspension move forward on the open shelf as hyperpycnal flows. Plunging would occur only if suspended sediment concentration (SSC) in the river exceeds the critical value of 35 45 kg m23 (Imran and Syvitski, 2000; see Mulder et al., 2003 for differences in values between equatorial and subpolar rivers).

6.5.4 Plume versus flow In practice, there is a tendency to equate the term “flow” with “plume.” These two terms are not one and not the same. In hydrodynamics, the term plume describes a condition when a column of one fluid moves through another fluid. To accommodate natural

Mass Transport, Gravity Flows, and Bottom Currents

6.5 Basic concepts

225

variability in plume types, a broad definition of plume is adopted in this article. Accordingly, a plume is a fluid enriched in sediment, ash, biological or chemical matter that enters another fluid. As it would be demonstrated later, there is a multitude of plume types in nature. Among them, the river plume is the most popular. NOAA Fisheries Glossary (2006, p. 42) defines a River Plume as “Turbid freshwater flowing from land and generally in the distal part of a river (mouth) outside the bounds of an estuary or river channel.” However, the term “flow” is used for a continuous, irreversible deformation of sediment-water mixture that occurs in response to applied shear stress, which is gravity in most cases (Pierson and Costa, 1987, p. 2). Not all plumes are flows. For example, floating hypopycnal plumes are not driven by gravity (Fig. 5.2A). However, both terms “flow” and “plume” are applicable to hyperpycnal type. The other practice is to employ terms “overflow,” “interflow,” and “underflow” for hypopycnal, homopycnal, and hyperpycnal plumes, respectively. Again, the term flow is not appropriate for hypopycnal plume that is unaffected by gravity.

6.5.5 Types of river-mouth flows In discussing river-mouth processes, geologists, geophysicists, and hydraulic engineers use process terms to represent hyperpycnal flows that are not consistent in meaning with each other, such as single-layer and multilayer hyperpycnal flows (Fig. 6.3). For example, the following concepts and terms are used in the literature: 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15.

density flow (Parker and Toniolo, 2007), underflow (Wright et al., 1986), hyperpycnal flow (Bates, 1953; Moore, 1966), hyperpycnal underflow (Wright et al., 1986), hyperconcentrated flow (van Maren et al., 2009), low-density hyperpycnal plume (Wright et al., 1986), high-density hyperpycnal plume (Wright et al., 1986), high-turbid mass flow (Fan et al., 2006), supercritical hyperpycnal flow (Yang et al., 2017b), tide-modulated hyperpycnal flow (Fig. 6.3D) (Wang et al., 2010), cyclone-induced hyperpycnal turbidity current (Liu et al., 2012), buoyancy-dominated flow (Dallimore et al., 2004), hyperpycnal turbidity current (Plink-Bjo¨rklund and Steel, 2004), ˇ turbidity front (Framinan and Brown, 1996), turbidity current (Kostic and Parker, 2003; Kostic et al., 2002; Lamb et al., 2010; Wright et al., 1986; Zavala and Arcuri, 2016), and 16. multilayer hyperpycnal flows (Morales de Luna et al., 2017). These 16 river-mouth processes, some with superfluous meanings, do not have a unifying principle of fluid dynamics as their foundation. It is confusing when geologists manufacture a plethora of superfluous names for a single process. In this tradition, the concept

Mass Transport, Gravity Flows, and Bottom Currents

226

6. Hyperpycnal flows

FIGURE 6.3 Variable types of hyperpycnal flows. (A) Single-layer hyperpycnal flow, Yellow River, China. Color concentration 5 suspended sediment concentration; H 5 flow thickness; τ t 5 upper surface; τ b 5 bed shear stress (Gao et al., 2015). (B) Bottom turbid layer with density and velocity stratification (i.e., debris flow with hydroplaning, red arrow added in this article, see text) Yellow River, China. Uw 5 wave orbital velocity; Uc 5 along shelf current magnitude; Ug 5 velocity of gravity current; NW/W 5 normal wind-induced wave velocity; TIW 5 typhoon-induced wave. The red line represents the downslope variation trend of the BTL (Gao et al., 2015). Additional labels by G. Shanmugam. (C) Multilayer hyperpycnal flow in numerical modeling (de Luna et al., 2017). Note that multilayer numerical modeling was also applied to hypopycnal flows. h 5 height of a fluid layer; u 5 velocity; Φ 5 particle concentration; p 5 density. See Morales de Luna et al. (2017) for details of various parameters and related equations. (D) Tide-modulated hyperpycnal flow, Yellow River (Wang et al., 2010), modified after Wright et al. (1988). Additional labels by G. Shanmugam. Note internal waves. Internal waves occur only along pycnoclines (Shanmugam, 2013a), but there is no indication of pycnoclines in this diagram. Source: From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 2 238. Springer, Open Access.

of “high-density turbidity currents” is the leader with 34 synonymous terms (Shanmugam, 2006a).

6.5.6 River currents versus turbidity currents The practice of equating subaqueous turbidity currents with subaerial river currents (Chikita, 1989) is confusing for many reasons (Table 6.2). River currents and turbidity

Mass Transport, Gravity Flows, and Bottom Currents

227

6.5 Basic concepts

TABLE 6.2 Comparison of subaerial river currents and subaqueous turbidity currents (partly based on Shanmugam, 1997a). Features

River currents

Turbidity currents

Ambient fluid

Air

Water

Rheology of fluid

Newtonian

Newtonian

Type of gravity influence

Fluid gravity

Sediment gravity

Nature of flow

Uniform, steady, and continuous

Nonuniform, unsteady, and episodic

Sediment concentration

Low (1 5 vol.%)

High (1 23 vol.%)

Dominant transport

Bedload

Suspended load of sand

Dominant structures

Cross-bedding

Normally graded bedding

currents are fundamentally different, although both are turbulent in state (Middleton, 1993). River currents are low in suspended sediment (1% 5% by volume; Galay, 1987), whereas turbidity currents (i.e., low-density turbidity currents) are relatively high in suspended sediment (1% 23% by volume; Middleton, 1993), although both currents are considered to be Newtonian in rheology (Table 6.2). River currents are fluid-gravity flows, whereas turbidity currents are sediment-gravity flows (Middleton, 1993), which is the most important distinction. To reiterate, a turbidity current is a sediment flow with Newtonian theology and turbulent state in which sediment is supported by fluid turbulence and from which deposition occurs through suspension settling (Dott, 1963; Middleton and Hampton, 1973; Sanders, 1965; Shanmugam, 1996a, 2006a; Talling et al., 2012). In addition, according to Bagnold (1962), typical turbidity currents can function as truly turbulent suspensions only when their sediment concentration by volume is below 9%. Therefore river currents should not be equated with turbidity currents. In the 1930s, density currents (Daly, 1936) and turbidity currents were considered to be one and the same. Since then, the domain of turbidity currents went through a remarkable period of revolution and evolution (Shanmugam, 2000a, 2016a). After 80 years of research, we have come full circle. Today, we once again consider density currents and turbidity currents to be one and the same. For example, Parker and Toniolo (2007, p. 690) defined a turbidity current as follows: “When the density difference is mediated by the presence of suspended mud in the water column of the river, the resulting underflow is termed a turbidity current.” However, the distinction is that all turbidity currents are density currents, but not all density currents are turbidity currents (e.g., thermohaline-density driven bottom currents or “contour currents” [Hollister, 1967]). It should be reiterated that all hyperpycnal plumes are density plumes, but not all density plumes are hyperpycnal plumes (e.g., hypopycnal and homopycnal plumes). This unnecessary confusion can be easily avoided by simply adhering to the established concepts available in sedimentologic literature (Bagnold, 1962; Dott, 1963; Middleton and Hampton, 1973; Sanders, 1965; Shanmugam, 1996a, 2006a; Talling et al., 2012).

Mass Transport, Gravity Flows, and Bottom Currents

228

6. Hyperpycnal flows

The first step in evaluating density plumes is to distinguish a “plume” from a “flow” and to differentiate a “river current” from a “turbidity current” (Table 6.2).

6.5.7 Transformation of river currents into turbidity currents Based on experimental (Kostic et al., 2002) and numerical simulation, Kostic and Parker (2003) suggested that river currents transform into turbidity currents at the plunge point (Fig. 6.1B). Because Kostic et al. (2002) used freshwater in their experiment as a standing body of water, care must be exercised in applying the experimental results (i.e., initiation of turbidity currents at the plunge point) to marine settings (sea or ocean), which is the focus of this chapter. There are concerns with the experimental/numerical model. For example: 1. Average density of seawater at the surface is 1.025 kg L21, whereas that of freshwater is 1.0 kg L21 at 4 C (39 F). This density difference is crucial for understanding the generation of a density flow, such as the hyperpycnal flow. 2. No one has documented the transformation of river currents into turbidity currents at a shallow plunge point in modern marine environments. 3. These river flow triggered turbidity currents in laboratory experiments, yet to be documented in modern marine settings, are muddy flows. Therefore they are of no consequence in transporting sand and gravel across the continental shelf and deliver the sediment into the deep sea for developing submarine fans. 4. Importantly, not all density flows are turbidity currents. For example, although both debris flows and turbidity currents are considered to be density flows, each one can be distinguished from the other by fluid rheology and flow turbulence (Dott, 1963; Sanders, 1965). Such a distinction is not considered in defining hyperpycnal flows. Hyperpycnal flows are defined solely on the basis of fluid density. Therefore it is misleading to equate turbidity currents with hyperpycnal flows (Kostic et al., 2002; Lamb et al., 2010; Steel et al., 2016; Zavala and Arcuri, 2016). 5. Lamb et al. (2010) applied the numerical model in their experimental model for hyperpycnal flows with emphasis on marine environments (Fig. 6.1C). It is worth noting that Lamb et al. (2010) also used fresh tap water in their experiments for standing body of water. Therefore their experimental results on hyperpycnal flows are applicable only to freshwater lakes, but not to marine bodies of water (sea or ocean). In order for the experimental/numerical model to be applicable to marine settings, the model needs to be tested in the real world by documenting the transformation of river currents into turbidity currents in marine settings such as the Yellow River in China (Wright et al., 1986) that plunges into the Bohai Bay (see Section 6.6).

6.5.8 Fine-grained deltas versus coarse-grained deltas In the geologic and engineering literature, the focus of discussion on hyperpycnal flows is centered on fine-grained deltas or common deltas. McPherson et al. (1987) distinguished fine-grained deltas from coarse-grained deltas (Fig. 6.4). The importance here is that braid

Mass Transport, Gravity Flows, and Bottom Currents

6.6 The Yellow River, China: a case study

229

FIGURE 6.4 A Comparison of fan-deltas, braid deltas, and fine-grained deltas based on distributary-channel patterns and stability, sediment load and size, stream gradient and velocity, and other properties. Fan-deltas and braid deltas are "coarse-grained" deltas that contrast in shape, size, and composition with "fine-grained" deltas. Source: From McPherson, J.G., Shanmugam, G., Moiola, R.J., 1987. Fan-deltas and braid deltas: varieties of coarse-grained deltas. GSA Bull. 99, 331 340.

(braided) deltas, kind of coarse-grained deltas, are typical of high-gradient settings with high-velocity river flows (Fig. 6.4). Because these braided rivers plunge into a standing body of water with multiple entry points, separated by braided bars, these rivers develop linear hyperpycnal plumes (Fig. 6.5). Distinguishing linear types is important because braid deltas are known to develop various types of sediment flows, including debris flows, in the subaqueous delta fronts (McPherson et al., 1987). At present, coarse-grained deltas are totally ignored in studying hyperpycnal flows. As a consequence, all published examples of hyperpycnal flows are from fine-grained deltas, such as the Yellow River delta in China.

6.6 The Yellow River, China: a case study The Yellow River, which is the second largest river in China, is regarded as the world’s largest contributor of fluvial sediment load to the ocean (Yu et al., 2011).

Mass Transport, Gravity Flows, and Bottom Currents

230

6. Hyperpycnal flows

FIGURE 6.5 Braid delta from Chignik Lake, southeastern coast, Alaska, showing multiple entry points with shooting “linear” hyperpycnal plumes into a standing body of water. Source: From McPherson, J.G., Shanmugam, G., Moiola, R.J., 1987. Fan-deltas and braid deltas: varieties of coarse-grained deltas. GSA Bull. 99, 331 340. Photo courtesy of R.D. Kreisa.

Historically, it contributed a sediment load of nearly 100 million tons per year (Milliman, 2001). The Yellow River’s average annual suspended-load concentration of 25,000 mg L21 and flood stage concentration of 220,000 mg L21 are the largest in the world by 1983 (Milliman and Meade, 1983). In September 1995, a cruise was undertaken to detect hyperpycnal flows off the Yellow River mouth (Wang et al., 2010). During the cruise (18 19 September), daily SSC was close to 50 kg m23 and daily average stream discharge was 2000 m3 s21. The critical concentration of suspended sediment ranges from 36 to 43 kg m23 for coastal waters depending on local salinity, temperature, and climatic conditions (Mulder and Syvitski, 1995). The Yellow River drains through that part of the world that is covered by extensive soft and easily erodible, wind-transported, loess deposits in China (Fig. 6.6A). The loess is intensively eroded during the monsoon rains, generating unusual suspended particle concentration at the Yellow River mouth, which generates hyperpycnal flows (Wright et al., 1986, 1990). Because the Yellow River is an ideal river for generating hyperpycnal flows, I focus attention on this river, which is rich in empirical data (Fig. 6.6).

Mass Transport, Gravity Flows, and Bottom Currents

6.6 The Yellow River, China: a case study

231

FIGURE 6.6 Data from the Yellow River. (A) The course of the Yellow River (Huanghe) draining the loess plateau before entering the Bohai Bay. Note river mouth is highlighted by blue open circle. Note Yangtze River mouth (blue circle) in the East China Sea. (B) Flow types and their nomenclature used for the Yellow River. (C) Data from the Yellow River. (D) Empirical and numerical data on tidal shear front and on tidal currents for the Yellow River. Source: (A) Image from Wang, Q., Guo, X.Y., Takeoka, H., 2008. Seasonal variations of the Yellow River plume in the Bohai Sea: a model study. J. Geophys. Res. Oceans 113, C08046, with additional labels by G. Shanmugam. (B 2 D) From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 2 238. Springer, Open Access.

6.6.1 Delta versus estuary One confusing aspect of the Yellow River literature is that some authors refer to the river mouth as a “delta” (Gao et al., 2014; Wang et al., 2017; among many others), whereas others refer to it as an “estuary” (e.g., Hu et al., 1998; Wang and Wang, 2010). The distinction between a delta and an estuary is not trivial (Dalrymple, 1992; Dalrymple et al., 1992; Shanmugam et al., 2000). The Yellow River cannot be both a delta and an estuary at the same time. According to the Oxford Dictionaries (2018), the term “estuary” is derived from a mid-16th century Latin word “aestuarium” meaning tidal part of a shore (“estus” 5 “tide”).

Mass Transport, Gravity Flows, and Bottom Currents

232

6. Hyperpycnal flows

Fairbridge (1980) defined an estuary as “an inlet of the sea reaching into a river valley as far as the upper limit of tidal rise.” Whether the Yellow River is a delta or an estuary is important here because estuaries are not ideal candidates for transporting hyperpycnite sediments offshore. This is because of ebb and flood tides and their bidirectional currents. In this study, the Yellow River is considered to be a river-dominated delta with tidal influence.

FIGURE 6.7 Maps showing changes in bathymetry of the Yellow River Delta through time. Note changes in shallowest (red) areas surrounding the river mouth within the circle. The old Yellow River course that existed during 1976 2 96 was abandoned and a modern river course was established since 1996 (Wang et al., 2015, their Fig. 6.5). Note a slight protrusion to the north in red area (arrow) for the year 1996 caused by the change in river course (see Fig. 5. 2E). Source: From Wang, N., Li, G.X., Qiao, L.L., Shi, J.H., Dong, P., Xu, J.S., et al., 2017. Long-term evolution in the location, propagation, and magnitude of the tidal shear front off the Yellow River mouth. Cont. Shelf Res. 137, 1 12, with additional labels by G. Shanmugam.

Mass Transport, Gravity Flows, and Bottom Currents

6.6 The Yellow River, China: a case study

233

6.6.2 Bathymetry Wang et al. (2017) obtained bathymetric data for the Yellow River and the western Laizhou Bay for the years 1976, 1980, 1985, 1991, 1996, and 2012, and presented maps with a spatial resolution of 300 500 m (Fig. 6.7). Maps show a clear change in bathymetry in front of the river mouth because of the change in river course. The change in river course from an abandoned south-flowing old river (1976 96) to the modern north-flowing river was illustrated by Wang et al. (2015, their Figs. 1 and 2E,F). Changes in river-mouth bathymetry are a reflection of changes in river courses and related types of sediment plumes.

6.6.3 River-mouth processes Wright et al. (1986) were the first authors to investigate hyperpycnal flows at the Yellow River mouth. Because the turbidite paradigm was in full force during the 1970s and 1980s, Wright et al. (1986) emphasized the similarity between hyperpycnal flows and turbidity currents. As discussed in Chapter 3, Gravity Flows: Debris Flows, Grain Flows, Liquefied/Fluidized Flows, Turbidity Currents, Hyperpycnal Flows, and Contour Currents, turbidity currents are defined on the basis of fluid rheology, flow state, and sediment concentration, whereas hyperpycnal flows are defined solely on fluid density. The following is a compilation of types of flows that have been used for the Yellow River. 1. 2. 3. 4. 5. 6. 7.

hyperpycnal underflow (Wright et al., 1986) low-density hyperpycnal plume (Wright et al., 1986) high-density hyperpycnal plume (Wright et al., 1986) tide-modulated hyperpycnal flow (Wang et al., 2010) hyperconcentrated flow (van Maren et al., 2009) high-turbid mass flow (Fan et al., 2006) turbidity current (Wright et al., 1986)

Clearly, there is no consistency in terms of fluid dynamics. From a practical point of view, none of these publications discusses the depositional characteristics of various types of hyperpycnal flows.

6.6.4 Bottom-turbid layers Wright et al. (2001) suggested the influence of ambient currents and waves on gravity-driven sediment flows, which are different from hyperpycnal flows. In this context, Gao et al. (2015) suggested that the Yellow River has undergone a regime shift in response to resuspension induced by tidal currents and waves. This shift has presumably resulted in the replacement of hyperpycnal flows by bottom-turbid layers. The difference between the two is that hyperpycnal flows behave as a single layer without vertical stratification in density or velocity (Fig. 6.3A), whereas bottom-turbid layers reveal vertical stratification in density and velocity (Fig. 6.3B). Such a vertical stratification in bottom-turbid layers is similar to the concept of “high-density turbidity current” (Postma et al., 1988, their Fig. 2). The problem here is that stratified highdensity turbidity currents are sandy debris flows because of their basal plastic layers induced by high sediment concentration (Shanmugam, 1996a). In support of a debris flow, Gao et al.

Mass Transport, Gravity Flows, and Bottom Currents

234

6. Hyperpycnal flows

(2015, their Fig. 5B) depicted a “detached point” where the flow front is lifted up from the seafloor (Fig. 6.3B, the red arrow). This phenomenon is identical to the experimental debris flow with a detached and lifted-up front due to “hydroplaning” (Mohrig et al., 1998, their Fig. 3; see also Shanmugam, 2000a, his Fig. 15). Although clearly implied, Gao et al. (2014, 2015) did not cite the pioneering work of Mohrig et al. (1998) on hydroplaning.

6.6.5 Multilayer hyperpycnal flows Morales de Luna et al. (2017) simulated numerically a multilayer model for hyperpycnal flows on theoretical/mathematical basis (Fig. 6.3C). By contrast, Gao et al. (2015, their Fig. 5A) considered hyperpycnal flows as a single-layer phenomenon on the empirical basis (Fig. 6.3A). The problem is that no one has ever documented multilayer hyperpycnal flows in natural environments. Another problem is that Morales de Luna et al. (2017) have applied the multilayer model to both hyperpycnal and hypopycnal plumes. Such applications of numerical modeling to both types of density plumes raise the question on the validity of numerical modeling when there are no empirical bases for the existence of multilayer hyperpycnal flows in nature. This numerical approach is akin to inventing medicine for a hypothetical disease that does not exist.

6.6.6 Tide-modulated hyperpycnal flows The term “tide-modulated hyperpycnal flow” (Fig. 6.3D; Wang et al., 2010) is confusing. The reason is that hyperpycnal flows are unidirectional (i.e., travel seaward), whereas tidal currents are bidirectional [i.e., travel both seaward (ebb tide) and landward (flood tide)]. In this scenario, it is incongruous to mix tidal currents with hyperpycnal flows in the same nomenclature. In maintaining clarity, any current generated by tides should be called a tidal current.

6.6.7 Internal waves Wang et al. (2010; see also Wright et al., 1986) suggested internal waves but did not provide empirical evidence for internal waves at the mouth of the Yellow River (Fig. 6.3D). Internal waves are a complicated oceanographic phenomenon (Shanmugam, 2013a). For example, internal waves occur only along pycnoclines (Shanmugam, 2013a), but there is no evidence of pycnoclines at shallow depths where hyperpycnal plumes develop in front of the Yellow River. In summary, publications on the river-mouth processes of the Yellow River have perpetuated unnecessary conceptual problems by proposing complex processes without empirical basis.

6.6.8 Velocity measurements In their study on the Yellow River, Wright et al. (1990) reported that strong (B1 m s21) parabathic tidal currents resuspended newly deposited muds and advected them alongshore. It appears that tidal currents are more powerful than hyperpycnal flows. The velocity values used in numerical modeling studies are from tidal currents (e.g., Wang et al., 2010). Disappointingly, there are no empirical data on velocity measurement of hyperpycnal flows from the Yellow River mouth (Wright et al., 1986).

Mass Transport, Gravity Flows, and Bottom Currents

6.6 The Yellow River, China: a case study

235 FIGURE 6.8 Tidal shear front. (A) Satellite image showing the sediment dispersal pattern at the Yellow River (Huanghe) mouth and estimated mean depth-integrated sediment flux at six stations in 1995 cruise. Note a tidal shear front (white dashed line). (B) Distributions of bulk density along a transect through the six stations at 2100 on September 18, 1995, demonstrating the spatial pattern of a flow extending seaward beneath the ambient seawater. Source: (A) Wang, H., Bi, N. Wang, Y. Saito, Y. and Yang. Z. 2010. Tide-modulated hyperpycnal flows off the Huanghe (Yellow River) Mouth, China. Earth Surf. Process. Landforms, 35 (11), 1315 1329, with additional labels by G. Shanmugam. (B) From Wang, H., Bi, N. Wang, Y. Saito, Y. and Yang. Z. 2010. Tide-modulated hyperpycnal flows off the Huanghe (Yellow River) Mouth, China. Earth Surf. Process. Landforms, 35 (11), 1315 1329. From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 2 238. Springer, Open Access.

6.6.9 Tidal shear front Perhaps the most significant contribution on the dynamics of the Yellow River sedimentation is pertaining to the recognition of tidal shear front (Wang et al., 2010) (Fig. 6.8A). Li et al. (2001), based upon in-situ measurements and Landsat scanning images, studied spatial temporal changes in the shear front and associated sedimentation in the subaqueous delta slope of the Yellow River. The results showed that the shear front is an important dynamic factor in controlling rapid accretion at the Yellow River mouth. Suspended sediment converges and is deposited rapidly along the shear front zone. This is because a low-velocity zone is formed between two inverse flow bodies. Qiao et al. (2008), by combining a three-dimensional (3D) tidal front numerical model and a sediment transport module, explained the formation of a tidal shear front that occurs off the Yellow River mouth. Wang et al. (2010) documented the position of the tidal front about 5 km seaward off the Yellow River mouth (Fig. 6.8A) and explained the tideinduced density flows on the shelf (Fig. 6.8B). The importance of these numerical

Mass Transport, Gravity Flows, and Bottom Currents

236

6. Hyperpycnal flows

experiments is that the topography with a strong slope off the Yellow River mouth was a determining factor on the generation of a shear front. The sedimentologic implication of the shear front is that it limits seaward transport of sediments (Li et al., 2001; Qiao et al., 2008; Wang et al., 2007, 2010, 2017). If so, the extent of sediment transport into the deep sea by hyperpycnal flows comes into question. In other words, the entire concept of hyperpycnal flows transporting sediment into the deep sea (Mulder et al., 2003; Steel et al., 2016; Warrick et al., 2013; Zavala and Arcuri, 2016) is unsupported by the Yellow River, which is considered to be a classic river for hyperpycnal flows.

6.6.10 M2 tidal dynamics in Bohai and Yellow Seas Wiseman et al. (1986) were one of the early workers who recognized the importance of M2 tidal constituents in the Bohai Sea. Yao et al. (2012) conducted a modeling study of M2 tidal dynamics in understanding the regional tidal mixing and tidal residual currents. There are four regions of low values of log10(h/U3): The inner shelf of Seohan Bay, Kyunggi Bay, the shelf area off the southwest Korean peninsula, and the China shelf area

FIGURE 6.9 Tidal data. (A) Map showing tidal mixing fronts at four locations: (1) the inner shelf of Seohan Bay, (2) Kyunggi Bay, (3) the shelf area off the southwest Korean peninsula, and (4) the China shelf area between 34 N and 35 N. (B) Map showing depth-averaged M2 residual currents. Note strong residual currents are seen off the Yellow River mouth. The tidal residual vectors are plotted every five grid points. Source: From Yao, Z.G., R.Y. He, X.W. Ban, and D.X. Wu. 2012. M2 tidal dynamics in Bohai and Yellow Seas: a hybrid data assimilative modeling study. Ocean Dyn. 62, 753 769 with additional labels by G. Shanmugam. From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 2 238. Springer, Open Access.

Mass Transport, Gravity Flows, and Bottom Currents

6.7 The Yangtze River, China: a case study

237

between 34 N and 35 N (Fig. 6.9A). All these mixing zones are confined in the Yellow Sea (Fig. 6.9A). Inside the Bohai Sea, strong residual currents are seen off the Yellow River mouth (Fig. 6.9B), near Liaodong Bay and north of the Bohai Strait (Fang and Yang, 1985). From the above empirical and numerical data, it is clear that the Yellow River mouth is part of a regional tidal setting that comprises both Bohai Sea and Yellow Sea (Fig. 6.9). In summary, hyperpycnal flows are not simple processes that begin their journey at plunge points, transporting sediment across the shelf, and end up in the deep sea. They are invariably affected by external controls (see Section 6.8). For example, the acute impact of tidal currents on hyperpycnal flows is well documented in the next case study, which is the Yangtze River.

6.7 The Yangtze River, China: a case study 6.7.1 Hyperpycnal and hypopycnal plumes The Yangtze River is the longest river (about 6300 km) in Asia. Satellite images show that the Yangtze River generates both hyperpycnal and deflected hypopycnal plumes (Fig. 5.14A). The Yangtze River mouth is a complex setting in which both ocean currents and tidal currents are affecting sediment dispersal.

6.7.2 Ocean currents Unlike the Yellow River that enters a protected Bohai Bay from major ocean currents, the Yangtze River enters the East China Sea affected by the warm, north-flowing Kuroshio Current (Fig. 5.14B). As a consequence, muddy sediments brought by the Yangtze River are redistributed and deposited as a mud belt on the inner shelf (Wu et al., 2016). This mud belt is evident on the satellite images (Fig. 5.14A). This mud belt is distinctly different from the fan-shaped or lobate deposits of hyperpycnal flows associated with the Yellow River (Fig. 6.2E). Liu et al. (2006) proposed a sediment dispersal model by ocean currents for sediments supplied into the East China Sea by the Yangtze River (Fig. 5.14C). Ocean currents are a global phenomenon (Talley, 2013) with implications for sediment distribution in the world’s oceans (Shanmugam, 2017b).

6.7.3 Tidal river dynamics Similar to the Yellow River, both terms “delta” and “estuarine” are used for the Yangtze River mouth (e.g., Liu et al., 1992). However, evidence for a tide-dominated estuary is compelling. • Hoitink and Jay (2016) reviewed tidal river dynamics of the world’s rivers and classified the Yangtze as a “tidal river.” • Guo et al. (2015) documented that the tidal influence (salt-wedge intrusion) can extend to Datong, which is 650 km upstream from the river mouth (Fig. 6.10A).

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 6.10 Tidal data. (A) The Yangtze River estuary and the location of the tidal gauge stations (red filled circles). The numbers in brackets are distance downward from Datong station (Red triangle). Datong is the maximum limit of salt-wedge intrusion, which is 650 km. See also other studies on the limit of tidal influence (Chen et al., 2007; Zhang et al., 2012, 2017). (B) Geometry and bathymetry of the Yangtze Estuary in 1997 showing river-mouth bars. Such sand bars are typical of tide-dominated estuarine systems (see Dalrymple, 1992; Dalrymple et al., 1992; Shanmugam et al., 2000). (C) Discharge of the Yangtze River for 2009 and 2010 at the head of the tide at Datong. (D) Simultaneous water levels at Datong, Nanjing, Zhenjiang (decreased by 2 m), Jiangyin (decreased by 4 m), Xuliujing, (decreased by 7 m), and Niupijiao (decreased by 11 m). The red lines indicate the daily (24 h) averaged mean water levels (MWL). Note the positive correlation between discharge and MWL in July at Datong. (E and F) Cartoon showing a river entering an ocean with three zones of interest: a normal flow area, where depth is constant along the channel, a transition zone where mean sea level influences river depth, and the offshore river plume in (E) cross section and (F) plan view. At low flow, the transitional region is a zone of backwater, where the water depth at the shoreline (hs) is greater than the normal flow depth (hn), and the water surface (blue) and bed (black) diverge downstream, resulting in deceleration (shown by length of arrows) and deposition. At high flow hn . hs, the water surface (red) is convex, resulting in spatial acceleration of flow and erosion. In both cases, the elevation of the water surface at the river mouth is relatively insensitive to discharge due to lateral spreading of the plume. Source: (A) From Guo, L.C., van der Wegen, M., Jay, D.A., Matte, P., Wang, Z.B., Roelvink, D., He, Q., 2015. River-tide dynamics: exploration of nonstationary and nonlinear tidal behavior in the Yangtze River estuary. J. Geophys. Res. Oceans 120, 3499 3521, with additional labels by G. Shanmugam. (B) From Guo, L.C., van der Wegen, M., Roelvink, J.A., He, Q., 2014. The role of river flow and tidal asymmetry on 1-D estuarine morphodynamics. J. Geophys. Res. Earth Surf. 119, 2315 2334 with additional labels by G. Shanmugam. (C and D) From Guo, L.C., van der Wegen, M., Jay, D.A., Matte, P., Wang, Z.B., Roelvink, D., He, Q., 2015. River-tide dynamics: exploration of nonstationary and nonlinear tidal behavior in the Yangtze River estuary. J. Geophys. Res. Oceans 120, 3499 3521, with additional labels by G. Shanmugam. (E and F) Adapted from Lamb, M.P., Nittrouer, J.A., Mohrig, D., Shaw, J., 2012. Backwater and river plume controls on scour upstream of river mouths: implications for fluvio-deltaic morphodynamics. J. Geophys. Res. Earth Surf. 117, F01002, after Lane (1957). Caption by Hoitink, A.J.F., Jay, D.A., 2016. Tidal river dynamics: implications for deltas. Rev. Geophys. 54, 240 272. From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 2 238. Springer, Open Access.

6.8 External controls

239

• Guo et al. (2014) documented river-mouth bars (Fig. 6.10B) that are analogous to tidal sand bars (Dalrymple et al., 1990). • Liu et al. (1992) reported the development of estuarine sand bars. In support of this observation, a 1997 bathymetric map reveals river-mouth bars, mimicking tidal sand bars typical of tide-dominated estuaries (see Dalrymple et al., 1990; Shanmugam et al., 2000). • Guo et al. (2015) documented the changes in mean water level at Datong with respect to discharge associated with tides (Fig. 6.10C). Tides in the Yangtze River Estuary are semidiurnal with the average tidal range of 2.76 m and the maximum of 4.62 m (Lu et al., 2015) or 5.0 m (Chen et al., 1998). • Tidal flow velocity at the river mouth was measured to be 1 m s21 (Milliman et al., 1985). • Hori et al. (2002) proposed a tide-dominated delta with sand mud couplets and bidirectional cross-laminations for the Yangtze Holocene succession. The differences between a common river and a tidal river affect sedimentation at plunge points (Fig. 6.10D). For example, unlike river-dominated deltas with unidirectional sediment transport (i.e., seaward), tide-dominated estuarine systems are prone to bidirectional transport of sediment (i.e., both seaward and landward) (Dalrymple, 1992; Dalrymple et al., 1992). Under such conditions, the idea of sediment transport by hyperpycnal flows from the river mouth to the deep sea, traveling across the shelf, is misleading. Although both the Yellow and the Yangtze Rivers develop hyperpycnal flows at their river mouths, transport of hyperpycnal sediments from the river mouth to the deep sea has been blocked or diverted by different external controls, such as tidal shear front and ocean currents (Fig. 5.14). This important oceanographic control has been overlooked in studies of hyperpycnites (e.g., Zavala and Arcuri, 2016). In this book, 22 external controls have been identified from global case studies (Fig. 5.20).

6.8 External controls External controls are allogenic in nature, which are external to the depositional system, such as uplift, subsidence, climate, eustacy, etc. However, external controls of density plumes are much more variable and include some common depositional processes (e.g., tidal currents). At least, 22 external controls of plumes have been recognized (Fig. 5.20). Selected examples are (Table 6.1): 1. 2. 3. 4. 5. 6. 7. 8.

tidal shear front (Fig. 6.8): the Yellow River (Wang et al., 2010); ocean currents (Fig. 5.14): the Yangtze River (Liu et al., 2006); tidal currents (Table 6.1): San Francisco Bay (Barnard et al., 2006; NASA, 2017); monsoonal currents (Jagadeesan et al., 2013); wave action (Hawati et al., 2017); cyclones (Table 6.1): Gulf of Mexico; U.S. Atlantic shelf (Shanmugam, 2008a); tsunamis (Table 6.1): Sri Lanka, Arabian Sea (Shanmugam, 2006b); braid delta and related high gradients and coarse sediments (Fig. 6.4): Alaska, Pacific Ocean (McPherson et al., 1987); 9. seiche in lakes (Table 6.1): Lake Erie (NASA, 2017). Seiche is a large standing wave that occurs when strong winds and a quick change in atmospheric pressure push

Mass Transport, Gravity Flows, and Bottom Currents

240

10. 11. 12. 13. 14. 15.

6. Hyperpycnal flows

water from one end of a body of water to the other. de Jong and Battjes (2004) discussed the atmospheric origin of seiche; upwelling (Table 6.1): Off Namibia (Shillington et al., 1992); fish activity (Table 6.1): the Great Bahama Bank (Broecker et al., 2000); volcanic eruptions (Table 6.1): Bering Sea (NASA, 2017); coral reef (Table 6.1): South Pacific Ocean (NASA, 2017); pockmarks: Carolina Continental Rise, North Atlantic Ocean (Paull et al., 1995); internal waves and tides (Masunaga et al., 2015).

FIGURE 6.11 Satellite images showing the difference in plume generation between winter and summer periods in the Yangtze River. (A) Index map. (B) Satellite image of MODIS data on November 9, 2017 showing welldeveloped winter tidal plumes (additional comments by G. Shanmugam). Image acquired on November 9, 2017. (C) Satellite image of MODIS data on July 31, 2015 showing no major extended summer plumes. Source: (A) Image credit: NASA Earth Observatory. (B) From Luo, Z., Zhu, J. Wu, H., Li., X., 2017. Dynamics of the sediment plume over the Yangtze Bank in the Yellow and East China Seas. J. Geophys. Res. Oceans 122, 10,073 10,090. https://doi.org/10.1002/ 2017JC013215, with additional labels by G. Shanmugam. From Shanmugam, G., 2018c. A global satellite survey of density plumes at river mouths and at other environments: plume configurations, external controls, and implications for deep-water sedimentation. Pet. Explor. Dev. 45 (4), 640 661. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

6.9 Recognition of ancient hyperpycnites

241

FIGURE 6.12

Bottom shear stress (units: N m22) in the YECS in winter (top) and summer (bottom). (A and D) Wave-induced shear stress; (B and E) tide-induced shear stress; (C and F) total shear stress. Source: From Luo, Z., Zhu, J. Wu, H., Li., X., 2017. Dynamics of the sediment plume over the Yangtze Bank in the Yellow and East China Seas. J. Geophys. Res. Oceans 122, 10,073 10,090. https://doi.org/10.1002/2017JC013215. AGU.

I have discussed the importance of external controls, such as tidal shear front and ocean currents earlier in Section 6.6.9. Future studies should consider external controls in developing meaningful depositional models for hyperpycnites.

6.9 Recognition of ancient hyperpycnites Recognition of ancient hyperpycnites is rare. However, there are studies that claim that hyperpycnites can be recognized using various criteria. In the following discussion, problems associated with recognizing ancient hyperpycnites are identified. In a recent study, Luo et al. (2017) recognized that extended density plumes tend to develop during winter months (Fig. 6.11B), which are absent during the summer months (Fig. 6.11C). In explaining the winter density plumes (Fig. 6.12), Luo et al. (2017) make the case for vertical mixing of sediments by tides in the winter months using numerical modeling. Although the tides are active all year round, Luo et al. (2017) suggest that their modeling shows that the sediment can only rise up to the surface in the winter, when temperatures and salinities at the sea surface and bottom are roughly the same. In the summer, however, an influx of freshwater from the Yangtze,

Mass Transport, Gravity Flows, and Bottom Currents

242

6. Hyperpycnal flows

FIGURE 6.13 Comparison of conceptual diagrams showing differences in depositional models between the Yellow River (A) and the Yangtze River (B). Note external controls on the distribution of muddy hyperpycnites. In the Yellow River (A), a tidal shear front prevents seaward transport of sediment. In the Yangtze River (B), ocean currents deflect plumes and deposit them as inner-shelf mud belt. Source: From Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 238. Springer, Open Access.

combined with heating of the surface layers of the sea, prevents vertical mixing and keeps the resuspended sediment in the depths. A comparison of depositional models shows the difference between the Yellow River and Yangtze River in terms of deltaic versus estuarine settings (Fig. 6.13). The Yellow River represents a river-dominated delta, whereas the Yangtze River represents a tidedominated estuary. They are totally different dynamic depositional systems from one another. Care must be exercised in studying hyperpycnal flows in these settings.

6.9.1 The hyperpycnite facies model Mulder et al. (2003) proposed a facies model for hyperpycnites (Fig. 6.14). This model is based on a hypothesis that hyperpycnite facies is a function of the magnitude of the flood at the river mouth. According to this hypothesis, hyperpycnites accurately record the rising and falling discharge of a flooding river in terms of sediment-size, inverse grading to normal grading in ascending order (Fig. 6.14), primary sedimentary structures, bed

Mass Transport, Gravity Flows, and Bottom Currents

243

6.9 Recognition of ancient hyperpycnites

FIGURE 6.14 Hyperpycnite facies model showing inverse to normal grading with erosional contact in the middle (Mulder et al., 2003). Color erosional surface symbol by G. Shanmugam.

Hyperpycnite Facies Model Progressive deposition of type 4 bed

Discharge

2

3

4 Critical discharge for the formation of a hyperpycnal turbidity current

1

Time

Hb

Hb Ha 1

2

3

Ha

LEGEND Erosional contact Sharp contact Horizontal lamination Climbing ripples Ripple cross lamination

4

C1 fs ms cs fsa msa

(Mulder et al., 2003)

thickness, and erosional contacts. Mulder et al. (2003) were the first authors to propose a facies model with an internal erosional surface (Fig. 6.14). In testing Mulder et al.’s (2003) hypothesis, Lamb et al. (2010) conducted laboratory flume experiments and concluded that the hypothesis is unsupported by experimental results. Furthermore, Clare et al. (2016) reported that the largest river discharges did not create hyperpycnal flows based on field monitoring of the Squamish Delta, British Columbia, Canada during 2011, thus disputing the hypothesis. Although ichnological signatures are claimed to be characteristic of hyperpycnites (Buatois et al., 2011) and contourites (Stow and Fauge`res, 2008), skepticism about these claims exists (Shanmugam, 2018e).

6.9.2 Inverse to normal grading Following the concept of Mulder et al. (2003) (Fig. 6.14), Wilson and Schieber (2017) and Yang et al. (2017a) recognized ancient hyperpycnites based on inverse to normal grading. However, the origin of inverse grading by waxing flows is an unresolved issue (Shanmugam, 2002b). For example, mechanisms that are commonly used to explain inverse grading are (1) dispersive pressure, caused by grain-to-grain collision which tends to force larger particles toward the zone of least rate of shear (Bagnold, 1954), (2) kinetic sieving, by which smaller particles tend to fall into the gaps between larger particles (Middleton, 1967), and (3) the lift of individual grains toward the top of flow with lower pressures (Fisher and Mattinson, 1968). Nevertheless, Mulder et al. (2001) did not consider any of these alternative mechanisms.

Mass Transport, Gravity Flows, and Bottom Currents

244

6. Hyperpycnal flows

Yang et al. (2017a) recognized normal grading in the Triassic Yanchang Formation in the Ordos Basin, Central China and interpreted normal grading as hyperpycnite. However, normal grading in the Yanchang Formation was previously interpreted as turbidites (Zou et al., 2012). Wilson and Schieber (2014, 2017) described a muddy unit with normal to inverse grading in ascending order from the Devonian Lower Genesee Group, Central New York, which they interpreted as hyperpycnites. These muddy units were previously interpreted as turbidites by other researchers. Muddy turbidity currents and hyperpycnal flows are one and the same, according to some authors (see Kostic and Parker, 2003; Lamb et al., 2010). It is wrong to equate hyperpycnal flows with turbidity currents on fluid dynamical principles without empirical data (Section 6.15).

6.9.3 Internal erosional surface Following Mulder et al. (2003), Yang et al. (2017a) claimed that internal erosional surfaces, which occur between basal inversely graded layer and upper normally graded layer, are the diagnostic criteria of hyperpycnites. Conventionally, a genetic facies model is designed for a single depositional event, without internal hiatuses. A classic example is the turbidite facies model or the “Bouma Sequence” (Bouma, 1962). In fact, Walther’s Law (Middleton, 1973) is not meaningful for sequences with internal hiatuses. This is because a hiatus can represent a considerable span of time (spanning millions of years) that is missing along an erosional surface (Howe et al., 2001). Therefore it is sedimentologically meaningless to relate layers above and below an erosional surface, with a break in deposition in the middle, to the same process (Fig. 6.14). Yang et al. (2017a) are not the only group of authors who promote this flawed concept (there are others, e.g., Wilson and Schieber, 2017). Importantly, no one has reproduced the entire inverse- to normally graded sequence with internal erosional surface (i.e., the hyperpycnite facies model) in laboratory flume experiments. Nor has anyone documented this sequence from modern settings. The conceptual hyperpycnite model exists only in theory in publications, not in the real world.

6.9.4 Traction structures Wilson and Schieber (2014, 2017) interpreted traction structures in mudstone as hyperpycnites. Traction structures are characteristic attributes of bottom currents reworking by contour currents, tidal currents, wind-driven currents, and baroclinic currents (Hollister, 1967; Martın-Chivelet et al., 2008; Shanmugam, 2008b, 2013a, 2016b, 2017b; Shanmugam et al., 1993a,b). Mutti (2009) attributed the origin of hummocky cross-stratification, normally associated with storm deposits, to deposition by hyperpycnal flows. Gao et al. (2014) studied the modern Yellow River Delta and proposed a facies distribution (Fig. 6.14C). The Yellow River Delta model exhibits many attributes of the classic Mississippi River Delta (Coleman and Prior, 1982), including traction currents. In summary, there is a wide range of opinions in interpreting traction structures. Distinguishing hyperpycnites from bottom-current deposits that are ubiquitous in the world’s oceans is a challenge. In particular, contourites are of importance (Hollister, 1967;

Mass Transport, Gravity Flows, and Bottom Currents

6.9 Recognition of ancient hyperpycnites

245

Rebesco and Camerlenghi, 2008). It is worth noting that facies models of both contourites (Stow and Fauge`res, 2008) and hyperpycnites (Fig. 6.14) exhibit inverse to normal grading in ascending order and both have internal hiatus (Shanmugam, 2016b, his Fig. 9.19; also see Chapter 8, Bottom Currents, in this book).

6.9.5 Massive sandstones Massive sandstones, considered to be a recognition criterion for hyperpycnites (Steel et al., 2016; Zavala and Arcuri, 2016), are not unique to deposits of hyperpycnal flows. There are alternative processes that can equally explain the origin of massive sands. Flume experiments, carried out at the St. Anthony Falls Laboratory of the University of Minnesota in Minneapolis during 1996 98, demonstrated that massive sands can be deposited by a sudden freezing of sandy debris flows (Shanmugam, 2000a, his Fig. 18A; Marr et al., 2001, their Fig. 7). In discussing hyperpycnites, Steel et al. (2016, p. 1720) stated that “Although scattered shelf-derived shell fragments suggest an initially turbulent hyperpycnal flow, abrupt lobe terminations, lack of tractional structures, and convolute bedding from rapid dewatering indicate en masse deposition.” En masse deposition is typical of debris flows (Dott, 1963; Enos, 1977; Hampton, 1972; Johnson, 1970; Middleton and Southard, 1977; Shanmugam and Benedict, 1978; Takahashi, 1981). In their global study, Stow and Johansson (2000) attributed the origin of massive sands to sandy debris flows and high-density turbidity currents. In their experimental study, Breien et al. (2010) demonstrated that massive sands can be deposited by laminar sediment flows. The massive sands in their experiments represent deposition from a “fluidized segment” of the flow. Breien et al. (2010, p. 977) considered fluidization as “. . .a mechanism where the mass moves like a fluid, and as the particles settle due to gravity, the pore fluid is displaced upwards, thus providing further grain support.” Steel et al. (2016) indeed reported that there is evidence for dewatering in the cores; so fluidization is a viable explanation for at least some of the cored intervals studied by Steel et al. (2016). Conventionally, massive sand intervals are interpreted as the Bouma Ta division of a turbidite bed (Middleton and Hampton, 1973). The Ta division has also been attributed to deposition from sandy debris flows (Shanmugam, 1997a). In summary, interpreting massive sands is one of the most controversial topics in sedimentology.

6.9.6 Lofting rhythmites Zavala and Arcuri (2016, their Fig. 18), in justifying their criteria for recognizing hyperpycnites, presented a core photograph showing rhythmites, which they called “lofting rhythmites.” The core photograph is from the modern Orinoco Fan, off Orinoco Delta in Eastern Venezuela (their Fig. 15). Such rhythmites are common in deep-water tidal deposits (Cowan et al., 1998; Shanmugam, 2003a). The Orinoco Delta is a classic tide-dominated delta (Chen et al., 2014). Tidal range at coast in the Orinoco Delta area is 2.6 m (Warne et al., 2002). Tidal rhythmites have been documented from the Orinoco Delta (Chen et al., 2017). It is common for tidal rhythmites to occur both at shallow-water and at deep-water environments in a tidal setting, such as the Krishna-

Mass Transport, Gravity Flows, and Bottom Currents

246

6. Hyperpycnal flows

Godavari Basin, Bay of Bengal (Shanmugam et al., 2009) and the Oriente Basin, Ecuador (Shanmugam et al., 2000). This is because tidal currents operate in both shallow-water and deep-water environments concurrently (Boyd et al., 2008; Shanmugam et al., 2009; Shepard et al., 1979). Importantly, tidal rhythmites can be explained by empirical data on daily tidal cycles (Visser, 1980). Zavala and Arcuri (2016) did not consider alternative tidal origins for the Orinoco fan deposits with rhythmites.

6.9.7 Plant remains Plant remains have been used as a criterion for recognizing hyperpycnites (Zavala and Arcuri, 2016). The problem is that shelf currents operate on many continental margins. For example, Imran and Syvitski (2000) studied the Northern California Margin near the mouth of the Eel River and suggested that hyperpycnal flows may be influenced by the along-shelf currents and be deflected northward away from the canyon. Under such conditions, shelf currents can carry the plant remains away from the river mouth and supply them to a site of initiation of “classic” turbidity currents near canyon heads. Such shelf-current transported plant remains would result in erroneous recognition of hyperpycnites in the deep sea. Saller et al. (2006) interpreted leaves in deep-water turbidites. Also, tidal currents are important in transporting plant remains. Abundant coaly carbonaceous fragments have been reported in deep-marine lithofacies with double mud layers, typical of tidal rhythmites (Visser, 1980), from a Pliocene submarine canyon in the Krishna-Godavari Basin, Bay of Bengal (Shanmugam et al., 2009). In addition to shelf currents and tidal currents, wind is a common factor that influences sediment transport in coastal areas. The movement of sediment has been basically triggered by breaking waves. This factor coupled with the geographic location of certain countries such as Indonesia where not only plant remains are abundant but also the region is affected by extreme monsoonal winds. For example, at Coastal Region of Timbulsloko Demak in Indonesia, maximum speed of wind can reach at 23 knots (11.83 m s21) from December to February, wind direction predominantly from North West direction. Correlation between breaking waves and sediment transport is linear (Hawati et al., 2017). McGowen et al. (1977) documented the role of longshore currents in the Gulf Shoreline of Texas in transporting sediment of all sizes. I already discussed the effects of warm ocean currents in redistributing hyperpycnites along the inner shelf of the East China Sea (Fig. 5.14B). In light of these oceanographic factors, plant remains are not a viable criterion for recognizing hyperpycnites.

6.9.8 Hyperpycnite fan models Bouma et al. (1985) documented characteristics of both modern and ancient submarine fans. Various aspects of submarine fans were also discussed by other authors (Mutti, 1992; Shanmugam and Moiola, 1988; Shanmugam, 2016a). Conventionally, submarine fans were related to deposition from turbidity currents (Mutti, 1992). Recently, origin of submarine

Mass Transport, Gravity Flows, and Bottom Currents

6.9 Recognition of ancient hyperpycnites

247

FIGURE 6.15 Hypothetical sandy and muddy hyperpycnal flows and their related deposits. The theoretical concept of sandy hyperpycnal flows transporting sand and gravel across the shelf has never been documented in modern shelf environments. Source: Diagram from Zavala, C., Arcuri, M., 2016. Intrabasinal and extrabasinal turbidites: origin and distinctive characteristics. Sediment. Geol. 337, 36 54, with additional labels by G. Shanmugam.

fans has been attributed to hyperpycnal flows (Warrick et al., 2013; Zavala and Arcuri, 2016). Zavala and Arcuri (2016) proposed two types of hypothetical hyperpycnal flows, namely, sandy and muddy types (Fig. 6.15). In this classification, the lofting plume (i.e., positively buoyant) in sandy hyperpycnal flows is of significance. Similar lofting models were also proposed by Steel et al. (2016). It is worth noting that the wake part of a turbidity current discussed by Allen (1985) is somewhat analogous to the lofting part. Zavala and Arcuri (2016) also used the classification of turbidity currents by Mutti et al. (1994, 1999) into low-density turbidity currents and high-density turbidity currents. Despite our poor understanding of the behavior of high-density turbidity currents and hyperpycnal flows in deep-water environments (Shanmugam, 2016a), Zavala and Arcuri (2016) proposed two types of hyperpycnal flows, namely, sandy and muddy types. Importantly, they proposed two types of turbidites, namely “intrabasinal turbidites” and “extrabasinal turbidites” (Fig. 6.16). Intrabasinal turbidites are those with sediments derived locally from adjacent shelf and got transported into the basin by “classic” turbidity currents. In contrast, extrabasinal turbidites are those with sediments derived from distant land and delta and got transported into the basin by

Mass Transport, Gravity Flows, and Bottom Currents

248

6. Hyperpycnal flows

FIGURE 6.16 Conceptual block diagram showing the occurrence of intrabasinal (I) and extrabasinal (E) turbidites. Note that extrabasinal turbidites receive a direct supply from rivers in flood, and can accumulate shelfal or deep marine deposits. This model ignores external controls such as tidal shear fronts (see Fig. 6.8) and ocean currents (see Fig. 5.14), which prevent transport of hyperpycnal sediments from the river mouth to the deep sea. This model also ignores types of submarine canyons (Figs. 6.17 and 6.18), which are critical for fan deposition (see text). Source: Block diagram from Zavala, C., Arcuri, M., 2016. Intrabasinal and extrabasinal turbidites: origin and distinctive characteristics. Sediment. Geol. 337, 36 54, with additional labels by G. Shanmugam.

“flood-triggered” turbidity currents or hyperpycnal flows (Fig. 6.16). In other words, large river delta fed submarine fans on passive continental margins, such as the Mississippi Fan and the Amazon Fan, would be classified as extrabasinal turbidite. Because that deposits of hyperpycnal flows are called “hyperpycnites” (Mulder et al., 2002), large submarine fans could be termed hyperpycnite fans. These conceptual fan models have inherent problems.

6.9.9 Flawed principles 1. The concept of lofting hyperpycnal flow is problematic (Fig. 6.15) because it defies basic principles of buoyancy. In discussing buoyancy effects in fluids, Turner (1980) explained that positively buoyant plumes cannot be hyperpycnal (i.e., plume with excess density cannot loft). By definition, hyperpycnal flows are negatively buoyant due to their excess density. 2. Zavala and Arcuri (2016) classified turbidity currents and fans based on provenance (internal source vs external source), which is in conflict with the conventional definition of turbidity currents based on Newtonian rheology and turbulent state (Dott, 1963; Sanders, 1965). 3. The hypothetical model of extrabasinal turbidites and related hyperpycnite fans is untenable for two reasons. First, the concept of high-density turbidity currents, which serves as the basis, is not only theoretically flawed (Shanmugam, 1996a), but also empirically undocumented in the world’s oceans (Shanmugam, 2016a). Second, the

Mass Transport, Gravity Flows, and Bottom Currents

6.9 Recognition of ancient hyperpycnites

249

model fails to take into account the most fundamental factor in developing submarine fans, which is a submarine canyon (see Section 6.9.12).

6.9.10 Grain size Modern and ancient submarine fans contain a complex blend of gravel, sand, and mud (Shanmugam and Moiola, 1988). However, hyperpycnal flows cannot be responsible for transporting gravel and sand from the land, carrying them 10 100s of kilometers across the shelf (Fig. 6.1), and delivering them to the deep sea. For example, no one has ever documented by direct measurements or observations of transport of gravel and sand by hyperpycnal flows in suspension from the shoreline to the deep sea in modern settings. Without acknowledging this fundamental lack of empirical data, Warrick et al. (2013) suggested formation of submarine fans by hyperpycnal plume-derived sediments in the Santa Barbara Channel, California. Shallowwater muddy hyperpycnal flows should not be confused with deep-water sandy turbidity currents (Shanmugam, 2012a). In a comprehensive review of hyperpycnal flows, Talling (2014, p. 179) concluded that “Weak and dilute flows generated by plunging hyperpycnal flood discharges most likely deposits thin (mm to ,10 cm) and fine-grained sediment layers, similar to those documented for hyperpycnal flows in lakes and reservoirs (his Fig. 8D and E). The available field observations suggest that they do not form meter-thick sand layers in deep-water settings, as has been previously proposed (Mulder et al., 2003).”

6.9.11 Modern analogs Steel et al. (2016) claimed that “. . . hyperpycnal flows became positively buoyant and lifted off the seabed, resulting in well-sorted, structureless, elongate sand lobes.” However, such positively buoyant hyperpycnal plumes have never been documented in modern shelf environments. Steel et al. (2016, p. 1717) also claimed that “Turbidity currents generated by plunging of sediment-laden rivers at the fluvial-marine interface, known as hyperpycnal flows, allow for cross-shelf transport of suspended sand beyond the coastline.” Because their Fig. 2B1 shows medium-grained sand, the authors imply that medium-grained sands in the Santa Barbara Channel were transported by lofted hyperpycnal flows (or lofted turbidity currents). But there are no literatures on modern analogs where researchers have documented by direct observations of medium-grained sands being lofted by hyperpycnal flows and those sands being transported across the shelf. Specific publications on modern analogs of lofted hyperpycnal flows with empirical data on their physical properties (e.g., flow velocity, fluid density, grain size, sediment concentration, etc.) are nil. Experimental studies by Kostic et al. (2002) and Lamb et al. (2010) showed that muddy “turbidity currents” or “hyperpycnal flows” are generated at plunge points. Gladstone and Pritchard (2010), who demonstrated lofting of turbidity currents in laboratory experiments, used the fine fraction with an average grain diameter of 12.8 µm and the coarse fraction of 36.5 µm. In other words, all these experiments revealed that hyperpycnal flows are strictly muddy flows and they do not carry medium- to coarse sand or gravel in suspension. Other experiments have also shown that turbidity currents composed of pure sand (medium- to

Mass Transport, Gravity Flows, and Bottom Currents

250

6. Hyperpycnal flows

coarse-grained) without the fines tend to collapse soon after initiation (Shanmugam, 2000a). Turbidity currents are capable of transporting mainly mud and very fine-grained sand in suspensions. Therefore any hypothetical model that advocates lofting of sandy hyperpycnal flows is incongruous with respect to what we know from experiments about the inability of turbidity currents to transport medium-grained sand in suspension. Finally, both Warrick et al. (2013) and Steel et al. (2016) invoked the origin of sands in the Santa Barbara Channel studies to hyperpycnal flows. But they totally ignored the significance of tidal currents. Tidal currents in the Santa Barbara Channel had been well documented (Mu¨nchow, 1998). Boyd et al. (2008) convincingly documented that at high sea level, southeast Australian deep-water sands are delivered by a wave-driven coastal transport system, interacting with estuarine ebb-tidal flows that transport sand over the shelf edge. Therefore one could explain deposition of the so-called “hyperpycnites” in the Santa Barbara Channel by tidal current activities.

6.9.12 Submarine canyons Submarine canyons play a critical role in serving as conduits for the transfer of sediments from the land to the sea (Shepard, 1981; Shepard and Dill, 1966). Submarine canyons are also important to understanding conceptual models of hyperpycnites (Fig. 6.16) because plant remains are used as a criterion to recognize large submarine fans. Therefore it is imperative to acknowledge some fundamental aspects of submarine canyons that are well established (see Appendix A). 6.9.12.1 Origin Lamb et al. (2010, p. 1398) in their attempt to explain the origin of submarine canyons by hyperpycnal flows stated that “In fact, hyperpycnal flows might erode the seabed, which offers a potentially interesting feedback between plunging hyperpycnal flows and submarine canyon formation (e.g., Pratson et al., 1994).” However, Pratson et al. (1994, p. 411) concluded that “As reviewed here, mass wasting initiated the subsea sediment flows that began canyon formation and enhanced canyon growth by widening the canyons through retrogressive seafloor failures, for example, the gullying of canyon walls observed by Farre and others (Farre et al., 1983).” Clearly, Pratson et al. (1994) were not referring to hyperpycnal flows because their term “sediment flows” was meant for sediment-gravity flows (Middleton and Hampton, 1973). Sediment-gravity flows are composed of grain flows, debris flows, fluidized sediment flows, and turbidity currents (Middleton and Hampton, 1973, their Fig. 1). Sediment flows are not hyperpycnal flows. Daly’s hypothesis for the origin of submarine canyons by density (turbidity) currents was quite popular in the 1950s and 1960s (Daly, 1936). F.P. Shepard, who devoted his professional life at the Scripps Institution of Oceanography in California to the study of submarine canyons, concluded that submarine canyons were formed not by a single mechanism, but by a combination of processes, such as subaerial erosion, submarine erosion, and faulting, over a long period of time. The point is that Shepard did not even consider the possibility of hyperpycnal flows in explaining the origin of submarine canyons (Shepard, 1981). Brine-related dense shelf-water cascading currents (Roveri et al., 2013)

Mass Transport, Gravity Flows, and Bottom Currents

6.9 Recognition of ancient hyperpycnites

251

should not be confused with hyperpycnal flows in eroding submarine canyons. Importantly, there are no empirical data to document the erosion of modern seafloor by genuine hyperpycnal flows. 6.9.12.2 Classification Submarine canyons are critical for understanding deep-sea sedimentation (Normark and Carlson, 2003; Shanmugam, 2003a; Shepard and Dill, 1966; among others). In this regard, Harris and Whiteway (2011), based on ETOPO1 bathymetric grid, compiled the first inventory of 5849 separate large submarine canyons in the world’s oceans. They classified canyons into three basic types: Type 1: shelf-incising canyons having heads with connection to a major river or estuarine system (Fig. 6.17); Type 2: shelf-incising canyons with no clear connection to a major river or estuarine system (Fig. 6.18); Type 3: slope-incising blind canyons with their heads confined to the continental slope (Fig. 6.18). Harris and Whiteway (2011) also reported that canyons exhibit an impressive array of statistics from their length and spacing to their slope, depth range, dendricity, and

FIGURE 6.17 Type 1 shelf-incising, river-associated Zaire (formerly Congo) Canyon. Zavala and Arcuri (2016) did not consider the role of submarine canyons in developing their “extrabasinal turbidite” model. Source: Compiled from Harris, P.T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86, with permission from Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

252

6. Hyperpycnal flows

FIGURE 6.18 Types 2 and 3 canyons near the Laurentian Channel, many of which incise the shelf, incised into the glacial trough mouth fan. Zavala and Arcuri (2016) did not consider the role of submarine canyons in developing their turbidite models. Source: Compiled from Harris, P.T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86, with permission from Elsevier.

sinuosity. Active continental margins contain 44.2% of all canyons (2586/5849) and passive margins contain 38.4% (2244/5849). Canyons are steeper, shorter, more dendritic, and more closely spaced on active than on passive continental margins. River-associated, shelfincising canyons are more numerous on active continental margins than on passive margins. They are most common on the western margins of South and North America where they comprise 11.7% and 8.6% of canyons, respectively. In the Mediterranean Sea, where 518 large submarine canyons have been identified (Harris and Whiteway, 2011), all three types of canyons are present. If one wishes to study the role of hyperpcnal flows in causing submarine canyons, one needs to apply these kinds of robust global datasets. Despite the critical role of submarine canyons in forming submarine fans (Bouma et al., 1985), Zavala and Arcuri (2016) totally ignored the significance of the three types of submarine canyons in their models for hyperpycnite fans. It is worth noting that the Type 1 and Type 3 submarine canyons are likely to serve as conduits for extrabasinal and intrabasinal turbidites, respectively. For example, the Zaire Canyon (Fig. 6.17), which is a Type 1 canyon, would serve as a conduit for transport of plant remains from land to the sea, irrespective of the process, be it turbidity currents, tidal currents, or hyperpycnal flows. One could easily misinterpret the Type 1 canyon-fill deposits with plant remains as hyperpycnites, although tidal currents could have transported those plant remains. Shanmugam et al. (2009) interpreted canyon-fill deposits with plant remains as tidalites in the Bay of Bengal.

Mass Transport, Gravity Flows, and Bottom Currents

6.10 Cyclone-induced hyperpycnal turbidity currents in canyons

253

6.10 Cyclone-induced hyperpycnal turbidity currents in canyons Liu et al. (2012) claimed four episodes of “clone-induced hyperpycnal turbidity currents” in the Gaoping Submarine Canyon, Taiwan, China (Fig. 6.19). These cyclones, named Kalmaegi, Fong Wong, Kammuri, and Nuri, occurred during the cyclone (typhoon) season between July 8 and September 11, 2008. Liu et al. (2012) stated that “Our findings verify turbidite sequences with the characteristics of suspended sediment carried by passing turbidity currents that displayed

FIGURE 6.19 Properties measured during episodes of four cyclones (i.e., typhoons), Kalmaegi, Fong Wong, Kammuri, and Nuri (red filled circles), in the Gaoping submarine canyon, Taiwan. (A) Hourly atmospheric pressure and precipitation during the deployment. DL rain gauge was located along the riverbank and SLC rain gauge was located on Siaoliouciou Island. (B) Daily river runoff and interpolated and observed SSC at Liling Bridge gauging station about 30 km from the mouth. (C) Hourly mean wave height and period recorded at a wave buoy SW off Siaoliouciou Island. (D) Stick diagram of the recorded flow by the SonTek current meter whose scale is given on the top right corner. The upward direction is the north. The single red “stick” is the flow recorded on 08:00 LT 18 July. The light brown curve is the acoustic backscatter (echo intensity) measured by the SonTek current meter that represents the SSC in the water. The three arrows point to the timing of the first three timer-discs in the nonsequential sediment trap. (E) The contoured temperature anomaly measured by each mini log on the taut-line. See Shanmugam (2008a) for meteorological and sedimentological aspects of cyclones. Source: From Liu, J.T., Wang, Y.H., Yang, R.J., Hsu, R.T., Kao, S.J., Lin, H.L., et al., 2012. Cyclone-induced hyperpycnal turbidity currents in a submarine canyon. J. Geophys. Res. Oceans 117, C04033, with additional labels by G. Shanmugam. AGU.

Mass Transport, Gravity Flows, and Bottom Currents

254

6. Hyperpycnal flows

distinct waxing and waning phases. Our study also confirms the direct link between typhoontriggered hyperpycnal flows in a small mountainous river and turbidity currents in a nearby submarine canyon that transport sediment to the deep-sea efficiently.” Cyclones do generate flows that travel in various directions, but they should not be equated with “hyperpycnal flows” for meteorological reasons (Shanmugam, 2008a). The other problem is that cyclonic flows can travel in any direction (upslope, downslope, alongslope, even within canyons), whereas hyperpycnal flows travel only downslope (seaward) because they are density flows, unless they are redirected by shelf currents. The appropriate process term here is “cyclone-induced density flows.” Palanques et al. (2006) documented the role of cyclones and dense-water cascading in the Gulf of Lions Submarine Canyons.

6.11 Configurations of density plumes Our current understanding of hyperpycnal flows is based on a skewed emphasis on rivermouth processes. However, a global survey of density plumes suggests a plethora of plume types and origins. For example, the U.S. National Aeronautics and Space Administration (NASA, 2017) has archived thousands of satellite images of density plumes in its online publishing outlet called “Earth Observatory” since 1999. There are, at least, 24 configurations of density plumes (NASA, 2017; Table 6.1): 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24.

simple lobate (Fig. 5.2E) horse’s tail (Fig. 5.2F) deflecting (Fig. 6.11B) dissipating (Fig. 5.8C) U-Turn (Fig. 5.20C) swirly (Fig. 6.20B) cloudy massive tidal lobate (Fig. 6.21C) cascading backwash meltwater coalescing irregular eolian blanketing dust linear (Fig. 6.5) anastomosing coalescing lobate (Fig. 6.22B) whitings (Fig. 6.23A) ring (Fig. 6.23B) tendril (Fig. 6.24B) eolian dust feathery volcanic ash gas hydrate

Mass Transport, Gravity Flows, and Bottom Currents

6.12 Global case studies

255

FIGURE 6.20

Influence of ocean currents on hyperpycnal plumes off the Rio de la Plata Estuary. (A) Schematic representation of the depth-averaged current circulation in the southwestern Atlantic region. The shelf (depths smaller than 200 m) is marked by white background. Ocean currents are primarily responsible for the dilution of sediment plumes at the mouth of the estuary (see figure C). (B) Seasonal composites of chlorophyll-a concentration (mg m23) for the study area (white circles). Red pattern represents trends of plumes (arrows). Note heavy plume trending North during the winter season. The implication is that not all plumes routinely move offshore across the continental shelf and reach the deep sea. Source: (A) From Matano, R.P., Palma, E.D., Piola, A.R., 2010. The influence of the Brazil and Malvinas Currents on the southwestern Atlantic shelf circulation. Ocean ˇ Sci. 6, 983 995. (B) From Gonzalez-Silvera, A., Santamaria-del-Angel, E., Milla´n-Nu´nez, R., 2006. Spatial and temporal variability of the Brazil Malvinas Confluence and the La Plata Plume as seen by SeaWiFS and AVHRR imagery. J. Geophys. Res. 111, C06010. AGU.



Each type is of significance in sedimentary record. However, there are no systematic studies of these plumes and their deposits in modern settings.

6.12 Global case studies In addition to the three case studies discussed earlier (i.e., braid delta from Alaska, river-dominated delta of the Yellow River, and tide-dominated estuary of the Yangtze River in China), I have selected the following case studies in understanding the complex factors that control plume types in estuaries, rivers, bays, and lakes.

Mass Transport, Gravity Flows, and Bottom Currents

256

6. Hyperpycnal flows

FIGURE 6.21 Zambezi delta. (A) Location of the Zambezi delta (red filled circle). (B) Satellite image showing coalescing lobate plume as a product of multiple river mouths of the Zambezi River, central Mozambique. Note the influence of wave actions and related beaches. NASA Earth Observatory images by Robert Simmon, using Landsat 8 data from the USGS Earth Explorer. https://earthobservatory.nasa.gov/IOTD/view.php?id 5 82361. Image acquired on August 29, 2013. Source: (A) Image credit: ETOPO1 Global Relief Model, C. Amante and B.W. Eakins, ETOPO1 Arc-Minute Global Relief Model: Procedures, Data Sources and Analysis, NOAA Technical Memorandum NESDIS NGDC-24, March 2009, with additional labels by G. Shanmugam. (B) Additional labels by G. Shanmugam. NASA. Public Domain.

6.12.1 Dissipating plume with irregular front: the Rı´o de la Plata Estuary, Argentina and Uruguay, South Atlantic Ocean The Rı´o de la Plata Estuary is located on the east coast of South America, bordering Argentina and Uruguay. It is one of the largest estuaries in the world (Acha, 2008; Fossati ˇ et al., 2014; Framinan and Brown, 1996; Sepu´lveda et al., 2004). It is 280 km long and 220 km wide at its mouth, and its water depth does not exceed 10 m (Fig. 5.8B). It receives water and sediment from both the Parana´ and Uruguay rivers with an annual mean discharge of 22,000 m3 s21. Satellite images show dissipating plume with an irregular front (Fig. 5.8C). Gonzalez-Silvera et al. (2006) studied ocean color (OCTS, sea-viewing wide fieldof-view sensor) and sea surface temperature (SST; advanced very high resolution radiometer) images and evaluated spatial and temporal variability of the

Mass Transport, Gravity Flows, and Bottom Currents

6.12 Global case studies

257

FIGURE 6.22 Tidal density plume, San Francisco, California. (A) Location map showing the mouth of the San Francisco Bay with the Pacific Ocean at the Golden Gate Bridge (circle). (B) Image showing a field of giant tidal sand waves and other bedforms at the mouth of San Francisco Bay. The view is from the northwest toward the Golden Gate Bridge (seen in the background), which is approximately 2-km long. More than 40 large (greater than 50-m wavelength) sand waves were mapped, with crest-to-crest lengths of as much as 220 m and heights of as much as 10 m. This computer-generated image by Patrick Barnard of sand waves is based on high-resolution multibeam imaging of the seafloor using research vessel VenTresca by the CSUMB Seafloor Mapping Laboratory. Vertical exaggeration: 3X. Geological details of the setting and sand waves are discussed by Barnard et al. (2006). The land was imaged using digital orthophotos draped over a U.S. Geological Survey digital elevation model. The Golden Gate Bridge model is courtesy of IVS 3Dr. (C) Satellite image showing tidal lobate plume at the mouth (circle; Golden Gate Bridg) of the San Francisco Bay. Image acquired on March 1, 2017. Source: (A and B) Image courtesy of USGS. http://soundwaves.usgs.gov/2006/09/ViewtoGateHGLG.jpg (accessed 14.07.11.). Additional labels by G. Shanmugam. (C) Image credit: NASA: The Operational Land Imager (OLI) on Landsat 8. Additional labels by G. Shanmugam. USGS and NASA. Public Domain.

Brazil-Malvinas Confluence and La Plata Plume (20  45 S and 40 65  W). The dataset covered the period from January 1997 to June 2003. Chlorophyll and SST data were compiled and analyzed. The results show a gradual increase of the northward intrusion of the La Plata Plume throughout the period lasting from summer to

Mass Transport, Gravity Flows, and Bottom Currents

258

6. Hyperpycnal flows

FIGURE 6.23 (A) Location map showing position of Tropical Storm Ida (red filled circle) that came ashore on November 10, 2009, Gulf of Mexico. (B) Satellite image showing cyclonic swirly plume occupying the entire shelf area in the northern Gulf of Mexico. Image acquired on November 10, 2009, just a few hours after the storm hit Alabama and Florida. Unlike riverine hyperpycnal plumes, transport of gravel, sand, and mud to deep-water environments by cyclone-related flows is possible. Source: (A and B) NASA images courtesy Jeff Schmaltz, MODIS Rapid Response Team at NASA GSFC. https://earthobservatory.nasa.gov/IOTD/view.php?id 5 41237&eocn 5 image&eoci 5 related_image. NASA. Public Domain.

winter; the summer shape of the La Plata Plume showed a stronger penetration over the shelf on the Argentinean side of the estuary mouth; and the seasonal migration of the Brazil-Malvinas Confluence (Fig. 6.25). The implications of this study are: 1. In some cases, both sediment plumes and planktonic plumes operate. 2. This dataset strongly suggests the direct control of the sediment plume by ocean currents (Fig. 6.25A). For example, sediment plumes are diluted and dissipated to virtually nothing at the estuary mouth (Fig. 5.8C). 3. The seaward transport of planktonic plumes is diverted northward along the inner shelf by seasonal variations (Fig. 6.25B). Therefore one should not assume that all plumes transport sediment across the continental shelf and deliver sediment into the deep sea.

Mass Transport, Gravity Flows, and Bottom Currents

6.12 Global case studies

259

FIGURE 6.24 Carbonate platforms and reefs. (A) Index map. (B) Satellite image showing a large swath of Whitings plume, just west of the Andros Island, Great Bahama Bank. Image acquired on February 12, 2009. (C) Satellite image showing ring plume in the South Pacific Ocean, near the Tagula Island. Image acquired on January 13, 2002. Source: NASA image created by Jeff Schmaltz, MODIS Rapid Response Team, Goddard Space Flight Center. NASA image by Jesse Allen and Rob Simmon, using data provided by the United States Geological Survey. NASA caption by Rebecca Lindsey. NASA. Public Domain.

4. In cases like this with multiple external controlling factors, use of plant remains as a criterion for recognizing ancient hyperpycnites (Zavala and Arcuri, 2016) is meaningless. 5. The Rı´o de la Plata plume, sourced primarily by the Parana´ River, is in direct conflict with a theoretical hypothesis by Mulder et al. (2003, their Table 5) who argued that the Parana´ River “cannot” generate hyperpycnal flows. Similarly, Mulder et al. (2003) also included the Zaire River in West Africa as one of those rivers that cannot trigger hyperpycnal flows. However, studies showed that the Zaire River indeed developed hyperpycnal flows (Migeon, 2000). Clearly, the field observations do not support theoretical models.

Mass Transport, Gravity Flows, and Bottom Currents

260

6. Hyperpycnal flows

FIGURE 6.25

Great Lakes, United States (image: Wikipedia; additional labels by G. Shanmugam). (A) Index map. (B) Satellite image showing a tendrill plume in Lake Michigan. Image acquired on December 17, 2010. (C) Satellite image showing swirly plume induced by seiche in Lake Erie. Suomi NPP is the result of a partnership between NASA, the National Oceanic and Atmospheric Administration, and the Department of Defense. https:// earthobservatory.nasa.gov/IOTD/view.php?id 5 87079 Image acquired on November 25, 2015. Source: (A) NASA image courtesy MODIS Rapid Response Team at NASA GSFC. https://earthobservatory.nasa.gov/IOTD/view.php? id 5 87079. Additional labels by G. Shanmugam. (B) NASA Earth Observatory image by Jesse Allen, using VIIRS data from the Suomi National Polar-orbiting Partnership. Additional labels by G. Shanmugam. NASA. Public Domain.

6.12.2 Coalescing lobate plume: Zambezi River, Indian Ocean The Zambezi River in Central Mozambique is a wave-dominated delta. It has developed a coalescing lobate plume due to multiple river mouths (Fig. 6.22B). The importance of longshore currents in modifying the delta is discussed by Mikhailov et al. (2015).

6.12.3 Tidal lobate plume: San Francisco Bay, Pacific Ocean The Golden Gate Bridge is located at the mouth of the San Francisco Bay connecting the Pacific Ocean (Fig. 6.21A). Barnard et al. (2006) reported a field of giant sand waves of tidal

Mass Transport, Gravity Flows, and Bottom Currents

6.12 Global case studies

261

origin beneath the Golden Gate Bridge at the mouth of San Francisco Bay in California (Fig. 6.21B). Repeated surveys demonstrated that the sand waves are active and dynamic features that move in response to tidally generated currents. The significance of the tidal sand waves at the mouth of the San Francisco Bay is that muddy lobate sediment plumes have been imaged by NASA at this area (Fig. 6.21C). The oceanographic significance here is that these tidal lobate plumes are identical in shape to classic river-mouth sediment plumes, such as the one observed at the mouth of the Yellow River (Fig. 5.2E). Does it mean that tidal lobes and river-flood lobes would generate identical depositional sequences?

6.12.4 Swirly cyclonic plume: Tropical Storm Ida, Northern Gulf of Mexico Analogous to the 1999 Hurricane Floyd, which resuspended muddy sediments over the entire width of nearly 100 km shelf on the U.S. Atlantic margin (Fig. 5.23B), the 2009 Tropical Storm resuspended muddy sediments and caused a swirly cyclonic plume that is nearly 150 km in maximum width in the northern Gulf of Mexico (Fig. 6.20B).

6.12.5 Whitings plume: the Great Bahama Bank, North Atlantic Ocean The most striking feature of the Great Bahama Bank is the swath of whitish-blue waters west of the Andros Island (Fig. 6.24B). This swath of waters is nearly 350 km in length in the northwest-southeast direction and about 100 km in width, just west of the Andros Island. This swath is labeled as “whitings plume” (Fig. 6.24B). The reason is that the region west of Andros Island on the Great Bahama Bank is dominated by fine lime mud and pellet mud sediments (Purdy, 1963). The seas overlying these muds are well known for episodic, highly turbid events that produce milky white waters, historically referred to as “whitings” (Cloud, 1962). The origin of whitings has been attributed to both direct precipitation of lime mud (Cloud, 1962; Shinn et al., 1989) and by resuspension by fish activities (Broecker et al., 2000), by wind (Dierssen et al., 2009), and by the Florida Current (Purkis et al., 2017).

6.12.6 Ring plume: South Pacific Ocean In the South Pacific Ocean, coral reef environments have developed a ring plume (Fig. 6.23C). In the South Pacific, the dark blue waters of the Coral and Solomon Seas, the coral reefs, and forested islands of the Louisiade Archipelago stretch southeastward from the tip of Papua New Guinea for over 350 km. The lower-left corner of the satellite image shows part of the northwestern coast of the largest island in the archipelago, Tagula Island (also called Vanatinai Island). The Louisiade Archipelago has reefs and islands that are in various stages of “evolution” from fringing reef to coral atolls, which result in ring plumes (Fig. 6.23C). Khanna and Yadav (2008) discussed reef development near the Tagula Island.

6.12.7 Tendril plume: Lake Michigan, United States A satellite image shows tendril configuration of plumes in Lake Michigan (Fig. 6.24B). In this example, suspended sediments transformed the southern end of Lake Michigan.

Mass Transport, Gravity Flows, and Bottom Currents

262

6. Hyperpycnal flows

Ranging in color from brown to green, the sediment filled the surface waters along the southern coastline and formed a long, curving tendril extending toward the middle of the lake, induced by wind activity.

6.12.8 Swirly plume: Lake Erie, United States Swirly plumes in Lake Erie (Fig. 6.24C), United States were attributed seiche (NASA, 2017). Seiche is a large standing wave that occurs when strong winds and a quick change in atmospheric pressure push water from one end of a body of water to the other. de Jong and Battjes (2004) discussed the atmospheric origin of seiche. In summary, given this complex variability, any attempt to advocate facies models for deposits of hyperpycnal flows, under the false assumption that river-mouths are the only type of hyperpycnal flows, is misleading.

6.13 Challenges Twenty-four types of plumes are broadly grouped into 14 common categories (Fig. 5.3). Configurations of density plumes are controlled not only by river floods, but also by tidal currents, ocean currents, upwelling, tsunamis, cyclones, seiche, volcanism, fish activity, coral reef, etc. Despite their wide natural variability in triggering mechanisms, only riverine plumes have received the primary attention thus far. The challenge in studying density plume is that a single type (e.g., swirly) can be generated by different mechanisms (e.g., cyclone, seiche, upwelling, etc.). To date, no one has investigated how these different types of density plumes are preserved in the sedimentary record. Amid these uncertainties, it is premature to propose a facies model for hyperpycnites based on the false notion that there is only one type of hyperpycnal flow, which is the river-mouth type. The other issue is that these different plumes are composed mostly of suspended mud and may be ephemeral. Therefore there is an immediate need to evaluate these long-ignored plume types and their deposits.

6.14 Future research directions Students’ research in the future could benefit from the following objectives and guidelines: 1. apply meaningful process terms in studying density plumes; 2. avoid equating hyperpycnal flows with turbidity currents; 3. conduct laboratory flume experiments by using natural seawater as standing body of water; 4. realize that density plumes originate not only at plunge points associated with rivers, but also in sites unrelated to plunge points (e.g., open marine, away from the shoreline); 5. measure physical properties of hyperpycnal flows at plunge points in modern marine environments (e.g., water depth, gradient change, flow velocity, sediment concentration, seafloor erosion, initiation of turbidity currents, etc.);

Mass Transport, Gravity Flows, and Bottom Currents

6.15 Academic discussions

263

6. acquire empirical data on hyperpycnal flows and their ability to transport sand and gravel in suspension across the modern continental shelf; 7. investigate the link between plume types and their depositional characteristics in various settings influenced by tidal currents, glacial meltwater, eolian dust, volcanic ash, tsunamis, cyclones, upwelling, etc.; 8. keep in mind that popular facies models of the 20th century, associated with turbidites (Shanmugam, 1997a; Van der Lingen, 1969), tsunamites (Shanmugam, 2006b), contourites (Shanmugam, 2016b), seismites (Shanmugam, 2016c), and hyperpycnites (Shanmugam, 2018b), are all problematic in the end. Learn from history and resist the temptation of building genetic facies models. Amid numerous obstacles that exist on studies of density plumes, opportunities also exist for initiating MS and PhD-level research projects. The reason is that an enormous number of satellite images are available from various modern marine and lacustrine environments (Shanmugam, 2018b).

6.15 Academic discussions Shanmugam (2018b) has identified inherent problems associated with hyperpycnites in deep-water environments. Such controversial papers are bound to generate debates, which is normal. Unmitigated academic debates are an integral and a necessary part of advancing science. During the past 36 years (1983 2019), I have participated in 38 academic debates, both written and oral, either by myself or in collaboration with others. These debates are on eustasy, tectonics, sedimentation, sediment deformation, provenance, diagenesis, petroleum source rocks, and 3D reconstruction of paleogeography (Table 6.3). I strongly believe that academic discussions are an integral component of scientific progress because debates serve as a platform for challenging the conventional thinking in a constructive way. These debates focus not only on hyperpycnal flows, but also on associated turbidity currents, contour currents, tidal currents, and baroclinic currents (Table 6.1). It is well known that deep-water environments are invariably sites of multiple processes (e.g., turbidity currents, contour currents, tidal currents, etc.) operating concurrently at a time. In other words, hybrid flows (i.e., two or more processes operating in tandem) are the norm. However, published articles on hyperpycnites tend to treat this depositional facies as an end-member type. This practice is nothing new because researchers have traditionally treated depositional facies such as turbidites, contourites, and tidalites as end-member types. Although debates on end-member types are popular, such discussions do not reflect the real-world scenario. Therefore any discussion of hyperpycnites as an end-member facies is purely an academic exercise, with no redeemable practical value.

6.15.1 Stages of scientific development Thomas Kuhn (1970, 1996), an American physicist, historian, and philosopher of science, argued that science is not a steady, cumulative acquisition of knowledge as

Mass Transport, Gravity Flows, and Bottom Currents

TABLE 6.3 Details of 38 academic debates, both written and oral, participated by G. Shanmugam during the past 36 years (1983-2019), either by himself or in collaboration with others. This averages one debate for every single year during the past 36 years. These debates are on eustasy, tectonics, sedimentation, provenance, diagenesis, petroleum source rocks, and 3D reconstruction of palaeogeography. Four debates (Serial numbers 22, 36, 37, and 38) have been on hyperpycnal flows.

Type of debate

Journal (written) and debate (oral)

1 and 2 Analogous tectonic evolution of the Ordovician foredeeps, southern and central Appalachians

Reply

3

Eustatic control of turbidites and winnowed turbidites

Reply

4

Serial number Geologic topic

References

Sedimentologic importance

Geology

Shanmugam and Lash (1983a,b)

Tectonics and sedimentation

Geology

Shanmugam and Moiola (1983)

Eustasy and sedimentation

A model for carbonate to terrigenous clastic Reply sequences

GSA Bulletin

Walker, Shanmugam, and Ruppel (1984)

General model

5

Ophiolite source rocks for Taconic-age flysch: Trace element evidence

Comment

GSA Bulletin

Shanmugam (1985b) Provenance

6

Sedimentary facies of the Nova Scotian upper and middle continental slope, offshore eastern Canada

Comment

Sedimentology

Shanmugam and Moiola (1985b)

Turbidity currents

7

Provenance of the Middle Ordovician Blount clastic wedge, Georgia and Tennessee

Comment

GSA Bulletin

Shanmugam et al. (1985b)

Provenance

8

Sedimentation in the Chile Trench: depositional morphologies, lithofacies, and stratigraphy

Comment

GSA Bulletin

Shanmugam and McPherson (1987)

Tectonics and sedimentation

9

Parameters influencing Porosity in sandstones: a model for sandstone porosity prediction

Comment

AAPG Bulletin

Shanmugam and Alhilali (1988)

Porosity

10

Fan deltas and braid deltas: varieties of coarse-grained deltas

Reply

GSA Bulletin

McPherson, Shanmugam, and Moiola (1988)

Coarse-grained deltas

11

Comment Diagenetic quartz arenite and destruction of secondary porosity: an example from the Middle Jurassic Brent sandstone of northwest Europe

Geology

Shanmugam (1990a) Diagenesis

12 and 13

Porosity enhancement from chert dissolution beneath Neocomian unconformity: Ivishak Formation, North Slope, Alaska

Reply

AAPG Bulletin

Shanmugam and Higgins (1990a,b)

Chert dissolution

14 18

Reinterpretation of depositional processes in a classic flysch sequence in the Pennsylvanian Jackfork Group, Ouachita Mountains

Reply to five discussions by: Bouma et al. (1997); Coleman (1997); D’Agostino and Jordan (1997); Lowe (1997); Slatt et al. (1997)

AAPG Bulletin

Shanmugam and Moiola (1997)

Sandy debris flows and bottom currents. See Fig. 7 for flow density and origin of clasts

19

Basin-floor fans in the North Sea: sequence- Reply to Discussion by Hiscott et al. (1997) stratigraphic models versus sedimentary facies

AAPG Bulletin

Shanmugam et al. (1997a)

Slumps, sandy debris flows, and bottom currents

20

Processes of deep-water clastic sedimentation and their reservoir implications: What can we predict?

Comment and Reply

AAPG Convention Debate (oral) (no formal publication)

1997 AAPG Annual Convention, Dallas, Texas Moderator: H.E. Clifton Panelists: A.H. Bouma, J.E. Damuth, D.R. Lowe, G. Parker, and G. Shanmugam

Mass-transport deposits (MTDs), turbidity currents, and bottom currents

21

Slope turbidite packets in a fore-arc basinfill sequence of the Plio-Pleistocene Kakegawa Group: their formation and sealevel changes

Comment

Sedimentary Geology

Shanmugam (1997b) Turbidity currents

22

Inversely graded turbidite sequences in the deep Mediterranean. A record of deposits from flood-generated turbidity currents?

Comment

Geo-Marine Letters

Shanmugam (2002b) Hyperpycnal flows

23

Reply Tide-dominated estuarine facies in the Hollin and Napo (“T” and “U”) Formations (Cretaceous), Sacha Field, Oriente Basin, Ecuador

AAPG Bulletin

Shanmugam et al. (2002)

Tidal currents

(Continued)

TABLE 6.3 (Continued) Serial number Geologic topic

Type of debate

24

A preliminary experimental study of turbidite fan deposits

Comment

25

Journal (written) and debate (oral)

References

Sedimentologic importance

Journal of Sedimentary Research

Shanmugam (2003b) Experimental turbidity currents

Leaves in turbidite sand: the main source of Comment oil and gas in the deep-water Kutei Basin, Indonesia

AAPG Bulletin

Shanmugam (2008c) Petroleum source rocks

26

Late Holocene rupture of the Northern San Andreas Fault and possible stress linkage to the Cascadia Subduction Zone

Comment

Bulletin of the Seismological Society of America

Shanmugam (2009b) Subduction zone

27

Turbidites and turbidity currents from Alpine “flysch” to the exploration of continental margins

Comment

Sedimentology

Shanmugam (2010b) Turbidity currents

28

Transport mechanisms of sand in deepmarine environments—insights based on laboratory experiments

Comment

Journal of Sedimentary Research

Shanmugam (2011a) Experimental turbidity currents and fluidized flow

29

Evidence of internal-wave and internal-tide Comment deposits in the Middle Ordovician Xujiajuan Formation of the Xiangshan Group, Ningxia, China

Geo-Marine Letters

Shanmugam (2012c) Baroclinic currents

30

Internal waves, an underexplored source of Comment turbulence events in the sedimentary record

Earth-Science Reviews

Shanmugam (2013b) Baroclinic currents

31

Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands

Reply

AAPG Bulletin

Shanmugam (2014a) Baroclinic currents

32

Review of research in internal-wave and internal-tide deposits of China

Comment

Journal of Shanmugam (2014b) Baroclinic currents Palaeogeography

33

3D paleogeographic reconstructions of the Phanerozoic versus sea-level and Sr-ratio variations

Comment

Journal of Shanmugam (2015b) 3D reconstruction of Palaeogeography paleogeography

34

The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India

Comment

Journal of Shanmugam (2017c) Deformation of Palaeogeography stromatolites by seismic shocks

35

Ichnological analysis of contourites: Past, present, and future

Comment

Earth-Science Reviews

Shanmugam (2018f) Ichnology and contour currents

36

Climatic and tectonic controls of lacustrine hyperpycnite origination in the Late Triassic Ordos Basin, Central China: implications for unconventional petroleum development

Comment

AAPG Bulletin

Shanmugam (2019b) Hyperpycnal flows

37 and 38

Reply to discussions by Zavala (2019) and by Van Loon, Hu¨eneke, and Mulder (2019) on Shanmugam, G. (2018b, Journal of Palaeogeography, 7(3): 197 238): “The hyperpycnite problem”

Replies to both discussions

Journal of Shanmugam (2019c) Hyperpycnal flows Palaeogeography

By Zavala (2019) and by Van Loon, Hu¨eneke, and Mulder (2019)

This averages one debate for every single year during the past 36 years. These debates are on eustasy, tectonics, sedimentation, provenance, diagenesis, petroleum source rocks, and 3D reconstruction of paleogeography. Four debates (Serial numbers 22, 36, 37, and 38) have been on hyperpycnal flows.

268

6. Hyperpycnal flows

FIGURE 6.26 (A) Conventional stages of scientific development. (B) Stages of development on hyperpycnal flows and hyperpycnites. At present, we are in a state of crisis and awaiting for development of “Revolution and Normal Science.” Source: (A) Diagram, based on concepts of Kuhn (1970 and 1996), from Shanmugam, G., 2016a. Submarine fans: a critical retrospective (1950 2 2015). J. Palaeogeogr. 5 (2), 110 2 184.

portrayed in the textbooks. Instead, it is a series of peaceful interludes punctuated by intellectually violent revolutions. In these revolutions, one conceptual world view is replaced by another more complex view. Kuhn (1996, p. 84 85) wrote that “The transition from a paradigm in crisis to a new one from which a new tradition of normal science can emerge is far from a cumulative process, one achieved by an articulation or extension of the old paradigm. Rather it is reconstruction of the field from new fundamentals, a reconstruction that changes some of the field’s most elementary theoretical generalizations as well as many of its paradigm methods and applications.” Kuhn’s (1970, 1996) stages of scientific development may be grouped into five steps (Fig. 6.26A): (1) early random observations, (2) first paradigm, (3) crisis, (4) revolution, and (5) normal science or new paradigm (Fig. 6.26). Once the final step or normal science is achieved (i.e., the new paradigm); however, scientists enjoy a sense of confidence as well as comfort. This comfort often leads to complacency. The normal science is influential in: (1) forcing scientists to force-fit nature into preconceived models of the paradigm, (2) encouraging scientists to ignore data or observations that do not fit the basic principles of the paradigm, (3) discouraging scientists from inventing new theories, and (4) making scientists intolerant of new theories invented by others (Kuhn, 1996, p. 24). There are ample examples of such influences on deep-water research (e.g., the Bouma Sequence, submarine fan models, sequence-stratigraphic models discussed by Shanmugam, 2006a, 2016a).

6.15.2 Hyperpycnites and the remaining unresolved issues In the case of hyperpycnal flows, it took 118 years to evolve from early random observations (Forel, 1885) to first paradigm (Mulder et al., 2003) (Fig. 6.26B). It took another 15

Mass Transport, Gravity Flows, and Bottom Currents

6.15 Academic discussions

269

years to reach the crisis stage (Fig. 6.26B). During the current crisis period, there is no clarity on the fundamentals of hyperpycnal flows. For example, there are 16 types of hyperpycnal flows without a unifying principle of fluid mechanics. Plus, there are 22 types of external controls and 24 types of plume configurations (Shanmugam, 2018b,c). These basic problems are not addressed by Van Loon et al. (2019) and by Zavala (2019) in their comments. As a consequence, the following fundamental issues still remain unresolved: 1. Zavala has chosen an awkward title for his discussion, which is “The new knowledge is written on sedimentary rocks”—a comment on Shanmugam’s paper “The hyperpycnite problem.” According to the Cambridge Dictionary, the term “knowledge” is defined as follows: “Awareness, understanding, or information that has been obtained by experience or study, and that is either in a person’s mind or possessed by people generally.” Cambridge Dictionary, URL: https://dictionary. cambridge.org/us/dictionary/english/knowledge (accessed 10.003.19). In his book entitled Use the Right Word, Hayakawa (1968) states that, “Knowledge is more than a store of facts in the mind; it also includes the contribution of the mind in understanding data, perceiving relations, elaborating concepts, formulating principles, and making evaluations.” Clearly, the word “knowledge” refers to the acquired trait by the humankind, and one cannot write “knowledge” on the rocks, as Zavala has falsely implied in his paper title. 2. What is the fluid rheology of hyperpycnal flows and how does it differ from that of turbidity currents? 3. What is the flow state of hyperpycnal flows and how does it differ from that of turbidity currents? 4. What is the sediment concentration of hyperpycnal flows by volume and how does it differ from that of turbidity currents? 5. Given that the Yellow River is a delta, how tide-modulated hyperpycnal flows differ from the concept of conventional hyperpycnal flows that was originally advocated by Bates (1953) for deltaic environments? 6. What are the fluid dynamical properties of tide-modulated hyperpycnal flows? 7. What are the sedimentary characteristics of deposits of tide-modulated hyperpycnal flows? 8. What are the sedimentary characteristics of deposits of baroclinic currents (Shanmugam, 2013a,b; 2014a,b) that are closely associated with internal waves in the Yellow River (Wang et al., 2010)? 9. What is difference in sedimentary characteristics between deposits of single-layer and double-layer flows? 10. What is difference in sedimentary characteristics between deposits of single-layer and multilayer flows? 11. What is difference in sedimentary characteristics between deposits of double-layer and multilayer flows? 12. What is difference in sedimentary characteristics between deposits of double-layer and tide-modulated hyperpycnal flows?

Mass Transport, Gravity Flows, and Bottom Currents

270

6. Hyperpycnal flows

13. What are the sedimentary characteristics of deposits of internal waves associated with tide-modulated hyperpycnal flows (Wang et al., 2010)?, to mention a few issues.

6.16 Synopsis Sedimentologic, oceanographic, and hydraulic engineering publications on hyperpycnal flows claim the following: 1. River flows transform into turbidity currents at plunge points near the shoreline. 2. Hyperpycnal flows have the power to erode the seafloor and cause submarine canyons. 3. Hyperpycnal flows are efficient in transporting sand across the shelf and can deliver sediments into the deep sea for developing submarine fans. 4. Importantly, these claims do have economic implications for the petroleum industry for predicting sandy reservoirs in deep-water petroleum exploration. However, these claims are based strictly on experimental or theoretical basis, without the supporting empirical data from modern depositional systems. Therefore the primary purpose of this study, which began nearly 30 years ago, was to rigorously evaluate the merits of these claims. A global evaluation of density plumes, based on 45 case studies [e.g., Yellow River, Yangtze River, Copper River, Hugli River (Ganges), Guadalquivir River, Rı´o de la Plata Estuary, Zambezi River, among others], suggests a complex variability in nature. For example: 1. Density plumes occur in six different environments (i.e., marine, lacustrine, estuarine, lagoon, bay, and reef). 2. They are composed of six different compositional materials (e.g., siliciclastic, calciclastic, planktonic, etc.). 3. They derive material from 11 different sources (e.g., river flood, tidal estuary, subglacial, etc.). 4. They are subjected to 22 different external controls (e.g., tidal shear fronts, ocean currents, cyclones, tsunamis, etc.). 5. They exhibit 24 configurations (e.g., lobate, coalescing, linear, swirly, U-Turn, anastomosing, etc.). Major problems associated with hyperpycnal flows and their deposits (i.e., hyperpycnites) are as follows. 1. There are at least 16 types of hyperpycnal flows (e.g., density flow, underflow, highdensity hyperpycnal plume, high-turbid mass flow, tide-modulated hyperpycnal flow, cyclone-induced hyperpycnal turbidity current, multilayer hyperpycnal flows, etc.), without an underpinning principle of fluid dynamics. 2. The basic tenet that river currents transform into turbidity currents at plunge points near the shoreline is based on an experiment that used fresh tap water as a standing body. In attempting to understand all density plumes, such an experimental result is inapplicable to marine waters (sea or ocean) with a higher density due to salt content.

Mass Transport, Gravity Flows, and Bottom Currents

6.16 Synopsis

271

3. Published velocity measurements from the Yellow River mouth, a classic area, are of tidal currents, not of hyperpycnal flows. Importantly, the presence of tidal shear front at the Yellow River mouth limits seaward transport of sediments. 4. Despite its popularity, the hyperpycnite facies model has not been validated by laboratory experiments or by real-world empirical field data from modern settings. 5. The presence of an erosional surface within a single hyperpycnite depositional unit is antithetical to the basic principles of stratigraphy. 6. The hypothetical model of “extrabasinal turbidites,” deposited by river-flood triggered hyperpycnal flows, is untenable. This is because high-density turbidity currents, which serve as the conceptual basis for the model, have never been documented in the world’s oceans. 7. Although plant remains are considered a criterion for recognizing hyperpycnites, the “Type 1” shelf-incising canyons having heads with connection to a major river or estuarine system could serve as a conduit for transporting plant remains by other processes, such as tidal currents. 8. Genuine hyperpycnal flows are feeble and muddy by nature, and they are confined to the inner shelf in modern settings. 9. Distinguishing criteria of ancient hyperpycnites from turbidites or contourites are muddled. 10. After 67 years of research since Bates (1953), our understanding of hyperpycnal flows and their deposits is still incomplete and without clarity. 11. Analagous to the turbidite paradigm (Chapter 4), the current trend is to hype the importance of hyperpycnites in the ancient sedimentary record without sound empirical data from modern settings.

Mass Transport, Gravity Flows, and Bottom Currents

C H A P T E R

7 Triggering mechanisms of downslope processes O U T L I N E 7.1 Definition

273

7.2 Origin 7.2.1 Earthquakes 7.2.2 Meteorite impact 7.2.3 Volcanic activity 7.2.4 Tsunami wave 7.2.5 Rogue waves 7.2.6 Cyclonic waves 7.2.7 Internal waves and tides 7.2.8 Ebb-tidal currents 7.2.9 Monsoon flooding 7.2.10 Groundwater seepage

274 274 277 277 280 288 288 292 292 296 298

7.2.11 7.2.12 7.2.13 7.2.14 7.2.15 7.2.16 7.2.17 7.2.18 7.2.19 7.2.20 7.2.21

Wildfire Human activity Tectonic oversteepening Glacial maxima and loading Salt movements Depositional loading Hydrostatic loading Ocean-bottom currents Biological erosion Gas-hydrate decomposition Sea-level lowstand

7.3 Synopsis

298 298 300 300 300 301 302 302 302 304 304 307

7.1 Definition A triggering mechanism or a trigger is defined in this chapter as “the primary process that causes the necessary changes in the physical, chemical, and geotechnical properties of the soil, which results in the loss of shear strength that initiates the sediment failure and movement.” Commonly, triggering processes are considered “external” with respect to the site of failure. In continental margins, several triggering mechanisms may work concurrently or in tandem (e.g., earthquake-triggered tsunamis). Sowers (1979) articulated the

Mass Transport, Gravity Flows, and Bottom Currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00007-2

273

© 2021 Elsevier Inc. All rights reserved.

274

7. Triggering mechanisms of downslope processes

challenge of identifying the single mechanism that is responsible for the failure as follows: “In most cases, several ‘causes’ exist simultaneously; therefore, attempting to decide which one finally produced failure is not only difficult but also technically incorrect. Often the final factor is nothing more than a trigger that sets a body of earth in motion that was already on the verge of failure. Calling the final factor the cause is like calling the match that lit the fuse that detonated the dynamite that destroyed the building the cause of the disaster.” Although more than one triggering mechanism can cause a single process (e.g., debris flows) at a given site, there are no objective criteria to distinguish either the triggering mechanism or the transport process from the depositional record yet (Shanmugam, 2006b, 2012b; Mulder et al., 2011). In addition, unsubstantiated claims on the triggers of turbidity currents also exist. In studying the Squamish Delta in British Columbia, for example, Hizzett et al. (2018, p. 855) claim that “Subaerial rivers and turbidity currents are the two most voluminous sediment transport processes on our planet.” And yet, Hizzett et al. (2018, p. 855) also state that “There are few direct observations of turbidity currents in action, and even fewer observations of multiple turbidity currents at one site with different triggers.” One wonders as to why turbidity currents being the most voluminous sediment transport process on our planet are also being the least observed process on Earth? The reason is twofold. First, Hizzett et al. (2018) use the terms “turbidity currents” and “landslides” casually without a clear definition in terms of fluid mechanics and soil mechanics (see Chapter 2: Mass Transport: Slides, Slumps, and Debris Flows). Second, since the 1960s, the importance of turbidity currents has been the subject of overzealous promotion without empirical data (see Chapter 4: A Paradigm Shift). Nevertheless, an understanding of the multitude of triggering mechanisms is imperative in evaluating sediment failures and related masstransport deposits (MTDs) and gravity flows (Locat and Lee, 2005; Masson et al., 2006; Feeley, 2007; Talling, 2014; Clare et al., 2016).

7.2 Origin There are at least 21 triggering mechanisms that can initiate sediment failures in subaerial and submarine environments (Table 7.1). I have classified these mechanisms into three major groups based on their duration of activity: (1) short-term events that last for only a few minutes to several hours, days or months, (2) intermediate-term events that last for hundreds to thousands of years, and (3) long-term events that last for thousands to millions of years (Shanmugam, 2012a, 2015a). Conceivably, some intermediate-term events may last for a longer duration. The importance here is that short-term events and longterm events are markedly different in their duration. I have discussed each triggering mechanism in some detail elsewhere (Shanmugam, 2012a). Examples are as follows:

7.2.1 Earthquakes Tectonic activity is the primary cause of earthquakes. Earthquakes affect the stability of slopes in two ways. First, the acceleration during the seismic ground motion affects the

Mass Transport, Gravity Flows, and Bottom Currents

TABLE 7.1 Types and duration of triggering mechanisms of sediment failures.

Types of triggering

Environment of sediment emplacement

Duration

1. Earthquake (Heezen and Ewing, 1952; Henstock et al., 2006)

Subaerial and submarine

Short-term events: a few minutes to several hours, days, or months

2. Meteorite impact (Claeys et al., 2002; Barton et al., 2009, 2010)

Subaerial and submarine

3. Volcanic activity (Tilling et al., 1990)

Subaerial and submarine

4. Tsunami waves (Shanmugam, 2006b)

Subaerial and submarine

5. Rogue waves (Dysthe et al., 2008)

Submarine

6. Cyclonic waves (Bea et al., 1983; Prior et al., 1989; Shanmugam, 2008a)

Subaerial and submarine

7. Internal waves and tides (Shanmugam, 2013a,b, 2014a,b) 8. Ebb-tidal current (Boyd et al., 2008)

Submarine Submarine

9. Monsoonal rainfall (Petley, 2012)

Subaerial

10. Groundwater seepage (Bro¨nnimann, 2011)

Subaerial and submarine

11. Wildfire (Cannon et al., 2001)

Subaerial

a

12. Human activity (Dan et al., 2007)

Subaerial and submarine

1. Tectonic eventsb: (1) tectonic oversteepening (Greene et al., 2006); (2) tensional stresses on the rift zones (Urgeles et al., 1997); (3) oblique seamount subduction (Collot et al., 2001), among others 2. Glacial maxima, loading (Elverhøi et al., 1997, 2002); glacial meltwater (Piper et al., 2012) 3. Salt movement (Prior and Hooper, 1999)

Subaerial and submarine

Submarine Submarine

4. Depositional loading (Coleman and Prior, 1982; Behrmann et al., 2006) 5. Hydrostatic loading (Trincardi et al., 2003)

Submarine

6. Ocean-bottom currents (Locat and Lee, 2002)

Submarine

7. Biological erosion in submarine canyons (Dillon and Zimmerman, 1970; Warme et al., 1978) 8. Gas-hydrate decomposition (Popenoe et al., 1993; Sultan et al., 2004; Maslin et al., 2004)

Submarine

1. Sea-level lowstand (Damuth and Fairbridge, 1970; Shanmugam and Moiola, 1982, 1988; Vail et al., 1991)

Submarine

a

Intermediate-term events: hundreds to thousands of years

Submarine

Submarine Long-term events: thousands to millions of years

Although human activity is considered to be the second most common triggering mechanism (next to earthquakes) for known historic submarine mass movements (Mosher et al., 2010), it is irrelevant for interpreting ancient rock record. b Some tectonic events may extend over millions of years. Compiled from several sources. Updated after Shanmugam (2012a, 2015a). The change in numbering is to reflect the change in duration of triggering events.

276

7. Triggering mechanisms of downslope processes

soil in a cyclic manner. Second, the cyclic strains induced by the earthquake loads may reduce the shear strengths of the soil (Duncan and Wright, 2005). Earthquakes are considered to be the single most common triggering mechanism for known historic submarine mass movements (Mosher et al., 2010). Based on the spatial distribution of MTDs on the U.S. Atlantic Margin, Twichell et al. (2009) suggested that earthquakes associated with rebound of the glaciated part of the margin or earthquakes associated with salt domes were the primary triggering mechanism of sediment failures. Seismic energy released during earthquakes not only can trigger sediment failures but also can promote long-runout MTDs (Fig. 7.1). FIGURE 7.1 Map showing (1) the location of the epicenter of the 1929 Grand Banks earthquake on the continental slope off New Foundland, Canada, (2) slump zone near the epicenter, (3) location and timing (minutes after main shock) of submarine cable breaks, and (4) the limit of “turbidite sand” (Fruth, 1965). Source: After Piper, D.J.W., Shor, A.N., Hughes Clarke, J.E., 1988. The 1929 ‘Grand Banks’ earthquake, slump, and turbidity current. In: Clifton, H.E. (Ed.), Sedimentologic Consequences of Convulsive Geologic Events. Geological Society of America Special Paper 229, pp. 77 92. Courtesy: the Geological Society of America.

Mass Transport, Gravity Flows, and Bottom Currents

7.2 Origin

277

Chillarige et al. (1997) compiled the following examples of earthquake-induced liquefaction slides during 1811 1980 in mostly submarine environments: 1811: 1899: 1899: 1908: 1911: 1912: 1922: 1923: 1923: 1929: 1953: 1954: 1957: 1960: 1964: 1964: 1964: 1965: 1980:

New Madrid earthquake, Mississippi River banks Yakutat, Alaska Valdez, Alaska Messina cone (M 5 7.5) Valdez, Alaska (M 5 6.9) Valdez, Alaska (M 5 7.25) Chile (M 5 8.3) Sagami Wan, Japan Kwanto, Tokyo (M 5 8.2) Grand Banks, off Newfoundland (M 5 7.2) Suva, Fiji (M 5 6.75) Orleansville, off Newfoundland (M 5 6.7) San Francisco, California (M 5 5.3) Puerto Montt, Chile (M 5 8.4) Valdez, Alaska (M 5 8.0) Seward, Alaska (M 5 8.3) Valdez, Alaska (M 5 8.3) Seattle, Washington (M 5 6.7) Klamath River delta, California (M 5 6.5)

7.2.2 Meteorite impact The role of meteorite impact on deep-water sedimentation has not received sufficient attention so far. The Chicxulub asteroid impact at the K-Pg boundary on northern Yucatan, Mexico is of interest (Figs. 7.2 and 7.3). This K-T event had generated not only major mass movements directly by the impact-induced seismic shocks, but also by the impact-triggered tsunamis (Smit et al., 1996; Claeys et al., 2002). The Chicxulub event triggered MTDs and other deposits have been documented at the K-Pg boundary all around the Gulf of Mexico (Bourgeois et al., 1988; Smit et al., 1996; Grajales-Nishimura et al., 2000; Claeys et al., 2002; Lawton et al., 2005). Seismic energy released during meteorite impacts not only can trigger sediment failures but also can promote long-runout MTDs.

7.2.3 Volcanic activity The Smithsonian Institution map reveals the importance of volcanoes worldwide along plate boundaries and some continental margins (Shanmugam, 2012a, his Fig. 5.4). Volcanoes are a major cause of sediment failures along modern and ancient continental margins. The 1980 Mount St. Helens eruption triggered subaerial MTDs (Fig. 7.4) with an estimated volume of 3,700,000,000 m3 (130,664,266,870 ft3) (Shanmugam, 2012a, his Table 2.1). Slope failures due to volcanic island growth are among the largest on Earth, involving material volumes of several cubic kilometers. For example, submarine mass movements have been attributed to volcanic activity in the Hawaiian Islands (Lipman

Mass Transport, Gravity Flows, and Bottom Currents

278

7. Triggering mechanisms of downslope processes

FIGURE 7.2 Map showing sites of meteorite impacts in North America, including the Chicxulub (K-Pg, previously K-T) in the Yucatan Peninsula. Source: Map credit: PASSC (Planetary and Space Science Centre), University of New Brunswick, Fredericton, New Brunswick, Canada. http://www.passc. net/EarthImpactDatabase/NorthAmerica.html (accessed 05.04.11.). After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production vol. 9, Elsevier, Amsterdam, p. 524. Elsevier.

FIGURE 7.3 Map showing the site of Chicxulub meteorite impact at the K-Pg boundary in Yucatan, Mexico. Stars represent approximate locations of mass-transport deposits and tsunami-related deposits associated with the Chicxulub impact at the K-Pg boundary (Bourgeois et al., 1988; Smit et al., 1996; Grajales-Nishimura et al., 2000; Takayama et al., 2000; Claeys et al., 2002; Lawton et al., 2005). Generalized outline of Lower Tertiary Wilcox Trend: From several sources (e.g., Lewis et al., 2007; Meyer et al., 2007). Source: After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production vol. 9, Elsevier, Amsterdam, p. 524. Elsevier.

Mass Transport, Gravity Flows, and Bottom Currents

279

7.2 Origin

FIGURE 7.4 Debris flow triggered by volcanic activity. (A) Photograph of Mount St. Helens showing a 3000 ft. (1 km) highsteam plume on May 19, 1982, 2 years after its major eruption. (B) Photograph showing larger pumice blocks near the front of a volcanic debris flow associated with the Eruption of the Mount St. Helens on May 18, 1980. This type of depositional mechanism would result in inverse grading in the rock record. Red arrow shows low direction. Source: (A) Credit: Lyn Topinka, Wikipedia. (B) Photo by T.A. Leighley, USGS, October 17, 1980. Credit: http://libraryphoto.cr.usgs. gov/cgi-bin/show_picture.cgi?ID 5 ID.CVOF.73ct&SIZE 5 large (accessed 27.02.11). USGS.

et al., 1988; Moore et al., 1989; Normark et al., 1993) and in the South Sandwich arc, South Atlantic (Leat et al., 2010), among others. Normark et al. (1993) described 17 giant volcanorelated landslides and the development of the Hawaiian Islands. The largest of these is the Nuuanu landslide that occupies an area of 23,000 km2 in NE Oahu. The Alika 2 landslide in W. Hawaii is 120,000 years old. It has a volume of about 500 km3, and it shows tongueshaped planform geometry (McMurtry et al., 2004, their Fig. 1A). It is 95 km long, 15 km wide, and occupies an area of about 1700 km2 (Moore et al., 1989). Lipman et al. (1988) considered Alika 2 event as a high-velocity debris avalanche. McMurtry et al. (2004) suggested the possibility of a tsunami caused by the giant Alika 2 landslide from nearby Mauna Loa volcano. On the southeast coast of Kilauea Volcano, Hawaii, when lava enters the Pacific Ocean, “lava delta” develops (Shanmugam, 2012a, his Fig. 5.5A). Upon entering the ocean, the lava commonly shatters into sand- to block-size fragments because of rapid cooling. These brecciated volcanic fragments accumulate on the submarine slope to form a loose foundation for the “lava delta.” Because of rapid accumulation of volcanic fragments on steep submarine

Mass Transport, Gravity Flows, and Bottom Currents

280

7. Triggering mechanisms of downslope processes

slopes, the leading edge of “lava delta” invariably collapses to form a series of submarine sandy slides (Shanmugam, 2012a, his Fig. 5.5B). In the Naples Bay (Italy), triggering of volcanic MTDs were associated with concomitant factors such as angle of slope margin, seismic activity, basement architecture, and high pore-fluid pressure (Milia et al., 2006).

7.2.4 Tsunami wave Tsunamis are oceanographic phenomena that represent a water wave or series of waves, with long wavelengths and long periods, caused by an impulsive vertical displacement of the body of water by earthquakes, landslides, volcanic explosions, or extraterrestrial (meteorite) impacts. Earthquakes commonly generate tsunamis through the transfer of large-scale elastic deformation associated with rupture to potential energy within the water column (Geist, 2005). The two prominent tsunamis of the 21st century were triggered by M59 earthquakes in (1) off West coast of Sumatra on December 26, 2004 (Fig. 7.5) and (2) offshore Honshu, Japan on March 11, 2011 (USGS, 2011). Any skeptic of tsunamis FIGURE 7.5 Upper image: Map showing propagating tsunami waves away from the epicenter (solid dot) of the 2004 Indian Ocean Earthquake on December 26, 2004. The epicenter was located 3.307oN 95.947oE off the west coast of Sumatra. Measurements of sea level were made from space using the Satellite (Jason-1) 2 h after the earthquake. Lower image: Plot of relative sea level along the transect X_X0 (see Upper image for location). Source: Modified after NOAA (National Oceanic and Atmospheric Administration), 2005. NOAA News Online, Story 2365. http://www.noaanews.noaa.gov/stories2005/ s2365.htm (accessed 15.06.05.).

Mass Transport, Gravity Flows, and Bottom Currents

7.2 Origin

281

as a real geologic force in controlling sediment transport and deposition must watch YouTube videos of the 2011 Honshu tsunami waters that destroy huge buildings in a matter of minutes and carry large buildings, trucks, and cars in suspension into the Pacific Ocean. Wave heights of the 2004 Indian Ocean Tsunami reached up to 15 m. The coastline of Sumatra, near the fault boundary, received waves over 10 m tall, while those of Sri Lanka and Thailand received waves over 4 m (NOAA, 2005). On the other side of the Indian Ocean, Somalia and Seychelles were struck by waves approaching 4 m in height. Wave height measured from space, 2 hours after the earthquake, reached 60 cm near the east coast of India. A tsunami wave can trigger a number of transportational processes, such as overwash surge, backwash flow, debris flow, turbidity current, and bottom current. These processes, in turn, will emplace sediment from a variety of depositional mechanisms, namely sudden freezing, settling from suspension, and bedload or traction (Shanmugam, 2012b). The transport of tsunami-induced sediment into the deep sea (Kastens and Cita, 1981), which includes mass transport, was discussed by Shanmugam (2006b). Tsunami-related deposition in deep-water environments may be explained in four progressive steps (Fig. 7.6): 1. triggering stage

FIGURE 7.6

Depositional model showing the link between tsunamis and deep-water deposition. (A): 1. Triggering stage in which earthquakes trigger tsunami waves. 2. Tsunami stage in which an incoming (uprun) tsunami wave increases in wave height as it approaches the coast. 3. Transformation stage in which an incoming tsunami wave erodes and incorporates sediment, and transforms into sediment flows. (B): 4. Depositional stage in which outgoing (backwash) sediment flows (i.e., debris flows and turbidity currents) deposit sediment in deep-water environments. Suspended mud created by tsunami-related events would be deposited via hemipelagic settling. Source: After Shanmugam, G., 2006b. The tsunamite problem. J. Sediment. Res. 76, 718 730.

Mass Transport, Gravity Flows, and Bottom Currents

282

7. Triggering mechanisms of downslope processes

FIGURE 7.7 An aerial image showing sediment-rich backwash flows during the 2004 Indian Ocean tsunami. Note that position of shoreline has retreated seaward nearly 300 m (arrow) during the tsunami. Also note the development of backwash fans along the position of former shoreline, Kalutara Beach (south of Colombo), southwestern Sri Lanka. Image collected on December 26, 2004. Source: Image credit: Courtesy: DigitalGlobe. After Shanmugam, G., 2006b. The tsunamite problem. J. Sediment. Res. 76, 718 730, with permission from SEPM.

2. tsunami stage 3. transformation stage 4. depositional stage During the triggering stage, earthquakes, volcanic explosions, undersea landslides, and meteorite impacts can trigger displacement of a large quantity of water either up or down, causing tsunami waves. During the tsunami stage, tsunami waves carry energy traveling through the water, but these waves do not move the water, nor do they transport sediment. During the transformation stage, the tsunami waves erode and incorporate sediment into the incoming wave. The enormous tsunami waves are important triggering mechanisms of sediment failures. The advancing wavefront from a tsunami is capable of generating large hydrodynamic pressures on the seafloor that would produce soil movements and slope instabilities (Wright and Rathje, 2003). The transformation stage is evident in sediment-rich backwash flows during the 2004 Indian Ocean tsunami at Kalutara Beach, southwestern Sri Lanka (Fig. 7.7). The incoming ocean waters are clearly blue in color (implying sediment-free), but these waters transform into brown in color near the coast because of their incorporation of sediment. The transformation to brown color is the result of the wave breaking, and the wave will break in different water depths according to its wavelength and seafloor irregularities. Frohlich et al. (2009) documented huge exotic boulders from the Tongatapu Island, southwest Pacific where the largest boulder has dimensions of 15 m39 m311 m (Fig. 7.8). Frohlich et al. (2009) estimated masses of boulders to be in the range of 70 1600 metric tons. Such boulder emplacement could be attributed to the transformation stage and related sediment emplacement. These tsunami-induced backwash flows should not be confused with hyperpycnal flows introduced by river waters (Bates, 1953). Published accounts of MTDs generated by tsunamis are common (Fig. 7.9). High frequency of tsunami has been well documented (Table 7.2).

Mass Transport, Gravity Flows, and Bottom Currents

283

7.2 Origin

FIGURE 7.8 Photograph showing an exotic limestone boulder, interpreted to be emplaced by tsunami-related processes, in Tongatapu Island, south-west Pacific. Dimensions of the boulder are 15 3 311 3 39 m. See Frohlich et al. (2009) for more details. Source: Photo courtesy: C. Frohlich.

FIGURE 7.9 Published sedimentological features, including MTDs, a, b, c, associated with tsunami-related deposits by other authors. These features are also associated with cyclone-related deposits. Source: After Shanmugam, G., 2012b. Processsedimentological challenges in distinguishing paleo-tsunami deposits. Nat. Hazards, 63, 5 30, with permission from Springer.

During the final depositional stage, the outgoing sediment could generate slides, slumps, debris flows, and turbidity currents (Fig. 7.6). Although many sedimentary features are considered to be reliable criteria for recognizing potential paleo-tsunami deposits, similar features are also common in cyclone-induced deposits. At present, paleo-tsunami deposits cannot be distinguished from paleo-cyclone deposits using sedimentological features alone, without historical information (Shanmugam, 2008a, 2012b).

Mass Transport, Gravity Flows, and Bottom Currents

284

7. Triggering mechanisms of downslope processes

TABLE 7.2 Number of tsunami events (128) during 416 2007 (highstand) in the Indian Ocean region. Year

Month

Day

416 1722

10

1737

10

11

1750

Country/ocean

Geographic location

BB

NR

Indonesia

Java-S, Java

2

Indonesia

Java, Indonesia

0

India

Calcutta

X

1

Myanmar (Burma)

Burma Coast

X

0

1757

8

24

Indonesia

Java, Indonesia

1762

4

12

Indian Ocean

Bay of Bengal: Northern end

X

1

1770

Indonesia

SW Sumatra

X

1

1773

Indonesia

Java, Indonesia

Indonesia

SW Sumatra

1799

Indonesia

SE Sumatra

1

1814

Timor Sea

Timor Sea

0

1797

2

10

0

7.5

0 X

1

1815

4

10

Indonesia

Tambora

4

1815

11

22

Indonesia

Bali Sea

1

1816

4

29

Malaysia

Penang Island, Malacca Peninsula

1

1818

3

18

Indonesia

Bengkulu, Sumatra

1818

11

8

Indonesia

Bali Sea

1819

6

16

India Indonesia

1821

EM

7.0

1

7.0

1

8.5

Kutch

1

7.7

Flores Sea

0 0

6.8

3

8.2

X

1823

9

9

Indonesia

Java, Indonesia

1833

11

24

Indonesia

SW Sumatra

1836

3

5

Indonesia

Flores Sea

0

1836

11

28

Indonesia

Flores Sea

0

7.5

1837

9

Indonesia

Banda Aceh

1

7.2

1840

1

4

Indonesia

S. Java Sea

0

7.0

1842

11

11

Indian Ocean

Bay of Bengal: Northern end

X

3

1843

1

5

Indonesia

SW Sumatra

X

3

7.2

1843

2

7

Indonesia

Java, Indonesia

0

6.0

1845

6

19

India

Rann of Kutch

3

1847

10

31

India

Little Nicobar Island

1851

5

4

Indonesia

Lampung Bay, Indonesia

1852

11

11

Indonesia

Sibolga, Sumatra

X

X

X

1 1

X

1

6.8

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

285

7.2 Origin

TABLE 7.2 (Continued) Year

Month

Day

Country/ocean

Geographic location

1852

11

23

Indonesia

Java, Indonesia

0

1855

4

14

Indonesia

Timor

0

1856

7

25

Indonesia

Java-Flores Sea

0

1857

5

13

Indonesia

Bali Sea

2

1859

10

20

Indonesia

S. Java Sea

1

1861

2

16

Indonesia

SW Sumatra

X

9

1861

2

21

Indonesia

SW Sumatra

X

0

1861

3

9

Indonesia

SW Sumatra

X

4

7.0

1861

4

26

Indonesia

SW Sumatra

X

1

7.0

1861

6

5

Indonesia

Java, Indonesia

1861

6

17

Indonesia

SW Sumatra

X

0

6.8

1861

9

25

Indonesia

SW Sumatra

X

1

6.5

1862

4

8

Indonesia

Java, Indonesia

1

1863

3

16

Indonesia

Java, Indonesia

0

1864

2

16

Indonesia

SW Sumatra

X

0

1868

8

19

India

Andaman Islands

X

1

1881

12

31

Indian Ocean

W. of Car Nicobar Islands

X

11

1882

1

Sri Lanka

Sri Lanka

X

1

1883

8

26

Indonesia

Krakatau

X

5

1883

8

26

Indonesia

Sumatra

X

0

1883a

8

27

Indonesia

Krakatau

X

75

1883

10

10

Indonesia

Java-S. Java, Indonesia

1884

2

Indonesia

Krakatau

X

0

1885

7

29

Indonesia

Ajerbangis, Sumatra

X

0

1885

12

14

Indonesia

Banda Aceh

X

0

Indian Ocean

Bay of Bengal

X

0

1886

BB

NR

EM

8.5

0

0

6.8

1889

8

16

Indonesia

Java-S. Java, Indonesia

0

1889

11

23

Indonesia

Java, Indonesia

0

6.0

1891

10

5

Timor Sea

Timor Sea

0

7.0

1892

5

17

Malaysia

Malay Peninsula

4

1896

10

10

Indonesia

SW Sumatra

X

1

6.8

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

286

7. Triggering mechanisms of downslope processes

TABLE 7.2 (Continued) Year

Month

Day

Country/ocean

Geographic location

BB

1904

9

7

Indonesia

S. Java, Indonesia

0

1905

4

4

India

Kangra, India

1

8.6

1907

1

4

Indonesia

SW Sumatra

X

7

7.6

1908

2

6

Indonesia

SW Sumatra

X

1

7.5

1908

3

23

Timor Sea

Timor Sea

0

6.6

1909

6

3

Indonesia

Sumatra

0

7.6

1914

6

25

Indonesia

Indonesia

0

7.6

1914

7

26

Indonesia

Lais, Sumatra

1

1917

1

21

Indonesia

Bali Sea

0

1917

3

16

Indonesia

Java, Indonesia

0

1919

2

13

Timor Sea

Timor Sea

0

1921

9

11

Indonesia

S. Java Sea

0

1922

2

22

Indonesia

Banda

0

1922

4

10

Indonesia

SW Sumatra

1922

7

8

Indonesia

Lhoknga, Aceh

1926

6

28

Indonesia

SW Sumatra

X

1928

3

26

Indonesia

Krakatau

X

1928

8

4

Indonesia

Flores Sea

2

1930

3

17

Indonesia

Java-S. Java, Indonesia

0

1930

5

5

Myanmar (Burma)

Myanmar Coast

1930

6

19

Indonesia

1930

7

19

1931

9

1934

X

X

NR

EM

6.6

7.5

0 0

X

0

6.7

0

1

7.3

Java-S. Java Sea

0

6.0

Indonesia

S. Java Sea

0

6.5

25

Indonesia

SW Sumatra

0

7.4

1

15

India

Bihar-Nepal, Ganges River

1

8.4

1935

5

31

India

India

1

7.5

1935

11

25

Indonesia

Celebes Sea

0

6.5

1935

12

28

Indonesia

SW Sumatra

0

7.9

1936

8

23

Malaysia

Malay Peninsula

0

7.3

1941

6

26

India

Andaman Sea, E. Coast India

2

7.6

1945

11

27

Pakistan

Makran Coast

6

8.3

1948

6

2

Malaysia

Malay Peninsula

0

6.3

X

X

X

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

287

7.2 Origin

TABLE 7.2 (Continued) Year

Month

Day

Country/ocean

Geographic location

1949

5

9

Malaysia

Malay Peninsula

1950

8

15

India

Brahmaputra River, India

1955

5

17

Malaysia

1957

9

26

1958

4

1963

BB

NR

EM

0

6.7

1

8.7

Malay Peninsula

0

7.3

Indonesia

S. Java Sea

0

5.5

22

Indonesia

SW Sumatra

0

6.5

12

16

Indonesia

Java, Indonesia

0

6.5

1964

4

2

Indonesia

Off Northwest Coast of Indonesia

0

7.0

1967

4

12

Malaysia

Malay Peninsula

3

6.1

1974

12

28

Pakistan

Pakistan

1

6.2

1977

8

19

Indonesia

Sunda Islands

9

8.0

1979

7

18

Indonesia

Lembata Island

1

1979

8

9

Indonesia

Lomblen Island

1

1981

12

31

Indian Ocean

Bay of Bengal

1982

2

24

Indonesia

Java Trench, Indonesia

0

5.4

1982

3

11

Indonesia

Sumbawa Island

0

6.6

1982

12

25

Indonesia

Flores Sea

0

6.0

X

X

X

0

1983

11

30

India

Indian Ocean

2

7.7

1985

4

13

Indonesia

Bali Island, Indonesia

1

6.2

1987

11

26

Indonesia

Timor

1

6.6

1988

8

6

Bangladesh

Jamuna River, Aricha, Bangladesh

0

7.2

1991

7

4

Indonesia

Kalabahi, Alor, Indonesia

0

6.5

1992

12

12

Indonesia

Flores Sea

24

7.8

1994

2

15

Indonesia

Southern Sumatra

0

6.9

1994

6

2

Indonesia

Java, Indonesia

25

7.8

1994

6

3

Indonesia

Java, Indonesia

1

6.6

1994

6

4

Timor Sea

Timor Sea

1

6.5

1995

5

14

Indonesia

Timor

1

6.9

2000

6

18

India

South Indian Ocean

1

7.9

2002

9

13

India

Andaman Islands, India

0

6.5

2004

11

11

Indonesia

Kepulauan Alor

0

7.5

12

26

Indonesia

Off W. Coast of Sumatra

717

9.0

b

2004

X

X

X

X

(Continued)

Mass Transport, Gravity Flows, and Bottom Currents

288

7. Triggering mechanisms of downslope processes

TABLE 7.2 (Continued) Year

Month

Day

Country/ocean

Geographic location

2005

3

28

Indonesia

2005

4

10

2006

7

2007

9

BB

NR

EM

Indonesia

11

8.7

Indonesia

Kepulauan Mentawai

1

6.7

17

Indonesia

Java, Indonesia

20

7.7

12

Indonesia

Sumatra

4

8.2

X

a

1883 Krakatau volcanic eruption with 75 tsunami runups. 2004 Indian Ocean tsunami with 717 runups. Notes: No data available for the period 417 1721. BB, Bay of Bengal; X, Tsunami events that affected the Bay of Bengal during 1737 2007 (48); NR, Number of runups (998); EM, Earthquake magnitude (valid values: 0.0 9.9). Compiled from NGDC (National Geophysical Data Center) Tsunami Event Database, 2007. Tsunami events in Indonesia. ,http://www.ngdc. noaa.gov/nndc/struts/results?bt_0 5 &st_0 5 &type_8 5 EXACT &query_8 5 None 1 Selected&op_14 5 eq&v_14 5 &type_15 5 EXACT&query_ 15 5 None 1 Selected&type_7 5 Like&query_7 5 &st_1 5 &bt_2 5 &st_2 5 &bt_1 5 &bt_10 5 &st_10 5 &ge_9 5 &le_9 5 &bt_3 5 &st_3 5 & type_19 5 EXACT&query_19 5 83&op_17 5 eq&v_17 5 &type_18 5 EXACT&query_18 5 None 1 Selected&bt_20 5 &st_20 5 &bt_13 5 & st_13 5 &bt_16 5 &st_16 5 &bt_6 5 &st_6 5 &bt_11 5 &st_11 5 &d 5 7&t 5 101650&s 5 7. (accessed 10.11.07). b

7.2.5 Rogue waves Oceanic rogue waves (also known as freak waves, killer waves, monster waves, extreme events, abnormal waves, etc.) are surface gravity waves whose wave heights are much larger than expected for the sea state (i.e., at least twice as large as the significant wave height, Dysthe et al., 2008). Since the introduction of the term “Freak” oceanic wave by Draper (1964), the concept of rogue waves has received considerable attention in the field of mathematics, physics, and ocean navigation (e.g., Lavrenov, 1998; Dysthe et al., 2008; Akhmediev and Pelinovsky, 2010). Furthermore, measurements of rogue waves affecting oil platforms in the North Sea are well documented (Kjeldsen, 1984; Haver, 2004). The Draupner platform in the North Sea was affected by a rogue wave on January 1, 1995 (Shanmugam, 2012a, his Fig. 5.12). It was recorded by a laser instrument at an unmanned satellite platform. The crest height was 18.5 m above the mean water level and the wave height was 26 m. A similar wave affected the 2 4A platform in the Ekofisk oil field on January 3, 1984 (Kjeldsen, 1984). The crest height of this rogue was estimated to be at least 21 m above mean sea level. Even so, the concept of rogue waves is still an alien topic to most petroleum geologists when it comes to the study of deep-water sandstones. Extremely large rogue waves have been observed to occur near the southeastern shore of South Africa and have caused numerous cases of catastrophic ship collisions (Lavrenov, 1998). These dangerous waves are associated with the Agulhas Current (see Appendix A). Although there is considerable debate on the origin and mechanics of rogue waves (e.g., Ruban et al., 2010), studies have documented the occurrence of rogue waves in coastal zones (Chien et al., 2002) and have discussed their influence in shallow water (Soomere, 2010). Therefore, the role of rogue waves in triggering sediment failures in shallow-water coastal regions, similar to tsunami waves, should not be underestimated.

7.2.6 Cyclonic waves Tropical cyclones are meteorological phenomena. Structurally, tropical cyclones are large, rotating systems of clouds, winds, and thunderstorms. In the Northern Hemisphere Mass Transport, Gravity Flows, and Bottom Currents

7.2 Origin

289

the rotation is counter-clockwise, but in the Southern Hemisphere the rotation is clockwise due to the Coriolis force. Aspects of Hurricane Floyd that affected the U.S. Atlantic Coast were discussed in Chapter 5, Density Plumes: Types, Deflections, and External Controls (Figs. 5.22 and 5.23). Maximum measured velocities of cyclone-triggered bottom flows are commonly in the range of 100 300 cm s21 (39 5 117 in s21) on the shelf and 200 7000 cm s 1 (782730 in s21) in submarine canyons and troughs (Table 7.3). At these high velocities, even gravel-size grains would be eroded and transported into the deep sea. Like tsunami waves (Wright and Rathje, 2003), cyclonic waves are capable of generating large hydrodynamic pressures on the seafloor that produce soil movements and slope instabilities (Henkel, 1970; Bea et al., 1975). Submarine mudflows, triggered by the 1969 Category 5 Hurricane Camille, destroyed the offshore South Pass 70B platform in the Mississippi Delta area (Sterling and Strohbeck, 1975). Mudflows and mudslides, triggered by the 2004 Category 5 Hurricane Ivan (Nodine et al., 2006), toppled the Mississippi Canyon 20 platform at a depth of 146 m near the shelf edge (Hooper and Suhayda, 2005) and damaged up to 17 pipelines (MMS, 2005; Alvarado, 2006). I have already discussed the Loop Current-tropical cyclone connection in causing sediment failures associated with the Hurricane Katrina in the northern Gulf of Mexico (Section 4.6.4). High frequency of cyclones in the Bay of Bengal and the Arabian Sea has been well documented (Table 7.4). On the U.S. Pacific margin, a large slump mass of about 100,000 m3 (3,531,467 ft3) in size was triggered in the Scripps Canyon (California) by the May 1975 cyclone (Marshall, 1978). A sediment gravity flow within the Monterey Canyon (California), which was triggered by a cyclone on December 20, 2001, carried a current-meter package 550 m (1804 ft) downcanyon from its deployment site in less than 10 minutes and buried the instrument package within its deposit (Paull et al., 2003). In understanding the depositional effect of this cyclone, a 178 cm-long (69 in) vibracore (VC-80), taken from a depth of 1297 m (4254 ft), was composed of a basal pebbly sand unit and an upper muddy unit. Paull et al. (2003) interpreted this pebbly sand unit as a deposit of sediment gravity flow. This pebbly sand could be interpreted as sandy debrite. Perhaps the best-documented case study of cyclone-induced sedimentation in the deepwater has been by Hubbard (1992) on Hurricane Hugo. This hurricane, which passed over St. Croix on September 17, 1989, generated winds in excess of 110 knots (204 km hour21, Category 3 in Saffir-Simpson Scale) and waves 6 7 m (20 23 ft) in height. In the Salt River Submarine Canyon [.100 m (.328 ft) deep] offshore St. Croix, a current meter measured net downcanyon currents reaching velocities of 200 cm s21 (78 in s21) and oscillatory flows up to 400 cm s21 (156 in s21). Hugo caused erosion of 2 m (7 ft) of sand in the Salt River Canyon at a depth of about 30 m. More than 2 million kg of sediments were flushed down the canyon into deep-water in a matter of hours. The transport rate associated with Hurricane Hugo was 11 orders of magnitude greater than that measured during the fairweather period. In the Salt River Canyon, much of the soft reef cover (e.g., sponges) was eroded away by the power of the hurricane. Debris composed of palm fronds, trash, and pieces of boats found in the canyon were the evidence for hurricane-generated debris flows. In light of the available data on sediment transport by cyclonic waves and related currents, new insights on cyclone-induced deep-water sedimentation have evolved (Shanmugam, 2008a). Aspects of sediment transport by nearshore processes were discussed by Komar Mass Transport, Gravity Flows, and Bottom Currents

290

7. Triggering mechanisms of downslope processes

TABLE 7.3 Measured velocity values of cyclone-induced bottom flows in various submarine settings during the present highstand. Meteorological event (date)

Submarine setting (bathymetry)

1. Category 2 Hurricane Isabel (September 18, 2003) 2. Tropical Storm Delta (September 1973)

Shelf (Onslow Bay), North Carolina 30 (98 ft) Shelf, Gulf of Mexico 21 m (69 ft) Shelf (Eel), Northern California 50 m (164 ft) Shelf (Eel), Northern California 60 m (197 ft) Shelf (Texas), Gulf of Mexico 70 m (230 ft) Shelf, New Jersey 12 m (39 ft) Shelf (Onslow Bay), North Carolina 24 33 m (79 108 ft) Shelf (Atchafalaya), Gulf of Mexico 4.5 m (15 ft) Shelf (Alabama), Gulf of Mexico 89 m (292 ft) Shelf (Columbia River Mouth), Oregon 35 m (115 ft) Shelf, Gulf of Mexico 10 m (33 ft) Shelf, Great Barrier Reef, Australia 12 m (39 ft)

3. Unnamed cyclone (December 13, 1995) 4. Unnamed cyclone (October 28, 1999) 5. Category 5 Hurricane Allen (August 1980) 6. Tropical Storm Floyda (September 18, 1999) 7. Category 3 Hurricane Diana (September 11 13, 1984) 8. Category 4 Hurricane Lili (October 3, 2002) 9. Category 5 Hurricane Ivan (September 16, 2004) 10. Category 2 Unnamed hurricane (March 3, 1999) 11. Category 5 Hurricane Camille (August 1969) 12. Category 3 Hurricane Joy (December 1990) 13. Unnamed cyclone (January 7 11, 1989) 14. Category 5 Hurricane Ivan (September 2004) 15. Category 2 Hurricane Georges (September 24 28, 1998) 16. Unnamed cyclone (October 28, 1999) 17. Unnamed cyclone (February 2004) 18. Unnamed cyclone (December 17 19, 2002) 19. Unnamed cyclone (November 24, 1968) 20. Category 3 Hurricane Hugo (September 1989) 21. Category 1 Hurricane Iwa (November 1982) 22. Unnamed cyclone (August 1990) a

Slope, Middle Atlantic Bight 500 m (1640 ft) Slope (west of the DeSoto Canyon), Gulf of Mexico 500 1000 (1640 3280 ft) Canyon (Mississippi), Gulf of Mexico 300 m (984 ft) Canyon (Eel), Northern California 120 m (394 ft) Canyon (Cap de Creus), Gulf of Lions 300 m (984 ft) Canyon (Monterey), Northern California 1300 m (4264 ft) Canyon (Scripps), southern California 44 m (144 ft) Canyon (Salt River), St. Croix, V.I. . 100 m (328 ft) Reentrant (Kahe Point), Oahu, Hawaii 220 m (722 ft) Trough (Suruga), Japan . 500 m (1640 ft)

Category 4 Hurricane Floyd weakened to a Tropical Storm strength offshore New Jersey.

Mass Transport, Gravity Flows, and Bottom Currents

Velocity [cm s21 (in s21)] (reference) . 50 (20) (Wren and Leonard, 2005) 50 75 (20 29) (Forristall et al., 1977) 80 (31) (Cacchione et al., 1999) 88 (34) (Puig et al., 2003) 80 90 (31 35) (Snedden et al., 1988) 80 100 (31 39) (Kohut et al., 2006) 125 (49) (Mearns et al., 1988) 140 (55) (Allison et al., 2005) 150 (59) (Stone et al., 2005) . 150 (59) (Moritz., 2004) 160 (62) (Murray, 1970) 140- .300 (55- . 117) (Larcombe and Carter, 2004) 40 (16) (Brunner and Biscaye, 1997) 150 (59) (Mitchell et al., 2005) 68 (27) (Burden, 2000) 78 (30) (Puig et al., 2003) 80 (31) (Palanques et al., 2006) 150 500 1 (59 195 1 ) (MBARI, 2003) 190 (74) (Inman et al., 1976). 200 400 (78 156) (Hubbard, 1992) 300 (117) (Dengler et al., 1984) 7000 (2730) (Mitsuzawa et al., 1993)

291

7.2 Origin

TABLE 7.4 Frequency of tropical cyclones per year during 1945 2000 (highstand) in the Bay of Bengal and in the Arabian Sea. Year

Arabian Sea

Bay of Bengal

Year

Arabian Sea

Bay of Bengal

1945

1

13

1973

5

11

1946

1

16

1974

4

8

1947

2

16

1975

5

15

1948

5

13

1976

3

11

1949

1

11

1977

2

4

1950

0

16

1978

0

4

1951

3

12

1979

2

6

1952

1

16

1980

1

4

1953

0

10

1981

0

3

1954

3

11

1982

1

4

1955

0

13

1983

1

3

1956

2

12

1984

1

3

1957

3

4

1985

1

5

1958

0

12

1986

1

2

1959

4

12

1987

1

7

1960

4

11

1988

1

4

1961

4

13

1989

1

1

1962

2

11

1990

0

4

1963

4

12

1991

1

3

1964

3

12

1992

3

9

1965

1

13

1993

1

1

1966

2

17

1994

2

3

1967

0

1

1995

1

3

1968

0

13

1996

3

5

1969

1

13

1997

2

2

1970

3

12

1998

4

4

1971

3

15

1999

1

4

1972

4

14

2000

0

4

Total

57

344

Total

48

137

Grand total for the Bay of Bengal 5 481. Grand total for the Arabian Sea 5 105. Data from the Joint Typhoon Warning Center (Chu et al., 2002).

Mass Transport, Gravity Flows, and Bottom Currents

292

7. Triggering mechanisms of downslope processes

(1976). Snedden et al. (1988) concluded that during fair-weather periods, winds cannot generate bottom-current velocity sufficient in strength to transport fine sand beyond the shoreface region with water depths of 0 10 m (0 33 ft) during the present highstand (Fig. 7.10A). In contrast, convincing empirical data suggest that deep-water sedimentation is accelerated during storm-weather periods. Cyclone-generated debris were transported to reach depths ranging from 500 m (1640 ft) (Harmelin-Vivien and Laboute, 1986) to 1297 m (4254 ft) (Paull et al., 2003). Based on empirical data on sediment transport, a sedimentological model for demonstrating cyclone-induced transport of gravel, sand, and mud into the deep sea during sealevel highstand has been proposed (Fig. 7.10B). In this model, sand-size grains are transported across the open shelf, over the shelf edge, and via submarine canyons into the deep sea by cyclone-induced bottom flows. Sediment particles of all sizes and composition are transported by cyclone-triggered mass movements (slides and slumps) and sediment flows (cascading sand fall and debris flows) during the present highstand.

7.2.7 Internal waves and tides Aspects of internal waves and tides, based on 51 modern case studies in the world’s oceans, are discussed in Chapter 8, Bottom Currents (Shanmugam, 2013a, 2013b, 2014a,b). Internal waves are gravity waves that oscillate along oceanic pycnoclines. Internal tides are internal waves with a tidal frequency. Internal solitary waves (i.e., solitons), the most common type, are commonly generated near the shelf edge [100 200 m (328 656 ft) in bathymetry] and in the deep ocean over areas of seafloor irregularities, such as mid-ocean ridges, seamounts, and guyots. Empirical data from 51 locations in the Atlantic, Pacific, Indian, Arctic, and Antarctic Oceans reveal that internal solitary waves travel in packets. Internal waves commonly exhibit (1) higher wave amplitudes [5 50 m (16 64 ft)] than surface waves [ , 2 m (6.56 ft)], (2) longer wavelengths [0.5 15 km (0.31 9 mi)] than surface waves [100 m (328 ft)], (3) longer wave periods (5 50 minutes) than surface waves (9 10 s), and (4) higher wave speeds [0.5 2 m s21 (1.64 6.56 ft s21)] than surface waves [25 cm s21 (10 in s21)]. Maximum speeds of 48 cm s21 (19 in s21) for baroclinic currents were measured on guyots. These oceanographic phenomena can cause sediment failures at various depths that are controlled by points of intersections of pycnoclines with the seafloor (Fig. 7.11) and by the bathymetry of shelf edges (Fig. 7.12). Rivera-Rosario et al. (2017) discussed bed failure induced by internal solitary waves.

7.2.8 Ebb-tidal currents Transport of coastal sand into the deep ocean by ebb-tidal currents during the present highstand has been documented in the offshore areas of Hervey Bay and Fraser Island, southeast Australia (Boyd et al., 2008). The Hervey Bay area is characterized by (1) mesotidal range [2 4 m (6.5 13 ft)] (Boyd and Leckie, 2004), (2) high-velocity (150 cm s21 or 59 in s21) ebb-tidal currents, (3) narrow (5 25 km or 3 16 mi) shelf widths, and (4) numerous slope gullies. Furthermore, ebb-tidal currents have been ascribed to initiation of sediment gravity flows in submarine canyons (Kottke et al., 2003; Boyd et al., 2008).

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 7.10 (A) Highstand sedimentological model showing calm shelf waters and limited extent of sediment transport in the shoreface zone (short green arrow) during fair-weather periods. Shoreface bottom-current velocities during fair weather are in the range of 10 2 20 cm s21 (Snedden et al., 1988). The shelf edge at 200 m water depth separates shallow-water (shelf) from deep-water (slope) environments. (B) Highstand sedimentological model showing sediment transport on the open shelf, over the shelf edge, and in submarine canyons during periods of tropical cyclones (storm weather) into deep-water (long red arrow). Mass-transport processes are commonly induced by intense hurricanes (e.g., Hurricane Katrina, 2005). Source: Modified after Shanmugam, G., 2008a. The constructive functions of tropical cyclones and tsunamis on deep-water sand deposition during sea level highstand: implications for petroleum exploration. AAPG Bull. 92, 443 471.

FIGURE 7.11 (A) Conceptual diagram showing the intersection of pycnoclines with sloping seafloor topography with increasing bathymetry. Note that most pycnoclines (76%) intersect the seafloor in water depths shallower than 200 m (656 ft.). Pycnoclines that intersect the sloping seafloor near the shelf edge are of significance. Hypothetical increase in the density of fluid layer with increasing bathymetry is shown by r1, r2, r3, r4, and r5. (B) The number of cases plotted in different bathymetric intervals is from Table 1 of Shanmugam (2013a). (C) Bathymetric intervals. These intervals are selected to be consistent with the intervals chosen for shelf edges (Fig. 7.12). Source: From Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811. AAPG.

FIGURE 7.12 Conceptual diagrams showing the frequency of occurrences of shelf edges with increasing bathymetry. (A) Twelve percent of shelf edges occur in water depths between 50 and 99 m (164 and 325 ft.). (B) Eighty-one percent of shelf edges occur in depths between 100 and 200 m (328 and 656 ft.). (C) Five percent of shelf edges occur in water depths between 201 and 500 m (659 and 1640 ft.). (D) Two percent of shelf edges occur between 501 and 1000 m (1644 and 3281 ft.). Bathymetric data are from Table 2 of Shanmugam (2013a). Source: From Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811. AAPG.

296

7. Triggering mechanisms of downslope processes

FIGURE 7.13 Sites of non-seismic MTDs of the world that occurred during 2006 08, with emphasis on the Asia-Pacific region. There is a strong influence from the South Asian monsoon on the high frequency of MTD in the Asia-Pacific region (Petley, 2010a). Other factors, such as tropical cyclones, earthquakes, and volcanic activity are also important in initiating MTD in this region. The background image is the ETOPO digital elevation model, with the darker colors indicating higher ground. Source: From Petley, D.N. 2010b. An analysis of fatal landslides in the Asia-Pacific region for 2006 to 2008. Dave’s Landslide Blog. http://daveslandslideblog.blogspot.com/2010/02/analysis-offatal-landslides-in-asia.html (accessed 19.11.10) and http://daveslandslideblog.blogspot.com/2010/02/maps-of-global-fatallandslides.html (accessed 18.11.10).

Prior et al. (1982) documented the initiation of sediment failures in fjords and coastal regions during periods of low tides. When the fall in water level is sufficiently rapid, then the elevated water table on the slope would lead to a hydraulic push downward, which destabilizes the slope and triggers sediment failures. This process is called “rapid drawdown” (Lambe and Whitman, 1969, p. 477).

7.2.9 Monsoon flooding In the Asia-Pacific region, the frequent occurrence of MTDs during 2006 08 was controlled by nonseismic factors, such as monsoonal rainfall and tropical cyclones (Figs. 7.13 and 7.14). Empirical data show that seasonal monsoon in July is the primary controlling factor that triggers maximum number of MTDs in East Asia. In the Bay of Bengal (Indian Ocean), the water body has been affected by reversal in current circulation twice a year due to double monsoon seasons (Gangadhara Rao and Shree Ram, 2005). In the Bay of Bengal, sand supply into the deep sea and related deep-water sedimentation (Weber et al., 1997) occurs during sea-level highstand due to increasing monsoon intensity (Goodbred, 2003). Empirical data show that heavy rainfall has been a major factor in initiating debris flows in central Chile (Sepu´lveda and Padilla, 2008). The 1996 Finneidfjord slide in Norway has been related to heavy rainfall combined with rock blasting (Longva et al., 2003). Finally, river discharge is an important trigger of sediment failures at river mouths (Calre et al., 2016).

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 7.14 The predictive model, developed at NASA’s Goddard Space Flight Center in Greenbelt, Maryland, estimates potential landslide activity triggered by rainfall. Rainfall is the most widespread trigger of MTD around the world. If conditions beneath Earth’s surface are already unstable, heavy rains act as the last straw that causes mud, rocks, or debris—or all combined—to move rapidly down mountains and hillsides. Source: NASA.

298

7. Triggering mechanisms of downslope processes

7.2.10 Groundwater seepage Bro¨nnimann (2011) has discussed the role of groundwater seepage in triggering MTDs. She has recognized seven triggering mechanisms associated with hydrogeology: (1) suction, (2) rising pore-water pressure, (3) seepage forces, (4) inner erosion, (5) liquefaction, (6) overpressure, and (7) mechanisms related to high plasticity. Although these processes are important in affecting slope stability, only groundwater seepage is considered a triggering mechanism (Table 7.1).

7.2.11 Wildfire Wildfire-initiated debris flows in subaerial environments have been documented in the Storm King Mountain, Colorado (Cannon et al., 2001). Stevens et al. (2008) discussed the importance of postwildfire debris-flow occurrences in the Three Lakes Watershed in Colorado. Wildfire leads to significant hydrological and geomorphological changes by weathering of bedrock surfaces and by changing of soil structure and related properties (Shakesby and Doerr, 2006). These changes would result in initiating mass movements in subaerial environments, which in turn extend into adjacent lacustrine and marine environments (e.g., California).

7.2.12 Human activity The second most common triggering mechanism (next to earthquakes) for known historic submarine mass movements has been human activity (Mosher et al., 2010). A welldocumented modern example is the 1979 extension of the international airport at Nice, France (Dan et al., 2007). In 1979, a catastrophic event occurred on the Nice continental slope (French Riviera) generating the lost of human lives and important material damages. Part of the new harbor constructed at the edge of the International Airport of Nice collapsed into the sea. The data show the existence of a sensitive clay bed between 30 and 45 m below the seafloor. Under high deviatoric load a sensitive clay layer underwent an important creep, which dramatically decreased its resistance and caused the slope failure. This hypothesis was supported by the good agreement between the maximum thickness of the failure surface and the depth of the sensitive clay layer (Fig. 7.15). Numerical calculations demonstrated that creeping of the sensitive clay layer could be at the origin of the 1979 slide (Dan et al., 2007). In addition, the exceptionally heavy rainfall which occurred before the accident and consequently the seepage of freshwater probably induced the decrease of the effective stress and accelerated sediment creeping and triggered the Nice slope failure. A progressive and relatively long-term creeping failure scenario is in good agreement with the official report mentioning cracks, settlements, failures, and embankment collapses occurred during landfilling operations (Fig. 7.15). Although human activity is included here for completeness (Table 7.1), it is of no consequence for interpreting ancient stratigraphic record in Earth Sciences.

Mass Transport, Gravity Flows, and Bottom Currents

Mediterranean Sea Clay High-permeability sand Clay

(A) Human activity

Embankment

Creeping

(B) Rainfall Embankment

Fresh water conduit

Failure surfa ce (primary gli de plane)

Increase in pore-water pressure (decrease in slope stability)

(C) Zone of sediment failure

FIGURE 7.15 Illustration of the 1979 sediment failure that occurred at the Nice International Airport in southern France. The Nice sediment failure has been attributed to a combination of both external and internal factors (Dan et al., 2007). (A) Internal (in situ) lithologic factor composed of clay and sand layers; (B) human factor involving the building of airport embankment; (C) external meteorological and internal geotechnical factors. See text for details. Diagram is based on the concept of Dan et al. (2007, their Fig. 20). Source: With permission from Elsevier.

300

7. Triggering mechanisms of downslope processes

7.2.13 Tectonic oversteepening Tectonic compression has elevated the northern flank of the Santa Barbara Basin and overturned the slope in southern California along the U.S. Pacific margin (Greene et al., 2006). Several thrust faults exist beneath the shelf, at the shelf break, and along the middle to upper slope. Uplift (thrusting) along these slopes has led to oversteepening and sediment failure. This sediment failure, known as the Goleta Slide, is composed of slumps and mud flows (Fig. 2.18). Samples from this failure contain gravel, sand, and mud lithofacies. This 300-year-old sediment failure occurred during the present highstand.

7.2.14 Glacial maxima and loading During most glacial maxima, ice tends to advance onto the continental shelf, reach the shelf break, and deliver sediments directly into the upper continental slope as MTDs. On glaciated margins (Dowdeswell and O’Cofaigh, 2002), a single event composed of submarine slides and debris flows typically last only a few hours (Elverhøi et al., 2002). Glacial loading is common and it operates on a wide range of scale, which varies from small-scale mass wasting processes in fjords to large-scale slides covering several thousand square kilometers on glaciated continental margins. Factors which are significant in glacial loading induced landslides are the flexing of crust due to the loading and unloading of a fluctuating ice front (e.g., Lindberg et al., 2004), variation in drainage and groundwater seepage, quick deposition of low-plasticity mud, and rapid formation of glacial sediments above interglacial hemipelagic sediments. Examples of glacial-maxima MTDs composed of glacigenic debrits have been documented (Dowdeswell et al., 1998; King et al., 1998). Morner (1991) discussed intense earthquakes and seismotectonics as a function of glacial isostasy. Submarine mass movements have been attributed to glacial loading and unloading along the Scotian Margin in the North Atlantic (Mosher et al., 2004). A review of known ages of submarine MTD along the margins of the Atlantic Ocean shows a relatively even distribution of large MTD with time from the last glacial maximum until about 5000 years after the end of glaciation (Lee, 2009). Based on the spatial distribution of MTD on the U.S. Atlantic Margin (Shanmugam, 2012a, his Fig. 5.20), Twichell et al. (2009) suggested that earthquakes associated with rebound of the glaciated part of the margin or earthquakes associated with salt domes were the primary triggering mechanism of sediment failures.

7.2.15 Salt movements In the Gulf of Mexico, salt structures are commonly the basis for developing attractive petroleum plays. Salt is an integral part of slope environments in the Gulf of Mexico (Fig. 7.16). Sigsbee Escarpment represents the base-of-slope where MTDs are ubiquitous (Fig. 7.16). Submarine mass movements along the flanks of intraslope basins have been related to mobilization of underlying salt masses in the Gulf of Mexico (Tripsanas et al., 2004). The importance of salt diapirism in triggering MTD in the Gulf of Mexico has been discussed by Prior and Hooper (1999). The Cape Fear slide complex on the U.S. Atlantic Margin (see Fig. 5.20) has been attributed to faulting and related upward migration of salt (Hornbach et al., 2007). A seismic transect across the Cape Fear diapir itself reveals the dramatic seafloor relief and associated slump (Fig. 7.17). The Cape Fear diapir protrudes as much as 275 m above the surrounding slide scar and is Mass Transport, Gravity Flows, and Bottom Currents

301

7.2 Origin

FIGURE 7.16

3D perspective of modern Gulf of Mexico, based on high-resolution bathymetric swath data, showing intraslope basins. The Sigsbee Escarpment marks the seaward edge of the salt deformation province. Salt diapirs have created numerous domes and isolated mini basins on the slope. Note highly irregular sea-floor topography controlled by salt tectonics. Basins commonly range in width from 10 to 20 km. GB 5 Garden Banks area, GC 5 Green Canyon area, MTD 5 mass-transport deposits (slide/slump/debris flow), MB 5 mini basin, LD 5 linear depression, MC 5 Mississippi Canyon, and KC 5 Keathley Canyon. Source: After Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476, with permission from Elsevier.

FIGURE 7.17 A transect showing the Cape Fear diapir and adjacent mass-transport deposit (slump) on the U.S. Atlantic margin. The Cape Fear diapir protrudes as much as 275 m above the surrounding slide scar and is bounded by faults. At the location of this crossing, the northern diapir is covered by slide rubble carried downslope. Source: Image and descriptions are courtesy of C. Ruppel. http://oceanexplorer.noaa.gov/explorations/03windows/logs/aug01/media/diapirline.html (accessed 04.04.11). After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Amsterdam, Elsevier, p. 524. Elsevier.

bounded by faults. These examples clearly demonstrate the genetic link between salt tectonics and triggering of MTD.

7.2.16 Depositional loading Delta-front mass movements (slides, slumps, and mudflows) have been related to rapid rate of sedimentation and depositional loading in the Mississippi River Delta area Mass Transport, Gravity Flows, and Bottom Currents

302

7. Triggering mechanisms of downslope processes

(Coleman and Prior, 1982). In this case, bottleneck slides, bar-front slumps, debris-flow chutes, and debris-flow “lobes” have been reported (Fig. 7.18). These are muddy MTD. In 1975, the South Pass mudflow moved completely across the shelf and onto the upper continental slope in water depths greater than 500 m (Coleman and Garrison, 1977). The failure of delta-front sands could also generate sandy mass-transport deposit (SMTD). The rapid sedimentation and related generation of sandy slides from a “lava delta” of the Kilauea volcano is an example of SMTDs. SMTDs are probably more common on modern slope environments than they have been reported. An example is the interpretation of delta-slope derived sandy debris flow in the Kitimat Arm fjord, British Columbia (Nemec, 1990, his Fig. 28). Modern mass movements have been attributed to depositional oversteepening near the mouth of the Magdalena River in Colombia (Heezen, 1956).

7.2.17 Hydrostatic loading Slope-failure deposits in offshore Cape Licosa and in the Paola slope basin originated during the late Quaternary sea-level rise on the eastern Tyrrhenian margin (Trincardi et al., 2003). These slope failures were attributed to rapid drowning of unconsolidated sediment, which resulted in increased hydrostatic loading. Such a loading would enhance pore pressure and result in sediment failure.

7.2.18 Ocean-bottom currents Vigorous ocean-bottom currents have caused erosion of the abyssal seafloor covering thousands of square kilometers (Berggren and Hollister, 1977; Tucholke and Embley, 1984; Shanmugam, 1988). In the Rockall Trough region, bottom currents associated with the North Atlantic Deep-Water have caused an erosive area extending over 8,500 km2 in water depths of 500 2,000 m (Howe et al., 2001). Such currents are capable of generating slope instability and related sediment failures due to oversteepening caused by current-induced scouring (Locat and Lee, 2002). The close association between the Storegga Slide and contourites on the Norwegian continental margin has been discussed by Solheim et al. (2005b). Although the Storegga Slide was triggered by earthquakes, Laberg and Camerlenghi (2008) observe the close association of the slide with contourites (see Fig. 2.32) and suggest that contour-current associated factors favor slope instability. These factors are (1) mounded geometry of contourites, (2) growth of contourite mounds on inclined seafloor, (3) well-sorted contourite sediment, and (4) environment rich in organic carbon. Such factors are prone to induce sediment failures because of associated high sedimentation rates, liquefaction (Sultan et al., 2004), excess pore pressure, and gas-hydrate dissociation on the Norwegian margin.

7.2.19 Biological erosion Dillon and Zimmerman (1970) investigated the Block and Corsair submarine canyons incised into the New England continental shelf. They suggested that biological erosion has been the primary cause of deterioration of canyon walls. In the two canyons, the sediment surface has been subjected to intense burrowing by organisms. The

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 7.18 Schematic diagram showing lobe-like distribution of submarine debris flows in front of the Mississippi River Delta, Gulf of Mexico. This shelf-edge deltaic setting, associated with high sedimentation rate, is prone to develop ubiquitous debris flows, mud diapers, and contemporary faults due to depositional loading. Mud diaper, intrusion of mud into overlying sediment causing dome-shaped structure; chute, channel. This region is subjected to frequent and intense cyclonic waves (e.g., Hurricane Katrina in 2005; Shanmugam, 2008a) that also generate mass movements resulting in the destruction of petroleum platforms and pipelines. Source: From Coleman, J.M., Prior, D.B., 1982. Deltaic environments. In: Scholle, P.A., Spearing, D. (Eds.), Sandstone Depositional Environments, AAPG Memoir 31, pp, 139 178. AAPG.

304

7. Triggering mechanisms of downslope processes

burrowing results in weakening of the canyon walls, which often leads to slumping and retreat of the canyon walls to the angle of repose. In addition, burrowing activity by organisms breaks down the semiindurated larger rock fragments into smaller grains. This continuous biologic activity has been an important triggering mechanism of MTD in the Block and Corsair canyons. Shepard and Dill (1966, their Fig. 21) photographed truncated animal burrows, gouge marks, and erosion scars on the rock wall of a tributary to Sumner Branch of the Scripps Canyon, California. Submarine mass movements have also been associated with erosion of the walls and floors of submarine canyons by invertebrates and fishes and boring by animals in offshore canyons, California (Warme et al., 1978).

7.2.20 Gas-hydrate decomposition Gas hydrates form when methane and water freeze at high pressures and relatively low temperatures. These conditions occur in the shallow part of marine sedimentary sections on many continental margins. Studies have indicated that gas hydrates, which lie beneath many submarine slopes, may have contributed to the triggering of submarine slides. An example is the Cape Fear slide on the U.S. Atlantic margin (Shanmugam, 2012a, his Fig. 5.20). Embley (1980) first recognized this slide complex on echo sounder profiles and cores. The main headwall scarp of this slide occurs on the lower slope at 2,600 m water depth (Popenoe et al., 1993). This slide is characterized by its association with extruded salt diapirs and sediments with gas hydrates (Shanmugam, 2012a, his Fig. 5.25). Popenoe et al. (1993, their Fig. 3) documented the strong bottom-simulating reflector (BSR) indicative of the base of the gas-hydrate interval (Shanmugam, 2012a, his Fig. 5.26). The gashydrate interval extends from the seafloor to the BSR. A combination of factors, which include (1) decomposition of gas hydrates and related high pore-fluid pressures, (2) intrusion of salt diapers and related fracturing and oversteepening, and (3) occurrence of earthquakes, has been suggested as possible triggering mechanisms of the Cape Fear slide (Popenoe et al., 1993). Paull et al. (2000) investigated the Ocean Drilling Program Leg 164 sediments with gas hydrates in the area of the Blake Ridge (Shanmugam, 2012a, his Fig. 5.25). Paull et al. (1996) suggested an increase in frequency of seafloor slumping on continental margins containing gas hydrates. Mienert et al. (2005) attributed the failure of the Storegga Slide on the mid-Norwegian margin to excess pore pressures, caused by gas-hydrate dissociation, due to changes in sea level and water temperature. During lowstands of sea level, groundwaters may discharge on the continental slope and gas-hydrate decomposition may occur in slope sediments (Kvenvolden, 1993). Gas-hydrate decomposition could also occur during highstands of sea level. Secular warming of bottom water during interglacials, caused by competition and deflection of water masses, could cause gas-hydrate decomposition and consequent slope failure (Driscoll et al., 2000).

7.2.21 Sea-level lowstand In the petroleum industry, the sea-level lowstand model is the perceived norm for explaining deep-water sands (Fig. 7.19). An example is the attribution of reservoir sands

Mass Transport, Gravity Flows, and Bottom Currents

FIGURE 7.19 (A) Generalized models showing eustatic control of sedimentation during periods of lowstand and highstand. (B) Conceptual highstand model showing accumulation of pelagic and/or hemipelagic muddy lithofacies over the shelf, slope, and basin. (C) Conceptual lowstand model showing the development of sandy lithofacies at the base of slope. In reality, however, both lowstand and highstand deposits are composed of mixed lithofacies (i.e., gravel, sand, and mud). Concepts are partly after Shanmugam and Moiola (1988). Source: After Shanmugam, G., 2008a. The constructive functions of tropical cyclones and tsunamis on deep-water sand deposition during sea level highstand: implications for petroleum exploration. AAPG Bull. 92, 443 471. Reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

306

7. Triggering mechanisms of downslope processes

in the Kutei Basin in the Makassar Strait (Indonesian Seas) to a lowstand of sea level (Saller et al., 2006). Nonetheless, the Kutei Basin’s location is frequently affected by earthquakes, volcanoes, tsunamis, tropical cyclones, monsoon floods, the Indonesian throughflow, and M2 baroclinic tides (Shanmugam, 2008b, 2013a). These daily activities of the solar system (e.g., earthquakes, meteorite impacts, tsunamis, cyclonic waves, etc.) do not come to a halt during sea-level lowstands. These short-term events are the primary triggering mechanisms of deep-water sediment failures. Contrary to the lowstand model, empirical data suggest that tropical cyclones and tsunamis are the two most underrated phenomena when it comes to understanding sediment transport into the deep sea during periods of sea-level highstands (Shanmugam, 2008a). Given the high frequency of cyclones and tsunamis during the present highstand, deposition of deep-water sand is not unique to periods of lowstands (Fig. 7.20). Thus, the lowstand model is inappropriate for explaining the timing of deep-water sands. For example, deep-sea sand deposition, triggered by tsunamis and cyclones can also occur during highstands (Shanmugam, 2008a). In discussing the La Jolla highstand fan in the California

FIGURE 7.20 (A) Conventional sea-level model showing the popular belief that deep-water deposition of sand occurs during periods of lowstand and deposition of mud occurs during periods of highstand (Shanmugam and Moiola, 1988; Vail et al., 1991). The present highstand is estimated to represent a period of 20,000 years. BP 5 before present. (B) 200,000 cyclones are estimated to occur during the present highstand in the Indian Ocean (Bay of Bengal) and the Atlantic Ocean based on data from Shanmugam (2008a) (C) 140,000 tsunamis are estimated to occur during the present highstand in the Pacific Ocean based on data from Shanmugam (2008a). The implication is that sands can be deposited during periods of sea-level highstands, thus making the conventional model obsolete. Source: After Shanmugam, G., 2008a. The constructive functions of tropical cyclones and tsunamis on deep-water sand deposition during sea level highstand: implications for petroleum exploration. AAPG Bull. 92, 443 471. AAPG.

Mass Transport, Gravity Flows, and Bottom Currents

7.3 Synopsis

307

borderland, Covault et al. (2007, p. 786) state, “Contrary to widely used sequence stratigraphic models, lowstand fans are only part of the turbidite depositional record, and this analysis reveals that a comparable volume of coarse clastic sediment has been deposited in California borderland deep-water basins regardless of sea level.” The importance of short-term events, such as tsunamis and cyclones, can occur irrespective of sea-level changes. But there are no criteria to distinguish tsunami-generated sands from cyclone-generated sands (Shanmugam, 2012b). More importantly, there are no criteria to distinguish tsunami-generated sands from earthquake-generated sands into the deep sea during long periods of sea-level lowstands or highstands (Shanmugam, 2007, 2012b).

7.3 Synopsis On continental margins, there are 21 possible triggers of sediment failures that generate downslope gravity flows. Short-term triggers are composed of earthquakes, meteorite impacts, volcanic activities, tsunamis, tropical cyclones, etc. Intermediate-term triggers are composed of tectonics, glacial maxima, etc. Long-term events are sea-level changes. However, the prevailing notion that deep-water deposits develop during periods of sealevel lowstands is a myth. In the petroleum industry, sea-level lowstand model is the perceived norm for explaining deep-water sands. An extreme example is the attribution of reservoir sands in the Kutei Basin in the Makassar Strait (Indonesian Seas) to lowstand of sea level (Saller et al., 2006), despite the basin’s location affected by earthquake belts, active volcanoes, tsunamigenic zones, cyclonic belts, monsoon floods, the Indonesian throughflow, and M2 baroclinic tides. However, the daily activities of the solar system (e.g., earthquakes, meteorite impacts, tsunamis, etc.) do not come to a halt during sea-level lowstands. These short-term events are the primary triggering mechanisms of deep-water sediment failures. The lowstand model is irrelevant for explaining the triggering of deepwater sandy mass-transport sands.

Mass Transport, Gravity Flows, and Bottom Currents

C H A P T E R

8 Bottom currents O U T L I N E 8.1 Introduction

310

8.6.9 Bedform-velocity matrix 346 8.6.10 Abyssal plain contourites 347 8.6.11 Sandy intervals of contourite facies models 348

8.2 Vertical continuum: surface currents, deep-water masses, and bottom currents 310 8.3 The thermohaline circulation

313

8.4 Four types of bottom currents

320

8.7 Wind-driven bottom currents 8.7.1 The Gulf Stream and the Loop Current 8.7.2 Current velocity 8.7.3 Ewing Bank 826 Field: a case study

8.5 Thermohaline-induced geostrophic bottom currents (i.e., contour currents) 321 8.5.1 Origin of the Antarctic bottom water 322 8.5.2 Seismic geometries of contourites 322 8.5.3 Current velocity 323 8.5.4 Traction structures 324 8.6 The contourite problem 327 8.6.1 Dual forcing of global ocean circulation 328 8.6.2 Continuum between turbidity currents and contour currents 331 8.6.3 Revision of the basic principle of contour currents 333 8.6.4 Hiatuses in contourites 336 8.6.5 Origin of erosional features 337 8.6.6 Gulf of Cadiz as the type locality 338 8.6.7 The contourite facies model 341 8.6.8 Traction structures and shale clasts 344

Mass transport, gravity flows, and bottom currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00008-4

349 349 349 352

8.8 Tidal bottom currents in submarine canyons 356 8.8.1 Types of submarine canyons 356 8.8.2 Current velocity 357 8.8.3 Identification 359 8.9 Baroclinic currents (internal waves and internal tides) 363 8.9.1 Basic concept 363 8.9.2 Empirical data 364 8.9.3 Depositional framework 365 8.10 Sediment provenance 8.10.1 Current directions 8.10.2 Detrital composition

366 366 369

8.11 Reservoir quality

370

8.12 Synopsis

373

309

© 2021 Elsevier Inc. All rights reserved.

310

8. Bottom currents

8.1 Introduction In order to provide a complete story on deep-water processes, the objective of this section is to discuss deep-water bottom currents and their deposits from oceanographic and sedimentological viewpoints. The domain of bottom currents has a long history of contributions on both physical oceanography and process sedimentology (Wu¨st, 1933; Stommel, 1958; Heezen and Hollister 1964; Hubert, 1964; Hsu¨, 1964; Heezen et al., 1966; Klein, 1966; Hollister, 1967; Hollister and Heezen, 1972; Pequegnat, 1972; Bouma and Hollister, 1973; Shepard et al., 1979; Stow and Lovell, 1979; Shanmugam et al., 1993a,b; Apel et al., 2006; Herna´ndez-Molina et al., 2006; Zenk, 2008; St. Laurent et al., 2012; Talley, 2013; Rebesco et al., 2014; Shanmugam, 2018a; Eberli and Betzler, 2019; Mulder et al., 2019; de Castro et al., 2020; Fonnesu et al., 2020; Fuhrmann et al., 2020, among others). The influence of the European research community on contourite research is evident in three Geological Society of London publications (Stow and Piper, 1984; Stow et al., 2002; Viana and Rebesco, 2007). The European influence is even more striking in a thematic volume “Contourites” edited by Rebesco and Camerlenghi (2008). Of the 25 chapters in the volume, 22 (88%) are from the European research community (Table 8.1). The U.S. Atlantic Margin is characterized by both downslope mass-transport deposit (MTD) and alongslope contour currents (Fig. 8.1). However, both processes are originally triggered by downslope gravity-driven mechanisms. These complexities need to be explained in developing a clear understanding of deep-water processes, which is one of the objectives of this volume. In achieving this objective, datasets from 35 case studies worldwide have been used (Fig. 8.2, Table 8.2).

8.2 Vertical continuum: surface currents, deep-water masses, and bottom currents A sound knowledge of global ocean surface currents is critical for understanding ocean bottom currents (Gill, 1982; Apel, 1987; Stewart, 2008; CIMAS, 2015). This is because surface currents and bottom currents are interrelated entities in the world’s oceans. For example, the three segments of the Southern Ocean, composed of (1) the upper surface currents, (2) the middle deep-water masses, and (3) the lower bottom currents, form a vertical continuum (Fig. 8.3). Surface ocean currents are commonly wind-driven entities that exhibit clockwise rotation in the Northern Hemisphere and counter-clockwise rotation in the Southern Hemisphere. Surface currents, which are mostly restricted to the upper 400 m of the ocean, make up nearly 10% of the water in the world’s oceans. The deep-water masses in the world’s oceans are caused by differences in temperature and salinity. When sea ice forms in the polar regions due to freezing of shelf waters, seawater experiences a concurrent increase in salinity due to salt rejection and a decrease in temperature. The increase in the density of cold saline (i.e., thermohaline) water directly beneath the ice triggers the sinking of the water mass down the continental slope (Fig. 8.3) and the spreading of the water masses to other parts of the ocean (Fig. 8.4). These are called thermohaline water masses.

Mass transport, gravity flows, and bottom currents

311

8.2 Vertical continuum: surface currents, deep-water masses, and bottom currents

TABLE 8.1 Contourite research contributions by country for the 25 chapters in the edited volume “Contourites” (Rebesco and Camerlenghi, 2008).

Chapter Contribution (chapter title)

Authorship

First author’s affiliated institution or residence by country

1

Contourite research: a field in full development

M. Rebesco, A. Camerlenghi, and A.J. Van Loon

Italy

2

Personal reminiscences on the history of contourites

K.J. Hsu¨

United Kingdom

3

Methods for contourite research

J.A. Howe

United Kingdom

4

Abyssal and contour currents

W. Zenk

Germany

5

Deep-water bottom currents and their deposits

G. Shanmugam

United States

6

Dynamics of the bottom boundary layer

S. Salon, A. Crise, and A.J. Van Loon

Italy

7

Sediment entrainment

Y. He, T. Duan, and Z. Gao

China

8

Size sorting during transport and deposition of fine sediments: sortable silt and flow speed

I.N. McCave

United Kingdom

9

The nature of contourite deposition

D.A.V. Stow, S. Hunter, D. Wilkinson, and F.J. Herna´ndez-Molina

United Kingdom

10

Traction structures in contourites

J. Martı´n-Chivelet, M.A. Fregenal-Martı´nez, and B. Chaco´n

Spain

11

Bioturbation and biogenic sedimentary structures in contourites

A. Wetzel, F. Werner, and D. A.V. Stow

Switzerland

12

Some aspects of diagenesis in contourites

P. Giresse

France

13

Contourite facies and the facies model

D.A.V. Stow and J.-C. Fauge`res

United Kingdom

14

Contourite drifts: nature, evolution and controls

J.-C. Fauge`res and D.A.V. Stow

France

15

Sediment waves and bedforms

R.B. Wynn and D.G. Masson

United Kingdom

16

Seismic expression of contourite depositional systems

T. Nielsen, P.C. Knutz, and A. Denmark Kuijpers

17

Identification of ancient contourites: problems and palaeoceanographic significance

H. Hu¨neke and D.A.V. Stow

Germany

18

Abyssal plain contourites

F.J. Herna´ndez-Molina, A. Maldonado, and D.A.V. Stow

Spain (Continued)

Mass transport, gravity flows, and bottom currents

312

8. Bottom currents

TABLE 8.1 (Continued)

Chapter Contribution (chapter title)

Authorship

First author’s affiliated institution or residence by country

19

Continental slope contourites

F.J. Herna´ndez-Molina, E. Llave, and D.A.V. Stow

20

Shallow-water contourites

G. Verdicchio and F. Trincardi Italy

21

Mixed turbidite contourite systems

T. Mulder, J.-C. Fauge`res and E. Gonthier

France

22

High-latitude contourites

T. van Weering, M. Stoker, and M. Rebesco

The Netherlands

23

Economic relevance of contourites

A.R. Viana

Brazil

24

Palaeoceanographic significance of contourite drifts

P.C. Knutz

Denmark

25

The significance of contourites for submarine slope stability

J.S. Laberg and A. Camerlenghi

Norway

Spain

Note that 22 chapters (88%) represent contributions from the European Research Community, and three chapters (5, 7, and 23) are from non-European countries (United States, China, and Brazil).

FIGURE 8.1 Comparison of downslope mass lows and their deposits (i.e., debrites, left map) (Embley, 1980; with alongslope contour currents and their deposits (i.e., contourites, right map) (Flood and Hollister, 1974) on the U.S. Atlantic Margin. Source: Shanmugam, G., 2017b. Contourites: physical oceanography, process sedimentology, and petroleum geology. Pet. Explor. Dev. 44 (2), 183 216. Elsevier. Mass transport, gravity flows, and bottom currents

8.3 The thermohaline circulation

313

FIGURE 8.2 Map showing the locations of case studies used In this review, which include 22 critical case studies by other researchers (locations A V), and locations of studies by other researchers that resulted in recent debates on deep-water processes (locations G, K, and N). Note 35 locations of core and outcrop descriptions of deep-water sandstones with traction structures that were interpreted by the present author as products of bottom-current reworking (Table 9.2). Source: Blank world map credit: http://upload.wikimedia.org/wikipedia/commons/8/83/ Equirectangular_projection_SW.jpg (accessed 24.01.16.).

8.3 The thermohaline circulation Stommel (1958) first developed the concept of the global circulation of thermohaline water masses and the vertical transformation of light surface waters into heavy deep-water masses in the oceans. Broecker (1991) presented a unifying concept of the global oceanic “conveyer belt” by linking the wind-driven surface circulation with the thermohalinedriven deep circulation regimes. The large-scale horizontal transport of water masses, which also sink and rise at select locations, is known as the “thermohaline circulation” or THC. Aspects of thermohaline circulation are discussed by Zenk (2008). The global overturning circulation has been presented by Talley (2013) (Fig. 8.4). Examples of selected deep-water masses in various parts of the world’s oceans and their acronyms are given below. 1. 2. 3. 4. 5. 6. 7. 8. 9.

AABW: Antarctic bottom water (Fig. 8.3) ABW: Arctic bottom water AAIW: Antarctic intermediate water (Brazilian margin) ACC: Antarctic circumpolar current (Antarctica) AW: Atlantic water (Mediterranean sea) BC: Brazil current BICC: Brazil intermediate counter current CDW or CPDW: Circumpolar deep-water DGSRF: Deep Gulf Stream return flow

Mass transport, gravity flows, and bottom currents

314

8. Bottom currents

TABLE 8.2 Summary of 21 locations of published case studies on deep-marine bottom currents by other researchers that are used in this chapter (locations: A U shown by filled squares, see Fig. 8.2).

Location symbol and number in Fig. 8.2

Case studies

Data: thickness of core and outcrop described by the author (not applicable to studies by other researchers)a

Comment (this chapter)

A. Case study: Blake Modern contour Plateau and Blake-Bahama currents Outer Ridge (Heezen et al., 1966)

Echo sounding, bottom Introduction of basic photographs, sediment cores concept of contour currents

B. Case study: Straits of Florida (Mullins et al., 1980)

Modern sandy carbonate contourites

Seismic profiles, cores, rocks Porosity and permeability recovered by dredging and data (Table 8.4) in situ sampling

B. Case study: Straits of Florida (Lu¨dmann et al., 2016)

Bottom currents

Hydroacoustic data, highresolution multichannel seismic reflection data, conductivity, temperature, and depth (CTD) casts and sampling

C. Case study: Argentine Basin (Klaus and Ledbetter, 1988)

Modern muddy contourites

High-resolution seismic Sheet-like sediment waves records (3.5 kHz echograms)

D. Case study: Eirik Drift (Stanford et al., 2011)

Deep western boundary CTD and lowered acoustic current (DWBC) Doppler current profiler measurements E. Case study: Weddell Sea Cyclonic circulation of Current-meter records (Michels et al., 2002) the Weddell Gyre

Modern carbonate mounds

Eirik Drift, South of Greenland Velocity: 24 cm s21

F. Case study: Gulf of Cadiz (Fauge`res et al., 1984; Gonthier et al., 1984; Stow and Fauge`res, 2008)

Modern Faro contourite 3.5 kHz seismic profiles, drift sediment cores

Discussion of problematic contourite facies model (discussed in this chapter)

F. Case study: Gulf of Cadiz (Herna´ndez-Molina et al., 2006; Garcia et al., 2009)

Modern Faro contourite Seismic profiles, bottom Discussion of complex drift photographs, sediment cores origin of erosional features (discussed in this chapter)

F. Case study: Gulf of Cadiz (Mulder et al., 2013)

Modern Faro contourite Sediment cores, grain-size Discussion of problematic drift analysis, thin-section studies contourite facies model in terms of velocity (discussed in this chapter)

F. Case study: Gulf of Cadiz (Stow et al., 2013)

Modern Cadiz Channel

2 gravity cores and over Discussion of problematic 3000 submarine photographs origin contourite sands (Stow et al., 2013) (discussed in this chapter)

G. Case study: NE Spain (Pomar et al., 2012)

Ricla Section, Upper Jurassic

1 outcrop section (Ba´denas et al., 2012; Pomar et al., 2012)

Discussion of problematic internal-wave and internaltide deposits (Shanmugam, 2013a,b) (Continued)

Mass transport, gravity flows, and bottom currents

315

8.3 The thermohaline circulation

TABLE 8.2 (Continued)

Location symbol and number in Fig. 8.2

Case studies

Data: thickness of core and outcrop described by the author (not applicable to studies by other researchers)a

Comment (this chapter)

H. Case study: Southern Adriatic Sea (Chiggiato et al., 2016)

Dense water

Various measurements and sampling

Velocity of bottom currents: 40 50 cm s21

I. Case study: Israel (Bein and Weiler, 1976)

Ancient sandy carbonate contourite (Cretaceous Talme Yafe Formation)

Outcrop and core

Sediment prism

J. Case study: Southeast of Agulhas Current South Africa (Bryden et al., 2005)

K. Case study: China (He et al., 2011)

Ningxia, Middle Ordovician

Mooring deployment off Fort Edward from coast out to 203 km offshore; maximum depth of measurement: 2200 m

69.7 Sverdrups (1 Sv or Sverdrup 5 106 m3 s21) at 31 S. Poleward velocity is .100 cm s21 at or above 100m depth on mooring B Several outcrop sections (He Discussion of problematic et al., 2011) internal-wave and internaltide deposits (Shanmugam, 2012b, 2014b)

L. Case study: South China Modern contourites Sea (Yu et al., 2014)

Bottom simulating reflectors

Gas hydrates

M. Case study: West Philippine Sea (Lien et al., 2015)

Kuroshio and Luzon Undercurrent

Field experiment

The annual Kuroshio transport is 16 6 4 Sv

N. Case study: Makassar Strait (Saller et al., 2006)

Kutei Basin, Miocene

2 wells? (Saller et al., 2006, 2008a,b)

Discussion of deep tidal currents (Shanmugam, 2008c)

N. Case study: Makassar Strait (Dunham and Saller, 2014)

Kutei Basin, Miocene

2 wells (Saller et al., 2006, 2008a, b)

Reply to a discussion on the reservoir quality of bottomcurrent reworked sands (Shanmugam, 2014a)

O. Case study: Off Fraser Island, SE Australia (Boyd et al., 2008)

Deep-marine tidal bottom currents

Regional bathymetry and multibeam echo sounding

Highstand transport of coastal sand to the deep ocean

P. Case study: Canterbury Drifts, SW Pacific Ocean (Carter, 2007)

Subantarctic mode water (SAMW), Antarctic intermediate water (AAIW) Modern sandy volcaniclastic contourites

ODP 1119

Planar-bedded units up to several meters thick

Seismic and side-scan sonar data, seafloor photo, grab samples, piston core

Sheet contourites

Q. Case study: Offshore of the Pennell Coast, Antarctica (Rodriguez and Anderson, 2004)

(Continued)

Mass transport, gravity flows, and bottom currents

316

8. Bottom currents

TABLE 8.2 (Continued)

Location symbol and number in Fig. 8.2

Case studies

Data: thickness of core and outcrop described by the author (not applicable to studies by other researchers)a

R. Case study: Meiji Drift Emperor seamount chain (Kerr et al., 2005)

Thermohaline circulation (THC)

S. Case study: Horizon Guyot, Mid-Pacific Mountains (Lonsdale et al., 1972)

Baroclinic currents reworking sediments on flat tops of towering guyot terraces

Narrow-beam echosounding system, a pair of side-looking sonars, a 3.5kHz seismic profiler and a proton magnetometer, and deep-sea cameras

T. Case study: Equatorial Pacific (Dubois and Mitchell, 2012)

Large-scale sediment redistribution by bottom currents

Digital seismic reflection data and wireline logging data

Bottom-current induced resedimentation

U. Case study, Monterey Canyon, U.S. Pacific Margin (Shepard et al., 1979)

Deep-marine tidal bottom currents

Current meter

Velocity of bottom currents: 30 cm s21 (both upcanyon and downcanyon)

V. Santos Basin, SW Atlantic Ocean (Duarte and Viana, 2007)

Miocene erosional channels

3D seismic reflection profiles Erosional channel geometry

1. Gulf of Mexico, United States (Shanmugam et al., 1988b)

1. Mississippi Fan, Quaternary, DSDP Leg 96

B500 m DSDP core (selected intervals described)

Modern submarine fan

1067 m Conventional core and piston core 25 wells

Sandy mass-transport deposits and bottom-current reworked sands common

650 m Conventional core 3 wells

Sandy mass-transport deposits and bottom-current reworked sands

1. Gulf of Mexico, United 2. Green Canyon, late States (Shanmugam et al., Pliocene, 1993a,b; 1995b; 3. Garden Banks, Shanmugam and Zimbrick, middle Pleistocene 1996) 4. Ewing Bank 826, Pliocene-Pleistocene 5. South Marsh Island, late Pliocene 6. South Timbalier, middle Pleistocene 7. High Island, late Pliocene 8. East Breaks, late Pliocene-Holocene 2. California (Shanmugam 9. Midway Sunset and Clayton, 1989; Field, upper Shanmugam, 2006a, 2012a) Miocene, onshore

Seismic data

Comment (this chapter) Dimensions: Thickness: 1800 m Length: .1000 km Width: B350 km Asymmetrical dunes and ripples Bathymetry of bedforms: 1630 32 m

(Continued)

Mass transport, gravity flows, and bottom currents

317

8.3 The thermohaline circulation

TABLE 8.2 (Continued)

Location symbol and number in Fig. 8.2

Case studies

Data: thickness of core and outcrop described by the author (not applicable to studies by other researchers)a

Comment (this chapter)

3. Ouachita Mountains, 10. Jackfork Group, Arkansas and Oklahoma, Pennsylvanian United States (Shanmugam and Moiola, 1995)

369 m 2 outcrop sections

Sandy mass-transport deposits and bottom-current reworked sands common

4. Southern Appalachians, Tennessee, United States (Shanmugam, 1978; Shanmugam and Benedict, 1978; Shanmugam and Walker, 1978, 1980)

11. Sevier Basin, Middle Ordovician

2152 m 5 outcrop sections

Ancient submarine fan

5. Brazil (Shanmugam, 2006a, 2012a)

12. Lagoa Parda Field, 200 m Conventional core lower Eocene, 10 wells Espirito Santo Basin, onshore 13. Fazenda Alegre Field, upper Cretaceous, Espirito Santo Basin, onshore 14. Cangoa Field, upper Eocene, Espirito Santo Basin, offshore 15. Peroa´ Field, lower Eocene to upper Oligocene, Espirito Santo Basin, offshore 16. Marlim Field, Oligocene, Campos Basin, offshore 17. Marimba Field, upper Cretaceous, Campos Basin, offshore 18. Roncador Field, upper Cretaceous, Campos Basin, offshore

Sandy mass-transport deposits and bottom-current reworked sands common

6. North Sea (Shanmugam et al., 1995a)

19. Frigg Field, lower Eocene, Norwegian North Sea

Sandy mass-transport deposits and bottom-current reworked sands common

3658 m Conventional core 50 wells

(Continued)

Mass transport, gravity flows, and bottom currents

318

8. Bottom currents

TABLE 8.2 (Continued)

Location symbol and number in Fig. 8.2

7. U.K. Atlantic Margin (Shanmugam et al., 1995a)

8. Norwegian Sea and vicinity (Shanmugam et al., 1994)

9. French Maritime Alps, Southeastern France (Shanmugam, 2002a, 2003a)

Case studies

Data: thickness of core and outcrop described by the author (not applicable to studies by other researchers)a

20. Harding Field (formerly Forth Field), lower Eocene, U.K. North Sea 21. Alba Field, Eocene, U.K. North Sea 22. Fyne Field, Eocene, U.K. North Sea 23. Gannet Field, Paleocene, U.K. North Sea 24. Andrew Field, Paleocene, U.K. North Sea 25. Gryphon Field, upper Paleocenelower Eocene, U.K. North Sea 26. Faeroe area, Paleocene, west of the Shetland Islands 27. Foinaven Field, Paleocene, West of the Shetland Islands

Thickness included in the N. Sandy mass-transport Sea count deposits and bottom-current 1 well reworked sands common; Conventional core contourites have been 1 well reported (Damuth and Olson, 2001)

28. Mid-Norway region, Cretaceous, Norwegian Sea 29. Agat region, Cretaceous, Norwegian North Sea 30. Annot Sandstone, Eocene-Oligocene

10. Nigeria (Shanmugam, 31. Edop Field, 1997b; Shanmugam, 2006a, Pliocene, offshore 2012a) 11. Equatorial Guinea 32. Zafiro Field, (Famakinwa et al., 1996; Pliocene, offshore Shanmugam, 2006a, 2012a) 33. Opalo Field, Pliocene, offshore

Comment (this chapter)

500 m Conventional core 14 wells

Sandy mass-transport deposits and bottom-current reworked sands common

610 mb 1 outcrop section (12 units described)

Sandy mass-transport deposits and bottom-current reworked sands common (deep tidal currents)

875 m Conventional core 6 wells

Sandy mass-transport deposits and bottom-current reworked sands common (deep tidal currents)

294 m Conventional core 2 wells

Sandy mass-transport deposits and bottom-current reworked sands common

(Continued) Mass transport, gravity flows, and bottom currents

8.3 The thermohaline circulation

319

TABLE 8.2 (Continued)

Location symbol and number in Fig. 8.2 12. Gabon (Shanmugam, 2006a, 2012a)

13. Bay of Bengal, India (Shanmugam, et al., 2009)

Case studies 34. Melania Formation, lower Cretaceous, offshore (includes four fields) 35. Krishna-Godavari Basin, Pliocene

Total thickness of rocks described by the author

Data: thickness of core and outcrop described by the author (not applicable to studies by other researchers)a

Comment (this chapter)

275 m Conventional core 8 wells

Sandy mass-transport deposits and bottom-current reworked sands common

313 m Conventional core 3 wells 11,463 m

Sandy debrites and tidalites common

a

The rock description of 35 case studies of deep-water systems comprises 32 petroleum-producing massive sands worldwide. Description of core and outcrop was carried out at a scale of 1:20 1:50, totaling 11,463 m, during 1974 2011, by G. Shanmugam as a Ph.D. student (1974 78), as an employee of Mobil Oil Corporation (1978 2000), and as a consultant (2000 11). Global studies of cores and outcrops include a total of 7832 m of conventional cores from 123 wells, representing 32 petroleum fields worldwide (Shanmugam, 2015a). These modern and ancient deep-water systems include both marine and lacustrine settings. b The Peira Cava outcrop section was originally described by Bouma (1962), and later by Pickering and Hilton (1988, their Fig. 62), among others. Note conventional core and outcrop description carried out by the present author worldwide (locations: 1 13, filled circles, see Fig. 8.2). Traction structures of bottom-current origin are common in all 35 case studies carried out by the author.

10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28.

DWBUC or DWBC: Deep western boundary undercurrent IDW: Indian deep-water ITF: Indonesian throughflow LCDW: Lower circumpolar deep-water LIW: Levantine intermediate water (Mediterranean sea) MOW: Mediterranean outflow water MUC: Mediterranean undercurrent NADW: North Atlantic deep-water NAdDW: North Adriatic dense water NPDW: North Pacific deep-water (Japan) NSDW: Norwegian sea deep-water PDW: Pacific deep-water SACW: South Atlantic central water (Brazilian margin) SOW: Sea overflow water UCDW: Upper circumpolar deep-water WBUC or WBU: Western boundary undercurrent WDW: Warm deep-water (Antarctica) WSBW: Weddell sea bottom water (Antarctica) WSDW: Weddell sea deep-water (Antarctica)

The thermohaline circulation and related deep-marine bottom currents in modern oceans became popular when Heezen et al. (1966) reported deep-water masses and related contour currents along the continental rise in the U.S. Atlantic margin. An example of

Mass transport, gravity flows, and bottom currents

320

8. Bottom currents

FIGURE 8.3 A conceptual model of the Southern Ocean showing three vertical segments, composed of the upper surface currents, the middle deep-water masses, and the lower bottom currents, forming a vertical continuum (left). Note the origin of AABW by freezing of shelf waters (right). As a consequence, the increase in the density of cold saline (i.e., thermohaline) water triggers the sinking of the water mass down the continental slope and the spreading of the water masses to other parts of the ocean. Source: Modified after Hannes Grobe, April 7, 2000. http://en. wikipedia.org/wiki/File:Antarctic_bottom_water_hg.png (accessed 18.05.11.). Figure from Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524, with permission from Elsevier.

such deep-water mass is the Antarctic bottom water (AABW). AABW was first identified by Brennecke (1921) in the northwest corner of the Weddell Sea in the Antarctic region.

8.4 Four types of bottom currents The general term “bottom current” is used because it covers a variety of bottom currents of different origins, flow directions, and velocities (Shanmugam et al., 1993a, p. 1242). In deep-water environments, the family of bottom currents is distinguished from gravitydriven downslope turbidity currents on both oceanographic and sedimentologic aspects (Shanmugam, 2012a). Southard and Stanley (1976) recognized five types of bottom currents at the shelf break based on their origin. These currents are generated by (1) thermohaline differences, (2) wind forces, (3) tidal forces, (4) internal waves, and (5) surface waves. In

Mass transport, gravity flows, and bottom currents

8.5 Thermohaline-induced geostrophic bottom currents (i.e., contour currents)

321

FIGURE 8.4 Map showing the global overturning circulation (GOC). The location of Gulf of Cadiz is added in this article. This site served as the type locality for the contourite facies model (see Section 8.6.7 in the text). The global circulation is not important in interpreting the primary sediment provenance at a given site. A simpler version of thermohaline circulation (THC) pattern was first published by Broecker (1991); later simplified by Rahmstorf (2002, 2006). Source: Modified after Talley, L.D., 2013. Closure of the global overturning circulation through the Indian, Pacific, and Southern Oceans: schematics and transports. Oceanography 26 (1), 80 97, with permission from the Oceanography Society.

addition, tsunami-related traction currents have been speculated to occur in bathyal waters (Yamazaki et al., 1989), but the mechanics of such currents are not yet understood (Shanmugam, 2008a, 2012b). I have selected four basic types of deep-water bottom currents, namely (1) thermohaline-induced geostrophic bottom currents (i.e., contour currents), (2) wind-driven bottom currents, (3) deep-water tidal bottom currents, and (4) baroclinic currents associated with internal waves and internal tides for discussion here.

8.5 Thermohaline-induced geostrophic bottom currents (i.e., contour currents) The deep-water component of these water masses that winnow, rework, and deposit sediment on the seafloor for a sustained period of time is called thermohaline-induced bottom

Mass transport, gravity flows, and bottom currents

322

8. Bottom currents

currents. These currents are also known as contour currents because of their tendency to follow bathymetric contours of continental slope and rise. In addition to thermohaline-induced bottom currents or contour currents, there are three other major types, namely wind-driven, tide-driven, and internal tide-driven bottom currents.

8.5.1 Origin of the Antarctic bottom water As discussed in Chapter 3, Gravity Flows: Debris Flows, Grain Flows, Liquefied/ Fluidized Flows, Turbidity Currents, Hyperpycnal Flows, and Contour Currents, the origin of thermohaline water masses are best studied using the AABW (Gordon, 2001, 2019, 2013; Purkey et al., 2018, among five others). The AABW is initiated as downslope gravity flows on the continental slope (Fig. 3.59). The AABW has a density of 0.03 g cm23 with temperatures ranging from 20.8 C to 2 C (35 F), salinities from 34.6 to 34.7 psu. Being the densest water mass of the oceans (Purkey et al., 2018), AABW is found to occupy the depth range below 4000 m. The AABW is formed in the Weddell and Ross Seas, off the Ade´lie Coast and by Cape Darnley from surface water-cooling in polynyas and below the ice shelf. A unique feature of AABW is the cold surface wind blowing off the Antarctic continent (Fig. 3.59). The surface wind creates the polynyas (i.e., an area of open water surrounded by sea ice), which opens up the water surface to more wind. This Antarctic wind is stronger during the winter months and thus the AABW formation is more pronounced during the Antarctic winter season.

8.5.2 Seismic geometries of contourites Rebesco et al. (2014) discussed the similarity in seismic geometry between turbidite channel-levee systems and contourite drifts. The difference is that contourite drifts show asymmetric moat and mound geometry, whereas turbidites exhibit symmetric gull-wing geometry (Fig. 8.5). But in the rock record, it would be difficult to distinguish one from the other on seismic profiles. Seismic channel geometry (Fig. 8.6) has been reported for the Santos Channel, a 100 km-long channel-like gutter at the foot of an intraslope escarpment, in the Santos Basin in offshore Brazil by Duarte and Viana (2007). This channel was excavated by the strong northward-flowing Southern Ocean current (SOC) during early Miocene. Analogous to the present-day AABW, the SOC was an alongslope-flowing contour current. The SOC not only formed the channel by erosion, but also deposited the adjacent contourite drifts on both sides. The Santos Channel and adjacent contourite drifts (Fig. 8.6) resemble those of turbidite channels with adjacent levee complexes. In the stratigraphic record, distinguishing contourite channels that align parallel to the slope from turbidite channels that generally align perpendicular to the regional slope is a challenge. However, turbidity currents may flow parallel to the strike of the regional slope along trench floors (Underwood and Bachman, 1982). Rebesco et al. (2014) reviewed contourite drift geometries, such as mounded (Fig. 8.7) and sheet. Sheet geometries with continuous and parallel reflections, for example, are associated with contourite deposits in offshore Brazil (Fig. 8.6) and in offshore Norway (see Fig. 2.32).

Mass transport, gravity flows, and bottom currents

8.5 Thermohaline-induced geostrophic bottom currents (i.e., contour currents)

323

FIGURE 8.5 Schematic model showing ideal, large-scale differences between contourite drifts (A) and channel-levee systems (B). Source: From Rebesco, M., Herna´ndez-Molina, F. J. Van Rooij, D., Wa˚hlin, A., 2014. Contourites and associated sediments controlled by deep-water circulation processes: State-of-the-art and future considerations. Marine Geology 352, 111 154, with permission from Elsevier.

8.5.3 Current velocity Maximum current velocities of bottom currents in different parts of the world’s oceans are summarized in Table 8.3. Measured current velocities usually range from 1 to 20 cm s21 (Hollister and Heezen, 1972); however, exceptionally strong, near-bottom currents with maximum velocities of up to 300 cm s21 were recorded in the Strait of Gibraltar (Gonthier et al., 1984). Bottom-current velocities of 73 cm s21 were measured at a water depth of 5 km on the lower continental rise off Nova Scotia (Richardson et al., 1981). Heezen and Hollister (1971) suggested that at extremely high bottom velocities of over 100 cm s21, relict pockets of sand and gravel may occur on the barren seafloor. According to Bulfinch and Ledbetter (1983/ 1984), a Deep western boundary undercurrent (DWBUC) flows south along the North American continental slope and rise between 1000 and 5000 m. The DWBUC has a 300-km

Mass transport, gravity flows, and bottom currents

324

8. Bottom currents

FIGURE 8.6 Seismic block diagram showing channel geometry of the Santos Channel caused by the erosion of seafloor by the along-slope flowing (northward) SOC during early Miocene. The paleocurrent direction is same as the present-day direction of the AABW. Cross-sectional channel geometry in seismic profiles could be misinterpreted as turbidite channels in other areas. Note levee-like geometry on both sides of the channel. Also note continuous and parallel seismic reflections of contourite deposits on the right-hand side showing sheet geometry. After Duarte, C.S.L., Viana, A.R., 2007, Santos Drift System: stratigraphic organization and implications for late Cenozoic palaeocirculation in the Santos Basin, SW Atlantic Ocean. In: Viana, A.R., Rebesco, M. (Eds.), Economic and palegeographic significance of contourite deposits. Geological Society, London, Special Publications, 276, 171 198, with permission from the Geological Society of London. Figure from Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524. Elsevier.

wide high-velocity zone, with a maximum measured velocity of 73 cm s21, which winnows both fine and very fine silt and results in deposition of medium and coarse silt.

8.5.4 Traction structures The genetic term contourite was originally introduced for deposits of thermohalineinduced contour currents in the deep oceans (Hollister, 1967). Traction structures, such as ripple cross laminae (Fig. 8.8A) and associated sharp upper contacts, are common in contourites. My studies worldwide have shown that traction structures are ubiquitous in bottom-current deposits of all four kinds (Fig. 8.9) (Shanmugam et al., 1993a,b). MartınChivelet et al. (2008) have also documented the common occurrence of traction structures in contourites (Fig. 8.10). This has created problems in distinguishing deep-sea sediments with traction structures either as turbidites, using the Bouma Sequence, or as contourites (Bouma and Hollister, 1973). However, the importance of traction structures in contourite has been documented worldwide (Hsu¨, 1964; Hubert, 1964; Klein, 1966; Hollister, 1967;

Mass transport, gravity flows, and bottom currents

8.5 Thermohaline-induced geostrophic bottom currents (i.e., contour currents)

325

FIGURE 8.7 (A) Location map showing Expedition 339 sites (yellow solid circles) in the Gulf of Cadiz and West Iberian margin. Note red line through Sites U1386 and U1387 showing the position of seismic profile shown in Fig. 8.7B. (B) North-South seismic profile of the Pliocene-Quarternary Faro Drift showing parallel reflections, Gulf of Cadiz. See position of this profile in Fig. 8.7A. See calibrated core intervals in two Sites U1386 and U1387. BQD 5 Lower Quaternary seismic reflector; M 5 Messinian seismic reflector. Source: (A) Map from Herna´ndezMolina et al. (2016b) (245). (B) Modified after Alonso et al. (2016) (203). After Shanmugam, G., 2017b. Contourites: physical oceanography, process sedimentology, and petroleum geology. Pet. Explor. Dev. 44 (2), 183 216. Elsevier.

Mutti, 1992; Stanley, 1993; Ito, 2002; Martın-Chivelet et al., 2008; Mutti and Carminatti, 2011; Shanmugam, 2008a,b, 2012a). Bottom currents may flow parallel to the strike of the regional slope, may flow in circular motions (gyres) unrelated to the slope, or may flow up and down submarine canyons (tidal), whereas turbidity currents commonly flow downslope (Fig. 8.11). In areas where both downslope sandy debris flows and alongslope bottom currents operate concurrently (Fig. 8.12), the reworking of the tops of sandy debrites by bottom currents may be expected. Such a scenario, common on continental margins, could generate a basal massive sand division and an

Mass transport, gravity flows, and bottom currents

326

8. Bottom currents

TABLE 8.3 Maximum current velocities of bottom currents in 33 different parts of the world’s oceans. Study area 1. Straits of Gibraltar, Mediterranean outflow water (MOW) (Gonthier et al., 1984; see also Herna´ndezMolina et al., 2013) 2. Upper slope, Offshore Brazil, Equatorial Atlantic (Viana et al., 1998) 3. Gulf of Mexico, Loop Current (Cooper et al., 1990) 4. Green Canyon 166 area, Gulf of Mexico. Drilling operations were temporarily suspended in August of 1989 because of high current velocities that reached 153 cm s21 (Koch et al., 1991) 5. Faeroe Bank Channel, North Atlantic (Crease, 1965) 6. Southeast of South Africa, Agulhas Current (Bryden et al., 2005) 7. Rise, Off Nova Scotia, North Atlantic (Richardson et al., 1981) 8. Base of North American Continental Rise (Bulfinch and Ledbetter, 1983/1984) 9. Trench, Ryukyu Trench, Japan (Tsuji, 1993) 10. Samoan Passage, Western South Pacific (Hollister et al., 1974) 11. Southern Adriatic Sea (Chiggiato et al., 2016) 12. Hebrides Slope, North Atlantic (Howe and Humphrey, 1995) 13. Faeroe-Shetland Channel, North Atlantic (Akhurst, 1991) 14. Rise, near Hatteras Canyon, North Atlantic (Rowe, 1971) 15. Carnegie Ridge, Eastern Equatorial Pacific (Lonsdale and Malfait, 1974) 16. SE of Iceland, North Atlantic (Steele et al., 1962)

Depth (m) (dominant driving mechanism, this chapter)

Maximum current velocity (cm s21)

400 1400 (thermohaline)

300

200 (thermohaline)

300

100 (wind-driven)

204

45 (wind-driven)

153

760 (thermohaline)

109

,100 (thermohaline)

100

5000 (thermohaline)

73

5022 (thermohaline)

73

340 (tidal)

51

(?)

50

100 130 (dense water triggered by 40 50 February 2012 severe cold spell) 403 468 (thermohaline)

48

900 (thermohaline)

33

(thermohaline)

33

1000 2000 (?)

.30

2100 slope (thermohaline)

30

(thermohaline)

30

4000 4600 (thermohaline)

30

3000 5000 (thermohaline)

26.5

4300 5200 (thermohaline)

26

17. Argentine Basin, Western South Atlantic (Ewing et al., 1971) 18. Amirante Passage, Western Indian Ocean (Johnson and Damuth, 1979) 19. Rise, Off New England, North Atlantic (Zimmerman, 1971) 20. Blake-Bahama Outer Ridge, North Atlantic (Amos et al., 1971) 21. Off North Carolina, North Atlantic (Rowe and Menzies, 1968) 22. Weddell Sea, Antarctica (Michels et al., 2002)

1500 4000 (thermohaline)

25

4440 (thermohaline)

24

23. Off Cape Cod, North Atlantic (Volkman, 1962)

10 3200 (thermohaline)

21.5 (Continued)

Mass transport, gravity flows, and bottom currents

327

8.6 The contourite problem

TABLE 8.3 (Continued) Study area

Depth (m) (dominant driving mechanism, this chapter)

Maximum current velocity (cm s21)

24. Off Cape Hatteras, North Atlantic (Barrett, 1965)

(thermohaline)

21

25. Greater Antilles Outer Ridge, North Atlantic (Tuholke et al., 1973) 26. Off Blake Plateau, North Atlantic (Swallow and Worthington, 1961) 27. Kuroshio Current, off East Coast of Taiwan (Rudnick et al., 2015) 28. Tonga Trench and vicinity, Western South Pacific (Reid, 1969) 29. Western North Atlantic (Wust, 1950)

5300 5800 (thermohaline)

20

3300 3500 (thermohaline)

20

20 (thermohaline)

20

.4800 (?)

19

2000 3000 (thermohaline)

17

30. West Bermuda Rise, North Atlantic (Knauss, 1965)

5200 (thermohaline)

17

31. Scotia Ridge , Antarctic Circumpolar Current, Antarctica (Zenk, 1981)

3008 (wind-driven) (Howe et al., 1997)

17b

32. Greenland-Iceland-Faeroes Ridge, North Atlantic (Worthington and Volkman, 1965) 33. Antillean-Caribbean Basin (outer), North Atlantic (Wust, 1963)

2000 3000 (thermohaline)

12

4000 8000 (thermohaline)

10

a

a

Antarctic circumpolar current has both wind-driven and thermohaline-driven components (CIMAS, 2015). One-year vector averaged speed. Note 1: Miramontes and Penven et al. (2019) reported current measurements obtained from moorings at 3400 4050 m water depth in the Zambezi and Tsiribihina valleys show periods of intense currents at the seafloor with peaks of 40 50 cm s 2 1 that last up to 1 month and are not related to turbidity currents. These strong bottom-current events are correlated with a change in current direction and an increase in temperature. The periods of current intensification may be related to eddies. Note 2: Velocity measurements for some examples (e.g., Agulhas and Kushiro) were made on shallow-water settings. b

upper reworked division, mimicking a partial Bouma Sequence (Fig. 8.12B). Such offspring deposits of two flow types, namely sandy debris flows and contour currents (i.e., hybrid flows), are termed as “hybridites.” These genuine hybrid flows should not be confused with the usage of the term “hybrid flows” by Houghton et al. (2009) for flow transformation from one gravity flow into another. The distinction is that flow transformation represents a transitional stage between two flows, whereas hybrid flows represent two hydrodynamically different flows without flow transition. The other difference is that hybrid flows travel at right angle to each other (i.e., downslope vs alongslope, Fig. 8.12A). Analogous to the concept of Houghton et al. (2009), Talling (2013) also proposed a transitional hybrid-flow model, using flow transformation, between turbidity currents and cohesive debris flows. Hybridites are common in the geologic record and could be easily misinterpreted as turbidites. Bidirectional tidal currents in submarine canyons are likely to develop hybridites through bottom-current reworking as well.

8.6 The contourite problem Shanmugam (2016a) discussed this topic in great detail. Although there are numerous contourite problems, the following 10 fundamental issues have been selected for discussion.

Mass transport, gravity flows, and bottom currents

328

8. Bottom currents

FIGURE 8.8 (A) Core photograph showing well-sorted fine-grained sand and silt layers (light gray) with interbedded mud layers (dark gray). Note sand layers with sharp upper contacts, internal ripple cross-laminae, and mud-offshoots. Also note lenticular nature of some sand layers. Pleistocene, continental rise off Georges Bank, Vema 18_374, 710 cm, water depth 4756 m. (B) Core photograph showing rhythmic layers of sand and mud, inverse grading, and sharp upper contacts of sand layers (arrow), interpreted as bottom-current reworked sands, Paleocene, North Sea. Source: (A) After Hollister, C.D., 1967. Sediment distribution and deep circulation in the western North Atlantic. Unpublished Ph.D. dissertation. Columbia University, New York, p. 467 (his Fig. VI-1, p. 208); and Bouma, A.H., Hollister, C.D., 1973. Deep ocean basin sedimentation. In: Middleton, G.V., Bouma, A.H. (Eds.), Turbidites and Deep-Water Sedimentation, Anaheim, SEPM Pacific section Short Course, California, pp. 79 118, reproduced with permission from SEPM.

This is somewhat analogous to the author’s previous review of “Ten turbidite myths” in identifying fundamental problems (Shanmugam, 2002a). For example, the concept of highdensity turbidity currents (HDTCs) in explaining gravelly and sandy turbidites (Shanmugam, 1996a), somewhat analogous to the concept of irrational numbers in Mathematics (Havil, 2014), is incommensurable (Shanmugam, 2016a,b). Furthermore, despite the constant promotion of the turbidite-fan link (Grotzinger et al., 2007), the number of documented cases of the existence of gravelly and sandy turbidity currents in modern deep-water environments is zero!

8.6.1 Dual forcing of global ocean circulation The paradigm of global ocean circulation has been the thermohaline forcing of two independent water masses, namely the North Atlantic deep-water (NADW) or the “great Mass transport, gravity flows, and bottom currents

329

8.6 The contourite problem

Climbig-ripple cross-bedding

Mud offshoots Suspension (mud-offshoot)

Traction and suspension

5 cm

5 cm

Traction

Lenticular bedding

Flaser bedding

Suspension Traction

Suspension

FIGURE 8.9 Summary of traction features interpreted as indicative of deep-water bottomcurrent reworking by all types of bottom currents. Each feature occurs randomly and should not be considered as part of a vertical facies model. Source: From Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deep-marine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, with permission from AAPG.

5 cm

5 cm

Traction

Rhythmic bedding

Horizontal bedding

Sharp upper contact

Traction

5 cm

1 cm

Traction Suspension

Sharp upper contact

Cross-bedding

Sharp upper contact

Inverse grading Erosion

10 cm

5 cm

Traction

Fine sand

Gradational lower contact

Mud

ocean conveyor” (Broecker, 1991) and the AABW (Gordon, 1986). The global ocean circulation is initiated in the Southern Ocean (Antarctica) as the cumulative result of (1) winddriven (adiabatic) upwelling, (2) surface buoyancy flux, and (3) deep-water formation by cooling and saline rejection (i.e., thermohaline) (Fig. 8.3). Both atmospheric forcing (i.e., surface-wind stress) and thermohaline forcing (i.e., bottom-water formation) are necessary

Mass transport, gravity flows, and bottom currents

330

8. Bottom currents

FIGURE 8.10 Main types of sedimentary structures in contourite deposits. Source: From Martı´n-Chivelet, J., Fregenal-Martl-Ma, M.A., Chaco´n, B., 2008. Traction structures in contourites. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites. Dev. Sedimentol. 60, 159 182., with permission from Elsevier.

to induce and maintain global ocean circulation (Talley, 2013) (Fig. 8.13). For example, the Antarctic circumpolar current (ACC) is widely accepted as being dominantly a winddriven current (Howe et al., 1997). Therefore, a sound knowledge of global ocean surface currents is critical for understanding ocean bottom currents (Gill, 1982; Apel, 1987; Stewart, 2008; CIMAS, 2015). In light of the dual forcing of most water masses, it is inappropriate to classify an ancient layer as a contourite routinely, with a skewed emphasis on thermohaline forcing and with a total avoidance of the role of atmospheric forcing. The term “contourite drift” is used commonly in the geologic literature (see book chapters by Fauge`res and Stow, 2008 and Fauge`res and Mulder, 2011). The ACC

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

331

FIGURE 8.11

Conceptual model showing the spatial relationship between downslope turbidity currents and along-slope contour currents, which result in hybrid flows. This scenario, which is an unlikely scenario for developing a process continuum between the two types, is ideal for developing hybrid flows. Note that turbidity currents transport sediment downslope from the primary sediment source to basin along with mass-transport processes, whereas contour currents are reworking agents and as such they are unrelated to the primary sediment provenance. Source: After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deep-marine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, with permission from AAPG.

produces drifts at great depths of over 3000 m (Howe et al., 1997; Pudsey and Howe, 1998). These drift sediments are products of currents that follow bathymetric contours, and therefore they could be classified as contourites. However, these drift sediments are not genuine “contourites” because they are products of mostly wind-driven currents, not thermohaline-driven currents. In other words, contourites could be generated by more than one type of bottom currents. The problem here is that there are no sedimentological criteria for distinguishing deposits of purely wind-driven bottom currents from those of thermohaline-driven bottom currents. Therefore, the application of the term “contourites” to the ancient stratigraphic record, with little information on forcing mechanisms, should proceed with caution. A solution is to replace the genetic term “contourite drift” with a nongenetic term “sediment drift.”

8.6.2 Continuum between turbidity currents and contour currents Rebesco et al. (2014, their Fig. 1) begin their review with a ternary diagram with three end members composed of contourites, turbidites, and pelagites. The ternary diagram is based on the continuum principle of these three basic deep-sea sediment types that was advocated nearly 35 years ago by Stow and Lovell (1979). It is difficult, however, to reconcile a process continuum between turbidity currents and contour currents. By definition, the term “continuum” refers to a gradual transition from one end member to the other, without any abrupt changes. The continuum principle is unsustainable for the following reasons:

Mass transport, gravity flows, and bottom currents

332

8. Bottom currents

FIGURE 8.12 (A) Conceptual model of hybrid flows showing reworking the tops of downslope sandy debris flows by alongslope bottom currents. Such complex deposits would generate a sandy unit with a basal massive division and upper reworked divisions with traction structures (ripple laminae), mimicking the “Bouma sequence.” (B) The turbidite facies model (i.e., the Bouma Sequence) showing Ta, Tb, Tc, Td, and Te divisions. Conventional interpretation is that the entire sequence is a product of a turbidity current (Bouma, 1962; Walker, 1965; Middleton and Hampton, 1973). According to Lowe (1982), the Ta division is a product of a high-density turbidity current and Tb, Tc, and Td divisions are deposits of low-density turbidity currents. In this article, the Ta division is considered to be a product of a turbidity current only if it is normally graded, otherwise it is a product of a sandy debris flow; the Tb, Tc, and Td divisions are considered to be deposits of bottom-current reworking. Source: (A) From Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476, with permission from Elsevier. (B) From Shanmugam, G., 1997a. The Bouma Sequence and the turbidite mind set. Earth-Sci. Rev. 42, 201 229, with permission from Elsevier.

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

333

FIGURE 8.13 Schematic diagram showing the wind-driven and thermohaline-driven mechanisms in the Southern Ocean (Antarctica) in initiating global ocean circulation. Source: From Talley, L. D., 2013. Closure of the global overturning circulation through the Indian, Pacific, and Southern Oceans: schematics and transports. Oceanography 26 (1), 80 97, with permission from the Oceanography Society.

• Downslope-flowing turbidity currents and alongslope flowing contour currents are almost at right angles with each other (Fig. 8.11). Even if the two interact with each other, the interaction would be ephemeral and is of no sedimentological significance. • Turbidity currents are local or regional in transport, whereas most contour currents are global in scale. • Turbidity currents are episodic (Kuenen and Migliorini, 1950) or surge-type events that fail to develop equilibrium conditions (Allen, 1985), whereas contour currents persist for long periods of time and can develop equilibrium conditions. In addition, the ternary diagram totally ignores the importance of MTDs (Mosher et al., 2010), tidal currents in submarine canyons (Shepard et al., 1979), baroclinic currents (Shanmugam, 2013a), and bottom currents associated with cyclones and tsunamis in the deep ocean (Shanmugam, 2008a).

8.6.3 Revision of the basic principle of contour currents The basic principle of contour currents was introduced first by Heezen and Hollister (1964) to the marine geologic community and later by Heezen et al. (1966) to the general scientific community. Their studies were based on a regional study of the continental rise off eastern United States in the Atlantic Ocean, covering the Blake Plateau and Blake-Bahama Outer Ridge (Fig. 8.2, location A). Their seminal study was based on a robust dataset composed of echo sounding, bottom photographs, and sediment cores.

Mass transport, gravity flows, and bottom currents

334

8. Bottom currents

Therefore, it is useful to revisit the following three fundamental points from Heezen et al. (1966): • Page 502: “Geostrophic contour-following bottom currents involved in the deep thermohaline circulation of the world ocean appear to be the principal agents which control the shape of the continental rise and other sediment bodies.” • Page 504: “Pressure gradients indicated by the inclined isopycnals must be opposed by an opposite and equal force which would seem to be provided by a current in which the Coriolis forces are acting normal to the direction of motion (to the right in the northern hemisphere). These currents flow along isopycnals which are approximately parallel to the bathymetric contours. We refer to these currents as contour currents.” • Page 507: “In marked contrast to the steady, low velocity (2 to 20 cm/sec) contourfollowing geostrophic currents which never flow downslope, turbidity currents are intermittent, high-velocity (up to 2500 cm/sec) downslope movements. . .” As envisioned by Heezen et al. (1966), the basic principle of contour currents was scientifically sound, and there is no need to revise it. Nevertheless, other authors have broadened the meaning. For example: • Johnson et al. (1980) applied the term “contourites” to sediments in the Lake Superior, United States. • Lovell and Stow (1981, p. 349) conclude that “Contourite: a bed deposited or significantly reworked by a current that is persistent in time and space and flows along slope in relatively deep-water (certainly below wave base). The water may be fresh or salt; the cause of the current is not necessarily critical to the application of the term.” I have italicized the last phrase to emphasize the point that contourites can be produced by any kind of bottom current (Fig. 8.14), irrespective of their origin (i.e., thermohaline, wind, tide, or baroclinic). • The last phrase in Lovell and Stow (1981) (see above) has served as the foundation and a continuum for a subsequent paper by Stow et al. (2008) in which they expanded the meaning of the term “contourite.” For example, Stow et al. (2008, p. 144) explicitly state that “Bottom (contour) currents are those currents that operate as part of either the normal thermohaline circulation or wind-driven circulation systems. . .” • Furthermore, Stow et al. (2008, p. 145) state that “Bottom currents are highly variable in location, direction and velocity over relatively short time scales (from hours to months). Velocity increase, decrease and flow reversal occur as a result of deep tidal effects (e.g. Shanmugam, 2008).” (i.e., Shanmugam, 2008a in this book). Although Stow et al. (2008) were justified in searching for a broad term to represent all bottom currents, their choice of the term “contour currents” for all types is inappropriate. As noted earlier, there are four basic types of bottom currents, namely (1) thermohalinedriven contour currents, (2) wind-driven bottom currents, (3) tidal bottom currents, and (4) baroclinic currents. Major problems associated with the broadening the meaning of the term “contour currents” are as follows: • Unlike thermohaline-driven contour currents, the other three types do not originate due to thermohaline forcing. The Loop Current in the Gulf of Mexico, for example, is strictly

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

335

FIGURE 8.14 Four types of bottom currents and their depositional facies. The facies term “contourites” is appropriate only for deposits of thermohaline-driven geostrophic contour currents in deep-water environments, but not for deposits of other three types of bottom currents (i.e., wind, tide, or baroclinic). Note that BCRS represent only sandy lithofacies, but may also be applicable to silty lithofacies. Source: From Shanmugam, G., 2016b. The contourite problem. In: Mazumder, R. (Ed.), Sediment Provenance. Elsevier, pp. 183 254 (Chapter 9), with permission from Elsevier.

a wind-driven current (Pequegnat, 1972; CIMAS, 2015); no thermohaline forcing is involved. It would be incorrect to classify deposits of the Loop Current as contourites. • Unlike thermohaline-driven contour currents, the other three types commonly do not follow bathymetric contours. The wind-driven Loop Current in the Gulf of Mexico, for example, does not follow bathymetric contours (Pequegnat, 1972; Mullins et al., 1987; Shanmugam et al., 1993a). The Loop Current also triggers eddies that fail to follow bathymetric contours. • Deep-marine tidal currents flow up and down submarine canyons (Shepard et al., 1979). • In some cases, baroclinic tidal currents flow across the canyon and in a direction parallel to the shelf break (Allen and Durrieu de Madron, 2009). Rebesco et al. (2008, p. 6) argued that a strict adaptation of the basic definition of Heezen et al. (1966) would prevent the application of the contour-current concept to ancient deposits, where both depth and direction of the currents can rarely be precisely reconstructed. Although interpretation of ancient deep-water strata will always remain a challenge, one should not compromise the basic principles of contour currents for the sake of convenience and simplicity. A solution is to adopt the general term “bottom currents” for all four types. As a continuation of the above problem, the original meaning of the term “contourite” has been broadened. The tradition of genetic nomenclature in sedimentary geology began with the introduction of the term “turbidite” for a deposit of a turbidity current in deepwater environment (Kuenen, 1957). Shanmugam (2006b) presented a detailed review of

Mass transport, gravity flows, and bottom currents

336

8. Bottom currents

the problems associated with genetic nomenclatures. The term “contourite” was first introduced in a publication for deposits of contour currents by Hollister and Heezen (1972), although Hollister (1967) discussed contourites earlier in his unpublished Ph.D. dissertation. In these early contributions, the term “contourites” solely referred to deposits of contour currents. But other researchers have widened the definition to include deposits of a variety of bottom currents that include wind-driven currents and tidal currents (Stow et al., 2008). Such a broad application of the term contourite undermines the very basic tenet of process sedimentology, which is to distinguish deposit of one specific process from that of the other. In acknowledging this conceptual-nomenclatural problem, Rebesco et al. (2008, p. 7) state, “This implies the risk of an excessively wide application of the term “contourite,” and consequently of a loss of significance.” Although the original contourite concept was designed solely for deep-water deposits (Hollister and Heezen, 1972), it has been expanded to include shallow-water deposits (e.g., Verdicchio and Trincardi, 2008), causing additional confusion. These problems can be alleviated by simply being faithful to the original definition of the term “contourite” as envisaged by the founding fathers of the concept: the late B.C. Heezen and the late C.D. Hollister. In discussing gravity-driven downslope processes, Middleton and Hampton (1973) proposed four types of “sediment-gravity flows,” namely grain flow, fluidized flow, debris flow, and turbidity current, based on sediment-support mechanisms. No one would classify deposits of all four types of sediment-gravity flows as “turbidites”! Similarly, one should not classify all four types of bottom currents as “contourites.” Such classifications defeat the very purpose of distinguishing one type of deposit from the other and such practices render the field of process sedimentology irrelevant.

8.6.4 Hiatuses in contourites In nonmarine and shallow-marine clastic environments, hiatuses (breaks in sedimentation) are ubiquitous. For example, Miall (2014) reported that only 10% of elapsed time is represented by sediment in these environments, the remainder (90%) is nothing but hiatuses. In deep-marine environments, regional erosion throughout thousands of square kilometers of seafloor has been attributed to bottom currents (Berggren and Hollister, 1977; Tucholke and Embley, 1984). In the Gulf of Cadiz, the lower core of the Mediterranean outflow water (MOW) tends to cause more erosion (Herna´ndez-Molina et al., 2014). In the Rockall Trough region, bottom currents associated with the NADW have caused an erosive area extending over 8500 km2 in water depths of 500 2000 m (Howe et al., 2001). This erosive phase, which eroded approximately 150 m of sediment and lasted nearly 35 Ma (Early Oligocene-Holocene), existed through four supercycles (second order) and 23 cycles (third order) of sea-level rise and fall in the global chronostratigraphic chart of Haq et al. (1988). Viana (2008) cautioned on the potential dangers of misinterpreting regional unconformities at the base of contourites as “sequence boundaries” on seismic profiles using examples from the Santos Drift, offshore Brazil (Duarte and Viana, 2007). Clearly, there is no simple correlation between current-induced erosional surfaces (unconformities) and eustasy. These practical challenges exist because there are no objective criteria to recognize erosional surfaces, caused by deep-marine bottom currents versus by other processes, on seismic profiles (Shanmugam, 1988, 2007).

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

337

8.6.5 Origin of erosional features Pe´rez et al. (2015) discussed erosional and depositional features associated with contourites on seismic data. However, there are conceptual and sedimentological problems. • In defining the contourite depositional system (CDS), Herna´ndez-Molina et al. (2008, p. 350) state, “An association of various drifts and related erosional features has been termed a “contourite depositional system” (CDS). . .” This inclusion of erosional features under the term “contourite depositional system” is conceptually confusing. It is useful to maintain a distinction between erosion and deposition. A solution is simply to group both erosion and deposition under “contourite system” instead. • Following Herna´ndez-Molina et al. (2006, 2008), Garcı´a et al. (2009) attributed the origin of four types of erosive features, including contourite channel, to erosion exclusively by the MOW in the Gulf of Cadiz. However, these authors did not consider the alternative possibility of erosion by baroclinic currents in the Gulf of Cadiz, where internal waves and internal tides are active oceanic phenomena (Cairns, 1980; Armi and Farmer, 1988; LaViolette and Lacombe, 1988; Apel, 2000; Morozov et al., 2002; Vargas-Yaˇnez et al., 2002; Che´rubin et al., 2003, 2007; Serra, 2004; Pavec et al., 2005; Ambar et al., 2008; Huthnance et al., 2008; Sa´nchez-Roma´n et al., 2008; Vsemirnova et al., 2009; Leo´n et al., 2014). The other problem is that there are no detailed measurements and observations on the velocities and erosive power of baroclinic currents on the deep seafloor. This is a potential topic for future research. • Stow et al. (2013, p. 112) state, “In this paper, we have detailed the development and characteristics of a contourite channel, which is as long, wide and deep as many turbidity current channels, but which has been cut and shaped by bottom currents, and by their interaction with a bottom topography influenced by neotectonics. In places it is floored by contourite sands and gravel.” If the channel was cut and shaped by “bottom currents” that include four types (Shanmugam, 2008b), it is misleading to classify any channel a “contourite channel” with a skewed emphasis on contour currents, ignoring the other three bottom currents. • There are no sedimentological criteria to distinguish deep-sea channels cut by turbidity currents from those cut by contour currents. This problem is further complicated when similar depositional features, such as mud drapes, are associated with channels of different origins. For example, mud drapes have been reported from (1) turbidite channels (Miocene) exposed at the San Clemente State Beach, California (Walker, 1975) and (2) estuarine tidalite channels (Cretaceous) in the subsurface conventional cores, Oriente Basin, Ecuador (Shanmugam et al., 2000). • Erosion by strong bottom currents tends to cause lag deposits in submarine environments. Various aspects of contourite lag deposits were discussed by other authors (Hu¨eneke and Stow, 2008; Martın-Chivelet et al., 2008; Stow and Fauge`res, 2008; Wetzel et al., 2008). The grain size of the lag deposits merely indicates which grain-size fractions could not be transported. Besides, a lag represents a gap in the sedimentary record, which may cause problems with the construction of highresolution age models of sediment cores. By nature, erosion does not leave behind any clue in the rock record for establishing the type of process that caused the erosion. Furthermore, modern unfilled submarine channels

Mass transport, gravity flows, and bottom currents

338

8. Bottom currents

and canyons are a testimony to the fact that the processes that created these erosional features in the past are probably not the same processes that will fill them in the future. Therefore, there is a need to develop criteria for distinguishing erosional features cut by contour currents from those cut by other processes, such as turbidity currents.

8.6.6 Gulf of Cadiz as the type locality Herna´ndez-Molina et al. (2013) characterized the Gulf of Cadiz as “the world’s premier contourite laboratory.” The modern Gulf of Cadiz has served as the center for contourite research activities since the 1970s (Fig. 8.2, location F). For example: • The Gulf of Cadiz is the birthplace of the first contourite facies model (Fauge`res et al., 1984; Gonthier et al., 1984). • The MOW (Fig. 8.15) and related properties have been well studied (Zenk, 1975; Ambar and Howe, 1979; Zenk and Armi, 1990; Pinardi and Masetti, 2000; Criado-Aldeanueva et al., 2006; Herna´ndez-Molina et al., 2003, 2006, 2014; Garcia et al., 2009; Alves et al., 2011; Mulder et al., 2013; Stow et al., 2013), with salinity, temperature, and velocity measurements (Price et al., 1993; Baringer and Price, 1999). FIGURE 8.15 Map showing the main water-mass circulation in the Gulf of Cadiz. Note the trajectory of the Mediterranean Outflow Water (MOW) flowing westward in the gulf and turning northward as it enters the Atlantic Ocean at Cape Sa˜o Vicente (San Vicente cp.). The initial black and white version was published by HernandezMolina et al. (2003); modified by Llave et al. (2011) and Stow et al. (2013). Source: With permission from the Geological Society of America.

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

339

• Internal waves and internal tides have been documented in the Gulf of Cadiz (Cairns, 1980; Armi and Farmer, 1988; LaViolette and Lacombe, 1988; Apel, 2000; Bruno et al., 2006; Alvarado-Bustos, 2011; Sanchez-Garrido et al., 2011; Quaresma and Pichon, 2013). • Sedimentary bedforms on the seafloor were documented using side-scan sonar images (Kenyon and Belderson, 1973) and submarine photographs (Stow et al., 2013). • The Gulf of Cadiz was the site of the Integrated Ocean Drilling Program (IODP) Expedition 339 (Herna´ndez-Molina et al., 2013). The Gulf of Cadiz, despite its popularity, has its limitations. Although the MOW in the Gulf of Cadiz is a thermohaline-driven water mass (Alves et al., 2011), it is not a genuine contour current. For example, Zenk (2008, p. 45) characterizes the behavior of MOW as follows: “The warm and salty Mediterranean outflow water (MOW) in the Gulf of Cadiz of the eastern North Atlantic represents an excellent example for the transition (italicized for emphasis) between a purely bottom-following current to a genuine contour current. . .” Empirical data indeed support the transition of the MOW in the Gulf of Cadiz. The MOW undergoes three progressive stages of evolution during its journey from the Strait of Gibraltar where it enters the Gulf of Cadiz to Cape Sa˜o Vicente where it exits the gulf before entering the Atlantic Ocean (Fig. 8.16):

FIGURE 8.16 Schematic diagram showing the location of Gulf of Cadiz and complex transport nature of the Mediterranean Outflow Water (MOW), involving three stages of evolution: (1) channel-current stage, (2) mixing and spreading (i.e., transition) stage, and (3) genuine contour-current stage (see Zenk, 2008, his Fig. 4.10). Velocity at the Strait of Gibraltar is from Heezen and Johnson (1969). Velocity near Cape Sa˜o Vicente is from Prater and Sanford (1994) and Baringer and Price (1999). Other velocity values, Froude numbers, and MOW widths are from Baringer and Price (1997, 1999). Details on IODP Expedition 339 cores are discussed by Herna´ndez-Molina et al. (2013) who reported 300 cm s21 (118.11 in. s21) velocity at the Strait of Gibraltar (see also Gonthier et al., 1984) and B80 100 cm s21 near Cape Sa˜o Vicente. The popular Faro contourite drift (Fauge`res et al., 1984) is located just south of the town of Faro offshore. C.S. Vicente 5 Cape Sa˜o Vicente 5 Cape St. Vincent (in some publications). Sill 5 Camarinal Sill (Sa´nchez-Roma´n et al. 2008). Source: From Shanmugam, G., 2016b. The contourite problem. In: Mazumder, R. (Ed.), Sediment Provenance. Elsevier, pp. 183 254 (Chapter 9), with permission from Elsevier.

Mass transport, gravity flows, and bottom currents

340

8. Bottom currents

8.6.6.1 Channel-current stage Price et al. (1993), based on the 1988 Gulf of Cadiz Expedition that included 99 fulldepth profiles of temperature and salinity and 56 horizontal current profiles, characterized the MOW in the Gulf of Cadiz as a “steady channel flow” near the Strait of Gibraltar (Fig. 8.16). At this first stage, the current was highly turbulent and the Froude number was above 1. The transport was downslope from east to west (Fig. 8.16); however, the descent was asymmetric and occurred in two preferred modes or cores (Baringer and Price, 1997). 8.6.6.2 Mixing and spreading Stage Mixing and spreading of MOW represents the second transition stage (Fig. 8.16). Within 100 km downstream from the Strait of Gibraltar, the MOW was affected by the Coriolis force. Due to mixing, the MOW lost its density and increased its transport volume westward. The velocity progressively decreased westward from 150 cm s21 at the strait to 10 30 cm s21 near Cape Sa˜o Vicente (Fig. 8.16). At this turning point, the MOW became neutrally buoyant in the lower portion of the North Atlantic thermocline (Baringer and Price, 1999). In the western Gulf of Cadiz, where the entrainment was much weaker, Froude numbers were consistently below 1 (Baringer and Price, 1997). 8.6.6.3 Contour-current Stage After making a 90-degree turn to the right (north) in the open Atlantic Ocean due to the full effect of the Coriolis force, the MOW attains total geostrophic balance and flows northward nearly parallel to the bottom topography of the Atlantic Ocean, off the western Iberian margin (Zenk, 2008, his Fig. 4.10; Herna´ndez-Molina et al., 2011, their Fig. 8.4). At this final stage, the MOW is considered a genuine contour current (Fig. 8.16). In summary, the Gulf of Cadiz is a highly complex oceanographic location for studying depositional and erosional aspects of genuine contour currents because the deep-sea sediments in this gulf are controlled by the following factors (Fig. 8.16): • • • • • • • • • • • •

transitory MOW (Zenk, 2008) internal waves and tides (Apel, 2000; Alvarado-Bustos, 2011) sediment-gravity flows (Herna´ndez-Molina et al., 2013) pelagic and hemipelagic settling tsunamis (Lario et al., 2010) cyclones (Lario et al., 2010) mud volcanism (Pinheiro et al., 2003) methane seepage (Magalha˜es et al., 2012) sediment supply (Mulder et al., 2013) pore-water venting and hydraulic pumping (Leo´n et al., 2014) channels and ridges (Stow et al., 2013) the Camarinal Sill (Go´mez-Enri et al., 2007)

Complex localities such as the Gulf of Cadiz requires an understanding of all processes in concert with each other because “deep-water” processes are tightly intertwined with “shallow-water” processes by oceanic wave phenomena, such as internal waves and

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

341

tsunamis. Therefore, the archaic notion of dealing with a particular “deep-water” process (e.g., contour currents) in a vacuum is over. The 21st century necessitates the rigor of holistic process sedimentology.

8.6.7 The contourite facies model Fauge`res et al. (1984) explained the role of MOW in developing the first muddy contourite facies model from the Gulf of Cadiz (Fig. 8.17). Students (Brackenridge, 2014; Lathrop, 2015) and researchers (Rebesco et al., 2014) use this model routinely. Nevertheless, the vertical facies model suffers for the following reasons. 8.6.7.1 Five internal divisions Fauge`res et al. (1984) developed the original facies model without internal divisions. Stow and Fauge`res (2008: their Fig. 13.9), however, revised the original model with five internal divisions (C1, C2, C3, C4, and C5) (Fig. 8.17), analogous to the Bouma turbidite model (Bouma, 1962). In their most recent version, Fauge`res and Mulder (2011, their Fig. 3.18) have reverted back to the 1984 version, without the five internal divisions. Reasons for such back-and-forth fundamental changes to the facies model, by the same group of authors, need to be explained in the literature for the benefit of the international research community. If recognized in the ancient rock record, these five divisions would

FIGURE 8.17 Left: Revised contourite facies model with five divisions (C1 C5) proposed by Stow and Fauge`res (2008). Right: Original contourite facies model by Fauge`res et al. (1984). Note that the original authors of this model did not include the five internal divisions (Fauge`res et al., 1984). The most recent version of this model by Fauge`res and Mulder (2011) contains neither the five internal divisions nor the hiatuses in the C3 division (red arrow inserted in this article). This figure was originally from Fauge`res et al. (1984), with permission from the Geological Society of America. Source: From Shanmugam, G., 2016b. The contourite problem. In: Mazumder, R. (Ed.), Sediment Provenance. Elsevier, pp. 183 254 (Chapter 9). Elsevier.

Mass transport, gravity flows, and bottom currents

342

8. Bottom currents

reveal nothing about deposition from thermohaline-driven geostrophic contour currents in deep-water environments. 8.6.7.2 Current velocities The vertical facies model, composed of a basal upward-coarsening interval followed by an upward-fining interval (Fig. 8.17), has been attributed to a successive increase and decrease in contour-current velocity and competency (Fauge`res et al., 1984). However, Mulder et al. (2013) suggest that the origin of this vertical sequence is much more complex than due to a simple velocity variation. Mulder et al. (2013, p. 357) state that “. . . the contourite sequence is only in part related to changes in bottom current velocity and flow competency, but may also be related to the supply of a coarser terrigenous particle stock, provided by either increased erosion of indurated mud along the flanks of confined contourite channels (mud clasts), or by increased sediment supply by rivers (quartz grains) and downslope mass transport on the continental shelf and upper slope. The classical contourite facies association may therefore not be solely controlled by current velocity, but may be the product of a variety of depositional histories.” No further explanation is necessary. 8.6.7.3 Internal hiatuses In the original contourite facies model, Fauge`res et al. (1984, their Fig. 8.4) did not include internal hiatuses. However, Stow and Fauge`res (2008, their Fig. 13.9) included hiatuses in the middle C3 division of their revised contourite facies model (Fig. 8.17, see horizontal red arrow). In the most recent 2011 version of the model (Fauge`res and Mulder, 2011, their Fig. 3.18), the hiatuses are absent once again. How can a natural, observed, sedimentary feature (i.e., hiatus) simply vanish? The authors need to explain this puzzle. Wetzel et al. (2008, p. 189) state, “When bottom currents prevent deposition for a considerable time span, and/or erode sediments, submarine hiatuses develop, represented by semi-consolidated firm- or hard grounds or stable cohesive partially dewatered muddy substrates.” Because hiatuses occur in the C3 division (Fig. 8.17), the lower and upper intervals must represent two different depositional events. Conventionally, a genetic facies model is designed for a single depositional event, without internal hiatuses (e.g., the turbidite facies model, Bouma, 1962). In fact, Walther’s Law (Middleton, 1973) is not meaningful for sequences with internal hiatuses. This is because a hiatus can represent a considerable span of time (spanning millions of years) that is missing in the sedimentary record (Howe et al., 2001). 8.6.7.4 Bioturbation A characteristic feature of the contourite facies model is the bioturbation (Fig. 8.17), which has generated debates (Shanmugam, 2002b; Mulder et al., 2002). Conventionally, a genetic facies model (e.g., the turbidite facies model, Bouma, 1962) is based on vertical disposition of primary physical sedimentary structures. This is because physical structures can be used to interpret a particular physical process in the rock record. But bioturbation cannot be used as a criterion for interpreting deposit of a single process (i.e., contour currents). The bioturbation criterion is defective because ancient deep-water turbidites (e.g., in the Late Cretaceous Point Loma Formation near San Diego, California) are also

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

343

extensively bioturbated and even contain the trace fossil Ophiomorpha (Nilsen and Abbott, 1979). Furthermore, convincing cases of “contourites” without bioturbation have been documented in the rock record (Dalrymple and Narbonne, 1996). In describing the Canterbury Drifts from SW Pacific Ocean Carter (2007, p. 129) state that “Bioturbation is moderate and rarely destroys the pervasive background, centimetre-scale, planar or wispy alternation of muddy and sandy silts displayed by Formation Micro-Scanner imagery.” The muddy contourite facies model with emphasis on bioturbation defies the very first principle of process sedimentology, which is to interpret the fluid mechanics of depositional processes using primary physical sedimentary structures (Sanders, 1963). • The above two basic issues are still unresolved (Shanmugam, 2016b). Amid this knowledge vacuum, Rodrı´guez-Tovar and Herna´ndez-Molina (2018a) have published a paper entitled “Ichnological analysis of contourites: Past, present and future.” However, the same authors (Rodrı´guez-Tovar and Herna´ndez-Molina, 2018b) conceded that “Nowhere in our manuscript did we present bioturbation as an exclusive feature of contourites with respect to other deposits such as turbidites, debrites, etc.” Clearly, bioturbation and trace fossils are not diagnostic features of contourites (Shanmugam, 2018d,e). • Importantly, Hollister (1967, Appendix C, his p. 392) did not even include “bioturbation” as a basic sedimentary feature in the “Sediment Core Logs” of sediments that formed the very foundation for introducing the concept of contourites (his Fig. 1). • All four types of bottom currents are characterized by traction structures (Fig. 8.11). The contourite facies model with emphasis on bioturbation (Fig. 8.17) defies the very first principle of process sedimentology, which is to interpret the fluid mechanics of depositional processes using primary physical sedimentary structures (Sanders, 1963), not bioturbation. The reason is that bioturbation can occur after deposition. Clearly, bioturbation and trace fossils are not diagnostic features of contourites. In short, bioturbation is of no process sedimentological significance for interpreting ancient deep-water contourite facies (Shanmugam, 2018e). 8.6.7.5 Multiple interactive processes The muddy contourite facies model was based on the notion that a single process, namely deposition from contour currents, was solely responsible for the deposit (Fauge`res et al., 1984). But Stow et al. (2013) have demonstrated that multiple interactive processes are operating in the Gulf of Cadiz. In 1984, prior to detailed velocity measurements of MOW (Price et al., 1993) and numerous other investigations of internal waves and internal tides in the Gulf of Cadiz, it was reasonable for Fauge`res et al. (1984) to propose a contourite facies model at a time when we were grappling with complex deep-water processes, without much data. But today, a great wealth of empirical data (see references in Stow et al., 2013) is available. The Gulf of Cadiz is an extremely complex setting in terms of physical oceanography with multiple processes (e.g., MOW, internal waves, and internal tides) and bottom topography with channels, ridges, and sills. The physical, chemical, and sedimentological aspects of the MOW are equally complex (Ambar et al., 2002; CriadoAldeanueva et al., 2006). Rebesco et al. (2014, p. 139) acknowledge that “Regardless, the previous research on this issue holds two important lessons: firstly, that there is no unique

Mass transport, gravity flows, and bottom currents

344

8. Bottom currents

facies sequence for contourites; and secondly, that traction sedimentary structures are also common within contourites. . .” Deep-water depositional processes are variable in time and space. Furthermore, extensive bioturbation caused by influx of prolific oxygen in deep-sea currents obliterates physical structures. From a practical viewpoint of interpreting ancient deposits as contourites on land, there is no way of knowing the contours of the paleo-seafloor (Stow et al., 1998). In summary, the global applicability of the contourite facies model is dubious. 8.6.7.6 Grain-size data A fundamental aspect of many sedimentological studies is the documentation of detailed vertical grain-size variation that is plotted on a sedimentological log. It is so vital that the present author has allotted the maximum space for grain size (i.e., expanded column widths for silt, very fine sand, medium sand, etc.) in sedimentological logs (e.g., Shanmugam et al., 2009). But such sedimentological logs illustrating vertical grain-size variations and other sedimentological details for sandy contourite intervals are absent in publications by Stow and Fauge`res (2008) and by Stow et al. (2008). In fact, none of the 19 core photographs (six from the Gulf of Cadiz, eight from the Brazilian margin, and five from the U.K. margin) has associated sedimentological logs in Stow and Fauge`res (2008). Consequently, the reader is left with core photographs of sandy contourites without the fundamental grain-size data. During the IODP Expedition 339, five sites were drilled in the Gulf of Cadiz and two sites off the West Iberian margin (Herna´ndez-Molina et al., 2013). The total length of recovered core is 5447 m, with an average recovery of 86.4% (Expedition 339 Scientists, 2012). Published results of the IODP 339 core studies, although preliminary, are useful in testing the contourite facies model. • Core photographs labeled as “bigradational sequences” (Fig. 8.18A) and “sandy contourite” (Fig. 8.18C) do not show vertical grain-size variations based on measurements. • Specific sedimentological criteria used for distinguishing base-cut-out contourites with normal grading (Fig. 8.18B) from turbidites with normal grading (Fig. 8.18D) are not discussed. • The five internal divisions of the contourite facies model are not evident in any of the published core intervals. Even in the core interval U1390A-8H-6A, which is labeled as “bigradational grading,” which presumably represents the entire contourite sequence, the five internal divisions are not evident (Fig. 8.18A).

8.6.8 Traction structures and shale clasts The presence of traction structures in cores and outcrops have long been recognized as evidence for bottom-current reworked sands by contour currents, wind-driven currents, and tidal currents in deep-water strata (Hsu¨, 1964, 2008; Hubert, 1964; Klein, 1966; Hollister, 1967; Natland, 1967; Piper and Brisco, 1975; Shanmugam et al., 1993a,b; Shanmugam, 2008b; Martın-Chivelet et al., 2008; Mutti and Carminatti, 2011). As noted earlier, ripples and dunes

Mass transport, gravity flows, and bottom currents

8.6 The contourite problem

345

FIGURE 8.18 Core photographs showing sedimentary facies of contourites (A, B, C, and E), turbidites (D), debrites (F), and slumps (G) recovered during IODP Expedition 339. Note that vertical grain-size variations showing grading are schematic (red arrows), not factual using the Wentworth grain-size class on the abscissa Source: From Herna´ndez-Molina, F. J., Stow, D. A. V., Alvarez-Zarikian, C., 2013. IODP Expedition 339 in the Gulf of Cadiz and off West Iberia: decoding the environmental significance of the Mediterranean outflow water and its global influence. Sci. Drill. 16, 1 11, with permission from IODP Expedition 339; Scientific Drilling. The application of "cut-out" logic of Walker (1965), which was originally introduced for the turbidite facies model (i.e., the Bouma Sequence) with sharp basal contacts and sandy divisions containing well-developed primary traction structures (Fig. 8.12B), to muddy contourites with gradational basal contacts (Fig. 8.17) and an absence of primary traction structures (Fig. 8.18B and E) is misleading.

have been associated with internal tidal currents (Lonsdale and Malfait, 1974). In other words, traction structures and bedforms have been associated with all four types of bottom currents. The challenge is how to distinguish a traction structure (e.g., ripple or parallel laminae) formed by contour currents from those formed by wind-driven bottom currents in the ancient stratigraphic record. In discussing the origin of shale clasts in muddy and sandy contourites, Stow and Fauge`res (2008, p. 231) state, “The shale clasts are generally millimetric in size, and occur with long axes sub-parallel to bedding and, presumably, also sub-parallel to the current direction.” Alternatively, the planar clast fabric (i.e., alignment of long axis of clasts parallel to the bedding surface) could be interpreted as evidence for laminar debris flow (Fisher, 1971; Enos,

Mass transport, gravity flows, and bottom currents

346

8. Bottom currents

1977; Shanmugam and Benedict, 1978). In short, there are no reliable sedimentological criteria that one can apply in interpreting the ancient rock record as sandy contourites.

8.6.9 Bedform-velocity matrix Van Rooij (2013) used the bedform-velocity matrix (Fig. 8.19) of Stow et al. (2009) in discussing the challenges associated with processes and products of deep-water bottom currents. Problems associated with the bedform-velocity matrix are as follows: • Stow et al. (2009) proposed a bedform-velocity matrix (Fig. 8.19) for deep-water bottom currents. This matrix diagram is a slightly modified version of Figs. 3.1 and 3.2 in Belderson et al. (1982). Stow et al. (2009) applied the bedform-velocity matrix, developed by Belderson et al. (1982) for shelf tidal currents, to all types of deep-water bottom currents. But shallow-water tidal currents and deep-water bottom currents are not one and the same hydrodynamically. As mentioned earlier, at least four different types of deep-water bottom currents exist (Shanmugam, 2008b). The underpinning assumption of the matrix, which is that all four deep-water bottom currents hydrodynamically behave the same as the shallow-water tidal currents, is incongruous. FIGURE 8.19 Bedform-velocity matrix for deep-water bottom currents. Source: From Stow, D.A. V., Hernandez-Molina, F.J., Llave, E., Sayago-Gil, M., Diaz del Rio, V., Branson, A., 2009. Bedformvelocity matrix: the estimation of bottom current velocity from bedform observations. Geology 37, 327 330, with permission from the Geological Society of America. Color version from Rebesco, M., Herna´ndez-Molina, F. J. Van Rooij, D., Wa˚hlin, A., 2014. Contourites and associated sediments controlled by deep-water circulation processes: state-of-the-art and future considerations. Mar. Geol. 352, 111 154.

Mass transport, gravity flows, and bottom currents

347

8.6 The contourite problem

• Stow et al. (2009) acknowledged that (1) although the velocity data presented by them were for near-bottom flow, they did not define the exact height above seafloor; (2) they did not address the variable nature of the benthic boundary layer that will also complicate how flow velocity affects seafloor bedform; (3) for most of their data sets it was impossible to know the precise flow velocity (mean or peak) that created the observed bedform; (4) they rarely had the opportunity of witnessing the development of deep-water bedforms in situ; and (5) they did not consider the effects of sediment supply and bed roughness on bedform development. In other words, the matrix was built without the necessary empirical data. • The concept of bedform-velocity matrix became popular in the 1960s with the advent of matrix diagrams of alluvial sedimentary structures based on empirical data derived from flume experiments (Simons et al., 1965). However, the matrix diagram proposed by Stow et al. (2009) is not based on experiments; meaning that their “data” are neither verifiable nor reproducible independently. • In commenting on the problems with the bedform-velocity diagram of Stow et al. (2009), Dykstra (2012, his Fig. 14.2 caption) states, “Note that this Figure 8 does not take into account either the duration of a current or sediment availability, both of which are important controls on the development of bedforms. . .” Given the above uncertainties, it is unreliable to estimate current velocities for modern bedforms using the bedform-velocity matrix.

8.6.10 Abyssal plain contourites A field of large migrating mud waves, approximately 1.0 3 106 km2 in extent, occurs on sediment drifts in the Argentine Basin (Fig. 8.20). The mud waves range in amplitude up FIGURE 8.20 Distribution of contourite sediment waves in the Argentine Basin formed by AABW (modified after Klaus and Ledbetter, 1988). Source: Image after Shanmugam, G., 2017b. Contourites: physical oceanography, process sedimentology, and petroleum geology. Pet. Explor. Dev. 44 (2), 183 216. Elsevier.

Mass transport, gravity flows, and bottom currents

348

8. Bottom currents

to 137 m and average 26 m; wavelengths are about 3 7 km. The migrating mud waves are current-controlled features that form in areas of relatively weak AABW benthic circulation (Klaus and Ledbetter, 1988). Herna´ndez-Molina et al. (2008) discussed “abyssal plain contourites.” Conventionally, the term “abyssal plain” refers to a flat region of the ocean floor, usually at the base of a continental rise, where slope is less than 1:1000 (Heezen et al., 1959). It represents the deepest and flat part of the ocean floor that occupies between 4000 and 6500 m in the U.S. Atlantic Margin. A more general term “basin plain” is commonly used in referring to ancient examples. However, Herna´ndez-Molina et al. (2008) consider abyssal plains or basin plains to include up to 10 distinct morphological elements: (1) continental rise; (2) abyssal plains; (3) oceanic rises; (4) distal fans and their distributary channels; (5) sediments drifts; (6) abyssal hills; (7) seamounts; (8) transfer fracture zones; (9) mid-ocean channels; and (10) oceanic trenches. This reclassification of abyssal plains, ignoring the basic principles of classification of continental shelf, slope, rise, and plain based on the position of seafloor depths, is confusing. This reclassification defies the basic concept of “contour currents” that was introduced for contour-following bottom currents along continental slope and rise, not for bottom currents flowing over flat abyssal plains.

8.6.11 Sandy intervals of contourite facies models • Stow and Smillie (2020, their Fig. 14) proposed large-scale cross bedding in medium- to coarse sands for “The sandy contourite family” based on the study of Brackenridge et al. (2018). However, Brackenridge et al. (2018) did not document cross bedding in sands. • de Castro et al. (2020) reported starved ripples in the sandy intervals, but were unable to document with the empirical photographic evidence due to poor preservation of primary sedimentary structures in unconsolidated sediment intervals. It is important, however, to note that these authors were able to document wispy and lenticular laminae (de Castro et al., 2020, their Figs. 7A and 9D), indicating bottom-current reworking in a sand-starved system. • Stow and Smillie (2020, their Fig. 13) presented the original muddy contourite facies model of Gonthier et al. (1984) with a sandy interval in the middle in portraying basecut-out contourites and top-cut-out contourites. Walker (1965) originally applied this “cut-out” logic to the Bouma Sequence. This approach assumes that the entire sequence was present at the time of deposition, but portions were cut-out later by some mysterious forces. If this were true, one could assume anything to arrive at a desired interpretation, such as turbidite, contourite, tidalite, or seismite! Process interpretations must be based on observations and not on assumptions. • Historically, the contourite facies model of Gonthier et al. (1984) has been represented as follows. • Gonthier et al. (1984, their Fig. 12) originally published the model without five internal divisions of C1-C5 and without hiatus at the bottom of the sandy interval in the middle. • Stow and Fauge`res (2008, their Fig. 13. 9) published the model with five internal divisions of C1-C5 and with hiatus at the bottom of the sandy interval in the middle. • Fauge`res and Mulder (2011, their Fig. 3.18) published the model without five internal divisions of C1-C5 and without hiatus at the bottom of the sandy interval in the middle.

Mass transport, gravity flows, and bottom currents

8.7 Wind-driven bottom currents

349

• Rebesco et al. (2014, their Fig. 24) published the model with five internal divisions of C1-C5 and with hiatus at the bottom of the sandy interval in the middle. • Most recently, Stow and Smillie (2020, their Fig. 13) published the model with five internal divisions of C1-C5 and without hiatus at the bottom of the sandy interval in the middle. • The reason for such inconsistent and confusing portrayal of the same model has never been explained by the authors. Consequently, if the hiatus were truly present in the middle of the contourite facies model, then the contourite facies model (Fig. 8. 17) is obsolete. This is because the model cannot function as a predictor, a requirement of a facies model (Walker, 1992). As noted earlier, hiatuses preclude application of the Walther’s Law for predicting facies (Middleton, 1973).

8.7 Wind-driven bottom currents 8.7.1 The Gulf Stream and the Loop Current The Gulf Stream is a powerful, warm, and swift Atlantic Ocean current that originates at the tip of Florida and follows the eastern coastlines of the United States and Newfoundland before crossing the Atlantic Ocean. The Gulf Stream proper is a westernintensified current, largely driven by wind stress (Wunsch, 2002). The Loop Current in the eastern Gulf of Mexico is a wind-driven surface current (Fig. 8.21A), and it is genetically linked to the Gulf Stream in the Atlantic Ocean. The Loop Current enters the Gulf of Mexico through the Yucatan Strait as the Yucatan Current; it then flows in a clockwise loop in the eastern Gulf as the Loop Current, and exits via the Florida Strait as the Florida Current (Nowlin and Hubert, 1972; Mullins et al., 1987). Finally, this current merges with the Antilles Current to form the Gulf Stream (Fig. 8.21A). The Loop Current also propagates eddies into the north-central Gulf of Mexico, where the Ewing Bank area is located, a case study used in this chapter (Fig. 8.21B). The eastern part of Gulf of Mexico is strongly influenced by the Loop Current (Fig. 8.22) and in some cases the Loop Current influences the intensity of cyclones causing related bottom currents (Fig. 8.23).

8.7.2 Current velocity Velocities in eddies that have detached from the Loop Current have been recorded as high as 200 cm s21 at a depth of 100 m (Cooper et al., 1990). Current-velocity measurements, bottom photographs, high-resolution seismic records, and GLORIA side-scan sonar records indicate that the Loop Current influences the seafloor at least periodically in the Gulf of Mexico (Pequegnat, 1972). Computed flow velocities of the Loop Current vary from nearly 100 cm s21 at the sea surface to more than 25 cm s21 at 500 m water depth (Nowlin and Hubert, 1972). This high surface velocity suggests a wind-driven origin for these currents. Flow velocities measured using a current meter reach up to 19 cm s21 at a depth of 3286 m (Pequegnat, 1972). Kenyon et al. (2002b) reported 25 cm s21 current velocity measured 25 m above the seafloor. Such currents are capable of reworking fine-grained sand on the seafloor. Current ripples, composed of sand at a depth of 3091 m on the seafloor (Fig. 8.24), are the clear evidence of deep bottom-current activity in the Gulf of

Mass transport, gravity flows, and bottom currents

350

8. Bottom currents

FIGURE 8.21 (A) Sea surface temperature (SST) image showing the Loop Current in the Gulf of Mexico and the axis of the Gulf Stream in the Atlantic Ocean along the U.S. Continental margin on March 12, 2011. Darker orange to red color enhancement represents temperatures in the upper 70 s F (upper 20 s C). (B). Location map of the Ewing Bank and adjacent areas in the Northern Gulf of Mexico. Solid dots show locations of cores. Source: NOAA’s Cooperative Institute for Meteorological Satellite Studies (CIMSS), University of Wisconsin-Madison, USA, http://cimss.ssec.wisc.edu/goes/blog/wpcontent/uploads/2011/03/MODIS_SST_20110312_1615_largescale.png (accessed 29.03.11). Figure from Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524. with permission from Elsevier. (B) After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deep-marine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, with permission from AAPG. Mass transport, gravity flows, and bottom currents

FIGURE 8.22 A rendering of present-day sea-floor topography of the Gulf of Mexico showing the distribution of the Yucatan Current, the Loop Current, and the Florida Current. Deep thermohaline driven water masses are unable to enter the Gulf of Mexico from the Atlantic Ocean due to complex sea-floor topographic barriers in the Caribbean basins. Source: After Figure from Shanmugam, G., 2006a. DeepWater Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476. Elsevier.

FIGURE 8.23 Image of sea surface height (cm) showing that the warm waters of the Loop Current (red) were 50 75 cm higher than the surrounding water when Hurricane Katrina passed through the Loop Current during August 26, 27, 28, and 29, 2005. Note that the wind speeds (mi hr21) of Hurricane Katrina increased dramatically as it passed over the warm waters of the Loop Current toward the Gulf Coast. Hurricane Katrina’s wind speed is highlighted by Saffir Simpson Scale categories 2 5 (see Tropical cyclone section in Appendix A). The image was produced by a University of Colorado at Boulder team, and processed at CU-Boulder’s Colorado Center for Astrodynamics Research (CCAR). Source: http://www.colorado.edu/news/releases/2005/358.html (accessed 31.05.11). After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524. Elsevier.

352

8. Bottom currents

FIGURE 8.24 Undersea photograph showing possible muddraped (arrow) current ripples at 3091 m water depth in the Gulf of Mexico. Similar mud drapes may explain the origin of mud offshoots observed in the core (see Fig. 8.25). A current measuring nearly 18 cm s21 was recorded on the day this photograph was taken. Current flow was from upper left to lower right. Bar scale is 50 cm. Alaminos Cruise 69-A-13, St. 35. Source: Photograph originally published by Pequegnat, W.E., 1972. A deep bottom-current on the Mississippi Cone. In: Capurro, L.R.A., Reid, J.L. (Eds.), Contribution on the Physical Oceanography of the Gulf of Mexico. Texas A&M University Oceanographic Studies 2. Gulf Publishing Co., Houston, pp. 65 87. From Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524, with permission from Elsevier.

Mexico today (Pequegnat, 1972). These current ripples are draped by thin layers of mud. If these mud drapes on sand ripples were preserved in the rock record, they would be termed “mud-offshoots” (Shanmugam et al., 1993a).

8.7.3 Ewing Bank 826 Field: a case study Deposits of the Loop Current have been interpreted in the cores from the Ewing Bank 826 Field, Plio-Pleistocene, Gulf of Mexico (Shanmugam et al., 1993a,b). The Ewing Bank Block 826 Field is located nearly 100 km off the Louisiana coast in the northern Gulf of Mexico (Fig. 8.21B). It contains hydrocarbon-producing clastic reservoir sands that have been interpreted as bottom-current reworked sands (Shanmugam et al., 1993a,b). Cores from the Gulf of Mexico show a variety of traction structures that include horizontal layers (Fig. 8.25A), low-angle cross-laminae (Fig. 8.25B), ripple cross laminae, flaser bedding (Fig. 8.26), mud offshoots in ripples (Fig. 8.27), eroded and preserved ripples, and inverse grading (Fig. 8.28) (Shanmugam et al., 1993a,b). Most of the features are interpreted as the products of deposition by traction or combined traction and suspension. As with deposits of contour-following thermohaline currents, it is impossible to establish that a given sedimentary structure in the rock record was originated by wind-driven bottom currents,

Mass transport, gravity flows, and bottom currents

353

8.7 Wind-driven bottom currents

FIGURE 8.25 (A) Core photograph showing rhythmic layers of sand and mud. Middle Pleistocene, Gulf of Mexico. (B) Core photograph showing discrete thin sand layers with sharp upper contacts (top arrow). Traction structures include horizontal laminae, low-angle cross-laminae, and starved ripples. Dip of cross-laminae to the right suggests current from left to right. Note rhythmic occurrence of sand and mud layers. Middle Pleistocene, Gulf of Mexico. Source: (A) From Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, p. 524, with permission from Elsevier. (B) After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deep-marine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, with permission from AAPG.

FIGURE 8.26 Photograph showing flaser bedding. Note the presence of mud in ripple troughs (arrow). Upper Pliocene, Ewing Bank Block 826, Gulf of Mexico. Source: After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deep-marine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

Mass transport, gravity flows, and bottom currents

354

8. Bottom currents

FIGURE 8.27 Core photograph showing discrete sand units with current ripples and mud offshoots (arrow). Note sigmoidal configuration of ripples and truncated tops. Middle Pleistocene, Ewing Bank Block 826, Gulf of Mexico. Source: After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deepmarine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

FIGURE 8.28 Photomicrograph of a fine-grained sand layer (arrow) showing microscopic inverse size grading and sharp upper contact. Note largest quartz grain (tip of arrow) at the top of this sand layer. Middle Pleistocene, Gulf of Mexico. After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deepmarine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

Mass transport, gravity flows, and bottom currents

355

8.7 Wind-driven bottom currents

FIGURE 8.29 Common primary traction structures recognized in cores from the Ewing Bank area, Gulf of Mexico.

FIGURE 8.30 Seismic profile showing sheet geometry with continuous and parallel reflection patterns. Note position of cored wells through two sand units examined in the Ewing Bank area, Gulf of Mexico. Core from the L-1 sand shows a dominance of bottom-current reworked sands. Source: After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deep-marine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

without establishing the paleo-water circulation pattern independently. Again, the general term “bottom-current reworked sands” is appropriate in many cases. Sand layers with traction structures occur in discrete units, but not as part of a vertical sequence of structures formed by a single process (Fig. 8.29). Because traction structures are also observed in deposits of different bottom currents, caution must be exercised in classifying a deposit as a “contourite” based solely on traction structures without

Mass transport, gravity flows, and bottom currents

356

8. Bottom currents

FIGURE 8.31 A schematic hybrid flow depositional model for the Pliocene and Pleistocene sands in the Ewing Bank Block 826 area. Note that bottom-current reworked sands in the interchannel area constitute a distinctly different facies from channel-levee facies. Compare the position of No. 3 and No. 3ST wells in the model with their position in the seismic profile. Cored intervals are shown by thick lines. Source: After Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deep-marine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259, reprinted by permission of the American Association of Petroleum Geologists whose permission is required for further use.

independent evidence for contour-following bottom currents in the area. A seismic profile through the study area (Fig. 8.30) and a depositional model (Fig. 8.31) of the Ewing Bank area show the complexity of deep-water processes and deposits.

8.8 Tidal bottom currents in submarine canyons 8.8.1 Types of submarine canyons Deep-marine tidal bottom currents in submarine canyons and in their vicinity are one of the best-studied and most extensively documented modern geologic processes (e.g., Shepard et al., 1969, 1979; Shepard, 1976; Beaulieu and Baldwin, 1998; Petruncio et al., 1998; Xu et al., 2002). During the past five decades, an understanding of deep-marine tidal bottom currents has been achieved by synthesizing a great wealth of information that includes direct observations from deep-diving vehicles, dredging from canyon floors, underwater photographs of canyon floors, photographs and X-radiographs of box cores, seismic profiles, and current-velocity measurements (Shepard and Dill, 1966; Shepard et al., 1969, 1979; Dill et al., 1975; Shepard, 1976). Selected examples of studies that dealt with tidal processes and/ or their deposits in modern and ancient deep-water environments have been reviewed by Shanmugam (2003a, 2008b). In order to discuss deep-marine tidal bottom currents, one must first establish a clear framework on submarine canyons. This is necessary because tidal currents tend to focus their energy within submarine canyons. As discussed earlier, Harris and Whiteway (2011), based on ETOPO1 bathymetric grid, have compiled the first inventory of 5849 separate large submarine canyons in the world’s oceans. They have classified canyons into three types, namely Type 1: shelf-incising canyons having heads with a clear Mass transport, gravity flows, and bottom currents

8.8 Tidal bottom currents in submarine canyons

357

FIGURE 8.32 (A) Proposed classification of canyons for understanding tidal energy. This is a modified scheme of Harris and Whiteway (2011). Type 1A: land-incising canyon into estuary or river system with tidal influence. Type 1B: shelf-incising canyons having heads with connection to a major river or estuarine system, but they do not incise onto the land. Type 2: shelf-incising canyons with no clear connection to a major river or estuarine system. Type 3: slope-incising blind canyons with their heads confined to the continental slope. (B) Increasing influence of surface (barotropic) tides from Type 3 to Type 1A canyons. (C) Increasing influence of internal (baroclinic) tides from Type 1A to Type 3 canyons.

bathymetric connection to a major river system, Type 2: shelf-incising canyons with no clear bathymetric connection to a major river system, and Type 3: blind canyons that incise onto the continental slope. In order to differentiate the role of tidal currents in different types of submarine canyons, I have further subdivided the Type 1 into Type 1A and Type 1B based on the position of canyon head (Fig. 8.32).

8.8.2 Current velocity Shepard et al. (1979) measured current velocities in 25 submarine canyons worldwide at water depths ranging from 46 to 4200 m by suspending current meters, usually 3 m above the sea bottom (Fig. 8.33A). Shepard et al. (1979) also documented systematically that upand downcanyon currents closely correlated with timing of tides (Fig. 8.33B). These canyons include the Hydrographer, Hudson, Wilmington, and Congo (Zaire) in the Atlantic Ocean; and the Monterey, Hueneme (Fig. 8.33B), Redondo, La Jolla/Scripps, and Hawaii canyons in the Pacific Ocean. Maximum velocities of up- and downcanyon currents commonly ranged from 25 to 50 cm s21 (Fig. 8.33A) (Table 8.4). Keller and Shepard (1978) Mass transport, gravity flows, and bottom currents

358

8. Bottom currents

FIGURE 8.33

(A) Conceptual diagram showing a cross-section of a submarine canyon with ebb and flood tidal currents (opposing arrows). Shepard et al. (1979) measured current velocities in 25 submarine canyons at water depths ranging from 46 to 4200 m by suspending current meters commonly 3 m above the sea bottom. Measured maximum velocities commonly range from 25 to 50 cm s21. (B) Time-velocity plot from data obtained at 448 m in the Hueneme Canyon, California, showing excellent correlation between the timing of up- and downcanyon currents and the timing of tides obtained from tide tables (solid curve). 3mAB 5 Velocity measurements were made 3 m above sea bottom. Source: (A) Figure from Shanmugam, G., 2003a. Deepmarine tidal bottom currents and their reworked sands in modern and ancient submarine canyons. Mar. Pet. Geol. 20, 471 491, with permission from Elsevier. (B) Modified after Shepard, F.P., Marshall, N.F., McLoughlin, P.A., Sullivan, G.G., 1979. Currents in Submarine Canyons and Other Sea Valleys. AAPG Studies in Geology, No. 8, p. 173, with permission from AAPG.

reported velocities as high as 7075 cm s21, velocities sufficient to transport even coarsegrained sand, from the Hydrographer Canyon. In the Niger Delta area of West Africa, five modern submarine canyons (Avon, Mahin, Niger, Qua Iboe, and Calabar) have been recognized. In the Calabar River, the tidal range is 2.8 m and tidal flows with reversible currents are common (Allen, 1965). In the Calabar Estuary, maximum ebb-current velocities range from 60 to 280 cm s21, and flood current velocities range from 30 to 150 cm s21. These velocities are strong enough to transport particles of sand and gravel size. The Calabar Estuary has a deep-water counterpart that cuts through sediments of the outer shelf and slope, forming the modern Calabar Submarine canyon. Thus, as they do in the

Mass transport, gravity flows, and bottom currents

359

8.8 Tidal bottom currents in submarine canyons

TABLE 8.4 Selected examples of maximum velocities of up- and downcanyon currents measured at varying water depths by suspending current meters 3 m above sea bottom. Water depth (m)

Submarine canyon and location

Mean tidal range (m)

Upcanyon current velocity (cm s21)

Downcanyon current velocity (cm s21)

348

Hydrographer, Atlantic, United States

1.6

39

52

375

La Jolla, Pacific, United States

2.5

19

18

400

Congo (Zaire), West Africa 1.3

22

13

448

Hueneme, Pacific, United States

2.4

32

32

458

Santa Monica, Pacific, United States

2.7

27

30

914

Wilmington, Atlantic, United States

1.8

20

21

1445

Monterey, Pacific, United States

2.0

30

30

1737

Kaulakahi, Pacific Islands

0.9

26

24

1904

Rio Balsas, Mexico

0.7

21

21

Compiled from Shepard, F.P., Marshall, N.F., McLoughlin, P.A., Sullivan, G.G., 1979. Currents in submarine canyons and other sea valleys. In: AAPG Studies in Geology, vol. 8, p. 173.

Zaire Canyon to the south, tidal currents are likely to operate in the Calabar and Qua Iboe Canyons. Unlike the robust datasets on measured current velocities of deep-marine tidal currents in submarine canyons worldwide (Shepard et al., 1979), there has not been a single published current velocity of HDTC in submarine canyons.

8.8.3 Identification Sedimentary features indicative of tidal processes in shallow-water environments have been well established (e.g., Reineck and Wunderlich, 1968; Klein, 1970; Visser, 1980; Terwindt, 1981; Allen, 1982; Banerjee, 1989; Nio and Yang, 1991; Dalrymple, 1992; Archer, 1998; Shanmugam et al., 2000). Traction structures that develop in shallow-water estuaries also develop in deep-water canyons and channels with tidal currents (Shanmugam, 2003a). Klein (1975), based on studies of Deep Sea Drilling Project (Leg 30, Sites 288 and 289) cores, suggested that current ripples, micro-cross-laminae, mud drapes, flaser bedding, lenticular bedding, and parallel laminae reflect alternate traction and suspension deposition from tidal bottom currents in deep-marine environments. Perhaps the single most diagnostic structure is the double mud layers (DMLs) in deep-water strata that clearly suggest deposition by tidal bottom currents in offshore Nigeria (Figs. 8.34 and 8.35), Bay of Bengal (Fig. 8.36), and France (Fig. 8.37). Visser (1980) explained the origin of DMLs (Fig. 8.38). Canyon-fill facies tend to be composed of a mix of MTD, DML, and other traction structures (Figs. 8.38 and 8.39).

Mass transport, gravity flows, and bottom currents

360

8. Bottom currents

FIGURE 8.34 Core photograph showing double mud layers (DML), indicative of deposition by deep-marine tidal currents, in a submarine-canyon setting. Pliocene strata, Edop Field, offshore Nigeria. Source: From Shanmugam, G., 2003a. Deep-marine tidal bottom currents and their reworked sands in modern and ancient submarine canyons. Mar. Pet. Geol. 20, 471 491, with permission from Elsevier.

FIGURE 8.35 Core photograph showing thick-thin sand layers with double mud layers (DML) in Pliocene sand. Edop Field, Pliocene, offshore Nigeria.

Mass transport, gravity flows, and bottom currents

361

8.8 Tidal bottom currents in submarine canyons

FIGURE 8.36 (A) Sedimentological log showing alternation of sand (lithofacies 3) and mudstone (lithofacies 4) intervals with continuous presence of double mud layers (DML). Note muddy debrite facies (Lithofacies 2) near the bottom. Wentworth grain-size classes: C 5 clay; S 5 silt; VFS 5 very fine sand; FS 5 fine sand; MS 5 medium sand. (B) Lithofacies 3 core photograph showing rhythmic bedding (rhythmites) and double mud layers (DML, arrows) in sand. N 5 neap (thin) bundle; S 5 spring (thick) bundle. Note that one could designate the DML intervals as Tb and the massive sand unit (between scale divisions 2 and 4 cm) as Ta using the Bouma Sequence; however, Shanmugam et al., (2009) did not. Source: Core photograph from Shanmugam et al., (2009), with permission from SEPM. See Figures 3.35, 3.36, and 3.37 for the stratigraphic and geographic position of this well and cored intervals in a submarine canyon. FIGURE 8.37 Outcrop photograph of Annot Sandstone (Eocene-Oligocene) showing double mud layers (DML, arrows). Note mud-draped ripples between upper and middle arrows. Source: After Shanmugam, G., 2003a. Deep-marine tidal bottom currents and their reworked sands in modern and ancient submarine canyons. Mar. Pet. Geol. 20, 471 491, with permission from Elsevier.

Mass transport, gravity flows, and bottom currents

362

8. Bottom currents

FIGURE 8.38 Schematic diagram illustrating stages of development of double mud layers by tidal currents. Source: From Visser, M.J., 1980. Neap-spring cycles reflected in Holocene subtidal large-scale bedform deposits: a preliminary note. Geology 8, 543 546. GSA.

Submarine canyons are not only unique for providing a protected environment for focusing tidal energy from shallow-marine estuaries to deep-marine canyons but also prone to generating mass movements (e.g., slides, slumps, and debris flows) due to failure of steep canyon walls. Recognition of tidal facies in deep-water sequences is important in understanding sand distribution because deposits of tidal processes and mass-transport processes (i.e., slides, slumps, and debris flows) characterize fills of modern and ancient submarine canyons (Fig. 8.39). This complex facies association (Fig. 8.39), mimicking both shallow-water and deep-water deposits, has been recognized in the modern La Jolla canyon box cores (offshore California), ancient Qua Iboe Canyon conventional cores (Pliocene, Edop Field, offshore Nigeria), ancient Pliocene canyon conventional cores [Krishna-Godavari (KG) Basin, Bay of Bengal], and ancient Annot Sandstone outcrops (Eocene-Oligocene, onshore SE France), among others. It appears that the association of tidal and mass flow facies is unique to canyon environments. Therefore, this facies association may be used as a criterion for inferring submarine canyon settings in the rock record where direct evidence for canyon filling is lacking. Because MTD (i.e., deposits of slides, slumps, and

Mass transport, gravity flows, and bottom currents

8.9 Baroclinic currents (internal waves and internal tides)

363

FIGURE 8.39 Submarine canyon-fill facies dominated by MTD and DML. Source: From Shanmugam, G., 2003a. Deepmarine tidal bottom currents and their reworked sands in modern and ancient submarine canyons. Mar. Pet. Geol. 20, 471 491, with permission from Elsevier.

debris flows) can occur both inside and outside submarine canyons, the correct identification of tidal facies in deep-water sequences is extremely critical in establishing the facies association. In channel-mouth environments, downslope turbidity currents are likely to develop depositional lobes, whereas bidirectional tidal bottom currents are likely to develop elongate bars. Turbidite lobes are aligned perpendicular to channel axis, whereas tidal bars are aligned parallel to channel axis (Shanmugam, 2003a). Depositional lobes are likely to be much larger than channel width, whereas tidal sand bars are thought to be much smaller than channel width. Deep-water elongate tidal bars are speculated to be analogous to tidal bar sands that develop in shallow-water estuarine environments (see Shanmugam et al., 2000). In frontier exploration areas, an incorrect use of a turbidite-lobe model (with sheet geometry) instead of a tidal bar model (with bar geometry) will result in an unrealistic overestimation of sandstone reservoirs.

8.9 Baroclinic currents (internal waves and internal tides) 8.9.1 Basic concept Gill (1982) discussed the basic differences between barotropic (surface) waves that develop at the air water interface and baroclinic (internal) waves that develop at the water water interface (Fig. 8.40). Fluid parcels in the entire water column move together in the same direction and with same velocity in a surface wave, whereas fluid parcels in shallow and deep layers of the water column move in opposite directions and with different velocities in an internal wave (Fig. 8.40). The surface displacement and interface displacement are the same for a surface wave, while the interface displacements are large for internal wave. Although the free surface movement

Mass transport, gravity flows, and bottom currents

364

8. Bottom currents

FIGURE 8.40 (A) Top panel: barotropic (surface) wave showing the movement of fluid parcels in the entire water column (H) in the same direction with same velocity (short horizontal arrows). (B) Bottom panel: baroclinic (internal) wave showing the fluid-parcel movement in upper (H1) and lower (H2) layers in opposite directions (short horizontal arrows) with different velocities (U1 and U2). (C) Explanation of symbols. Source: Modified after Gill, A. E., 1982., Atmosphere-Ocean Dynamics. International Geophysics Series, vol. 30. Academic Press, An Imprint of Elsevier, San Diego, 662 p. After Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811. AAPG.

FIGURE 8.41 Conceptual diagram showing internal waves along pycnoclines. Source: From Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deepmarine baroclinic sands. AAPG Bull. 97, 767 811. AAPG.

associated with the baroclinic mode is only 1/400 of the interface move movement, this is still sufficient for baroclinic motions to be detectable by sea-surface changes (Wunsch and Gill, 1976).

8.9.2 Empirical data Apel (2002), Apel et al. (2006), and Jackson (2004) have published comprehensive accounts of internal waves and tides worldwide. A sedimentologic and oceanographic review is provided

Mass transport, gravity flows, and bottom currents

8.9 Baroclinic currents (internal waves and internal tides)

365

FIGURE 8.42 (A) Map showing 51 locations (red dots) of observed oceanographic internal waves and tides in coastal seas and in the open ocean (after Apel, 2002; Jackson, 2004a). (B) Explanation of symbols and numbers. Yellow triangles and numbers represent locations of internal waves used for physical properties in a study by Shanmugam (2013a, his Table 2). Source: Base map courtesy of C.R. Jackson, Global Ocean Associates. From Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811, with permission from AAPG.

by Shanmugam (2013a,b). Internal waves are gravity waves that oscillate along oceanic pycnoclines (Fig. 8.41). Internal tides are internal waves with a tidal frequency. Internal solitary waves (i.e., solitons), the most common type, are commonly generated near the shelf edge (100 200 m in bathymetry) and in the deep ocean over areas of seafloor irregularities, such as mid-ocean ridges, seamounts, and guyots. Empirical data from 51 locations (Fig. 8.42) in the Atlantic, Pacific, Indian, Arctic, and Antarctic oceans reveal that internal solitary waves travel in packets (Fig. 8.43). Internal waves commonly exhibit (1) higher wave amplitudes (5 50 m) than surface waves (,2 m), (2) longer wavelengths (0.5 15 km) than surface waves (100 m), (3) longer wave periods (5 50 minutes) than surface waves (9 10 seconds), and (4) higher wave speeds (0.5 2 m s21) than surface waves (25 cm s21). Maximum speeds of 48 cm s21 for baroclinic currents were measured on guyots.

8.9.3 Depositional framework A preliminary depositional framework is proposed for continental slopes, submarine canyons, and guyots (Fig. 8.44A). To date, the only convincing photographic documentation of sedimentary bedforms, attributed to internal waves and internal tides, in modern deep-marine environments has been from seamounts and guyots in the Pacific Ocean (Menard, 1952; Lonsdale et al., 1972). These photographs are the only empirical foundation for gaining insights into depositional processes (e.g., traction vs suspension) associated

Mass transport, gravity flows, and bottom currents

366

8. Bottom currents

FIGURE 8.43 (A) Index map showing the study area in the Sulu Sea (arrowhead). (B) Satellite image showing three packets (P1, P2, and P3) of internal waves, forming a wave train (three yellow arrows) in the Sulu Sea between the Philippines (to the northeast) and Malaysia (to the southwest). This true-color Aqua ModerateResolution Imaging Spectroradiometer (MODIS) image was acquired on April 8, 2003. Source: (A) Map credit: http://www.ngdc.noaa.gov/mgg/topo/img/globeco3.gif (accessed 01.10.12.). (B) Image courtesy of Jacques Descloitres, AQ11 MODIS Land Rapid Response Team at the National Aeronautics and Space Administration/GSFC. http://earthobservatory. nasa.gov/Newsroom/NewImages/images.php3?img_id1415334 (accessed 01.10.12.).

with internal waves and internal tides. A sketch of these baroclinic traction bedforms is illustrated in Fig. 8.44B. However, core-based sedimentological studies of modern sediments emplaced by baroclinic currents on continental slopes, in submarine canyons, and on submarine guyots are lacking. There are no cogent sedimentologic or seismic criteria for distinguishing ancient counterparts. Therefore, a few outcrop-based facies models of these deposits are unsustainable. Not surprisingly, interpretations of ancient strata using outcrops as deposits of internal waves and internal tides have resulted in a series of lively debates (Dunham and Saller, 2014; Shanmugam, 2014a). These problems are further complicated by a myriad of propagation directions of internal waves with respect to the shelf edge (Fig. 8.45), which complicates current directions in interpreting the ancient rock record.

8.10 Sediment provenance 8.10.1 Current directions Commonly, primary sedimentary structures and related current directions are used in deciphering sediment provenance (Pettijohn, 1975; Potter and Pettijohn, 1977; Zuffa, 1985).

Mass transport, gravity flows, and bottom currents

8.10 Sediment provenance

367

FIGURE 8.44 (A) Conceptual oceanographic and sedimentologic framework showing deposition from baroclinic currents on continental slopes, in submarine canyons, and on guyots. On continental slopes and in submarine canyons, deposition occurs in three progressive stages: (1) incoming internal wave and tide stage, (2) shoaling transformation stage, and (3) sediment transport and deposition stage. Continental slopes and submarine canyons are considered to be environments with high potential for deposition from baroclinic currents. In the open ocean, baroclinic currents can rework sediments on flat tops of towering guyot terraces, without the need for three stages required for deposition on continental slopes. In this model, basin plains are considered unsuitable environments for deposition of baroclinic sands. Not to scale. (B) Crossprofile showing asymmetrical dunes and asymmetrical ripples observed from side-looking sonar and photographic evidence obtained from the terrace of the Horizon Guyot, Mid-Pacific Mountains. Bathymetry of bedforms: 1630 1632 m. Dune heights (H) were estimated from the length of acoustic shadows. Source: (A) From Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811, with permission from AAPG. (B) Redrawn from Lonsdale, P., Nornaark, W.R., Newman, W.A., 1972. Sedimentation and erosion on Horizon Guyot. GSA Bull. 83, 289 316 (their Fig. 10), with permission from the Geological Society of america. After Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811. AAPG.

However, complex current directions associated with all four types of bottom currents pose immense challenges in inferring the primary sediment source. For example (Fig. 8.46): • Contour currents are global in circulation pattern and flow parallel to the strike of the regional slope (Fig. 8.4). • Wind-driven bottom currents are complex in circulation pattern in the Gulf of Mexico (Fig. 8.22), which include circular motions (gyres) unrelated to the slope. Such bottom

Mass transport, gravity flows, and bottom currents

368

8. Bottom currents

FIGURE 8.45 Maps showing the variable directions of propagation of internal waves with respect to shoreline or shelf edge seen as surface manifestations on satellite images. (A) Internal waves propagating toward the shoreline of Palawan Island in the Sulu Sea. (B) Internal waves propagating away from the shoreline or shelf edge in the Yellow Sea (Hsu et al., 2000, their Fig. 8). (C) Internal waves propagating nearly parallel to the shoreline of northern Somalia in the Indian Ocean (Jackson, 2004b, his Fig. 3). (D) Internal waves propagating parallel to the strait or channel axis in the Strait of Messina. (E) Internal waves propagating in the same direction on both sides of the Strait of Gibraltar. Note the position of the Camarinal Sill at the point of origin of internal waves (Go´mez-Enri et al., 2007). (F) Internal waves propagating in opposite directions from the point of origin, which is a sill in the Lombok Strait (Susanto et al., 2005). Baroclinic currents, associated with internal waves and tides, are reworking agents and as such they are unrelated to the primary sediment provenance. Features shown are schematic and not to scale. Source: From Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811, with permission from AAPG.

currents have been reported beneath the Gulf Stream Gyre at a depth of nearly 4 km in the northern Bermuda Rise (Laine, 1978). Laine and Hollister (1981) suggest that the deep Gulf Stream return flow entrains suspended sediment in a deep gyre and may be responsible for the deposition at the base of the continental rise. • Deep-marine tidal currents are bidirectional in nature and they flow up and down submarine canyons (discussed below). • Baroclinic currents are extremely variable in propagation directions with respect to sediment source (Fig. 8.45).

Mass transport, gravity flows, and bottom currents

8.10 Sediment provenance

369

FIGURE 8.46 Four conceptual models showing the physical relationship between primary sediment provenance and current directions (red arrows) in deep-marine environments. (A) Downslope, unidirectional, turbidity currents. Current ripples in turbidites are reliable indicators of sediment provenance. (B) Alongslope, thermohaline-driven contour currents. Current ripples and cross-beddings in contourites are not reliable indicators of sediment provenance. (C) Circular, wind-driven bottom currents. Current ripples in these deposits are not reliable indicators of sediment provenance. (D) Bidirectional, tide-driven bottom currents are common in submarine canyons (Shepard et al., 1979). Current ripples in deep-marine tidalites are also not reliable indicators of sediment provenance. Some sites, such as the Gulf of Cadiz that served as the type locality for the contourite facies model, are also affected by bottom currents associated with internal waves, cyclones, and tsunamis, causing complex current directions. Source: From Shanmugam, G., 2016b. The contourite problem. In: Mazumder, R. (Ed.), Sediment Provenance. Elsevier, pp. 183 254 (Chapter 9). Elsevier.

• Because bottom currents are strictly a reworking agent, their sedimentary structures do not reflect the true direction of the primary sediment source (Fig. 8.46). Therefore, the conventional approach of inferring source directions (i.e., sediment provenance), using current ripples and cross beddings, is unreliable when dealing with deepmarine bottom currents and their deposits (Fig. 8.46)

8.10.2 Detrital composition The other important criterion in interpreting sediment provenance is the detrital composition (Zuffa, 1985; Arribas et al., 2007). However, reworking by bottom currents may not

Mass transport, gravity flows, and bottom currents

370

8. Bottom currents

alter the original composition of the sediment derived from the primary provenance. For example, in understanding the compositional difference between contourites and turbidites in the Bounty Submarine Fan, New Zealand, cored intervals from the Ocean Drilling Program (ODP) Site 1122 on Leg 181 have been studied. In discussing the results, Shapiro et al. (2007, p. 277) state that “. . . there are no significant trends among thickness, grain size, composition, and depth of Site 1122 sand samples, except that thicker beds tend to contain slightly more metamorphic rock fragments. The generally homogeneous composition of Site 1122 sand indicates that it may have had a relatively uniform source back into the early Miocene. Thus, the up-section change from sandy contourite to turbidite deposits at Site 1122 is not reflected in sand composition. This suggests that the sand provenance remained constant while the depositional processes of sand at Site 1122 changed.” Distinguishing compositional variations caused by variations in deep-sea depositional processes is a potential area of future research on sediment provenance.

8.11 Reservoir quality Perhaps the first application of the contourite concept to a major petroleum reservoir was in the Frigg Field, North Sea (Heritier et al., 1979). These authors interpreted a wavy surface, between wells 25/1 1 and 25/1 5, on a seismic profile as evidence for contour currents. The Frigg Field was considered as one of the largest gas fields in the world in the 1970s. Despite numerous published contourite reservoirs (Shanmugam et al., 1993a, 1995b; Moraes et al., 2007; Viana, 2008; Mutti and Carminatti, 2011; Shanmugam, 2006a, 2012a, 2014a; Maslin, 2015), some petroleum geologists still believe that reservoir quality of bottom-current reworked sands, which include contourites, is poor in comparison to that of turbidites. In discussing the reservoir quality of deep-water Miocene sands in the Kutei Basin, Makassar Strait (Fig. 8.2, location E), Dunham and Saller (2014) claim that “The key point from the perspective of the Exploration-Geologist is that bottom currents did not transport or redistribute these Kutei basin reservoir-sands from their original-depositional locations. If significant redistribution of sand had occurred, our exploration-model would have failed, and we would not have found thick high-quality reservoir sands in our prospects. We based our interpretations (Saller et al., 2006, 2008b) on evidence from seismic data, cores, and exploration discoveries.” Contrary to the above claim, published data do show that bottom-current reworked sands have good porosity and permeability. Selected examples are the following: • Off the Great Bahama Bank (Fig. 8.47) (see also Eberli and Betzler, 2019; Mulder et al., 2019), sandy calciclastic contourites (Middle Miocene to Pleistocene) have a measured maximum porosity of 40% and a maximum permeability of 9880 mD (Mullins et al., 1980) (Table 8.5). The high permeability has been attributed to the winnowing away of muds from the intergranular primary pores by vigorous contour currents (Fig. 8.47B). These carbonate sandy contourite drifts are hemiconical-shaped bodies that are up to 600 m in thickness and nearly 60 km in length (Fig. 8.48). • In the Ewing Bank Block 826 area (Fig. 8.21B), bottom-current reworked sands (PlioPleistocene) show 25% 40% measured porosity and 100 1800 mD permeability (Shanmugam et al., 1993a, their Table 1). Individual reworked sand layers commonly range in thickness from 5 to 10 cm, but the entire unit reached up to 6 m in total

Mass transport, gravity flows, and bottom currents

371

8.11 Reservoir quality

FIGURE 8.47 (A) Image showing the Great Bahama Bank and the Florida Current (see also Fig. 8.21A). (B) SEM photograph showing mud-free nature of primary intergranular pores in foraminifera-rich carbonate sandy contourites, northern Straits of Florida. (C) SEM photograph showing primary intragranular porosity. Note absence of mud in pores due to winnowing by bottom currents, which results in high porosity and permeability. Source: (A) NASA. (C) Samples courtesy of H.T. Mullins. After Shanmugam, G., 2017b. Contourites: physical oceanography, process sedimentology, and petroleum geology. Pet. Explor. Dev. 44 (2), 183 216. Elsevier.

TABLE 8.5 Measured properties of modern sandy and muddy contourites.

Sample

Location (bottom current)

Porosity (%)a

Air permeability (mD)

Weight loss (%)

Total organic carbon (%)

24

0.66

C5: Sandy carbonate contourite

Off Great Bahama Bank (Florida Current)

C6: Sandy carbonate contourite

Off Great Bahama Bank (Florida Current)

37.5

4809

84

0.20

C7: Sandy carbonate contourite

Off Great Bahama Bank (Florida Current)

35.0

688

85

0.18

C8: Sandy carbonate contourite (top hardground part)

Off Great Bahama Bank (Florida Current)

26.1

550

89

0.07

C8: Sandy carbonate contourite (bottom main part)

Off Great Bahama Bank (Florida Current)

39.7

9881

89

0.07

C9: Muddy siliciclastic contourite

Off Bermuda Rise (Gulf Stream)

33.2

186

87

0.35

a Samples were dried at 170 F for 48 hours. Measurements were made at the Field Research Laboratory, Mobil Research and Development Corporation, Dallas, Texas, in 1979. Samples C5 C8: Courtesy of H.T. Mullins (Mullins et al., 1980). Sample C9: Courtesy of E.P. Laine (Laine and Hollister, 1981).

Mass transport, gravity flows, and bottom currents

372

8. Bottom currents

FIGURE 8.48 Sediment prism consisting of carbonate sandy contourites, reworked by the Florida Current, northern Straits of Florida (after Mullins et al., 1980). Source: After Shanmugam, G., 2017b. Contourites: physical oceanography, process sedimentology, and petroleum geology. Pet. Explor. Dev. 44 (2), 183 216. Elsevier.

thickness. Hybrid flow model has been proposed previously for the Ewing Bank area in the Gulf of Mexico (Fig. 8.31). • In the Bay of Bengal (Fig. 8.2, core location 13), high-quality Pliocene petroleumproducing reservoir sands formed by deep-marine sandy debris flows and tidal currents have been documented in the KG Basin. Tidalite sands show measured porosity values of 34% 41% and permeability values of 525 5977 mD (Shanmugam et al., 2009, their Table 4). Individual tidalite units vary from a few centimeters to nearly a meter in thickness (Fig. 8.36B). • In the Gulf of Cadiz, a 10-m thick sheet sand has been interpreted as “contourites” (Stow et al., 2011). Brackenridge et al. (2018) discussed textural characteristics of sandrich contourite systems. • According to Fonnesu et al. (2020), the deep-water Coral and Mamba gas fields were discovered by Eni between 2011 and 2012 in Area 4 offshore Northern Mozambique (Fig. 8.49). These fields host over 80 trillion cubic feet of natural gas in-place distributed in Paleocene to Oligocene reservoirs, that together with neighboring Area 1, represent one of the most prolific exploration areas of the world. Seismic and core data from these fields provides a unique opportunity to understanding the architecture of deep-water sands subjected to hybrid flows, which is the interaction between downslope gravity flows and alongslope bottom currents (Fig. 8.50). Integration of high-quality seismic and extensive well data from both fields shows that very clean (clay matrix-poor) sandstone reservoirs, with thickness .100 m and extended over 10s of km can be formed by the syndepositional interaction of downslope high-density turbulent gravity flows and across-slope bottom currents (Fig. 8.51). In the appraisal well Coral-D, results are: Gross thickness:

72 m

Net to gross:

82%

Average porosity:

16.9%

In summary, bottom-current reworked sands have better reservoir quality than turbidites in many cases (Shanmugam, 2012a, 2014a).

Mass transport, gravity flows, and bottom currents

8.12 Synopsis

373

FIGURE 8.49

Ocean floor bathymetry and circulation in the modern Rovuma basin. (A) The basin is located in the offshore of northern Mozambique, within the Mozambique Channel, a tropical gateway for strong southerly surface currents and rotational eddies, with deeper undercurrents flowing northwards along the continental slope (modified from Breitzke et al., 2017). (B) Shaded relief map overlain by bathymetry of modern depositional systems and circulation pattern. (C) 3D seabottom visualization of the bathymetry highlighting the distribution of modern deep-water fans. (D) Detail of seafloor features including large-scale scour marks (flutes and crescent) created by northerly directed bottom currents. MC 5 Mozambique Current; EACC 5 East African Coastal Current; NADW 5 North Atlantic Deep Water; MCE 5 Mozambique Channel eddy. Source: After Fonnesu, M., Palermo, D., Galbiati, M., Marchesini, M., Bonamini, E., Bendias, D., 2020. A new world-class deep-water play-type, deposited by the syndepositional interaction of turbidity flows and bottom currents: the giant Eocene Coral Field in northern Mozambique. Mar. Pet. Geol. 111, 179 201. Elsevier.

8.12 Synopsis The four basic types of deep-marine bottom currents are: (1) thermohaline-induced geostrophic contour currents, (2) wind-driven bottom currents, (3) tide-driven bottom currents, mostly in submarine canyons, and (4) internal wave/tide-driven baroclinic currents. Contourites are deposits of thermohaline-driven geostrophic contour currents. Hybrid flows are common and their deposits may mimic the Bouma Sequence (i.e., turbidites). Contourites can be muddy or sandy in texture and siliciclastic or calciclastic in composition. Traction structures are common in deposits of all four types of bottom currents.

Mass transport, gravity flows, and bottom currents

374

8. Bottom currents

FIGURE 8.50 Conceptual model of a depositional lobe influenced by bottom currents, showing the distribution of sedimentological log patterns of the different facies tracts, observed in the Coral cores (grain size: P 5 pebbles, G 5 granules, C 5 coarse sand, M 5 medium sand, F 5 fine sand, S 5 silt). After Fonnesu, M., Palermo, D., Galbiati, M., Marchesini, M., Bonamini, E., Bendias, D., 2020. A new world-class deep-water play-type, deposited by the syndepositional interaction of turbidity flows and bottom currents: the giant Eocene Coral Field in northern Mozambique. Mar. Pet. Geol. 111, 179 201. Elsevier.

FIGURE 8.51 Hybrid flow depositional model depositional model showing downslope gravity flows and alongslope bottom currents (compare with Figs. 8.11, 8.12A, and 8.31). Source: After Fonnesu, M., Palermo, D., Galbiati, M., Marchesini, M., Bonamini, E., Bendias, D., 2020. A new worldclass deep-water play-type, deposited by the syndepositional interaction of turbidity flows and bottom currents: the giant Eocene Coral Field in northern Mozambique. Mar. Pet. Geol. 111, 179 201. Elsevier.

However, there are no diagnostic sedimentological or seismic criteria for distinguishing ancient contourites from other three types. DMLs are a reliable criterion for recognizing deep-marine tidalites in cores and outcrops. The Gulf of Cadiz is the type locality for the contourite facies model based on muddy lithofacies. However, there are no genuine contour currents in the Gulf of Cadiz. This site is affected only by transitory contour currents associated with the MOW. Furthermore, this site is affected by other complicating factors such as internal waves and tides, turbidity currents, tsunamis, cyclones, mud volcanism, methane seepage, sediment supply, pore-water venting, and bottom topography. Published

Mass transport, gravity flows, and bottom currents

8.12 Synopsis

375

IODP 339 cores todate (August 21, 2020) from the Gulf of Cadiz do not show well-developed primary sedimentary structures, which are necessary for interpreting depositional processes. Therefore, the contourite facies model is sedimentologically obsolete. Bottomcurrent reworked sands of all four types have the potential for developing petroleum reservoirs. Sandy carbonate contourites, off Great Bahama Bank in the Straits of Florida, have a measured maximum porosity of 40% and a maximum permeability of 9880 mD due to the winnowing away of muds from the intergranular primary pores by vigorous contour currents. These carbonate contourites are hemiconical-shaped bodies that are up to 600 m in thickness and are nearly 60 km in length. Empirical data of modern contourites also show potential for seal and source-rock development. Also, deep-water sandy debrites and associated tidalites are known to produce petroleum in the KG Basin, Bay of Bengal (India). Therefore, future petroleum exploration and development should focus attention on these often overlooked siliciclastic and calciclastic deep-marine contourite and tidalite petroleum reservoirs. The contourite problems, composed of conceptual, nomenclatural, empirical, and methodological issues, have effectively hindered progress on contourite research during the past six decades. Failure to acknowledge and rectify these issues will only further muddle the problem. Because the real-world oceans are ubiquitously affected by multiple processes concurrently, the grand ingrained principle of “one deposit for one flow type” is nothing more than a misplaced optimism. The contourite problem is not just incidental, but it is fundamental to the basic understanding of all deep-water sediments.

Mass transport, gravity flows, and bottom currents

C H A P T E R

9 Soft-sediment deformation structures O U T L I N E 9.1 Introduction

377

9.2 Datasets

395

9.3 Definition

395

9.4 Origin

396

9.5 Classification

401

9.6 Advances

402

9.7 Geological implications based on case studies 405

9.7.1 9.7.2 9.7.3 9.7.4 9.7.5 9.7.6

Breccias Lateral extent Ocean bottom currents Mass-transport deposits Future research Unresolved issues 9.7.6.1 The seismite problem 9.7.6.2 The breccias problem 9.7.6.3 The tombolith problem

9.8 Synopsis

405 405 406 407 413 414 414 417 436 437

9.1 Introduction Since the detailed treatment of the topic of soft-sediment deformation structures (SSDS; also known as penecontemporaneous or synsedimentary deformation structures) in terms of physics and sedimentology by Allen (1984), SSDS have received considerable attention worldwide (e.g., Maltman, 1994a,b,c; Collinson, 1994; Alfaro et al., 2016; Feng et al., 2016; Shanmugam, 2016c, 2017a,c,d). Examples of SSDS are convolute bedding, slump folds, load casts, dish-and-pillar structures, breccias, sand injections (dikes and sills), etc. Although SSDS commonly develop in deep-water flysch sequences by various sedimentgravity flows, such as turbidity currents (Kuenen, 1957; Dzulynski et al., 1959; Bouma, 1962; Sanders, 1965; Van der Lingen, 1969; Lowe, 1975, 1976a,b), SSDS are also common in deposits of other environments (e.g., fluvial, eolian, shelf, etc.) (Table 9.1). The two factors

Mass transport, gravity flows, and bottom currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00009-6

377

© 2021 Elsevier Inc. All rights reserved.

TABLE 9.1 A sedimentological compendium of 110 case studies (see Fig. 9.1, shown as filled black squares), published between 1863 and 2017, on SSDS. Serial number (case study)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

Comments (Shanmugam, 2017a)

1 (Logan, 1863)

Sandwiched occurrence of deformed beds between undeformed beds (limestone)

Synsedimentary origin

Devonian, limestones, Gaspe´ Peninsula, Quebec, Canada

Marine environment

Synsedimentary activity

2 (Seilacher, 1969)

The ideal vertical sequence Earthquakes of “seismites” with four internal divisions: 1. Soupy zone (top) 2. Rubble zone 3. Segmented zone 4. Undisturbed sediment (bottom)

Miocene, shale, Monterey Formation, California, United States

Deep-marine environment

Seilacher (1969) introduced the genetic term “seismite” for deformed beds by earthquakes. However, there are no reliable criteria for recognizing paleoseismicity (Shanmugam, 2016c)

3 (Helwig, 1970)

Slump folds

Early deformation

Late Ordovician, sandstone and shale, Newfoundland Appalachians, Canada

Submarine slope

Hybrid tectonic and sedimentary activities

4 (Kirkland and Anderson, 1970)

Microfolds Sandwiched occurrence of deformed layers between undeformed layers in anhydrites

Structural origin

Permian, anhydrites, castile evaporites, Texas and New Mexico, United States

Deep-water lagoon

Tectonic activity

5 (Fossen, 2010)

Deformation bands

Salt tectonics

Jurassic, Navajo Eolian environment Sandstone, Southeast Utah, United States

Salt tectonics

6 (Shanmugam et al., 1988a)

Duplex-like structures

Sediment loading

Pennsylvanian, sandstone and shale, Ouachita Mountains, Arkansas, United States

Deep-marine environment

Submarine channel-wall collapse

7 (Shanmugam et al., 1994)

Slump folds, contorted layers, brecciated clasts, clastic injections with ptygmatic folding and offshoots

Sediment loading

Cretaceous, sandstone and mudstone, Agat region, Norwegian North Sea

Deep-marine slope

Sediment loading

8 (Shanmugam, 2002a, 2003a)

Basal unit of contorted layers beneath the main sand with the “Bouma Sequence” (i.e., turbidites)

Sediment loading

Eocene 2 Oligocene, sandstone and shale, Annot Sandstone, French Maritime Alps, Southeastern France

Deep-marine environment

Sediment deformation is commonly associated with normal deep-water depositional processes (e.g., mass transport and turbidity currents)

9 (Shanmugam, 1997a, 2006a, 2012a)

Slump folds (Fig. 9.11), microfolds, clastic injections

Sediment loading

Pliocene, sand and mudstone, Edop Field, offshore Nigeria

Submarine canyon

Sediment loading

10 18 (Shanmugam et al., 1995a; Shanmugam, 2006a, 2012a)

Slump folds, contorted layers, steeply dipping layers, brecciated mud clasts, inclined dish structures, basal shear zone, secondary glide plane, abrupt change in fabric, injected sands (dikes and sills)

Sediment loading

Multiple case studies: Early Eocene, sand and mudstone, 10—Frigg Field, Norwegian North Sea 11—Harding Field (formerly Forth Field), Early Eocene, United Kingdom North Sea 12—Alba Field, Eocene, United Kingdom North Sea 13—Fyne Field, Eocene, United Kingdom North Sea 14—Gannet Field, Paleocene, United Kingdom North Sea 15—Andrew Field, Paleocene, United Kingdom North Sea 16—Gryphon Field, Late Paleocene-Early Eocene, United Kingdom North Sea 17—Faeroe area, Paleocene, west of the Shetland Islands

Deep-marine basin-floor fans

Sediment loading

(Continued)

TABLE 9.1 (Continued) Serial number (case study)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

Comments (Shanmugam, 2017a)

18—Foinaven Field, Paleocene, West of the Shetland Islands, United Kingdom Atlantic Margin 19 (Shanmugam and Zimbrick, 1996)

Slump folds, brecciated mud clasts, clastic injections

Sediment loading

Pliocene and Pleistocene, sand and mudstone, Gulf of Mexico, United States

Deep-marine slope

Sediment loading

20 (Shanmugam et al., 2009)

Slump folds, clastic injections, brecciated mud clasts, stretched clasts in slump folds

Sediment loading

Pliocene, sand and mudstone, KrishnaGodavari (KG) Basin, Bay of Bengal, India

Submarine canyon

Sediment loading

21 (Metz, 2010; Kilometer-scale convolute Metz et al., 2010) beds (Fig. 9.22). Small-scale folds, possible shrinkage cracks, and ripup clasts

Impact- or Sand and sandstone. seismically induced Mars liquefaction

Eolian and sublacustrine fan environments

Relative age of deformed strata on Mars is in dispute. Direct examination of sediments is still necessary to validate prelithification deformation and liquefaction

22 (VelascoVillareal et al., 2011)

Breccias

Meteorite impact

Chicxulub crater, K Pg boundary, Yucatan, Mexico

Marine environment

Meteorite impact

23 (Gregory, 1969)

Slumps and dikes

Earthquakes

Early Miocene, sandstone and siltstone, Waitemata Group, Whangaparaoa Peninsula, Auckland, North Island, New Zealand

Deep-marine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

24A (Odonne et al., 2011)

Folds

Submarine sliding

Cretaceous, limestone, Ayabacas Formation, Andes of Peru

Submarine environment

Submarine sliding

24B (Callot et al., 2008)

Megabreccia

Turonian Coniacian transition, Ayabacas Formation, Andes of Peru

Submarine environment

Most of the Ayabacas Formation was displaced during the giant submarine collapse of a regional carbonate platform

25A (Matsumoto Truncated flame structures Tsunami et al., 2008)

Indian Ocean Tsunami, 2004, sand, Bang Sak, Southwest of Thailand

Coastal environment

25B (Meshram et al., 2011)

Tsunami

2004 Indian Ocean Tsunami, sand, West coast of India

Coastal environment

25C (Gelfenbaum Rip-up clasts of muddy and Jaffe, 2003) soil

Tsunami

Align horizontally (move up) 17 July, 1998 Papua New Guinea Tsunami, sand

Coastal environment

Tsunami activity See a discussion of challenges in distinguishing tsunami deposits from storm deposits in the ancient record (Shanmugam, 2012b)

25D (Le Roux et al., 2008)

Dikes and sills

Ancient tsunami

Move up Pliocene, Ranquil Formation, Southern Chile

Coastal environment

25E (Sarkar et al., 2011)

Imbricated slump folds

Ancient tsunami

Move up Coastal environment Neoproterozoic, sandstone, Central India

26 (Chen and Lee, 2013)

Sand volcanoes and pillows

Storm wave

Cambrian, siliciclastic and carbonate sediments, Shandong Province, China

Shallow-marine mixed siliciclastic and carbonate environments

Cyclonic activity

27 (Greb and Archer, 2007)

Flow rolls

Tidal activity

Modern, sand and silt, Turnagain Arm, Alaska, United States

Hypertidal estuarine environment ( . 9 m tidal range)

Tidal activity

28 (Van Loon et al., 2016a; see also Mazumder et al., 2006)

Tomboliths (stromatolitic bioclasts of Shanmugam, 2017c)

Earthquakes

Paleoproterozoic, stromatolite mats on finegrained sandstone and mudstone, Chaibasa Formation, East India

Shelf environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c) See a lively debate on “Tomboliths”

Complex load structures and convolutions

Submarine collapse of a regional carbonate platform

(Continued)

TABLE 9.1 (Continued) Serial number (case study)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

Comments (Shanmugam, 2017a) (Shanmugam, 2017c; Van Loon et al., 2017)

29 (Nogueira et al., 2003)

Load casts

Rapid change from icehouse to greenhouse conditions

Neoproterozoic, Puga cap carbonate, southwestern Amazon craton, Brazil

Carbonate platform

Earth’s natural climatechange model

30 (El Taki and Pratt, 2012)

Convolute bedding, loop bedding, and dikes

Earthquakes

Late Ordovician, carbonates and evaporites, Red River strata, Williston Basin, Southern Saskatchewan, Canada

Low-energy marine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

31 (Korznikov, 2014; see also Prytkov et al., 2014)

Mud volcanoes

Complex origin

Modern, Sakhalin Islands, Russian Far East, Pacific Ocean

Coastal environment

Complex origin, but commonly associated with gas hydrates

32 (Kundu et al., 2015)

Folds of variable geometry, pseudonodules, water-escape features, flame structures, and chaotic laminae

Earthquakes

Mio-Pliocene, sandstones, mudrocks, heterolithic units, and conglomerates, Middle Siwalik Subgroup, Lish River Section, Darjeeling District of the Eastern Himalaya, India See also Middle Siwalik sediments of Arunachal Pradesh, Northeast Himalaya, India (Bhakuni et al., 2012)

Alluvial fan environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

33 (Rudersdorf et al., 2015)

Water-escape structures, folds, convolution, dish structures, load casts, small-scale intraformational normal faults, and reverse faults

Earthquakes

Pleistocene Holocene, sands and silts, Northeastern Ejina Basin, Inner Mongolia

Lake environment

Recognition of earthquakeinduced structures in modern environments is possible

34 (Mohindra and Thakur, 1998)

Oversteepening of Earthquakes sedimentary strata, folds, graben-type faults, plumelike intrusion, flame structure, slumping related to disrupted bedding, and pear-drop structure

Affected by 1720, 1803, and 1905 earthquakes, sand, Baldi Nadi in the Doon valley of the Garhwal Himalaya, India

Braided river

35 (Olabode, 2014)

Recumbent folds, clastic Earthquakes dikes, convolute lamination, subsidence lobes, pillars, cusps, and sand balls

Maastrichtian, sand, silt, Shallow-marine wave to tide and clay, Ajali Formation, influenced regressive sediments Western Flank of Anambra Basin, Southern Nigeria

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

36A (Fichman, 2013; Fichman et al., 2015)

Natural raindrop imprints

Impact by rain showers

Carboniferous, conglomerates, sandstones, siltstones, shales, and coal, Narragansett Basin, Massachusetts, United States

Alternative origins of raindrop imprints are possible (see Section 6.12)

36B (Zhao et al., 2015)

Experimental water-drop cratering that closely mimics natural raindrop imprints (Fig. 9.3)

Experimental impact cratering

Experimental studies Laboratory setting carried out at the University of Minnesota, Minneapolis, United States

Move up, align horizontally See discussion on similarities between experimental water-drop cratering and asteroidimpact cratering (Section 4.1)

37 (Li et al., 2008)

seismic microfractures, micro-corrugated laminations, liquefied veins, “vibrated liquefied layers,” deformed crosslaminations, convolute laminations, load structures, flame structures, breccias, slump structures, seismodisconformity

Earthquakes

Paleogene, sandstone and mudstone, East China Sea

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

Alluvial fan system with meandering streams and floodplains coupled with lowenergy interchannels and swamp areas

Shelf, tidal flat, delta, and river environments

Recognition of earthquakeinduced structures in modern environments is possible

(Continued)

TABLE 9.1 (Continued) Serial number (case study)

Comments (Shanmugam, 2017a)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

38 (Ackermann et al., 1995)

Dewatering structures, convolute bedding, kink bands, convergent fault fans

Subsurface dissolution of intrastratal karst

Late Triassic, sandstone and mudstone, Lower Blomidon Formation, Fundy rift basin, Nova Scotia, Canada

Fluvial, Eolian, playa lake

See case study 69 on surface dissolution of carbonate rocks in karst regions of Southern China and related collapse breccia

39 (Alfaro et al., 2002)

Load casts, ball-and-pillow Storm activity structures and pipes

Messinian, marls, sandstones, diatomites, and selenite gypsum, Bajo Segura Basin, Betic Cordillera, Southern Spain

Shelf environment

Cyclonic activity

40 (Agnon et al., 2006)

Intraclast breccias

Late Pleistocene, sand, breccia, marl, Dead Sea, Middle East

Continental environment

Recognition of earthquakeinduced structures in modern environments is possible

41 (Kang et al., 2010)

Load casts, ball-and-pillow Earthquakes structures, dish-and-pillar structures, clastic dikes, sills, convolute folds, slump structures, syndepositional faults, dislocated breccia

Cretaceous, sandstones, shales, mudstones, and conglomerates, Southeastern Gyeongsang Basin, Korea

Fluvial plain and lacustrine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

42 (Rossetti, 1999)

Convolute folds, ball-and- Earthquakes pillow structures, concaveup paths with consolidation lamination, recumbently folded crossstratification, irregular convolute stratification, pillars, dikes, cusps, subsidence lobes, fractures, and faults

Late Albian to Cenomanian, sandstones, shales, conglomerates, and limestone, Sa˜o Luı´s Basin, Northern Brazil

Middle to upper shoreface, foreshore, tidal channel, and lagoon/washover environments of a barred coast

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

Earthquakes

43 (Larsen, 1986) Imbricated flame structures and pseudonodules

Syndepositional secondary features from high-density turbidity currents

Silurian, fine-grained sandstone, North Greenland See Surlyk and Hurst (1984) for geologic history of North Greenland

Deep-marine environment

See a critical appraisal of high-density turbidity currents in Section 6.10

44 (Martı´nChivelet et al., 2011)

Boudins, pinch-and-swell layering, folds, normal (listric) dm-scale faults, subhorizontal detachment surfaces, slump structures

Earthquakes

Late Jurassic, limestones, Neuquen Basin, Argentina

Open-marine microbialites

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

45 (Su and Sun, 2012)

Diapirs, clastic dikes, convolute bedding, accordion folds, platespine breccias, moundand-sag structures, loopbedding

Earthquakes

Mesoproterozoic, sandy dolorudite, dolomite with siliceous concretions, and siliceous layer with dolomite breccia, Wumishan Formation, Yongding River Valley, Beijing, China

Shallow peritidal environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

46 (Kahle, 2002)

Convolute layers, homogenization of bedding, breccia clasts, stromatolitic clasts, boudinage structures, faults (normal and reverse), and breccias (discordant and concordant)

Earthquakes

Silurian, Lockport Dolomite, Northwestern Ohio, United States

Shallow-water, subtropical carbonate banks with reefs

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

47 (Bachmann and Aref, 2005)

Convolutions, folds, cusps, Earthquakes domes, offshooting dikes, breccias, faults

48 (Bhattacharya and Bandyopadhyay, 1998)

Deformed cross-bedding, convolute laminations, synsedimentary faults, graben-like structures,

Earthquakes

Triassic, Gypsum deposits, Shallow-marine environment Grabfeld Formation, Ladinian, southwestern Germany

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016a,c)

Proterozoic, sandstone, Singhbhum, Bihar, India

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

Tidal environment

(Continued)

TABLE 9.1 (Continued) Serial number (case study)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

Comments (Shanmugam, 2017a)

sandstone dikes, pseudonodules, and slump folds 49 (Wizevich et al., 2016)

Large-scale load casts, wedge-shaped load structures, deformation bands

Earthquakes

Late Cretaceous, sandstone, Wahweap Formation, Kaiparowits basin, southern Utah, United States

Fluvial environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

50 (Calvo et al., 1998)

Loop-bedding

Earthquakes

Late Miocene, laminites and marlstones, Hı´jar Basin, Southeast Spain

Lacustrine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

51 (Seth et al., 1990)

Breccias, and sand volcanoes

Earthquakes

Late Jurassic, sandstone and shale, Lower Member of the Katrol Formation, Kutch, India

Submarine fan complex

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

52 (Fortuin and Dabrio, 2008)

Funnel-shaped depressions filled with conglomerate, liquefaction dikes, chaotic intervals, flame structures, syndepositional faults, fault-grading, sandstone dikes, slumping and sliding of sandstone beds, convolute bedding, pillars, and flame structures

Earthquakes

Messinian, evaporites, Feos Formation, Nijar Basin, Southeast Spain

Alluvial, hypersaline and nonmarine coastal plain environments

Messinian salinity crisis Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

53A and 53B (Obermeier et al., 2005)

Dikes, sills, bowl-shaped intrusion, brecciated mud clasts

Earthquakes

Modern Examples from 1811 12 New Madrid (Missouri) (53A) earthquakes, from the Wabash Valley of Indiana Illinois, and from

Continental and estuarine environments

Recognition of earthquakeinduced structures in modern environments is possible

islands in the lower Columbia River of Oregon Washington (53B) 54 (McLean et al., 2016)

Meter-scale loading structures (simple and pendulous load casts, detached pseudonodules), volcaniclastic breccias

Volcanic activity

Miocene, volcanic rocks, Owyhee Mountains, Southwest Idaho, United States

Rhyolite lava

Not typical SSDS

55 (Rodrı´guezPascua et al., 2000)

Pseudonodules, mushroom-like silts protruding into laminites, mixed layers, disturbed varved lamination, and loop bedding

Earthquakes

Late Miocene, sand, mud, and gravel, Prebetic Zone, Southeast Spain

Lacustrine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

56 (Enzel et al., 2000)

Faults, liquefaction features, slumps

Earthquakes

Holocene, sandy and silty sediments, Dead Sea Graben, Nahal Darga, Israel

Fan-delta environment

Recognition of earthquakeinduced structures in modern environments is possible

57 (Topal and ¨ zkul, 2014) O

Load structures, flame structures, clastic dikes (sand and gravely sand dikes), disturbed layers, laminated convolute beds, slumps, and synsedimentary faults

Earthquakes

Late Pliocene, sand, gravelly sand, siltstone, and marl, Kolankaya Formation, Denizli Basin, Southwest Turkey

Fluvio-lacustrine environments

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

58A 58D (Douillet et al., 2015)

Pipes, overturned or oversteepened laminae, vortex-shaped laminae, folds, faults, potatoids, dishes, diapiric flame-like structures

Pyroclastic density currents. Triggers from granular shear, dynamic pore pressure, ballistic impacts, and shock waves

Soufrie`re Hills (Montserrat) (58A), Tungurahua (Ecuador) (58B), Ubehebe craters (United States), LaacherSee (Germany), and Tower Hill (58C), Purrumbete lakes (58D) (both Australia)

Syn-eruptive volcanic environment

The authors also discuss the importance of traction carpets associated with pyroclastic flows See Shanmugam (1996a,b, 2006a, 2012a, 2016b) on controversies associated with traction carpets and high-density turbidity currents (Continued)

TABLE 9.1 (Continued) Serial number (case study)

Comments (Shanmugam, 2017a)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

59 (Moretti et al., 2011)

Large-scale collapse features comprising irregularly elongated conical zones of sinking Small-scale load structures

Karstic sinkhole collapse

Late Pliocene Early Pleistocene, Calcarenite di Gravina Formation, Cala Corvino area, Northern Monopoli, Adriatic sector of the Apulian Foreland, Southern Italy

Shallow-marine environment

See case study 69 on surface dissolution of carbonate rocks in karst regions of Southern China and related collapse breccia

60 (Ozcelik, 2016)

Load structures, flame structures, slumps, Neptunian dikes, drops, pseudonodules, ball-andpillow structures, minor faults, water-escape structures

Earthquakes

Eocene, sands, silts, and clay layers, Kayıko¨y Formation, Fethiye Burdur Fault Zone, Southwest Turkey

Lacustrine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

61 (Ettensohn et al., 2011)

Accordion folds, sag-andmound structures, dikes, diapirs

Earthquakes

Early Mesoproterozoic Wumishan Formation, Jumahe region, about 90km southwest of Beijing, China

Peritidal environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

62 (Oliveira et al., 2011)

Flame structures, pseudonodules, and balland-pillow structures

Earthquakes

Permian, sandstone and siltstone, Ecca Group, Karoo Basin, South Africa

Marine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

63 (Berra and Felletti, 2011)

Ball-and-pillow structures, Earthquakes plastic intrusion, disturbed laminations, convolute stratifications, slumps, sand dikes, autoclastic breccias, and chicken-wire structures

Alluvial fan, continental clastics Early Permian, conglomerates, sandstones, with intercalations of ignimbritic flows and tuffs and siltstones, Southern Alps, Northern Italy

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

64 (Roy and Banerjee, 2016)

Slumps, folds, convolute bedding or lamination, load and flame structures,

Late Eocene to Oligocene, sandstone and shale, Andaman Flysch Group,

Interpretation of paleoseismicity using

Earthquakes

Deep-marine environment

Andaman Island, Bay of Bengal, India

pseudonodules, ball-andpillow structures, dishand-pillar structures, slumps, folds, convolute bedding or lamination encased by undeformed beds, sand volcanoes, load casts, pseudonodules

SSDS is unreliable (Shanmugam, 2016c)

65 (Massari et al., 2001)

Water-upwelling pipes, large dishes

Upwelling of overpressured groundwater

Early Pleistocene calcarenites, Salento, Southern Italy

Subtidal environment

Upwelling of overpressured groundwater

66 (Chen et al., 2010)

Intrastratal cracks and boudins

Storm or earthquake

Cambro-Ordovician strata, Early diagenetic deformation, Furongian ribbon rocks in Shandong Province of China

Low-energy subtidal environment (below fairweather wave base), early diagenetic structures

Complex origin

´ balos and 67 (A Elorza, 2011)

Cone-in-cone structures

Earthquakes

Early Maastrichtian, carbonate rocks, BasqueCantabrian Basin, North Spain

Deep-marine environment

Cone-in-cone structures could also represent early diagenetic features. Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

68 (Nanfito, 2008)

Chicken-wire structures

Early diagenetic deformation (?)

Carboniferous, gypsum and anhydrites, Spitsbergen

Coastal sabkha environment

Controversy exists over the origin of chicken-wire texture (i.e., primary syndepositional versus secondary diagenetic origin) (see Section 6.15)

Modern karst dissolution and collapse

Devonian Late Carboniferous carbonates, Guilin, Southern China

Li River, Southern China

Analogous to carbonate karst dissolution, chert dissolution has been attributed to porosity development beneath the Neocomian unconformity in the petroleum reservoir

69 (Shanmugam, Collapse breccia 1990b, 2017a)

(Continued)

TABLE 9.1 (Continued) Serial number (case study)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

Comments (Shanmugam, 2017a) of the Prudhoe Bay Field, Alaska, United States (Shanmugam and Higgins, 1988)

Marine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

70 (Gavrilov, 2017)

Neptunian dikes and sills

Earthquakes

Mesozoic Cenozoic, sand and clay, Northern Caucasus, European Russia

71 (Alfaro et al., 1997)

Load casts, pillows, and water-escape structures

Earthquakes

Pliocene, sands, silts, clays, Lacustrine environment and rare gravels, Guadix Baza Basin, Central Betic Cordillera, Southern Spain

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

72 (GindreChanu et al., 2014)

Enterolithic structures

Secondary diagenetic changes

Aptian evaporites, Namibe Deep-burial environment Basin, South West Angola

See Section 6.15 for controversies on the origin

73 (Jones and Omoto, 2000)

Sheet of convolutions showing soft-sediment overturned asymmetrical open/isoclinal/tight folding

Earthquakes

Late Pleistocene, sand and clay, Onikobe and Nakayamadaira Basins, Northeastern Japan

Lacustrine environment

Recognition of earthquakeinduced structures in modern environments is possible

74 (Passchier, 2000)

Slumps, dikes, anastomosing dilated laminae, crackle breccia, mosaic breccia, and contorted bedding

Hydrostatic fracturing, subglacial shearing, slumping, and gashydrate formation

Late Oligocene Early Miocene, diamictite, sandstone, mudstone, core from CRP-2/2A, Victoria Land Basin, Antarctica

Polar interglacial environment

Complex origin

75 (Lunina and Gladkov, 2016)

Clastic injections

Earthquakes

Modern, Lake Baikal region, Southern Siberia, Russian Siberia

Alluvial, alluvial swampy lacustrine, lacustrine swampy, lacustrine, and eolian environments

Recognition of earthquakeinduced structures in modern environments is possible

76 (Weathley et al., 2016)

Clastic pipes

Earthquakes

Jurassic, sand, Colorado Plateau, United States

Eolian dune, interdune, sabkha, Interpretation of and fluvial environments paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

77 (Rodrı´guezPascua et al., 2016)

Sand dikes and explosive sand-gravel craters (Fig. 9.21B)

Earthquakes

Earthquake-induced liquefaction, 4th Century CE Roman Complutum, Madrid, Spain

Continental environment

Recognition of earthquakeinduced structures in modern environments is possible

78 (Gruszka et al., 2016)

Load casts

Glacial activity

Modern, oscillations of the ice front over an esker near Ryssjo¨n, Southern Sweden

Glacial environment

Glacial activity

79 (Menzies et al., 2016)

Crenulation foliations, sills, dikes, folds, discrete shear lines, water-escape structures, necking structures, boudins, kink bands, rotational structures, strain shadows and caps

Glacial activity

Modern, subglacial microshear deformation

Glacial environment

Some of the SSDS are theoretical, and yet to be documented in the sedimentary record

80 (Alsop et al., 2016)

Slumps

Earthquakes

Modern, sand and gravel, Dead Sea Basin, Middle East

Continental environment

Recognition of earthquakeinduced structures in modern environments is possible

81 (Ezquerro et al., 2016)

Clastic dikes, load Earthquakes structures, diapirs, slumps, nodulizations, mudcracks

Pliocene, silty carbonates, marls, limestones, coal beds, mudstones, sandstones, and conglomerates, Concud Fault, Eastern Spain

Alluvial palustrine lacustrine environments

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

82 (Gladkov et al., 2016)

Load casts, flame structures, convolute laminations, ball-andpillow structures, folds, and slumps

Late Pleistocene, sand, mud, gravel, Issyk-Kul lake, Kyrgyzstan, Central Asia

Lacustrine environment

Recognition of earthquakeinduced structures in modern environments is possible

Earthquakes

(Continued)

TABLE 9.1 (Continued) Serial number (case study)

Comments (Shanmugam, 2017a)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

83 (Jiang et al., 2016)

Clastic dikes, ball-andpillow structures, flame structures, microfaults, and slump folds

Earthquakes

Late Pleistocene, metasandstone, Lixian, Eastern Tibetan Plateau

Lacustrine environment

Recognition of earthquakeinduced structures in modern environments is possible

84 (Onorato et al., 2016)

Sand dikes, convolute laminations, load structures, faults

Earthquakes

Holocene, Magallanes Fagnano Fault System, Isla Grande de Tierra del Fuego, Argentina

Glaciofluvial and glaciolacustrine environments

Recognition of earthquakeinduced structures in modern environments is possible

85 (Snyder and Waldron, 2016)

Load structures, neptunian Earthquakes and dikes, intraformational overpressured breccia conditions

Mississippian, conglomerate, sandstone, and mudstone, Maritime Basin, Canada

Fluvial lacustrine environments with possible local marine influence

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

86 (Sherrod et al., 2016)

Pockmarks (deformation craters)

Water gas escape

Modern, Glen Canyon National Recreation Area, Lake Powell Delta, Utah, United States

Lacustrine environment

Pockmarks are recognized on radar images. Conventionally, SSDS are recognized by the direct examination of the rocks. See Section 4.3

87 (To¨ro˝ and Pratt, 2016)

Folds, load structures, microfaults, breccias, and dikes

Earthquakes

Eocene, mudstone, siltstone, and sandstone, Green River Formation, Bridger Basin, Wyoming, United States

Lacustrine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

88 (Bryant et al., 2016)

Contorted cross-bedding

Paleohydrology, intra-dune deformation

Jurassic, Navajo Sandstone, United States

Eolian environment

Paleohydrologic control See Horowitz (1982) for conventional origin of deformation in the Navajo Sandstone by earthquakes

89 (Ford et al., 2016)

Dune collapse features, sand blows, clastic injections

Collapse

Jurassic, Navajo Sandstone, Zion National Park, Utah, United States

Eolian environment

See Horowitz (1982) for conventional origin of deformation in the Navajo Sandstone by earthquakes

90 (Liesa et al., 2016)

Fluid-escape structures

Earthquakes

Pliocene dune field of the Teruel half-graben basin, Eastern Spain

Eolian environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

91 (Hilbert-Wolf et al., 2016)

Balloon-shaped inflation structure (Fig. 9.21A)

Gas expulsion

Cretaceous, sandstones of the Rukwa Rift Basin, Southwestern Tanzania

Braided fluvial and lacustrine environments

Additional examples are needed to understand the importance of this new structure

92 (Rana et al., 2016)

Pendulous load structures, ball structures, flame and diapir structures, slump folds, water-escape structures, ground fissures, sedimentary faults

Sudden loading by flash floods

Quaternary, Alaknanda Fluvial environment Valley, Garhwal Himalaya, India

Fluvial activity

93 (Chiarella et al., 2016)

Deformed crosslaminations, folds, fluidescape structures

Increase in porewater pressure induced by depositional overloading

Early Pleistocene, tidally dominated Catanzaro strait-fill succession, Calabrian Arc, Southern Italy

Tidal environment

Sediment loading

94 (Kopf et al., 2016)

Micro-slumping and folding

Mass transport

Pliocene Holocene, submarine slope, France

Submarine slope

Mass transport

95 (Basilone et al., 2016)

Folds, faults, disturbed layers, clastic dikes, and slumps

Earthquakes

Late Triassic, cherty limestone, Southern Tethyan rifted continental margin, Central Sicily

Deep-water carbonate environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

96 (Mazumder et al., 2016)

Load casts, flame Earthquakes structures, dish-and-pillar structures, synsedimentary faults, convolutions, chaotic structures, sedimentary veins, dikes

Mio-Pliocene, Misaki Formation within alternating deep-sea clays and volcanic ashes, Miura Peninsula, Japan

Deep-marine environment

Interpretation of paleoseismicity using SSDS is unreliable (Shanmugam, 2016c)

97 (Ito et al., 2016)

Injections

Late Pliocene, trench-slope Deep-marine environment basin on the southern Boso Peninsula, Japan

Turbulent and laminar flows of high-density turbidity currents

High-density turbidity currents have never been documented in modern environments (Shanmugam, 2016b). See Section 6.10 (Continued)

TABLE 9.1 (Continued) Serial number (case study)

Comments (Shanmugam, 2017a)

Types of SSDS

Origins of SSDS

Stratigraphy and location

Environment

98 (Ortner and Kilian, 2016)

Slump structures

Sediment creep

Late Jurassic, pelagic limestones, Northern Calcareous Alps, Austria

Pelagic carbonate environment

Creep is a process with velocity connotations, and therefore is controversial when applied to ancient examples (Section 6.11; see also Shanmugam, 2015a)

99 (Sobiesiak et al., 2016)

Siltstone rafts and sandstone blocks in masstransport deposits (MTDs)

Mass transport

Carboniferous, sandstone blocks, Northwest Argentina

Open-marine environment

Mass transport

100A and 100B (Tinterri et al., 2016)

Convolute laminations and Flow reflections of load structures in high-density turbidites turbidity currents

Miocene, sandstone and shale, Marnoso Arenacea Formation, Northern Italy (100A) Eocene Oligocene, Annot Sandstones, Southeastern France (100B)

Deep-marine environment

High-density turbidity currents have never been documented in modern environments (Section 3.7.6; see also Shanmugam, 2016b)

Note 1: There are no reliable criteria exist for distinguishing SSDS formed by earthquakes from SSDS formed by other 20 triggering mechanisms in the ancient sedimentary record (Shanmugam, 2016c). Note 2: Selected serial numbers comprise more than one case study: 24, 25, 36, 53, 58, and 100. The grand total of all case studies used in this review is 140 (Table 9.1 with 110 and Table 9.2 with 30). Additional 30 case studies from scientific drilling at sea are listed in Table 9.2 (see Fig. 9.1, shown as filled purple squares). Note that at least 120 different types of SSDS have been recognized (excluding duplicate entries). Note that all 26 case studies from a special issue of the Sedimentary Geology on SSDS (Alfaro et al., 2016) are included here (case studies: 75 100). Also, numerous studies of SSDS have been published in Chinese language (see Feng et al., 2016; Feng, 2017a, 2017b). Note 2: Selected other studies on SSDS are by Alsop et al. (2019), Bond and Lebit (2020), Cossey (2011), Dasgupta (2008), Campbell et al. (2006), and Dasgupta and Chatterjee (2019). From Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320.

9.3 Definition

395

that control the origin of SSDS are prelithification deformation and liquidization (Allen, 1984; Shanmugam, 2016c, 2017a). Palladino et al. (2019) cautioned that distinguishing structures formed by soft-sediment deformation during mass transport from those produced by contractional tectonics can be problematic. For example, deformation occurring along detachment levels may completely obliterate the original sedimentary fabric. Furthermore, the importance of SSDS as depositional features has been severely compromised by the introduction of the “seismite” concept, which advocated the origin of SSDS by seismic shaking (Seilacher, 1969). Shanmugam (2017a) addressed basic problems associated with the recognition of SSDS formed by seismicity. The link between SSDS and earthquakes is still an unresolved issue. In this regard, scientific drilling at sea offers a great wealth of vital information in terms of depositional setting, tectonic framework, distribution of SSDS, and their possible triggering mechanisms. In terms of data, ODP (2007) reported that during its operation from 1983 to 2007, the drilling program recovered 222,430 m of core from 1797 holes. Disappointingly, researchers of SSDS have neglected to utilize this robust dataset so far. Therefore, the purpose of this chapter is to combine 110 case studies worldwide with 30 case studies from scientific drilling (Deep Sea Drilling Project/Ocean Drilling Program/Integrated Ocean Drilling Program sites) and to discuss the origins of SSDS in various settings in understanding the origin of SSDS with important geological implications.

9.2 Datasets Table 9.1 presents 110 case studies of outcrop and core (Fig. 9.1). Table 9.2 presents 30 case studies of deep-sea drilling sites (Fig. 9.1).

9.3 Definition Allen (1984, II, p. 343) provided an accurate account of soft-sediment deformation in terms of physics. The two factors that control the origin of SSDS are prelithification deformation and liquidization. Allen (1984) recognized the following basic structures as SSDS: • • • • • • • • •

convolute lamination load casts heavy mineral sags passively deformed beds dish structures folds and sand mounds sheet slumps imbricate structure deformed cross-bedding

Mass transport, gravity flows, and bottom currents

396

9. Soft-sediment deformation structures

FIGURE 9.1 Map showing locations of case studies of soft-sediment deformation structures (SSDS) (filled black squares) listed in Table 9.1. Scientific drilling sites (see Table 9.2) are shown by filled purple squares. Countries such as Spain and Italy have a high number of published case studies (see Table 9.1), but not all of them are shown here due to limited space on the map. Note that this review also includes all my previous studies of deep-water systems (filled yellow and red circles) that contain a variety of SSDS (see Table 3 in Shanmugam, 2016c). G. Shanmugam’s study localities are: (1) Gulf of Mexico, (2) California, (3) Ouachita Mountains, (4) Southern Appalachians, (5) Brazil, (6) U.K. North Sea, (7) U.K. Atlantic Margin, (8) Norwegian Sea and vicinity, (9) French Maritime Alps, (10) Nigeria, (11) Equatorial Guinea, (12) Gabon, and (13) Bay of Bengal. These modern and ancient deep-water systems include both marine and lacustrine settings (Shanmugam, 2012a, 2015a). Distribution of modern pockmarks in Lake Constance in Germany, shown by a filled blue square, is discussed by Wessels et al. (2010).

However, a recent inventory suggests a plethora of at least 120 different types of SSDS (Table 9.1) have been recognized in strata ranging in age from Paleoproterozoic to the present time (Shanmugam, 2017a).

9.4 Origin Although seismicity is commonly considered to be the origin of SSDS, there are at least 21 mechanisms that can generate SSDS (see Chapter 7: Triggering mechanisms of downslope processes). This is because 21 triggering mechanisms of sediment failures are also the cause of SSDS. In particular, SSDS are commonly associated with deep-water deposits.

Mass transport, gravity flows, and bottom currents

397

9.4 Origin

TABLE 9.2 Thirty case studies of SSDS from scientific drilling (SD) sites at sea (see Fig. 9.1, shown as filled purple squares), which comprise cores recovered from the Deep Sea Drilling Project (DSDP), Ocean Drilling Program (ODP; see Fig. 9.26 below for all ODP sites in the world’s oceans), Integrated Ocean Drilling Program (IODP), and International Ocean Discovery Program (IODP). SD case study number (Fig. 9.1)

Types of SSDS

Origins of SSDS, stratigraphy, and location

Drilling Leg, expedition, site, environment, and comment

SD 1

Slumps, breccia, steeply dipping laminae at 40 degrees

Submarine volcanic activity Cretaceous, hyaloclastic sediments, Nauru Basin, Western Equatorial Pacific (Moberly and Jenkyns, 1981)

DSDP Leg 61 Site 462 Deep-marine environment

SD 2

Contorted laminae

Submarine volcanic activity Cretaceous (?), tuffaceous sandstone, East of Tonga Trench, Western Pacific (Shipboard Scientific Party and Speeden, 1973b)

DSDP Leg 21 Site 204 Deep-marine environment

SD 3

Vein structures

Synsedimentary nontectonic origin DSDP Leg 67 Early Miocene Quaternary, Sites 496 and 497 hemipelagic muds, inner slope of Deep-marine environment the Middle America Trench, Off Guatemala (Cowan et al., 1982)

SD 4

Slumps, variable bedding dips, faults, shear bands

Subduction-related seismic activity Early Miocene Quaternary, sands, muds, and volcanic ash, Shikoku Basin, Nankai Trough, Sea of Japan (Taira et al., 1992)

ODP Leg 131 Site 808 Deep-marine environment

SD 5

Brecciated mud clasts (Fig. 9.4)

CE 1944 Tonankai (M 5 8.2) Earthquake Muds, Nankai Trough, SW Japan (Sakaguchi et al., 2011; Strasser et al., 2011)

IODP Expedition 316 Sites C0004 and C0008 Deep-marine environment

SD 6

Folded or truncated laminations

IODP Expedition 323 Mass transport in submarine Site U1345 canyon Holocene Late Pleistocene, muds, Shelf-edge environment Interfluve ridge near the large, broad head of the Navarin submarine channel off the Bering Sea shelf (Expedition 323 Scientists, 2011)

SD 7

Slumps

Synsedimentary origin Miocene, silty clay, Western Norwegian Greenland Sea (Thiede and Myhre, 1996)

ODP Leg 151 Site 909 Deep-marine environment (Continued)

Mass transport, gravity flows, and bottom currents

398

9. Soft-sediment deformation structures

TABLE 9.2 (Continued) SD case study number (Fig. 9.1)

Origins of SSDS, stratigraphy, and location

Types of SSDS

Drilling Leg, expedition, site, environment, and comment

SD 8

Contorted bedding, flowage folding, faulting, mud clasts, pockmarks

Mass transport, salt diapiric ODP Leg 164 activity, and gas hydrates Sites 991, 992, and 996 Pliocene Pleistocene, clay, Blake Deep-marine environment Ridge and vicinity, Atlantic Ocean (Paull et al., 2000)

SD 9

Volcanic breccia (Fig. 9.16), slumps

Submarine volcanic activity Middle Eocene, hyaloclastic sediments with basaltic fragments, pelagic calcareous biogenic oozes, chalk, Rio Grande Rise, South Atlantic Ocean (Fodor and Thiede, 1977; Perch-Nielsen et al., 1977; Thiede 1977)

DSDP Leg 39 Site 357 Deep-marine environment See also DSDP Leg 72 Site 516 (Shipboard Scientific Party, 1983)

SD 10

Folds, mud clasts

Submarine slumping DSDP Leg 40 Pliocene Middle Eocene, biogenic Sites 360 and 361 ooze and chalk, Cape Basin, Deep-marine environment Continental Rise, South Arica (Shipboard Scientific Party, 1978)

SD 11

Microfractures

Submarine volcanic activity Late Cretaceous, calcareous vitric tuff, Broken Ridge, Indian Ocean (Shipboard Scientific Party, 1989)

ODP Leg 121 Site 755 Deep-marine environment

SD 12

Deformed sands

Synsedimentary origin Late Eocene, matrix-supported pebbly coarse sand, braid delta, Prydz Bay, Antarctica (Shipboard Scientific Party, 2001)

ODP Leg 188 Site 1166 Shelf See McPherson et al. (1987) for introduction of a new depositional facies concept on “braid deltas”

SD 13

Slump fold, normal fault (Fig. 9.9)

Mass transport ODP Leg 145 Eocene Oligocene, chalk, North Site 884 Pacific (Shipboard Scientific Party, Deep-marine environment 1993, their Fig. 14, 16, and 17)

SD 14

Enterolithic vein, chicken-wire texture, and nodular anhydrite (Fig. 9.23)

Primary deposition Late Miocene, Messinian evaporates, the Mediterranean Sea (Garrison et al., 1978)

DSDP Leg 42A Site 371 Present-day drilling environment: deep-water Late Miocene depositional environment: Sabkha Note controversies on the origin of chicken-wire texture (see Section 6.15) (Continued)

Mass transport, gravity flows, and bottom currents

9.4 Origin

399

TABLE 9.2 (Continued) SD case study number (Fig. 9.1)

Types of SSDS

Origins of SSDS, stratigraphy, and location

Drilling Leg, expedition, site, environment, and comment DSDP Leg 23 Site 223 Deep-marine environment

SD 15

Core 10: microfolds and microfaults Core 24: chalk breccia

Slumping Late Miocene, chalk, Continental rise off the coast of Muscat and Oman (Shipboard Scientific Party et al., 1974)

SD 16

Steeply dipping (25 degrees) layers, volcanic breccia

Submarine volcanic activity DSDP Leg 59 Late Miocene, tuff, volcaniclastic Site 451 sediment, West Mariana Ridge, Deep-marine environment near the Mariana Trough, Western Pacific (Shipboard Scientific Party, 1981, their Fig. 15A D)

SD 17

Slumps

Mass movement Modern, foraminiferal sands, Horizon Guyot, Mid-Pacific Mountains (Heezen et al., 1971; Lonsdale et al., 1972)

SD 18

Slump folds

ODP Leg 134 Submarine volcanic activity Pleistocene, volcanic ash and silt, Site 832 Deep-marine environment central part of the North Aoba Basin, approximately 50 km northeast of the Queiros Peninsula of Espiritu Santo Island and 45 km due south of the smoking volcanic island of Santa Maria (Gaua) (Shipboard Scientific Party, 1992)

SD 19

Microfaulting, deformed ash layers

Submarine volcanic activity ODP Leg 115 Middle Eocene, chalk and volcanic Sites 712 and 713 ash, Chagos Ridge, Indian Ocean Deep-marine environment (Shipboard Scientific Party, 1988)

SD 20

Fault, deformed layers, steep layers dipping at 50 degrees (Fig. 9.5), and mud clasts

Mass-transport deposits (MTDs), rapid sedimentation, and salt tectonic activity Holocene Pleistocene, clay, Ursa Basin, Gulf of Mexico (Expedition 308 Scientists, 2006)

SD 21

Microfault, convolute layers (Fig. 9.6)

Slumping, Benguela Current ODP Leg 175 Unit II, Middle Miocene, clay-rich Site 1085 nannofossil ooze, Mid-Cape Basin, Deep-marine environment off South Africa, South Atlantic (Shipboard Scientific Party, 1998, their Fig. 9)

DSDP Site 44 Deep-marine environment The terrace top of the guyot is affected by internal tides (Lonsdale et al., 1972; Shanmugam, 2013a)

IODP Expedition 308 Site U1322 Deep-marine environment

(Continued)

Mass transport, gravity flows, and bottom currents

400

9. Soft-sediment deformation structures

TABLE 9.2 (Continued) SD case study number (Fig. 9.1)

Types of SSDS

SD 22

Origins of SSDS, stratigraphy, and location

Drilling Leg, expedition, site, environment, and comment

Bowed laminae (Fig. 9.7)

Slumping Late Pleistocene, mud, Eastern Aleutian Trench, Gulf of Alaska (Piper, 1975)

DSDP Leg 18 Site 180 Deep-marine environment

SD 23

Mud clasts (Fig. 9.8), breccia, contorted layers, normal fault. Common clast types in the diamict of Facies 9 are rhyolite, basalt (greenstone), siltstone, sandstone, argillite, quartz, gneiss, and granite (Jaeger et al., 2014b)

Mass transport Early Pleistocene, mud and diamict, Aleutian Trench area, Gulf of Alaska (Jaeger et al., 2014a)

IODP Expedition 341 Site U1418, Unit IV Deep-marine environment

SD 24

Volcanic breccia (Fig. 9.17)

Volcanic activity and mass transport Early Paleocene, claystone, Western margin of India, Arabian Sea (Pandey et al., 2016a,b)

IODPa Expedition 355 Site U1457, Unit V Deep-marine environment

SD 25A

Slump fold (Fig. 9.10), convolute bedding, tilted layers, shear zones

Mass transport Quaternary, mud, volcanic ash, sand, NanTroSEIZE complex, accretionary prism and Philippine Sea Plate (Expedition 333 Scientists, 2011, 2012b)

IODP Expedition 333 Site C0018 Deep-marine environment

SD 25B

Sand injection (dike) (Fig. 9.15)

Tectonic activity Quaternary, mud, Kashinosaki Knoll, subducting Philippine Sea Plate (Expedition 333 Scientists, 2011, 2012a)

IODP Expedition 333 Site C0012 Deep-marine environment

SD 26

Steeply dipping layers (Fig. 9.12)

Volcanic and volcaniclastic processes Oligocene, tuffaceous mudstone, claystone, sandstone, Amami Sankaku Basin, Philippine Sea region (Expedition 351 Scientists, 2015)

IODPa Expedition 351 Site U1438, Unit II Deep-marine environment

SD 27A

Volcanic breccia, load cast, steeply dipping layers, reverse grading (Fig. 9.13)

Volcanic activity and mass transport Late Miocene, sandstone, claystone, breccia, East Subbasin of the South China Sea (Expedition 349 Scientists, 2015a)

IODPa Expedition 349 Site U1431, Units VI and VII Deep-marine environment

(Continued)

Mass transport, gravity flows, and bottom currents

401

9.5 Classification

TABLE 9.2 (Continued) SD case study number (Fig. 9.1)

Types of SSDS

SD 27B

Origins of SSDS, stratigraphy, and location

Drilling Leg, expedition, site, environment, and comment

Faults and fractures

Volcanic activity and mass transport Late Miocene, sandstone, claystone, breccia, East Subbasin of the South China Sea (Expedition 349 Scientists, 2015b)

IODPa Expedition 349 Site U1434 Deep-marine environment

SD 28

Load cast, breccia, dike, convolute bedding

Subduction-related earthquakes IODP Expedition 344 Late Pliocene Late Pleistocene, Site U1380 sandstone, claystone, Deep-marine environment conglomerate. Middle slope of the Costa Rica margin (Expedition 344 Scientists, 2013)

SD 29

Load cast (Bouma, 1975, his Fig. 9), slump, microfault

Volcanic activity and earthquakes Late Oligocene Pleistocene, volcanic tuff and conglomerate, the west flank of the PalauKyushu Ridge adjacent to the Nankai Trough (Shipboard Scientific Party, 1975)

SD 30

Brecciated chalk clasts, folds, flow structures, faults (Fig. 2.4, this book)

Mass transport ODP Leg 150 Middle Miocene Pleistocene, silty Site 905 clay, white-colored chalk clasts, Deep-marine environment conglomerate. Upper Continental Rise, offshore New Jersey, U.S. Continental Margin (Shipboard Scientific Party, 1994)

DSDP Leg 31 Site 296 Deep-marine environment

a

SD 24, 26, 27A, and 27B are the cases of International Ocean Discovery Program (IODP), whereas all other IODPs used in this study are Integrated Ocean Drilling Program (IODP). From Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320.

9.5 Classification Although different authors have classified deformations based on various criteria in tectonic and depositional environments (e.g., Allen, 1984; Boggs, 2001; Ramsay, 1967; Waldron and Gagnon, 2011), the one by Collinson (1994) is considered the most comprehensive among 12 classifications considered in a review by Shanmugam (2017a). Collinson (1994) classified deformation structures based on both physical and chemical processes: 1. partial loss of strength and density inversion: load casts, flame structures, pseudonodules, and convolute bedding;

Mass transport, gravity flows, and bottom currents

402

9. Soft-sediment deformation structures

2. structures due to progressive loading of cohesive sediment: mud lumps or mud diapirs; 3. partial loss of strength and applied shear: Slump folds and overturned cross-bedding; 4. structures related to upward-escape of pore water and sediment-water mixtures: dishand-pillar structures, sheet dewatering structures, sediment-injection structures (dikes and sills), sand volcanoes, and extruded sand sheets; 5. synsedimentary faults: normal and reverse faults; 6. structures due to sediment shrinkage: desiccation cracks, syneresis cracks, and septarian nodules; 7. structures due to sediment wetting: lenses of deformed sand within the steeply dipping foresets of eolian dunes (i.e., the lee side); 8. deformation related to compaction: changes in channel shape, observed mainly on seismic reflection profiles; 9. deformation related to early chemical precipitation: concretions, cone-in-cone structures, and chicken-wire structures. In emphasizing the influence of tectonic environments on sediment deformations, Waldron and Gagnon (2011, their Fig. 3) recognized six different types of deformations using sand and mud lithofacies (Fig. 9.2). 1. Type 1 shows typical geometry of folds in clastic sedimentary strata deformed at low metamorphic grade. It exhibits dip isogons (lines joining points of equal dip on successive surfaces). Sandstone layers typically show tighter curvature on inner arcs (class 1 geometry; Ramsay, 1967) while mud layers tend to show tighter curvature on outer arcs (class 3 geometry; Ramsay, 1967). 2. Type 2 exhibits reversal of normal geometrical relationships, characteristic of folds formed while sand was liquefied. 3. Type 3 reveals undeformed configuration with (1) angular mud clasts surrounded by sand, (2) sand-filled dikes cross-cutting mud layers, and (3) folded liquefied sand layers. 4. Type 4 is basically Type 3 with superimposed simple shear parallel to bedding. 5. Type 5 is basically Type 3 with superimposed pure shear parallel to bedding. 6. Type 6 is basically Type 3 with superimposed arbitrary strain. This classification is useful to both structural geologists and sedimentologists.

9.6 Advances SSDS research is an actively developing field. Consequently, there are both advances and problems. The following advancements are of significance. The mechanics of impact cratering as a geologic process has long been a topic of interest (Holsapple, 1993; Melosh, 1989; Schoemaker, 1963). Although raindrop imprints have been interpreted in the geologic record (Fichman et al., 2015), the physics behind the impact of liquid drops on granular surface has not been fully understood. In this context, an experimental study conducted at the University of Minnesota by Zhao et al. (2015) is of relevance. When a granular material is impacted by a sphere, its surface deforms like a liquid yet it preserves a circular crater like a solid. Although the mechanism of granular impact cratering by solid spheres is well explored, our knowledge on granular impact cratering Mass transport, gravity flows, and bottom currents

9.6 Advances

403 FIGURE 9.2 Idealized diagrams of structures in six types of deformed sedimentary rocks. Sand: stippled. Mud: black. (A) Type 1: Typical geometry of folds in clastic sedimentary strata deformed at low metamorphic grade, showing dip isogons (lines joining points of equal dip on successive surfaces). Sandstone layers typically show tighter curvature on inner arcs (class 1 geometry; Ramsay, 1967), while mud layers have tighter curvature on outer arcs (class 3). (B) Type 2: Reversal of normal geometrical relationships characteristic of folds formed while sand was liquefied. (C) Type 3: Undeformed configuration with: angular mud clasts surrounded by sand; sand-filled dikes cross-cutting mud layers; and folded liquefied sand layers. (D) Type 4: Type 3 with superimposed simple shear parallel to bedding. (E) Type 5: Type 3 with superimposed pure shear parallel to bedding. (F) Type 6: Type 3 with superimposed arbitrary strain. Source: From Waldron, J.W.F., Gagnon, J.F., 2011. Recognizing soft-sediment structures in deformed rocks of orogens. J. Struct. Geol. 33, 271 279, with permission from Elsevier.

by liquid drops is still limited. In bridging this gap, Zhao et al. (2015) conducted experiments, which offered valuable geologic insights. By combining high-speed photography with high-precision laser profilometry (Fig. 9.3), Zhao et al. (2015) investigated liquid-drop impact dynamics on granular surface and monitored the morphology of resulting impact craters (Fig. 9.3I). Surprisingly, they found that despite the enormous energy and length differences, granular impact cratering by liquid drops follows the same energy scaling and reproduces the same crater morphology as that of asteroid-impact craters.

Mass transport, gravity flows, and bottom currents

404

9. Soft-sediment deformation structures

FIGURE 9.3 Experimental results of granular impact cratering by liquid drops (Zhao et al., 2015). I. Snapshots from three high-speed movies showing the impact of a 3.1-mm water drop with different E (5energy) levels on a granular surface. Top row (Movie S1): A E with E 5 7.8 3 1026 J (low energy); middle row (Movie S2): F J with E 5 6.0 3 1025 J (intermediate energy); bottom row (Movie S3): K O with E 5 2.3 3 1024 J (high energy). J 5 Joule. For the low E, the time elapsed after the initial impact is t 5 1.1 ms (A), 4.5 ms (B), 13.8 ms (C), 32.8 ms (D), and 84.0 ms (E); For the intermediate E, t 5 0.3 ms (F), 5.7 ms (G), 11.9 ms (H), 19.4 ms (I), and 56.8 ms (J); For the high E, t 5 0.3 ms (F), 1.0 ms (G), 1.9 ms (H), 6.4 ms (I), and 29.1 ms (J). Water in the liquid-granular mixtures gradually drains into the bed on the time scale of a second. II. Liquid-drop impact morphologies resembling those of “raindrop” imprints. Scale bars 5 3 mm. A and B: ring-shaped granular residues at low E; C: solid pellet-shaped granular residue at intermediate E; D and E: asymmetric granular residues; F: splash pattern at high E. III. Scaling of liquid-drop impact craters. A: Dc versus E for different drop sizes. Solid lines indicate the 0.17 scaling. The dashed line indicates the 1/4 scaling. B: Scaled Dc following the S-H scaling rule (see the text; Eq. (1)). The dashed line indicates 1.74. Source: Images from Zhao, R., Zhang, Q., Tjugito, H., Cheng, X., 2015. Granular impact cratering by liquid drops: understanding raindrop imprints through an analogy to asteroid strikes. Proceedings of the National Academy of Sciences, 112 (2), 342 347, with some labels added by G. Shanmugam. With permission from PNAS (Proceedings of the National Academy of Sciences of the United States of America). Public Domain.

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

405

Zhao et al. (2015) integrated the physical insight from planetary sciences, the liquid marble model from fluid mechanics, and the concept of jamming transition from granular physics into a simple theoretical framework that quantitatively describes all of the main features of liquid-drop imprints in granular media. Zhao et al. (2015) expressed that the 0.18 scaling of liquid-drop strikes on a granular surface is quantitatively similar to the Schmidt Holsapple (S H) scaling from hypervelocity impact cratering associated with asteroid strikes (Holsapple, 1993):





Dc 5 g20:17 D0:83 U 0:34 20:17 D0:32 E0:17 Dc 5 ρg





(9.1) (9.2)

where DC is the diameter of impact crater, D is the drop size, U is the impact velocity, ρ is the density of the projectile, g is the gravitational acceleration, and E is the energy. Zhao et al. (2015) characterized the size of an impact crater by measuring its diameter, Dc (Fig. 9.3II-C). By plotting Dc versus E (Fig. 9.3II), the authors have shown a power-law scaling with an exponent of 0.17 6 0.01 (Fig. 9.3II-A), consistent with Nefzaoui and Skurtys’s result (Nefzaoui and Skurtys, 2012). In summary, Zhao et al. (2015, p. 347) have shown that there is a convincing quantitative similarity between raindrop-impact cratering and asteroid-impact cratering in terms of both the energy scaling and the aspect ratio of their impact craters. Compared with extensively studied low-speed solid sphere impact cratering, liquid-drop impact cratering provides a better analogy to high-energy asteroid-impact cratering. The geologic implication of this breakthrough study is that future researchers of SSDS are compelled to consider asteroid-impact origin as a potential alternative for raindrop-like imprints seen on sedimentary strata.

9.7 Geological implications based on case studies 9.7.1 Breccias Breccias were generated by earthquakes from the Nankai Trough in Southwest Japan. IODP Expedition 316 cores from this trough contain brecciated mud clasts (Fig. 9.4).

9.7.2 Lateral extent IODP Expedition 308 of the Ursa Basin in the Gulf of Mexico has shown that Pleistocene mass-transport deposits (MTDs) contain not only common SSDS, such as deformed layers and steeply dipping layers (Fig. 9.5), but can also be correlated over a 12-km distance (Fig. 9.5B).

Mass transport, gravity flows, and bottom currents

406

9. Soft-sediment deformation structures

FIGURE 9.4

A case study showing earthquake-generated brecciated mud clasts in an active subduction zone. (A) Tectonic setting of Nankai Trough subduction zone. Large earthquakes occurred along Nankai Trough on several segments: Tokai (TK), Tonankai (TN), and Nankaido (ND). Orange shaded segment 5 CE 1944 Tonankai earthquake. NW 2 SE red line represents cross section shown in (B). (B) Cross section (red line in A). Two major fault systems, plate boundary and megasplay faults (after Park et al., 2002), are developed in this area (after Moore et al., 2007). (C) Locations of IODP Expedition 316 Sites C0004 (hanging wall, D) and C0008 (footwall, E). (D) X-ray computed tomography (X-CT) and schematic images of cores from hanging wall (sample from Site C0004). (E) X-ray computed tomography (X-CT) and schematic images of cores from footwall (sample from Site C0008) of megasplay fault. Interval from seafloor to 80 cm depth is shown. Source: From Sakaguchi, A., Kimura, G., Strasser, M., Screaton, E.J., Curewitz, D., Murayama, M., 2011. Episodic seafloor mud brecciation due to great subduction zone earthquakes. Geology 39, 919 922, with permission from GSA.

9.7.3 Ocean bottom currents Types of bottom currents are discussed in Chapter 8, Bottom currents. Convolute bedding is a common type of SSDS. They have been reported from the ODP Leg 175, off South Africa in the Mid-Cape Basin, South Atlantic (Fig. 9.6). The significance of this occurrence in association with a depositional slump is that it is likely to be related to the Benguela Current (Table 9.2, Case study 21), not earthquakes. The Benguela Current is an ocean current that flows northward along the west coast of Africa from Cape Point in the south to 16 S. These nutrient-rich waters cause the Benguela Upwelling. These robust and critical datasets have commonly been overlooked by most researchers of SSDS.

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

407

FIGURE 9.5 (A) Image showing locations of the Mississippi Canyon and Expedition 308 Sites U1324, U1323, and U1322 in the Ursa Basin, Northern Gulf of Mexico. (B) An East 2 West cross section, based on seismic reflection profiles, showing laterally extensive ( . 12 km) distribution of MTD (mass-transport deposits) in the three cored sites. (C) Core photograph and related sketch showing fault and deformed clay layers in MTD. Red arrow points to location of site. (D) Core photograph and related sketch showing steeply dipping clay layers in MTD. Source: From Expedition 308 Scientists (2006), with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

9.7.4 Mass-transport deposits 1. DSDP Site 180 is located on the Aleutian Trench slope (Fig. 9.7). Piper (1975), who emphasized sediment deformation in DSDP cores, reported bowed laminae (Fig. 9.7C) and suggested that these SSDS are natural in origin and associated with mass transport. 2. IODP Expedition Site U1418 is also located in the vicinity of the Aleutian Trench in the Gulf of Alaska (Fig. 9.8). Cores show mud clasts (Fig. 9.8), breccias, and contorted laminae in clay. Jaeger et al. (2014a,b) suggested that these SSDS are related to MTDs. 3. ODP Site 884 is located near the western tip of the Aleutian Islands in the North Pacific (Fig. 9.9). Cores show slump folds and normal faults (Fig. 9.9). Again, these SSDS are associated with MTDs (Shipboard Scientific Party, 1993). In other words, SSDS are associated with MTDs not only in tectonically active subduction zones, but also in other

Mass transport, gravity flows, and bottom currents

408

9. Soft-sediment deformation structures

FIGURE 9.6 (A) Image showing location of ODP Leg 175, Site 1085, Mid-Cape Basin, off South Africa, South Atlantic. Note north flowing Benguela Current. (B) A seismic profile with ODP Site 1085, Mid-Cape Basin. (C) Core photograph showing convolute layers. Red arrow points to position of photograph. Source: (A) http:// oceancurrents.rsmas.miami.edu/atlantic/img_topo2/benguela2.jpg. Labels added by G. Shanmugam. (B and C) From Shipboard Scientific Party, 1998. Site 1085. In: Wefer, G., Berger, W.H., Richter, C., et al., Proceedings of Ocean Drilling Program, Initial Reports, 134. College Station, Texas, pp. 385 428. https://doi.org/10.2973/odp.proc.ir.175.113.1998, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

basins, such as the Gulf of Mexico (Fig. 9.5). The implication is that SSDS are not unique to seismically active regions. The widespread distribution of MTD along the U.S. Atlantic margin has been well documented (Embley, 1980; Shipboard Scientific Party, 1994; Twichell et al., 2009). Geophysical and sedimentological studies of widespread MTD were also carried out on the Amazon Fan, Equatorial Atlantic (Damuth and Embley, 1981; Piper et al., 1997; Shanmugam, 2006a).

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

409

FIGURE 9.7 (A) Image showing the location of DSDP Site 180 (filled red circle) in the Gulf of Alaska. (B) A NW 2 SE cross section showing the location of Site 180 in the Aleutian Trench. Note Giacomni Seamount. (C) Core photograph and related sketch showing bowed laminae. Red arrow points to site location. Source: (A) NOAA. (B) From Shipboard Scientific Party, 1973a. 1. Introduction. In: Initial Reports of the Deep Sea Drilling Project, vol. 18. United States Government Printing Office, Washington, pp. 5 8. https://doi.org/10.2973/dsdp. proc.18.101.1973. (C) Photograph from Piper, D.J.W., 1975. Appendix II. Deformation of stiff and semilithified cores from legs 18 and 28. In: Deep Sea Drilling Project, Volume 28. pp. 977 979. https://doi.org/10.2973/dsdp.proc.28.app2.1975. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

However, sedimentological features such as slump folds, steeply dipping layers, and breccias are not only classified as SSDS but also as depositional features of MTD in numerous case studies (Table 9.2). Some obvious observations and their implications are: 1. Slump fold is associated with MTD not only in a seismically active environment near the Aleutian Islands (Fig. 9.9) and in the NanTroSEIZE Complex, Philippine Sea Plate (Fig. 9.10), but also with MTD, unrelated to earthquakes, in a submarine canyon environment in the Edom Field, offshore Nigeria (Fig. 9.11). 2. Steeply dipping layers, possibly caused by slumping, are associated not only with seismically active environment in the Amati Sankuru Basin, Philippine Sea (Fig. 9.12), but also with MTD in a passive depositional setting in the Gulf of Mexico (Fig. 9.5).

Mass transport, gravity flows, and bottom currents

410

9. Soft-sediment deformation structures

FIGURE 9.8 (A) Image showing the location of IOPD Expedition 341 Site U1418 (filled red circle) in the Gulf of Alaska. GS 5 Giacomini Seamount (see Figure 9B for cross section). (B) Seismic profile showing the position of Site U1418. (C E) Core photographs showing mud clasts in clayey matrix. Red arrow points to position of core photograph near the bottom of core. Source: (A) From Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 Scientists, 2014a. Expedition 341 Summary. In: Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 341, pp. 1 130. (B) From Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 Scientists, 2014b. Site U1418. In: Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 341. https://doi.org/10.2204/iodp.proc.341.104.2014. (C E) Core photographs from Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 Scientists, 2014b. Site U1418. In: Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 341. https://doi.org/10.2204/iodp.proc.341.104.2014. Additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

3. Load casts are not unique to a particular triggering mechanism because they are commonly associated with turbidities and MTD of different triggering mechanisms (Table 9.2; Fig. 9.13). 4. Similarly, faults (Fig. 9.14), sand injections (Fig. 9.15), brecciated mud clasts, and folds (Fig. 9.16) are present in both MTD and SSDS.

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

411

FIGURE 9.9 (A) Image showing location of ODP Site 884 (filled red circle) in the North Pacific. Note the site position near the western tip of Aleutian Islands. (B) Core photograph showing normal fault (white arrow). (C) Core photograph showing slump fold in mud. Source: (A) NOAA. (B and C) Core photographs from Shipboard Scientific Party, 1993. Site 884. In: Rea, D.K., Basov, I.A., Janecek, T.R., Palmer-Julson, A., et al., Proceedings of Ocean Drilling Program, Initial Reports, 134. College Station, Texas, pp. 209 302. https://doi.org/10.2973/odp.proc. ir.145.108.1993. Additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

5. Other examples of SSDS in various forms are mud clasts (Fig. 9.17), pockmarks (Figs. 9.18 9.20), balloon-shaped inflation structures (Fig. 9.21A), explosive sand-gravel crater (Fig. 9.21B), disharmonic fold on Mars (Fig. 9.22), and nodular anhydrites (Fig. 9.23).

Mass transport, gravity flows, and bottom currents

412

9. Soft-sediment deformation structures

FIGURE 9.10 (A) Image showing location of IODP Expedition 333, Site C0018 (filled red circle) in the NanTroSEIZE complex. White barbed line 5 position of deformation front of accretionary prism. Yellow arrow 5 estimated far-field vectors between Philippine Sea Plate and Japan. The Shikoku Basin was previously drilled during DSDP Leg 58 (Klein et al., 1980). (B) A NW 2 SE seismic profile showing position of Site C0018. Note the accretionary prism above the subducting Philippine Sea Plate. Also note location of Site C0012 near the SE end of seismic profile on a Knoll (see Fig. 9.15 for details). (C) Core photograph showing slump fold in mudstone. Red arrow points to Site C0018 location. (D) CT-Scan image of slump fold in core. Source: (A and B) From Expedition 333 Scientists, 2011. Integrated Ocean Drilling Program Expedition 333 Preliminary Report NanTroSEIZE Stage 2: Subduction inputs 2 and heat flow. Texas A&M University, College Station, Texas. https://doi.org/10.2204/iodp. pr.333.2011, with additional labels by G. Shanmugam. (B) Expedition 333 Scientists, 2011. Integrated Ocean Drilling Program Expedition 333 Preliminary Report NanTroSEIZE Stage 2: Subduction inputs 2 and heat flow. Texas A&M University, College Station, Texas. https://doi.org/10.2204/iodp.pr.333.2011. (D) From Expedition 333 Scientists, 2012b. Site C0018. In: Henry, P., Kanamatsu, T., Moe, K., the Expedition 333 Scientists, (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 333. Texas A&M University, College Station, Texas. https://doi.org/10.2204/iodp. proc.333.103.2012, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

413

FIGURE 9.11 (A) Map showing location of the Edop Field (filled red circle), offshore Nigeria. (B) A submarine canyon-fill model dominated by MTD in the Edop Field (Pliocene), offshore Nigeria. (C) Core photograph showing primary glide plane between overlying sand and underlying slump-folded mudstone, Pliocene, offshore Nigeria. Red arrow points to Well 25C from which the core was recovered. Source: (A) NOAA. (C) After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

9.7.5 Future research The problem is how to distinguish MTD from SSDS. After all, slump structures are not only classified as SSDS but are also considered a type of MTD. Attempting to separate SSDS from MTD is a distinction without a difference—a logical fallacy (Fig. 9.24). The other problem is the interrelationship between earthquakes and MTD. For example, an earthquake can trigger MTD and in turn MTD can trigger earthquakes, with both generating SSDS. In the real world, all these triggering mechanisms can and do occur simultaneously (Fig. 9.25). There are no objective criteria to resolve this problem yet. However, the wealth of available core data from scientific drilling at sea offers hope that the trigger and MTD problems may be resolved by undertaking a global research initiative (Fig. 9.26).

Mass transport, gravity flows, and bottom currents

414

9. Soft-sediment deformation structures

FIGURE 9.12 (A) Image showing location of IODP Expedition 351, Site U1438 (filled red circle) in the Philippine Sea Region. (B) Core photograph showing steeply dipping layers (yellow arrow) in tuffaceous mudstone, possibly caused by slumping. Red arrow points to site location. Source: From Expedition 351 Scientists, 2015. Site U1438. In: Arculus, R.J., Ishizuka, O., Bogus, K., and the Expedition 351 Scientists, (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 351. Texas A&M University, College Station, Texas. https://doi.org/10.14379/iodp. proc.351.103.2015. (B) From Expedition 351 Scientists, 2015. Site U1438. In: Arculus, R.J., Ishizuka, O., Bogus, K., and the Expedition 351 Scientists, (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 351. Texas A&M University, College Station, Texas. https://doi.org/10.14379/iodp.proc.351.103.2015, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

9.7.6 Unresolved issues 9.7.6.1 The seismite problem During a period of 88 years (1931 2013), 41 genetic terms were introduced for various deposits (Table 9.3). Of the 41 terms, only 11 are meaningful in understanding the true depositional origin (e.g., turbidities), the remaining 30 are just jargons (e.g., seatmates, tsunamites, etc.). The genetic term “seatmates,” introduced by Seilacher (1969) for recognizing paleoearthquakes in the sedimentary record, is a misnomer (Shanmugam, 2016a). The

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

415

FIGURE 9.13 (A) Image showing location of IODP Expedition 349, Site U1431 (filled red circle) in the East Subbasin of the South China Sea. (B) Lower part of the lithologic log of Site U1431 showing Miocene units. (C) Core photograph showing volcanic breccia. (D) Core photograph showing load cast. (E) Core photograph showing steeply-dipping layers. (F) Core photograph showing inverse grading. Red arrows point to positions of photographs. Source: All images are from Expedition 349 Scientists, 2015a. Site U1431. In: Li, C.-F., Lin, J., Kulhanek, D.K., the Expedition 349 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 349. Texas A&M University, College Station, Texas. https://doi.org/10.14379/iodp.proc.349.103.2015, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

term was introduced in haste, based on an examination of a single exposure of the Miocene Monterey Formation (10 m) in California, without a rigorous scientific analysis. The fundamental problem is that earthquake is a triggering mechanism, not a depositional process. Type of triggers cannot be recognized in the ancient sedimentary record because evidence for triggers is not preserved by nature. SSDS, commonly used as the criteria for interpreting seatmates, are a product of liquefaction. However, liquefaction can be induced by any one of 21 triggers, which include earthquakes, meteorite impacts, tsunamis, sediment loading, among others. Brecciated clasts, typically associated with earthquakeinduced deposits in the Dead Sea Basin, are also common depositional products of debris

Mass transport, gravity flows, and bottom currents

416

9. Soft-sediment deformation structures

FIGURE 9.14 (A) Regional bathymetric map showing seismic reflection profiles (red line, NDS3) and location of Site U1434 (filled red circle). (B) Seismic profile Line NDS3, with location of Site U1434. (C and D) Core photographs showing faults. (E) Line-scan image of a fault observed in the claystone. (F) White brush strokes are drawn on the image tracing the claystone laminations, whose orientation change slightly between hanging and foot walls. Source: From Expedition 349 Scientists, 2015b. Site U1434. In: Li, C.-F., Lin, J., Kulhanek, D.K., the Expedition 349 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 349. Texas A7M University, College Station, Texas. https://doi.org/10.14379/iodp.proc.349.106.2015.

flows (i.e., synsedimentary product unrelated to earthquakes). Also, various types of SSDS, such as duplex-like structures and clastic injections, can be explained by synsedimentary processes unrelated to earthquakes. Case studies of sandstone petroleum reservoirs worldwide, which include Gulf of Mexico, North Sea, Norwegian Sea, Nigeria, Equatorial Guinea, Gabon, and Bay of Bengal, reveal that there is compelling empirical evidence for sediment loading being the primary cause of SSDS. The Krishna-Godavari Basin, located on the eastern continental margin of India, is ideal for sediment failures by multiple triggering mechanisms where overpressure and liquefaction have led to multiorigin SSDS. Because tsunamis and meteorite impacts are important phenomena in developing extensive deposits, lateral extent of SSDS cannot be used as a unique distinguishing attribute of earthquakes. For these reasons, the genetic term “seatmates,” which has no redeemable scientific value, is obsolete for interpreting ancient geologic record.

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

417

FIGURE 9.15 (A) Detailed bathymetric map of Kashinosaki Knoll and Nankai Trough showing location of Site C0012 (filled red circle). (B) Sand injection (dike) in silty claystone. Yellow triangles show the trend of dike. Red arrow points to site location on the Knoll. Source: From Expedition 333 Scientists, 2012a. Site C0012. In: Henry, P., Kanamatsu, T., Moe, K., the Expedition 333 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 333. A&M University, College Station, Texas. https://doi.org/10.2204/iodp.proc.333.105.2012, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

9.7.6.2 The breccias problem At present, there are no criteria to distinguish SSDS formed by earthquakes (Fig. 9.27) from SSDS formed by the other 20 triggering mechanisms in the ancient sedimentary record (Shanmugam, 2017a). Even if one believes that earthquakes are the true triggering mechanism of SSDS in a given case, the story is still incomplete. This is because earthquakes (seismic shocks) are induced by a variety of causes: (1) global tectonics and associated faults (i.e., mid-ocean ridges, trenches, and transform faults), (2) meteorite-impact events, (3) volcanic eruptions, (4) postglacial uplift, (5) tsunami impact, (6) cyclonic impact, (7) landslides (MTDs), (8) tidal activity, (9) sea-level rise, (10) erosion, and (11) fluid pumping. These different causes are important for developing SSDS. Breccias are an important

Mass transport, gravity flows, and bottom currents

418

9. Soft-sediment deformation structures

FIGURE 9.16

(A) Image showing location of DSDP Site 357 on the Rio Grande Rise, off Brazil, South Atlantic. (B) Core photographs showing brecciated mud clasts of volcanic origin. Red arrow shows stratigraphic position of core photographs. (C) Simplified stratigraphy of Site 357. Source: (A) ETOPO1 Global Relief Model, C. Amante and B.W. Eakins, ETOPO1 Arc-Minute Global Relief Model: Procedures, Data Sources and Analysis, NOAA Technical Memorandum NESDIS NGDC-24, March 2009. (B) From Thiede, J., 1977. 10. Sedimentary structures in pelagic and hemipelagic sediments from the Central and Southern Atlantic Ocean. In: Initial Reports of the Deep Sea Drilling Project, Leg 39. United States Government Printing Office, Washington, pp. 407 421. https://doi.org/10.2973/dsdp.proc.39.110.1977, with additional labels by G. Shanmugam. (C) From Perch-Nielsen, K., Supko, P.R., Boersma, A., Carlson, R.L., Dinkelman, M.G., Fodor, R.V., et al., 1977. 6. Site 357: Rio Grande Rise. In: Supko, P.R., Perch-Nielsen, K., et al. (Eds.), Initial Reports of the Deep Sea Drilling Project, Leg 39. United States Government Printing Office, Washington, pp. 231 327. https://doi. org/10.2973/dsdp.proc.39.106.1977. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

group of SSDS. Although there are many types of breccias classified on the basis of their origin, five types are of significance, namely (1) fault (Figs. 9.28 9.30), (2) volcanic, (3) meteorite impact, (4) sedimentary-depositional, and (5) sedimentary-collapse (Fig. 9.31). Although different breccia types may resemble each other, distinguishing one type (e.g.,

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

419

FIGURE 9.17 (A) Image showing location of IODP Expedition 355, Site U1457 (filled red circle) in the Arabian Sea. (B) Seismic profile and its interpretation (C) showing location of Site U1457. Note MTD at the bottom of hole. (D) Lithologic log of Site U1457 showing volcanic breccia in Unit V. (E) Core photograph from Unit V showing brecciated mud clasts in claystone. Red arrow points to stratigraphic position of photograph. Source: (A D) From Pandey, D.K., Clift, P.D., Kulhanek, D.K., the Expedition 355 Scientists, 2016a. Expedition 355 summary. In: Proceedings of the International Ocean Discovery Program, vol. 355. https://doi.org/10.14379/iodp.proc.355.101.2016, with additional labels by G. Shanmugam. (E) From Pandey, D.K., Clift, P.D., Kulhanek, D.K., the Expedition 355 Scientists, 2016b. Site U1457. In: Proceedings of the International Ocean Discovery Program, vol. 355. https://doi.org/10.14379/iodp. proc.355.104.2016, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

meteorite breccias) from the other types (e.g., fault, volcanic, and sedimentary breccias) has important implications. (1) Meteorite breccias are characterized by shock features (e.g., planar deformation features in mineral grains, planar fractures, high-pressure polymorphs, shock melts, etc.), whereas sedimentary-depositional breccias (e.g., debrites) do not. (2) Meteorite breccias imply a confined sediment distribution in the vicinity of craters, whereas sedimentary-depositional breccias imply an unconfined sediment distribution, variable sediment transport, and variable sediment provenance. (3) Meteorite, volcanic, and fault breccias

Mass transport, gravity flows, and bottom currents

FIGURE 9.18 Multichannel seismic and 3.5-kHz profiles showing pockmark and associated chimney structures (also called pipes) on the northern flank of the Storegga Slide, offshore Norway. The section of profile outlined with the box is expanded in the lower left to illustrate a velocity pull-up. A BSR (bottom simulating reflector) is also indicated. The panel on the lower right is a 3.5-kHz record showing the same pockmark and associated chimney (pipe). This feature is B200 m across, and 8 m deeper than the surrounding seafloor. Note that this pockmark is hollow. Source: From Paull et al. (2008). After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

9.7 Geological implications based on case studies

421

FIGURE 9.19 Distribution of pockmarks in Lake Constance, Germany (see Fig. 9.1 for location). (A) Multibeam echo beam sounder data showing the general morphology of channels and sediment waves. Note two boxes (B and C) for enlarged areas showing details. (B) Individual pockmarks (white arrow) on top of sediment waves. (C) A series of pockmarks aligned along the side of a channel. Contour lines 5 2 m. Source: From Wessels, M., Bussmann, I., Schloemer, S., Schlu¨ter, M., Bo¨der, V., 2010. Distribution, morphology, and formation of pockmarks in Lake Constance, Germany. Limnol. Oceanogr. 55 (6), 2623 2633. https://doi.org/10.4319/lo.2010.55.6.2623, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

Mass transport, gravity flows, and bottom currents

422

9. Soft-sediment deformation structures

FIGURE 9.20 Origin of pockmarks by gas expulsion. (A) A gas reservoir forms when migrating gas from deeper horizons is collected at the top of topographic elevations such as sediment waves or shoulders of channels (see Fig. 9.19). (B) Once the hydrostatic pressure is exceeded, overlying sediment collapses and forms a pockmark, which still produces bubble streams of gas. Note that pockmarks are controlled by topographic highs associated with two different local environments, namely sediment waves and channels. Source: From Wessels, M., Bussmann, I., Schloemer, S., Schlu¨ter, M., Bo¨der, V., 2010. Distribution, morphology, and formation of pockmarks in Lake Constance, Germany. Limnol. Oceanogr. 55 (6), 2623 2633. https://doi.org/10.4319/lo.2010.55.6.2623, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

Mass transport, gravity flows, and bottom currents

9.7 Geological implications based on case studies

423

FIGURE 9.21 (A) Diagrams illustrating the genesis of the newly described "Balloon-shaped inflation structures" primarily by gas expulsion. (B) Diagram illustrating an explosive sand-gravel crater. Source: (A) From Hilbert-Wolf, H.L., Roberts, E.M., Simpson, E.L., 2016. New sedimentary structures in seismites from SW Tanzania: Evaluating gas- vs. water-escape mechanisms of soft deformation. Sediment. Geol. 344, 253 262. (B) From Rodrı´guezPascua, M.A., Silva, P.G., Perucha, M.A., Giner-Robles, J.L., Heras, C., Bastida, A.B., et al., 2016. Seismically induced liquefaction structures in La Magdalena archaeological site, the 4th century AD Roman Complutum (Madrid, Spain). Sediment. Geol. 344, 34 46, with permission from Elsevier.

Mass transport, gravity flows, and bottom currents

424

9. Soft-sediment deformation structures

FIGURE 9.22 (A) Mars Orbital Laser Altimeter (MOLA) of a colorized elevation base map with eight regions of deformation mapped and labeled. Four styles of deformation were observed, kilometer-scale convolute folds (purple), detached slabs (red), folded strata (yellow), and pull-apart structures (orange). Deformation was observed in Candor, Melas, and Ius Chasmata but not in Coprates or Tithonium Chasmata. White circle shows location of kilometer-scale convolute folded outcrops. (B) Kilometer-scale convolute folds that are observed only in two exposures in southern Melas basin. (C) Folded strata in Candor Chasma. Source: From Metz, M., Grotzinger, J., Okubo, C., Milliken, R., 2010. Thin-skinned deformation of sedimentary rocks in Valles Marineris, Mars. J. Geophys. Res. 115 (E11004). https://doi.org/10.1029/2010JE003593, with some labels added by G. Shanmugam. With permission from AGU. Note that these images were originally included as part of Metz’s (2010) Ph.D. Dissertation, which the author defended on March 30, 2010, at the California Institute of Technology. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

Mass transport, gravity flows, and bottom currents

FIGURE 9.23 Stratigraphic section showing positions of chicken-wire (filled red circles) and enterolithic structures (filled blue circle) observed in the Messinian (Late Miocene) evaporites recovered from the DSDP Leg 42A at Site 371 in the Mediterranean Sea. Anhydrite classification is after Maiklem et al. (1969). Source: From Garrison, R.E., Schreiber, B.C., Bernoulli, D., Fabricius, F.H., Kidd, R.B., Melieres, F., et al. 1978. Sedimentary petrology and structures of Messinian evaporitic sediments in the Mediterranean Sea. In: Initial Reports of the Deep Sea Drilling Project, Leg 42. United States Government Printing Office, Washington, pp. 571 611. https://doi.org/10.2973/dsdp.proc.42-1.123.1978, with additional labels by G. Shanmugam. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

FIGURE 9.24 Selected types of triggers, state of liquefaction, and SSDS. There are 21 triggers and they are all directly or indirectly responsible for sediment transport, deposition, and liquefaction (Shanmugam, 2016c). In reflecting published literature, earthquakes and tectonic activity are listed as two different types. However, earthquakes are an integral component of global tectonics (Kearey et al., 2009). Note that both tectonic and nontectonic triggers go through liquefaction in developing SSDS. Also note that earthquake is one of many triggers that develop SSDS. SSDS are not seismites. Thin blue arrows: one or more sediment transport processes with or without flow transformations (Fisher, 1983). Thick gray arrow: final deposition. SSDS 5 soft-sediment deformation structures. See Shanmugam (2006a, 2006b, 2012a, 2016c, 2017a) for discussion of examples of triggers shown here. Relevant references include: Basilone et al., 2014, Beck, 2009, Gradmann et al., 2012, Malkawi and Alawneh, 2000, Obermeier et al., 2002, Scholz et al., 2011. Source: After Shanmugam, G., 2016c. The seismite problem. J. Palaeogeogr. 5 (4), 318 362. Elsevier. Creative common.

FIGURE 9.25 Diagram illustrating complex interrelationships among the order of triggers, sediment transport, state of liquefaction, and deposition of SSDS. There are 21 triggers and they are all directly or indirectly responsible for transport processes, depositional mechanisms, and related liquefaction (Shanmugam, 2016c). Note that an earthquake can trigger tsunami waves that in turn can trigger mass movements. Thin red arrows: triggering of other triggers. Thin blue arrows: one or more sediment transport processes with or without flow transformations (Fisher, 1983). Thick gray arrow: final deposition. Note that mass movement can function both as a trigger and as a transport process. See Shanmugam (2006a, 2006b, 2012a) for discussion of examples of triggers shown here. Source: After Shanmugam, G., 2016c. The seismite problem. J. Palaeogeogr. 5 (4), 318 362. Elsevier. Creative common.

428

9. Soft-sediment deformation structures

FIGURE 9.26 Image showing potential areas (numbers in blue circles) of future research on triggering mechanisms of SSDS using continental, transitional, and marine environments. Examples (modern and ancient): (1) earthquakes: Aleutian Trench (Gulf of Alaska: boxed area in northwest part of image), Iceland, and Nankai Trough (Sea of Japan, Fig. 9.4). (2) Volcanic activity: Mount St. Helens, Yellowstone, Hawaiian Islands, Iceland, Montserrat (North Atlantic), Rio Grande Rise (South Atlantic), Ecuador, Mount Pinatubo (Philippines), and New Zealand (North Island). (3) Meteorite impact: Chicxulub (Yucatan, Mexico), Tunguska (Russia), Woodleigh (Western Australia), and Lake Acraman (Southern Australia). (4) Salt tectonics: Gulf of Mexico. (5) Mass-transport deposits (MTD): California Submarine canyons, U.S. Atlantic Continental Margin, Canary Debris Flow (Northwest Africa), and Agulhas Slump (Southeast Africa). (6) Tsunamis: Bay of Bengal and Hawaiian Islands. (7) Cyclones: Bay of Bengal and Gulf of Mexico. (8) Internal waves and tides: Hawaiian Islands, Mid-Atlantic Ridge (South Atlantic), and Sulu Sea. (9) Ocean bottom currents: Gulf of Cadiz, Faeroe-Shetland Channel (North Atlantic), and Argentine Basin (South Atlantic). (10) Glacial activity: Glacier Bay (United States), Russell Glacier (Greenland), Perito Moreno Glacier (Southwest Argentina), Norway 2 Sweden region, and Antarctica. (11) Wind: Sahara, Namibia, Saudi Arabia, and Taklamakan (Northwest China). (12) Lake: Lake Baikal (Siberia) and Lake Constance (Germany, Fig. 9.19). (13) Gas hydrate: Blake Ridge and vicinity (North Atlantic). (14) Karst dissolution: Guilin (Southern China). (15) Climate change: Antarctica. Note that in aiding research in marine settings, ODP sites (filled white circles) in the world’s oceans are shown. Ocean Drilling Program (ODP), which was in operation during 1983 2 2007 period, visited 669 sites, drilled: 1797 holes, and recovered a total of 222,430 m of cores. Maps are also available for DSDP and IODP sites from website: http://www-odp.tamu.edu/sitemap/sitemap.html (accessed 20.05.17.). Source: From ODP (2007), Texas A&M University, College Station, Texas. All DSDP/ODP/OPDP core photographs are available online free and can be downloaded from the ODP web page: http://www-odp.tamu.edu/. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common. Mass transport, gravity flows, and bottom currents

429

9.7 Geological implications based on case studies

TABLE 9.3 Lexicon of 41 genetic terms ending with “-ite.” Genetic terms

Comments (Shanmugam, 2006b)

References

1. Aeolianite

Implies the Aeolius (the god of the winds), not flow behavior

Sayles (1931); Bates and Jackson (1980)

2. Anastomosite

Implies river type, not flow behavior

Shanmugam (1984)

3. Atypical turbidite

Implies slumps, debris flows, and sand flows, not turbidity current

Stanley et al. (1978)

4. Baroclinite

Implies baroclinic currents

Shanmugam (2013a)

5. Braidite

Implies river type, not flow behavior

Shanmugam (1984)

6. Cascadite

Implies driving force (cascading), not depositional process

Gaudin et al. (2006)

7. Contourite

Implies current orientation, not flow behavior

Hollister (1967)

8. Debrite

Implies plastic debris flow

Pluenneke (1976)

9. Densite

Implies hybrid processes, not a single process

Gani (2004)

10. Diamictite

Implies pebbly mudstone, not flow (glacial) behavior

Flint et al. (1960)

11. Fluxoturbidite

Implies no discernible meaning (see Hsu¨, 1989)

Dzulynski et al. (1959)

12. Grainite

Implies grains, not flow behavior

Khvorova (1978)

13. Gravitite

Implies sediment gravity, not flow behavior

Natland (1967)

14. Gravite

Implies multiple processes, not a single process

Gani (2004)

15. Hemipelagite

Implies hemipelagic settling

Arrhenius (1963)

16. Hemiturbidite

Implies muddy turbidity current

Stow et al. (1990)

17. High-concentration sandy turbidite

Implies sandy debris flow, not turbidity current (Shanmugam, 1996a)

Abreu et al. (2003)

18. Homogenite

Implies uniform grain size (ungraded mud), not flow behavior

Kastens and Cita (1981)

19. Hybridite

Implies ntersection of two different flow types (hybrid flow)

Shanmugam (2020) (see Appendix A in this book)

20. Hyperpycnite

Implies river discharge, not flow behavior

Mulder et al. (2002)

21. Impactitea

Implies impact by meteorite, not flow behavior

Sto¨ffler and Grieve (2003)

22. Injectitea

Implies injection, not flow behavior

Vivas et al. (1988)

23. Internalite

Implies internal waves and tides, not flow behavior of baroclinic currents (Shanmugam, 2013b)

Pomar et al. (2012) (Continued)

Mass transport, gravity flows, and bottom currents

430

9. Soft-sediment deformation structures

TABLE 9.3 (Continued) Genetic terms

Comments (Shanmugam, 2006b)

References

24. Interpretite

A spoof on genetic terms!

Davies (1997)

25. Meanderite

Implies river type, not flow behavior

Shanmugam (1984)

26. Megaturbidite

Implies debris flow, not turbidity current

Labaume et al. (1987)

27. Pelagite

Implies pelagic settling

Arrhenius (1963)

Implies plumes, not flow behavior (see Hyperpycnites)

Mutti (2019)

29. Seismite

Implies triggering mechanism (i.e., seismic shocks), not flow behavior

Seilacher (1969)

30. Seismoturbidite

Implies mass flow (debris flow), not turbidity current

Mutti et al. (1984a)

31. Suspensite

Implies suspension settling

Lisitsyn (1986)

32. Tectonite

Implies tectonic deformation, not flow behavior

Turner and Weiss (1963)

33. Tempestite

Implies multiple processes, not a single process

Ager (1974)

34. tidalite

Implies deposition from tidal currents

Klein (1971, 1998)

35. Tillite

Implies pebbly mudstone, not flow (glacial) behavior

Harland et al. (1966)

36. Tractionite

Implies traction deposition by bottom current

Natland (1967)

37. Tsunamite

Implies multiple processes, not a single process (Shanmugam, 2006b)

Gong-Yiming (1988)

38. Turbidite

Implies turbulent turbidity current

Kuenen (1957)

39. Undaturbidite

Implies no discernible meaning

Rizzini and Passega (1964)

40. Unifite

Implies grain size (ungraded mud), not flow behavior

Feldhausen et al. (1981), Stanley (1981)

41. Winnowite

Implies winnowing action of bottom current

Shanmugam and Moiola (1982)

28. Plumite a

a Unrelated to depositional processes. Note that transportational processes may be different from depositional processes of a deposit due to flow transformation (Fisher, 1983). Note 1: Genetic terms in bold font are obsolete in depositional process sedimentology. Note 2: References include those that introduced the term, used the term early, or considered appropriate. Modified after Shanmugam, G., 2006b. The tsunamite problem. J. Sediment. Res. 76, 718 730. Note 3: The genetic term “Hybridite” is used for deposits of hybrid flows (see Appendix A).

Mass transport, gravity flows, and bottom currents

FIGURE 9.27 Schematic section showing the occurrence of seismic-liquefaction induced sand injections in the subsurface. This scenario is applicable to both subaerial and submarine environments that are subjected to seismic shaking. Note mud clasts in sand injections. Source: Originally from Obermeier, S.F., Olson, S.M., Green, R.A., 2005, Field occurrences of liquefaction-induced features: a primer for engineering geologic analysis of paleoseismic shaking. Eng. Geol. 76 (3 4), 209 234. After Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeogr. 6 (4), 251 320. Creative common.

FIGURE 9.28 (A) Geological map of the Pembrokeshire Peninsula, southwest Wales, United Kingdom, with 30 breccia localities associated with faults; (B) location and geological context of main map. Source: (B) From Woodcock, N.H., Miller, A.V.M., Woodhouse, C.D., 2014. Chaotic breccia zones on the Pembroke Peninsula, south Wales: evidence for collapse into voids along dilational faults. J. Struct. Geol. 69, 91 107, with permission from Elsevier.

FIGURE 9.29 (A) Outcrop photograph showing breccia zones related to vertical cross faults at Flimston Bay (locality 1, Fig. 9.28). (B) An example of chaotic breccia at Bullslaughter east (locality 3, Fig. 9.28). (C) An example of crackle breccia at Proud Giltar (locality 17, Fig. 9.28). Source: (A) Photo by Sid Howells. (A C) From Woodcock, N.H., Miller, A.V.M., Woodhouse, C.D., 2014. Chaotic breccia zones on the Pembroke Peninsula, south Wales: evidence for collapse into voids along dilational faults. J. Struct. Geol. 69, 91 107, with permission from Elsevier.

FIGURE 9.30 (A): (a) Diagrammatic map of folds and conjugate strike-slip faults formed by north 2 south Variscan shortening; (b) map of postulated post-Variscan north 2 south extensional reactivation of Variscan faults and steepened bedding; (c) cross-section across a dilational normal fault that steepens to parallel bedding at shallow depths. (B) Schematic diagram of four types of breccia formation mechanisms (see text for details). Breccia types are labeled in this article. Source: Both figures are from Woodcock, N.H., Miller, A.V.M., Woodhouse, C.D., 2014. Chaotic breccia zones on the Pembroke Peninsula, south Wales: evidence for collapse into voids along dilational faults. J. Struct. Geol. 69, 91 107, with permission from Elsevier.

FIGURE 9.31 Development of large caves in the Devonian 2 Upper Carboniferous carbonates as part of karst topography near Guilin, along the banks of Li River, Southern China. Note collapse breccia on the floor at water level. Source: Modified after Shanmugam, G., 1990c. Porosity prediction in sandstones using erosional unconformities. In: Meshri, I.D., Ortoleva, P.J. (Eds.), Prediction of Reservoir Quality through Chemical Modelling. AAPG Memoir, pp. 1 23, with permission from AAPG.

9.7 Geological implications based on case studies

435

FIGURE 9.32 Summary diagram showing that SSDS with breccias can be interpreted as five different types with different implications. Note that sedimentary 2 depositional breccias are the only type that preserves the depositional porosity and permeability at the time of deposition without diagenesis. However, sedimentary breccias are subjected to post-depositional diagenesis. For example, all five types are susceptible to develop secondary porosity (Shanmugam, 1985b). Source: From Shanmugam, G., 2017d. The fallacy of interpreting SSDS with different types of breccias as seismites amid the multifarious origins of earthquakes: implications. J. Palaeogeogr. 6 (1), 12 44. Elsevier. Creative common.

are invariably subjected to diagenesis and hydrothermal mineralization with altered reservoir quality, whereas sedimentary-depositional breccias exhibit primary (unaltered) reservoir quality. And finally, (4) sedimentary-collapse breccias are associated with economic mineralization (e.g., uranium ore), whereas sedimentary-depositional breccias are associated with petroleum reservoirs (Fig. 9.32). Based on this important group of SSDS with breccias, the current practice of interpreting all SSDS as “seatmates” is inappropriate. Ending this practice is necessary for enhancing conceptual clarity and for advancing this research domain.

Mass transport, gravity flows, and bottom currents

436

9. Soft-sediment deformation structures

FIGURE 9.33 Map of India showing the location of the study area (brown rectangle) used by Van Loon et al. (2016). The Palaeoproterozoic Chaibasa Fm., exposed in the study area, is the subject of this discussion. The boundary of Singhbhum Craton is from Pati et al. (2008). The Dhala impact structure is dated to be between 2.5 and 1.7 Ga (Pati et al., 2008) and the age of Chaibasa Fm. is 2.2 Ga (Sarkar et al., 1986). The purpose of including Dhala impact in this discussion is to illustrate that impact phenomena are well documented during the Palaeoproterozoic at the time of Chaibasa Fm. in northern India. See Shanmugam (2017d) for a review of meteorite or asteroid impact events worldwide. Source: From Shanmugam, G., 2017c. The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India: discussion and liquefaction basics. J. Palaeogeogr. 6 (3), 224 234. Elsevier. Creative common.

9.7.6.3 The tombolith problem This discussion of a paper by Van Loon et al. (2016), published in the Journal of Palaeogeography (2016, 5(4), 381 390), is aimed at illustrating that there are fundamental deficiencies (Shanmugam, 2017c), which include (1) incomplete etymological reasoning for proposing a new genetic term “tomboliths” for stromatolitic bioclasts in the Paleoproterozoic Chaibasa Formation, eastern India (Fig. 9.33), (2) omission of empirical data in documenting depositional facies using sedimentological logs, (3) misapplication of the stratigraphic concept of “angular unconformity” (Fig. 9.34), (4) failure to consider the multifarious origins of earthquakes, and (5) a dated view on the basic tenets of process sedimentology and triggering mechanisms of liquefaction (Figs. 9.35 and 9.36) that are the basis for forming SSDS. As a consequence, their conclusions are unconvincing.

Mass transport, gravity flows, and bottom currents

9.8 Synopsis

437

FIGURE 9.34 (A) Photograph showing two tomboliths. Features are labeled in this study for clarity. This photograph represents Fig. 1 in Van Loon et al. (2016a); (B) Hutton’s angular unconformity at Siccar Point, Berwickshire coast, Scotland, is known as the great unconformity. Note the discordance of dip between the Upper Old Red Sandstone (Devonian) and underlying vertical beds of Silurian. Source: (B) From Shanmugam, G., 1988. Origin, recognition and importance of erosional unconformities in sedimentary basins. In: Kleinspehn, K.L., Paola, C. (Eds.), New Perspectives in Basin Analysis. Springer-Verlag, New York, pp. 83 108. Springer. From Shanmugam, G., 2017c. The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India: discussion and liquefaction basics. J. Palaeogeogr. 6 (3), 224 234. Elsevier. Creative common

9.8 Synopsis SSDS, commonly associated with deep-water deposits, have been the focus of attention for over 150 years. Existing unconstrained definitions allow one to classify a wide range of features under the umbrella phrase “SSDS.” As a consequence, a plethora of at least 120 different types of SSDS (e.g., convolute bedding, slump folds, load casts, dish-and-pillar structures, pockmarks, raindrop imprints, explosive sand-gravel craters, clastic injections, crushed and deformed stromatolites, etc.) have been recognized in strata ranging in age from Paleoproterozoic to the present time. Two factors that control the origin of SSDS are prelithification deformation and liquidization. A sedimentological compendium of 140

Mass transport, gravity flows, and bottom currents

438

9. Soft-sediment deformation structures

FIGURE 9.35 Basic differences between a turbidite (A) and a seismite (B) bed. A turbidite represents deposition of a new bed on the seafloor, whereas a seismite represents deformation of an existing bed before lithification. Note that earthquake is only one of 21 triggering mechanisms that can generate SSDS. Source: From Shanmugam, G., 2017c. The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India: discussion and liquefaction basics. J. Palaeogeogr. 6 (3), 224 234. Elsevier. Creative common.

FIGURE 9.36 Diagram showing transformation of a solid state (preliquefaction) in which sand grains are in contact (A) to a liquid state (liquefaction) in which grains are loose under excess porewater pressure (B) due to earthquakes. Partly based on Youd (1992) and Greene et al. (1994, their Fig. 1). See also Allen (1984, Volume II, his Fig. 8.1 on p. 296 and Fig. 8.2 on p. 297). Source: From Shanmugam, G., 2017c. The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India: discussion and liquefaction basics. J. Palaeogeogr. 6 (3), 224 234. Elsevier. Creative common.

Mass transport, gravity flows, and bottom currents

9.8 Synopsis

439

case studies of SSDS worldwide, which include 30 case studies of scientific drilling at sea (DSDP/ODP/IODP), published during a period between 1863 and 2017, has yielded at least 31 different origins. Earthquakes have remained the single most dominant cause of SSDS because of the prevailing “seismite” mindset. Selected advances on SSDS research are: (1) an experimental study that revealed a quantitative similarity between raindropimpact cratering and asteroid-impact cratering; (2) IODP Expedition 308 in the Gulf of Mexico that documented extensive lateral extent ( . 12 km) of MTDs with SSDS that are unrelated to earthquakes; (3) contributions on documentation of pockmarks, recognition of new structures, and large-scale sediment deformation on Mars. Problems that hinder our understanding of SSDS still remain. They are: (1) vague definitions of the phrase “soft-sediment deformation”; (2) complex factors that govern the origin of SSDS; (3) omission of vital empirical data in documenting vertical changes in facies using measured sedimentological logs; (4) difficulties in distinguishing depositional processes from tectonic events; (5) a model-driven interpretation of SSDS (i.e., earthquake being the singular cause); (6) routine application of the genetic term “seatmates” to the “SSDS,” thus undermining the basic tenet of process sedimentology (i.e., separation of interpretation from observation); (7) the absence of objective criteria to differentiate 21 triggering mechanisms of liquefaction and related SSDS; (8) application of the process concept “high-density turbidity currents,” a process that has never been documented in modern oceans; (9) application of the process concept “sediment creep” with a velocity connotation that cannot be inferred from the ancient record; (10) classification of pockmarks, which are hollow spaces (i.e., without sediments) as SSDS, with their problematic origins by fluid expulsion, sediment degassing, fish activity, etc.; (11) application of the Earth’s climatechange model; and the most importantly, (12) an arbitrary distinction between depositional process and sediment deformation. Despite a profusion of literature on SSDS, our understanding of their origin remains muddled. A solution to the chronic SSDS problem is to utilize the robust core dataset from scientific drilling at sea (DSDP/ODP/IODP) with a constrained definition of SSDS.

Mass transport, gravity flows, and bottom currents

C H A P T E R

10 Epilogue: lessons learned O U T L I N E 10.1 Lessons learned

441

10.1 Lessons learned 1. The lesson learned from 540 case studies, involving theoretical, experimental, and empirical datasets, is that it is imperative to acknowledge the variability and complexity of gravity-driven processes in nature. Transitional and hybrid flows are ubiquitous. Criteria for recognizing their deposits are still evolving. However, hybrid flows invariably develop traction structures that may mimic the “Bouma Sequence” (Fig. 8.12). Perhaps the single most consequential lesson learned from case studies used in this book is that the use and abuse of the Bouma Sequence have misdirected interpretations of all deep-water facies. This is because a discrete rippled bed could be assumed to be "Tc" of the Bouma Sequence and could be interpreted to be a turbidite under the "base- and top-cut-out Bouma Sequence" logic of Walker (1965) (Fig. 10.1). However, based on strict observational methods, the same discrete rippled beds could alternatively be interpreted as contourites (Hollister, 1967), tidalites (Klein, 1975), baroclinites (i.e., deposits of internal waves and tides) Lonsdale et al., 1972), hybridites or bottom-current reworked sands (Shanmugam et al., 1993a) (Fig. 10.1). At present, the "cut-out" logic has infiltrated literature not only on turbidites, but also on contourites (e.g., Stow and Smillie, 2020, their Fig. 13). 2. The general term “mass transport” (i.e., slides, slumps, and debris flows) represents the failure, dislodgement, and downslope movement of either sediment or glacier under the influence of gravity. As a collective sedimentologic phenomenon, masstransport deposits (MTDs) reveal remarkable cosmic congruity in geometry on Mars,

Mass transport, gravity flows, and bottom currents DOI: https://doi.org/10.1016/B978-0-12-822576-9.00010-2

441

© 2021 Elsevier Inc. All rights reserved.

442

10. Epilogue: lessons learned

FIGURE 10.1 (A) The orthodoxy of the turbidite facies model (i.e., The Bouma Sequence) with five internal divisions. (B) My depiction of a discrete rippled bed that would be classified as "Tc" under the assumption-based "cut-out" logic of Walker (1965) and would be routinely interpreted as turbidites. (C) On the basis of conventional observations, discrete rippled beds in deep-water strata have been convincingly interpreted to be deposits of alternative processes, such as contour currents, tidal currents, wind-driven bottom currents, etc.

Jupiter, and Earth (i.e., subaerial, sublacustrine, and submarine environments). Most of them are a product of debris flows. 3. Gravity flows constitute the single most important sediment-transport mechanism on land, shelf, slope, and basin environments (Table 3.1). They play important roles not only in downslope, but also in alongslope directions (Fig. 3.61). In terms of transporting large volumes of coarse-grained sediment into the deep sea, debris flows and related mass movements are the most important of all other processes. Also, identification markers of debrites discussed in this book are of value for recognizing them in the ancient sedimentary record because sandy debrites are important petroleum reservoirs worldwide (Chapters 2 4). 4. Deep-water genetic facies models that were proposed for ideal end-member processes, such as the turbidite facies model (Bouma, 1962), the contourite facies model (Fauge`res et al., 1984), and the hyperpycnite facies model (Mulder et al., 2003), are obsolete. The popularity of genetic facies models is a manifestation of groupthink. For example, Walker (1965) proposed the concept of classifying discrete rippled beds as “Tc” under “base- and top-cut-out Bouma Sequence” categories, which allowed him to interpret rippled beds as turbidites, although contour currents can also develop rippled beds. Following Walker’s (1965) flawed convention, cored intervals from IODP Expedition 339 in the Gulf of Cadiz, with no distinct primary sedimentary structures, have been

Mass transport, gravity flows, and bottom currents

10.1 Lessons learned

443

classified as “base-cut-out” and “top-cut-out” contourites (see Fig. 8.18B and E). Our interpretations of depositional processes must always be based on what we can observe, not on what is missing. This is because process sedimentology is an observational science! It is time to bid farewell to genetic facies models of deep-water turbidites, contourites, and hyperpycnites. 5. This book has documented observational evidence from case studies, based on Amazon Fan (Equatorial Atlantic), Mississippi Fan (Gulf of Mexico), Monterey Fan (North Pacific), Krishna-Godavari Basin (Bay of Bengal), the Annot Sandstone (Eocene Oligocene, Peira-Cava Area, Maritime Alps, SE France), the Jackfork Group (Pennsylvanian, Ouachita Mountains, United States), basin-floor fans in the North Sea, upper Triassic Yanchang Formation (Ordos Basin, central China), among others, which compel a paradigm shift from turbidites to MTD and bottom-current reworked sands. All these case studies were previously considered as turbidites. 6. Satellite images have revealed that sediment (density) plumes are commonly deflected away from the normal downslope direction in 18 out of 29 cases at river mouths. These examples are Brisbane, Congo, Connecticut, Dart, Ebro, Eel, Elwha, Finesse, Guadalquivir, Krishna-Godavari, Mississippi, Morns, Rı´o de la Plata (Estuary), Pearl, Rhone, Tiber, Yellow, and Yangtze rivers. As a consequence, current directions change drastically and sediment distribution occurs on only one side of river mouths. Sediment transport is diverted by a plethora of 22 oceanographic, meteorological, and anthropogenic external factors. Empirical data show that wind forcing is the most dominant factor. Other factors are tidal currents, ocean currents, and coastal upwelling. Deflection of sediment plumes defies the conventional use of paleocurrent directions in determining sediment transport and provenance in the ancient sedimentary record. Failure to recognize deflected sediment plumes in the rock record could result in construction of erroneous depositional models with economic implications for reservoir prediction in petroleum exploration. 7. In advocating a rational theory for delta formation, Bates (1953) suggested three types of river-mouth sediment plumes: (1) hypopycnal plume for floating river water that has lower density than basin water, (2) homopycnal plume for mixing river water that has equal density as basin water, and (3) hyperpycnal plume for sinking river water that has higher density than basin water. Although Middleton and Hampton (1973) did not classify hyperpycnal flows as sediment-gravity flows in their classification, these flows are indeed sediment-gravity flows. A global evaluation of density plumes, based on 45 case studies [e.g., Yellow River, Yangtze River, Copper River, Hugli River (Ganges), Guadalquivir River, Rı´o de la Plata Estuary, Zambezi River, among others], suggests a complex variability in nature. Real-world examples show that density plumes (1) occur in six different environments (i.e., marine, lacustrine, estuarine, lagoon, bay, and reef), (2) are composed of six different compositional materials (e.g., siliciclastic, calciclastic, planktonic, etc.), (3) derive material from 11 different sources (e.g., river flood, tidal estuary, subglacial, etc.), (4) are subjected to 22 different external controls (e.g., tidal shear fronts, ocean currents, cyclones, tsunamis, etc.), and (5) exhibit 24 configurations (e.g., lobate, coalescing, linear, swirly, U-Turn, anastomosing, etc.). The major problem is that there are at least 16 types of hyperpycnal flows.

Mass transport, gravity flows, and bottom currents

444

10. Epilogue: lessons learned

Therefore the simplistic facies models of hyperpycnites, intrabasinal turbidites, and extrabasinal turbidites are obsolete (Fig. 6.16). 8. On continental margins, there are 21 possible triggers of sediment failures that generate downslope gravity flows. Short-term triggers are composed of earthquakes, meteorite impacts, volcanic activities, tsunamis, tropical cyclones, etc. Intermediateterm triggers are composed of tectonics, glacial maxima, etc. Long-term events are sealevel changes. However, the prevailing notion that deep-water deposits develop during periods of sea-level lowstands is a myth. The geologic reality is that frequent short-term events that last for only a few minutes to several hours or days (e.g., earthquakes, meteorite impacts, tsunamis, tropical cyclones, etc.) are more important in controlling deposition of deep-water sands than sporadic long-term events that last for thousands to millions of years (e.g., lowstand systems tract). 9. The four basic types of deep-marine bottom currents are: (1) thermohaline-induced geostrophic contour currents, (2) wind-driven bottom currents, (3) tide-driven bottom currents, mostly in submarine canyons, and (4) internal wave/tide-driven baroclinic currents. Contourites are deposits of thermohaline-driven geostrophic contour currents. Hybrid flows are common. Contourites can be muddy or sandy in texture, siliciclastic, or calciclastic in composition. Traction structures are common in deposits of all four types of bottom currents. However, there are no diagnostic sedimentological or seismic criteria for distinguishing ancient contourites from other three types. Double mud layers are a reliable criterion for recognizing deep-marine tidalites in cores and outcrops. The Gulf of Ca´diz is the type locality for the contourite facies model based on muddy lithofacies. However, there are no genuine contour currents in the Gulf of Cadiz. This site is affected only by transitory contour currents associated with the Mediterranean outflow water. Furthermore, this site is affected by other complicating factors such as internal waves and tides, turbidity currents, tsunamis, cyclones, mud volcanism, methane seepage, sediment supply, porewater venting, and bottom topography. Integrated Ocean Drilling Program (IODP) 339 cores from the Gulf of Ca´diz do not show well-developed primary sedimentary structures, which are necessary for interpreting depositional processes. 10. Soft-sediment deformation structures (SSDS), commonly associated with deep-water deposits, have been the focus of attention for over 150 years. Existing unconstrained definitions allow one to classify a wide range of features under the umbrella phrase “SSDS.” As a consequence, a plethora of at least 120 different types of SSDS (e.g., convolute bedding, slump folds, load casts, dish-and-pillar structures, pockmarks, raindrop imprints, explosive sand-gravel craters, clastic injections, crushed and deformed stromatolites, etc.) have been recognized in strata ranging in age from Paleoproterozoic to the present time. Two factors that control the origin of SSDS are prelithification deformation and liquidization. A sedimentological compendium of 140 case studies of SSDS worldwide, which include 30 case studies of scientific drilling at sea (Deep Sea Drilling Project/ODP/IODP), published during a period between 1863 and 2017, has yielded at least 31 different origins. Earthquakes have remained the single most dominant cause of SSDS because of the prevailing “seismite” mindset. This chapter provides a detailed account of documented case studies worldwide and

Mass transport, gravity flows, and bottom currents

10.1 Lessons learned

11. 12. 13. 14. 15.

16.

445

related problems. The remaining unresolved issue is in distinguishing SSDS formed from mass transport from those that formed by tectonics. Frequent academic debates on deep-water processes are necessary and an integral part of advancing science (Chapter 6: Hyperpycnal Flows). Despite advances made during the past century, the domain of deep-water sedimentation is still in a state of flux. In terms of future research, satellite images that are commonly ignored in studies of deep-water sedimentation are as important as other datasets in understanding sediment transport and provenance (Fig. 10.2). A summary chart of processes discussed in this book is shown in Fig. 10.3. Facts and nomenclature do matter. For example, 41 genetic terms were introduced for various deposits (Table 9.3). Of the 41 terms, only 11 are meaningful in understanding the true depositional origin (e.g., turbidities), the remaining 30 are just jargons (e.g., seismites). Transforming obstacles into opportunities in publishing research articles. I would like to end this book with an anecdotal experience on turning challenges into opportunities in publishing articles in peer-reviewed international journals. For example, in addition to my 38 published debates (Table 6.3), I also submitted a comment on a

FIGURE 10.2 (A) Index image of Mississippi River, United States. (B) Satellite image of the Mississippi River showing well-developed density plumes at river mouths. (C) An enhanced image of a dissipating density plume from the Mississippi River by Brian Romans published in his blog Clastic Detritus on April 20, 2009, using NASA’s image (additional labels by G. Shanmugam). Red dots show position of plumes. Source: (A) NOAA. (B) Image courtesy: Liam Gumley, Space Science and Engineering Center, University of Wisconsin-Madison and the MODIS science team. Image acquired on March 5, 2001. (C) Figure from Shanmugam (2018c), with permission from Elsevier.

Mass transport, gravity flows, and bottom currents

446

10. Epilogue: lessons learned

FIGURE 10.3 Summary chart showing types of processes under four major categories discussed. Arrows show superfluous types.

paper by Steel et al. (2016) to the GSA Bulletin on May 22, 2017 (MS # 831848). My comment, which dealt with hyperpycnal flows and hyperpycnites, was entitled “Highstand shelf fans: The role of buoyancy reversal in the deposition of a new type of shelf sand body: Comment .” On August 28, 2017, I contacted the GSA Bulletin office to find out the status of my manuscript. The journal office informed me that the editor-incharge (anonymous) was too busy with other matters and did not have a chance to send my manuscript out for a peer-review. Because most journals reach a decision to accept or reject in three months after submission, I promptly withdrew my manuscript from GSA from further consideration. This disappointing event was the sole incentive for me to publish a comprehensive review article entitled “The hyperpycnite problem” (Shanmugam, 2018b), which included my main points from the withdrawn manuscript. My review article, “The hyperpycnite problem,” has resulted in several offshoot publications. which include the following: 1. an article on satellite survey of density plumes by Shanmugam (2018c) 2. an article on bioturbation and trace fossils in turbidites, contourites, and hyperpycnites by Shanmugam (2018e), 3. a discussion on “The hyperpycnite problem” by Zavala (2019), 4. a discussion on “The hyperpycnite problem” by Van Loon et al. (2019), 5. an Encyclopedia chapter on “Slides, Slumps, Debris Flows, Turbidity Currents, Hyperpycnal Flows, and Bottom Currents” by Shanmugam (2019a),

Mass transport, gravity flows, and bottom currents

10.1 Lessons learned

447

6. a discussion on “Climatic and tectonic controls of lacustrine hyperpycnite origination...” by Shanmugam (2019b), 7. a “Words of the Editor-in-Chief — Some ideas about the comments and discussions of hyperpycnal flows and hyperpycnites” by Feng (2019), 8. a reply to discussions by Zavala (2019) and by Van Loon et al. (2019) on “The hyperpycnite problem” by Shanmugam (2019c), 9. an article on wind forcing of density plumes by Shanmugam (2019d), 10. Chapter 3 on gravity flows (This book), 11. Chapter 5 on density plumes (This book), and 12. Chapter 6 on hyperpycnal flows (This book). In short, we all encounter hallenges in publishing articles in one journal or another, but we also have the option to get neglected or rejected research works published elsewhere. That is the power of transforming obstacles into opportunities!

Mass transport, gravity flows, and bottom currents

A P P E N D I X

A Concepts, glossary, and methodology

This section is an expanded version of two previous publications (Shanmugam, 2012a, 2018a) Abyssal plain The term abyssal plain refers to a flat region of the ocean floor, usually at the base of a continental rise, where slope is less than 1:1000 (Heezen et al., 1959). It represents the deepest and flat part of the ocean floor that occupies between 4000 and 6500 m in the US Atlantic margin. A more general term “basin plain” is commonly used in referring to ancient examples. Examples of modern abyssal plains and their extent are shown in Fig. A.1 (Weaver et al., 1987). Agulhas Current The Agulhas Current is the western boundary current that flows down the east coast of Africa from 27 S to 40 S. It retroflects (turns back) due to interactions with the strong Antarctic Circumpolar Current. The Agulhas retroflection has been associated with generating immense “rogue waves.” Alluvial fan It is a subaerial triangular-shaped mass-transport deposit (MTD), composed mostly of debrites (i.e., deposits of debris flows). Alluvial fans develop immediately adjacent to a highland that is most commonly a major and active fault scarp (McPherson et al., 1987) (Fig. A.2). Ancient The term refers to deep-marine systems that are older than the Quaternary Period, which began approximately at 2.58 Ma. Antarctic bottom water (AABW) AABW originates in the northwest corner of the Weddell Sea in the Antarctic region by the formation of ice from surface freezing over the Antarctic continental shelves. Avalanche A large mass of snow, ice, soil, rock, or mixture of these materials moving downslope rapidly under the force of gravity. This term is not useful for interpreting ancient processes because it is difficult to quantify velocity (rapid vs slow) of processes accurately in the rock record. Back-analysis The method of determining the conditions and developing a suitable model of the slope from a failure (Duncan and Wright, 2005). Baroclinite Deposit of a baroclinic current induced by internal waves and tides (Shanmugam, 2013a). Basal shear zone The basal part of a rock unit that has been crushed and brecciated by many subparallel fractures due to shear strain. Bathyal Ocean floor that occupies depths between 200 (shelf edge) and 4000 m (656 and 13,120 ft.). Note that abyssal plains may occur at bathyal depths. Bathymetry The measurement of seafloor depth and the charting of seafloor topography. Bolide It refers to a bright fireball-like meteor. The term bolide is used synonymously with meteorite. Bottom-current reworking It refers to traction (bedload) processes associated with deep-water bottom currents. Bottom currents In deep-water environments, there are four types of bottom currents, namely (1) thermohalineinduced geostrophic bottom currents, (2) wind-driven bottom currents, (3) deep-water tidal bottom currents, and (4) internal waves and tides (baroclinic currents). These bottom currents should not be confused with turbidity currents.

449

450

Concepts, glossary, and methodology

FIGURE A.1 Sizes of modern abyssal plains. Source: Modified after Weaver, P.P.E., Thomson, J., Hunter, P.M., 1987. Introduction. In: Weaver, P.P.E., Thomson, J. eds., Geology and Geochemistry of Abyssal Plains: Geological Soc. London Special Publication 31, pp. vii xii; From Shanmugam, G., 2016d. Glossary: a supplement to “submarine fans: a critical retrospective (1950 2015).” J. Palaeogeogr. 5 (3), 258 277. Bouma Sequence refers to five divisions in a “turbidite” bed, namely Ta, Tb, Tc, Td, and Te (Bouma, 1962). Despite its popularity, this genetic facies model is fundamentally flawed (see Shanmugam, 1997a, 2016a). Brecciated clasts Angular mudstone clasts in a rock due to crushing or other deformations. Brecciated zone An interval that contains angular fragments caused by crushing or breakage of the rock. Clastic sediment Solid fragmental material (unconsolidated) that originates from weathering and is transported and deposited by air, water, ice, or other processes (e.g., mass movements). Continental margin The ocean floor that occupies between the shoreline and the abyssal plain. It consists of shelf, slope, and basin. Continental rise The seafloor that occupies between continental slope (_3000 m) and abyssal plain (_4000 m, U.S. Atlantic margin). Continental shelf The seafloor that occupies between the shoreline and the shelf-slope break ( . 200 m). Continental slope The seafloor that occupies between the shelf-slope break ( . 200 m) and the slope-rise break ( . 3000 m, U.S. Atlantic margin). Contorted bedding Extremely disorganized, crumpled, convoluted, twisted, or folded bedding. Its synonym is chaotic bedding. Contourite Deposit of thermohaline-induced geostrophic contour currents (Hollister, 1967). Core A cylindrical sample of a rock type extracted from underground or seabed. It is obtained by drilling into the subsurface with a hollow steel tube called a corer. During the downward drilling and coring, the sample

FIGURE A.2 Death Valley, California. The valley off Badwater Drive. (A) Typical debris flow dominated alluvial fan. The debris flows form narrow chutes as they flow downslope. A very recent flow is shown in the lower center of the image. The chutes form as a result of the coarsest clasts being moved to the margins of the flow as it travels downslope. These largest clasts are then left behind to form high levees that act as a wall to confine the flow. The older flows are distinguished by their reddish-brown weathering (oxidation) coating. (B) Sieve lobe fan: East side of the valley off Badwater Drive. A longitudinal profile view of a typical recent debris flow on an alluvial fan. This is a slice through the levee of a debris flow showing the characteristic poor sorting and large clasts. Notice how the largest clasts have been carried along near the top of the flow. This is because they get kicked upward by the smaller clasts as the flow moves downslope. Also notice how steep the slope angle is. This one is about 10 degrees. Source: Both photos are courtesy John. G. McPherson.

452

Concepts, glossary, and methodology

is pushed upward into the tube. After coring, the rock-filled tube is brought to the surface. In the laboratory, the core is slabbed perpendicular to bedding. Finally, the slabbed flat surface of the core is examined for geologic bedding contacts, sedimentary structures, grain-size variations, deformation, fossil content, etc. Gravity, box, piston, vibratory, percussion, and conventional cores are commonly used. The Ocean Drilling Program (ODP) uses core with a diameter of 6.8 cm. Debrite Deposit of a debris flow. Deep-lacustrine environments Deep-lacustrine environments are similar to deep-marine environments in terms of gravity-driven downslope processes and bottom currents. Deep and large lakes of the world were discussed by Herdendorf (1990). Some large lakes are remarkably deep with a maximum depth of 1637 m (Table A.1).

Lake Baikal is located in the central part of the tectonically active Baikal Rift Zone in SouthCentral Siberia. This lake, the world’s deepest (1637 m), is 636 km long and 79 km wide. This voluminous lake (23,000 km3) contains approximately 20% of the worldwide reserve of freshwater (salinity 0.76%) (Galaziy, 1993). Basin-fill deposits may be as old as late Cretaceous and up to 8000 m thick (Vanneste et al., 2001). Piston core obtained from 675 m water depth shows sandy slump (Charlet et al., 2005). Using high-resolution multibeam mapping system, a bathymetric survey of Lake Tahoe was made (Gardner et al., 2000). The lake has steep margins on the western (McKinney Bay), northern, and eastern sides, and a relatively gentle margin on the southern side. The lake is characterized by a narrow, flat nearshore zone, a steep slope that plunges more than 400 m, and a flat lake floor. On the western margin of the lake, off McKinney Bay, the slope varies from 30 to .70 . A major debris tongue, with a width of 7.5 km and a length of 9 km, occurs just immediately downdip of a scarp in the McKinney Bay area (Fig. 2.8). This debris tongue has a pronounced 5-m high arcuate toe. Downdip of the debris tongue, large debris blocks (1000 m long, 400 m wide, and 80 m high) are scattered on

TABLE A.1 The world’s deepest lakes. Rank

Lake

Location

Depth

1

Baikal

Siberia, Russia

5369 ft. (1637 m)

2

Tanganyika

Africa (Tanzania, Zaire, and Zambia)

4708 ft. (1435 m)

3

Caspian Sea

Iran and Russia

3104 ft. (946 m)

4

Nyasa

Africa (Mozambique, Tanzania, and Malawi)

2316 ft. (706 m)

5

Issyk Kul

Kyrgyzstan, Central Asia

2297 ft. (700 m)

6

Great Slave

Northwest Territories, Canada

2015 ft. (614 m)

7

Crater Lake

Oregon, United States

1943 ft. (592 m)

8

Lake Tahoe

California and Nevada, United States

1685 ft. (514 m)

9

Lake Chelan

Washington, United States

1419 ft. (433 m)

10

Great Bear

Northwest Territories, Canada

1356 ft. (413 m)

11

Lake Superior

Canada and United States

1333 ft. (406 m)

12

Titicaca

Peru

1214 ft. (370 m)

13

Pend Oreille

Idaho, United States

1150 ft. (351 m)

National Park Service, U.S. Department of the Interior. Uniform resource locator (URL): http://www.nps.gov/crla/brochures/deeplakes.htm (accessed 15.06.04.).

Concepts, glossary, and methodology

453

the flat lake floor. These blocks have been interpreted as part of a major debris avalanche. They were triggered by a failure on the western margin about 300 Ka. Also, coalescing of subaqueous debris cones (i.e., bajadas) has been observed (Gardner et al., 2000, their Fig. 13). These researchers stated, “Sediment transport processes following the debris avalanche created a series of sinuous channels that funneled sediment across the hummocky reaches of the upper debris-avalanche field toward the center of the basin. . .” The channels are 100 m wide, 1 2 m deep, and more than 3.8 km long. The exact origin of these sinuous channels is unclear. The Lucina Formation, offshore South Gabon, is composed of deep-water facies deposited in an early Cretaceous syn-rift lake (Smith, 1995). Deep-lacustrine sands in this basin constitute principal reservoirs in the Lucina and Lucina West Marine fields. Smith (1995, p. 201) recognized clean sandstone and interpreted it as “En masse deposition from high-concentration cohesionless sediment gravity flows due to intergranular friction.” This process is analogous to sandy debris flows with plastic rheology. Sammartini et al. (2019) have documented MTDs in lakes worldwide. Deep-marine sediments The term “deep marine” refers to bathyal sedimentary environments occurring in water deeper than 200 m (656 ft.), seaward of the continental shelf break, on the continental slope and the basin. The continental rise, which represents that part of the continental margin between continental slope and abyssal plain, is included under the broad term “basin.” On the slope and basin environments, sediment-gravity processes (slides, slumps, debris flows, and turbidity currents) and bottom currents are the dominant depositional mechanisms, although pelagic and hemipelagic deposition is also important. Deep-water (Fig. 2.25) The term “deep-water” refers to bathyal water depths ( . 200 m) that occur seaward of the continental shelf break on the slope and basin settings. I prefer the general term “deep-water” in order to include both marine and lacustrine environments. In the petroleum industry, the term deep-water is used with two different meanings. First, geologists use the term to denote deep-water depositional origin of the reservoir, even if the drilling for this reservoir commences from shallow-water shelf (e.g., Well A in Fig. A.3). Second, drilling engineers use the term to denote deep-water drilling depths (e.g., Well B in Fig. A.3), even if the target reservoir is of shallow-water origin (modified after Shanmugam, 2000a). Dish structures Concave-up (like a dish) structures caused by upward-escaping water in the sediment. Double mud layers (DMLs) Paired occurrence of mud layers. This is unique to tidal settings. Visser (1980) originally explained the origin of DMLs by alternating ebb and flood tidal currents with extreme time-velocity asymmetry in shallow-water subtidal settings. This feature also occurs in deep-water settings (Shanmugam, 2003a). Drained condition A condition under which water is able to flow into or out of soil in the length of time that the soil is subjected to some change in load (Duncan and Wright, 2005). ETOPO1 It is a 1 arc-minute (1/60 of one degree; 1 nautical mile 5 1852 km) global relief model of Earth’s surface that integrates land topography and ocean bathymetry (Fig. A.4). This dataset is a higher resolution version of ETOPO2, which is a 2 arc-minute global relief model of Earth’s surface. An arc-minute is 1/60 of a degree. Scientists use high-resolution maps such as ETOPO1 to improve accuracy in tsunami forecasting, modeling, and warnings, and also to enhance ocean circulation modeling and Earth visualization. For more information on the creation of the map, visit: http://sos.noaa.gov/datasets/Land/etopo1.html (accessed July 9, 2011). Floating mudstone clasts Occurrence of mudstone clasts at some distance above the basal bedding contact of a rock unit. Flow Continuous, irreversible deformation of sediment-water mixture that occurs in response to applied stress. Fluid A material that flows. Fluid dynamics A branch of fluid mechanics that deals with the study of fluids (liquids and gases) in motion. Fluid-gravity flows Middleton and Hampton (1973) distinguished sediment-gravity flows from fluid-gravity flows. In a fluid-gravity flow (e.g., river currents and some deep-ocean currents), fluid is directly driven by gravity, whereas in a sediment-gravity flow, the interstitial fluid is driven by the grains moving downslope under the influence of gravity. Fluid mechanics Study of the properties and behaviors of fluids.

454

Concepts, glossary, and methodology

FIGURE A.3 The term “deep-water” refers to bathyal water depths ( . 200 m) that occur seaward of the continental shelf break on the slope and basin settings. In petroleum exploration and production, the term deep-water is used with two different meanings. First, geologists use the term to denote deep-water depositional origin of the reservoir, even if the drilling for this reservoir commences from shallow-water shelf (e.g., Well A). Second, drilling engineers use the term to denote deep-water drilling depths (e.g., Well B), even if the target reservoir is of shallow-water origin. After Shanmugam, G., 2000a. 50 years of the turbidite paradigm (1950s 1990s): deep-water processes and facies models a critical perspective. Marine and Petroleum Geology, 17, 285 342. Reproduced with permission from Elsevier.

Geohazards Natural disasters (hazards) such as earthquakes, landslides, tsunamis, tropical cyclones, rogue (freak) waves, floods, volcanic events, sea-level rise, karst-related subsidence (sinkholes), geomagnetic storms, coastal upwelling, and deep-ocean currents. Glide plane A slip surface along which major displacement occurs, causing MTDs. Hemipelagite The term refers to deposits of hemipelagic settling of deep-sea mud in which more than 25% of the fraction coarser than 5 microns is of terrigenous, volcanogenic, and/or neritic origin. Heterolithic facies Thinly interbedded (millimeter- to decimeter-scale) sandstones and mudstones. High-density turbidity currents (HDTCs) HDTC is a euphemism for sandy debris flows (Shanmugam, 1996a). Hybrid flows According to the Cambridge Dictionary, the term "hybrid" represents the hybrid offspring byproducts of two different plants, animals, or other entities (https://dictionary.cambridge.org/dictionary/ learner-english/hybrid, accessed June 2, 2020). Accordingly, the term "hybrid flows" is defined in this book to represent the intersection of two different processes, such as alongslope bottom currents (e.g., contour currents) intersecting with downslope sediment-gravity flows (e.g., sandy debris flows, turbidity currents, etc.) in deep-water environments (e.g., continental slope) (Figs. 1.1, 8.11, 8.12A, and 8.51). Such an interaction commonly results in bottom-current-reworked sands with traction structures (Shanmugam et al., 1993a). Under such conditions, hybrid flows may generate partial "Bouma Sequence" (Fig. 8.12B). Fonnesu et al. (2020) and

455

Concepts, glossary, and methodology

FIGURE

A.4 ETOPO1-1 arc-minute global relief model of Earth’s surface that integrates land topography and ocean bathymetry. Source: Amante, C., Eakins, B.W. ETOPO1 1 arc-minute global relief model: procedures, data sources and analysis. In: NOAA Technical Memorandum NESDIS NGDC-24, 19 p., March 2009. ,http://www.ngdc.noaa.gov/ mgg/global/global.html. (accessed 26.07.11.).

FIGURE A.5 Difference between the terms "hybrid" and "flow transformation". Letters A, B, and C refer to different entities In animals, for example, a mule is the hybrid offspring of a male donkey and a female horse. By contrast, a debris flow often transforms downslope into a stratified flow with a lower debris-flow layer and an upper turbidity-current layer (see Fig. 3. 20; see also Norem et al., 1990). In other words, the concept of "hybrid" begins with two different parent species yielding a single hybrid offspring, whereas the concept of "flow transformation" begins with a single parent flow that transforms downslope into two sediment-gravity flows. In maintaining this conceptual clarity, the continued misapplication of hybrid concept for flow transformation must stop. Furthermore, although flow transformation is a real phenomenon, it cannot be recognized in the ancient sedimentary record with objective criteria (Shanmugam, 2012a).

Fuhrmann et al. (2020) have discussed the concept of hybrid flows using examples from Mozambique. By contrast, other authors have used the term "hybrid flows" for transitional flows associated with downslope flow transformation of sediment-gravity flows into stratified flows (Houghton et al., 2009; Talling, 2013; Fallgatter et al., 2017; Stow and Smillie, 2020). Such misapplications represent the opposite of the true etymological meaning of the term "Hybrid" (see Fig. A.5). In short, bottom-current reworking is the key function of hybrid flows. Hybrid flows and their deposits with traction structures are an important, but underrated, deep-water facies. Hybridite An amalgamated offspring deposit of two hydrodynamically different flow types, such as sandy debris flows and contour currents (i.e., hybrid flows). Hydrodynamics A branch of fluid dynamics that deals with the study of liquids in motion. Hyperpycnal flows In advocating a rational theory for delta formation, based on the concepts of Forel (1885), Bates (1953) suggested three major flow types: (1) hypopycnal flow for floating river water that has lower

456

Concepts, glossary, and methodology

density than basin water, (2) homopycnal flow for mixing river water that has equal density as basin water, and (3) hyperpycnal flow for sinking river water that has higher density than basin water. Hyperpycnite The term “hyperpycnite” (i.e., deposits of hyperpycnal flows) was first introduced by Mulder et al. (2002) in an academic debate with me (Shanmugam, 2002b) on the origin of inverse grading by hyperpycnal flows. The issue is still unresolved in 2020 (Shanmugam, 2018b). Injectite Injected material (usually sand) into a host rock (usually mudstone). Injections are common in igneous rocks. Internal tides Internal waves that correspond to periods of tides are called internal tides. Internal wave Internal waves are gravity waves that oscillate along the interface between two water layers of different densities (i.e., pycnocline). Gill (1982) illustrated that fluid parcels in the entire water column move together in the same direction and with the same velocity in a barotropic (surface) wave, whereas fluid parcels in shallow and deep layers of the water column move in opposite directions and with different velocities in a baroclinic (internal) wave (Shanmugam, 2013a). Inverse grading Upward increase in average grain size from the basal contact to the upper contact within a single depositional unit. Laminar flow Regular motion of fluid in parallel layers, without macroscopic mixing across the layers (Allen, 1970). Reynolds Number for laminar flow is less than 500. Debris flows are commonly laminar in state. See entry “Reynolds Number.” Lithofacies A rock unit that is distinguished from adjacent rock units based on its lithologic (i.e., physical, chemical, and biological) properties (see entry “Rock”). Liquefaction Allen (1984) used a general process term liquidization to describe mechanisms involving a change of state from solid-like to liquid-like (i.e., “quick”) in cohesionless grain mass. The two mechanisms of liquidization are liquefaction and fluidization. Liquefaction is a phenomenon commonly associated with earthquakes in which water-saturated sands behave like fluids. As seismic waves pass through water-saturated sands, void spaces (pores) between sand particles collapse, causing sediment deformation and ground failure. It occurs as a consequence of increased pore-fluid pressure. Liquefaction involves neither influx of external fluids into the grain mass nor volume change. Lobe A rounded, protruded, wide frontal part of a deposit in map view. Mass transport Mass transport represents the failure, dislodgment, and downslope movement of sediment under the influence of gravity. Continental margins provide an ideal setting for slope failure, which is the collapse of slope sediment from the shelf edge. Following a failure, the failed sediment moves downslope under the pull of gravity when the shear stress exceeds the shear strength. Since the 1970s, I have made contributions to this domain since my Ph.D. work (Shanmugam, 1978, 1990b, 2002c,d, 2008d, 2009a,c, 2010a, 2013c, 2018a, 2019a). Methodology Four methods are in use for recognizing slides, slumps, debris flows, and turbidity currents and their deposits. Method 1: Direct observations—Deep-sea diving by a diver allows direct observations of submarine mass movements. The technique has limitations in terms of diving depth and diving time. These constraints can be overcome by using a remotely operated deep submergence vehicle, which would allow observations at greater depths and for longer time. Both remotely operated vehicles and manned submersibles are used for underwater photographic and video documentation of submarine processes. Method 2: Indirect velocity calculations—A standard practice has been to calculate velocity of catastrophic submarine events based on the timing of submarine cable breaks. The best example of this method is the 1929 Grand Banks earthquake (Canada) and related cable breaks. This method is not useful for recognizing individual type of mass movement (e.g., slide vs slump). Method 3: Remote sensing technology—In the 1950s, conventional echo sounding was used to construct seafloor profiles. This was done by emitting sound pulses from a ship and by recording return echoes from the sea bottom. Today, several types of seismic profiling techniques are available depending on the desired degree of resolutions. Although popular in the petroleum industry and academia, seismic profiles cannot resolve subtle sedimentologic features that are required to distinguish turbidites from debrites. In the 1970s, the most significant progress in mapping the seafloor was made by adopting multibeam sidescan sonar survey. The SeaMARC 1 (Seafloor Mapping and Remote Characterization) system uses up to 5 km broad swath of the seafloor. The GLORIA (Geological Long-Range Inclined Asdic) system uses up

Concepts, glossary, and methodology

457

to 45 km broad swath of the seafloor. The advantage of GLORIA is that it can map an area of 27,700 km2 day21. In the 1990s, multibeam mapping systems were adopted to map the seafloor. This system utilizes hull-mounted sonar arrays that collect bathymetric soundings. The ship’s position is determined by global positioning system. Because the transducer arrays are hull-mounted, rather than towed in a vehicle behind the ship, the data are gathered with navigational accuracy of about 1 m and depth resolution of 50 cm. Two of the types of data collected are bathymetry (seafloor depth) and backscatter (data that can provide insights into the geologic makeup of the seafloor). An example is a bathymetric image of the US Pacific margin with MTDs (Fig. 2.3). The US National Geophysical Data Center (NGDC) maintains a website of bathymetric images of continental margins. Although morphological features seen on bathymetric images are useful for recognizing mass transport as a general mechanism, these images may not be useful for distinguishing slides from slumps. Such a distinction requires direct examination of the rock in detail. Method 4: Examination of the rock—Direct examination of core and outcrop is the most reliable method for recognizing individual deposits of slide, slump, debris flow, and turbidity current. This method, known as process sedimentology, is the foundation for reconstructing ancient depositional environments and for understanding sandstone petroleum reservoirs. Details of rock description Description of conventional core for the petroleum industry involves some special techniques. Thus, the philosophy and methodology of core description are reiterated here (see Shanmugam, 2006a). 1. Be accurate, be precise, and be consistent in describing the rocks. 2. Make direct observations on the core (and on the outcrop). Required information for interpreting fluid rheology, flow state, and sediment-support mechanism can be obtained only by examining the rocks directly for the presence of intricate small-scale sedimentary features. Extracting details from the rocks is a tedious and timeconsuming endeavor and there are no short-cut approaches. 3. Use slabbed cores. Unslabbed cores are prone to cause misleading observations. 4. Clean the core (and outcrop in some cases) surface. Before describing cores that have been in storage for a long period of time, scrape the surface with a knife and expose the fresh rock surface. Otherwise, surface chemical alteration and fungus growth can mimic sedimentary features. Weathered outcrops also need cleaning to expose the fresh rock surface. 5. Wet the slabbed core surface. Wetting the core surface with a sponge or washing the entire core section tends to reveal subtle sedimentary features. Use these procedures only for consolidated sandstone intervals, not for unconsolidated sands. 6. Be mindful of poor core recovery in unconsolidated sand. Carefully determine missing core intervals. 7. Calibrate core depth with log depth for each cored section. 8. Be mindful of artificial core disturbance. Poor handling of cores of semiconsolidated or unconsolidated sediment may result in artificial contortion of sediment layers. Failure to recognize this problem would result in an erroneous description and related misinterpretation. This problem can be remedied by comparing the core with core photograph taken immediately after slabbing. 9. Begin description at the stratigraphic bottom (oldest) and move upward to the top (youngest). 10. Describe the core (and outcrop) at a scale of 1:20 or in greater detail. Details are the underpinning of process sedimentology. Detailed megascopic and microscopic examinations of both the resinated 1/3 slabs and unresinated 2/3 cuts of cores should be carried out to ensure complete coverage. 11. Maintain an objective distinction between observation and interpretation. 12. Avoid using facies models during description. The purpose of describing sedimentary rocks is to seek the truth about their depositional origin, not to validate an existing facies model (e.g., the Bouma Sequence). 13. Plot details on sedimentological logs during description (Fig. A.6). Avoid the method of describing the rocks in the field using a field notebook and then transferring details onto sedimentological logs later in the office. 14. Identify depositional contacts. Establish bottom and top contacts of each depositional unit. 15. Use expanded grain-size scale on sedimentological logs (Fig. A.6). Expanded scale allows sandy intervals to be accentuated (protruded) in comparison to muddy intervals (Fig. A.7). Expanded scale is vital during calibration of wireline logs with sedimentological logs. Use Wentworth-size classes on the abscissa using either millimeter or phi scale (see Folk, 1968; Carver, 1971). Standard class divisions are: mud (,0.0625 mm), very fine

FIGURE A.6 Sedimentological log sheet with expanded grain-size scale for core and outcrop description. Note expanded grain-size scale in which each size class is given enough space for plotting grain-size variations accurately. K.B. 5 Kelly Bushing. K.B. is the journal-box insert in the rotary table of a rotary drilling rig and its upper surface is the zero-depth reference for wireline logs and other downhole measurements. Comments column 5 use this space for core-related details, such as (1) type of hole (e.g., straight vs deviated), (2) type of core liner (e.g., fiberglass), (3) type of core impregnation (e.g., resin), (4) diameter of core barrel, (5) core recovery (%), (6) type of core storage (e.g., frozen vs room temperature), (7) type of cut examined (e.g., 1/3 vs 2/3), (8) core-log shift, (9) oil staining, (10) fossils, (11) structural complications, (12) fractures, and (13) sample locations. Suggested logging scale is 1:20. Scale can be modified to suit individual’s needs based on the degree of detail desired, total thickness of core to be described, and the time available. Suggested scale for each vertical division on the log is 2 ft. or 0.5 m. An enlarged version of this log (11 3 17 in. size paper) is recommended. Source: After Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476, with permission from Elsevier.

Concepts, glossary, and methodology

459 FIGURE A.7 A graphic sedimentological column with expanded grainsize scale in the abscissa showing quantitative distribution of facies. Laminated mud facies 5 6 ft. (18%). Massive sand facies 5 34 ft. (50%). Normally graded sand facies 5 11 ft. (32%). Note sandy intervals are protruded in comparison to muddy intervals. C 5 coaly fragments; . . . 5 normally graded beds; solid black lines through massive sand represent internal glide planes (Zafiro Field, Pliocene, Equatorial Guinea). Source: After Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476, with permission from Elsevier.

sand (0.0625 0.125 mm), fine sand (0.125 0.25 mm), medium sand (0.25 0.50 mm), coarse sand (0.50 1.0 mm), very coarse sand (1.0 2.0 mm), and gravel ( . 2.0 mm). If necessary, gravel grade can be further subdivided into granule (2 4 mm), pebble (4 64 mm), cobble (64 256 mm), and boulder ( . 256 mm). Standard practice is to plot average grain size with additional remarks on maximum grain size. 16. Document grain-size scale as part of graphic lithologic columns. Graphic lithologic column and grain-size column should be combined and documented as a single column. In other words, grain-size scale should be the abscissa for lithologic columns (Fig. A.6). A separate column for grain-size plot, independent of lithologic column, is inefficient for communicating nature of grading (Fig. A.8). For example, ODP Leg 180 Initial Reports published a separate grain-size column for Site 1108 (Taylor et al., 2000). For Section 6 of Site 1108 (Fig. A.8), the description reads, “Section has normally graded and inversely graded beds.” However, it is not obvious from the graphic column the nature of grading trends. In contrast, a graphic lithologic column combined with expanded grain-size scale shows not only the nature of graded beds but also the exact position of graded bed contacts (Fig. A.7). 17. Record primary (i.e., depositional) sedimentary structures. Be familiar with basic sedimentary structures and their origin (e.g., Pettijohn and Potter, 1964; Middleton, 1965; Middleton and Bouma, 1973; Middleton and

460

Concepts, glossary, and methodology

FIGURE A.8 A published graphic sedimentological log. Section 6 at the bottom is described as having normally graded and inversely graded beds. However, it is unclear from the graphic column the exact stratigraphic positions of lower and upper contacts of graded beds. Note that the graphic lithology column does not show protruded sandy intervals (compare with Fig. A.7). These problems can be alleviated by combining graphic column with grain-size column and by using an expanded grain-size scale. Note that expanded grain-size column in Fig. A.7 is nearly four times wider than the grain-size scale in this figure. Mbsf 5 meters below seafloor [Credit: Core Description Section, ODP Leg 180, Site 1108 (Taylor et al., 2000)]. Source: After Shanmugam, G., 2006a. DeepWater Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476, with permission from Elsevier.

Concepts, glossary, and methodology

461

Southard, 1977; Reineck and Singh, 1980; Allen, 1984; Collinson and Thompson, 1982; Harms et al., 1982). In addition to depositional structures, record postdepositional features, such as sand injections. 18. Record position and fabric of outsize grains, dense grains, and mudstone clasts. 19. Identify reservoir facies in cores: This involves integration of basic reservoir lithologies (e.g., usually gravel and sand), measured reservoir properties (e.g., porosity and permeability), and wireline-log properties for a given cored interval. This also requires visual estimation of sand percent for a given. Because each reservoir facies is linked to its depositional environment, one can understand its three-dimensional geometry and connectivity. Pay special attention to grain size and sorting because texture influences porosity and permeability (Beard and Weyl, 1973). While describing cores, record changes in framework composition, cement, matrix, oil shows (Swanson, 1981), and porosity types (Shanmugam, 1985b). During core description, make sure that a petrographic microscope, thin sections of cored intervals, UV light source, and measured core porosity and permeability data of cored intervals are readily available for establishing reservoir facies. 20. Document sedimentary features with sketches and photographs. If a feature is unfamiliar or too complex to classify, sketch it in detail and describe it objectively. Photographs of core and outcrop are the permanent record because core and outcrop are liable to undergo deterioration with time. 21. Use appropriate timing for core photography. The timing of core photography is critical. Ideally, cores should be photographed immediately after they are slabbed and cleaned. Cores must be clearly marked with core depths. After photography is completed, cores should be preserved with solvent coating resin (e.g., mixture of acetone and resin) for long-term storage. 22. The importance of natural sunlight in core photography. One of the basic requirements of process sedimentology is the documentation of sedimentary features using close-up photographs. The reason is that the closer the observation the clearer the evidence for understanding the mechanics of sediment emplacement. In providing the clarity of image and documenting the intricacy of details, almost all core photographs included in this book were taken under natural sunlight with a professional scale (5 or 15 cm). In taking perspective photographs of long cored intervals under natural sunlight, forklifts were used in some cases (Fig. A.9).

FIGURE A.9 The importance of natural sunlight in core photography. Natural sunlight photography yields better results than flashlight photography. Note the author on a forklift in order to take photographs of long core intervals (arrow) under natural sunlight, September 1983, Salt Lake City, Utah. Source: After Shanmugam, G., 2012a. New perspectives on deep-water sandstones: origin, recognition, initiation, and reservoir quality. In: Handbook of Petroleum Exploration and Production, vol. 9. Elsevier, Amsterdam, 524 p., with permission from Elsevier.

462

Concepts, glossary, and methodology

FIGURE A.10 Satellites used by NASA for imaging density plumes in the world’s oceans and lakes. A. Aqua Mission. 9. Image of NASA’s Satellite Aqua, which was launched on May 4, 2002 (B). (C) Terra satellite. See text for details. Source: NASA image by Marit Jentoft-Nilsen. Credit: https://earthobservatory.nasa.gov/Features/Water/page4. php.

23. Use a professional scale. Use of coins, lens caps, hammers, toothbrushes, hats, and human beings as scale in photographs can be confusing for documenting small-scale features. 24. Use X-radiography for massive sand intervals. This technique may be helpful for resolving subtle amalgamation surfaces, internal contorted layering, and buried clasts. Formation microimager may also be useful. 25. Quantify geological data (Sorby, 1908; Griffiths, 1960). Quantification is vital for demonstrating the relative importance of depositional processes through time (Fig. A.7).

Principles of rock interpretation 1. Interpret each bed in terms of the physics of the flow. Only through bed-by-bed interpretation, an accurate quantification of facies can be made. 2. Apply James Hutton’s (1788) principle of Uniformitarianism. Emphasize modern analogs for interpreting ancient systems. 3. Apply Johannes Walther’s (1894) Law (i.e., vertical disposition of depositional facies, without erosional breaks, represents their lateral disposition of depositional environments) with restraints. This is because of frequent occurrences of internal erosional events and glide planes in mass movements. 4. Apply experimental results for interpreting processes (Allen, 1985).

Concepts, glossary, and methodology

463

FIGURE A.11 Other satellites used by NASA in studying oceans. (A) Topex/Poseidon satellite in space. (B) Graphic illustration of measurement system. Source: (A) Courtesy NASA/JPL. Caltech. http://sealevel.jpl.nasa.gov/files/ archive/missions/topex-in-space.jpg (accessed 09.08.11). (B) Courtesy NASA/JPL. http://en.wikipedia.org/wiki/File:Poseidon. graphic.jpg (accessed 11.08.11).

5. Discern and discriminate published process interpretations. Carefully evaluate each publication in terms of its data for sufficiency, relevancy, and credibility. Avoid publications that promote model-driven process interpretations. 6. Curb any compromising of process interpretations for the sake of consensus. 7. Avoid using triggering mechanisms for classifying depositional processes. 8. Avoid using seismic geometry for interpreting depositional processes of ancient systems. 9. Avoid using wireline-log motifs for interpreting depositional processes. 10. Avoid using geometry of deep-water sandbodies as evidence for interpreting depositional processes. For example, one may infer sheet-like sandbodies by correlating wireline-log motifs and then argue that these inferred sheet sands are evidence for turbidite lobes. Sand geometries may be inferred from processes, but not vice versa. 11. Interpret processes in a regional geologic context. 12. Reconstruct depositional environments based on process sedimentology. 13. Infer sand distribution, sandbody geometry, and reservoir quality based on process sedimentology. Modern The term refers to present-day deep-marine systems that are still active or that have been active since the Quaternary Period that began approximately 2.58 Ma. Mud offshoot It refers to mud drapes on ripples. NASA satellites The National Aeronautics and Space Administration (NASA) is an independent agency of the U.S. Government. NASA is the principal source of satellite imagery and other scientific information pertaining to the climate and the environment. NASA Aqua and Terra NASA’s Aqua (EOS PM-1) (Fig. A.10A) was a multinational scientific research satellite in that was in orbit around the Earth, studying the precipitation, evaporation, and cycling of water. It is the second major component of the Earth Observing System (EOS) preceded by Terra (launched in 1999) and

464

Concepts, glossary, and methodology

FIGURE A.12 A conceptual diagram showing the importance of process sedimentology. By knowing the process of origin, one might be able to predict the nature of sandbody dimension, geometry, and reservoir property away from the well bore. Source: From Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476.

followed by Aura (launched in 2004). Aqua is of importance in studying density plumes. Aqua satellite was launched on May 4, 2002, with six Earth-observing instruments: 1. the atmospheric infrared sounder, 2. the advanced microwave sounding unit, 3. the humidity sounder for Brazil, 4. the advanced microwave scanning radiometer for the EOS (AMSR-E), 5. the moderate resolution imaging spectroradiometer (MODIS), and 6. Clouds and the Earth’s radiant energy system (CERES).

In this article, most satellite images used were taken by MODIS. NASA Topex/Poseidon Topex/Poseidon satellite (Fig. A.11A) “Topex” stands for Ocean TOPography EXperiment. It was a joint satellite mission between NASA (National Aeronautics and Space Administration), the U.S. space agency, and CNES (The Centre National d’Etudes Spatiales), the French space agency, to map ocean surface topography. It was the first major oceanographic research vessel to sail into space. The U.S. portion of the mission was managed by NASA’s Jet Propulsion Laboratory, Pasadena, California, and the French portion was managed by CNES’ Toulouse Space Center, Toulouse, France.

During the period August 1992 to January 2006, the mission: • Completed nearly 62,000 orbits of Earth (1330 km above Earth). • Covered 95% of ice-free oceans (between 66_N and 66_S latitude) every10 days.

Concepts, glossary, and methodology

465

FIGURE A.13 Airgun seismic reflection profile showing two branches of rugged, slump-dominated walls of Zhemchug Canyon, Bering Sea (see Fig. A.14 for location). Note v-shaped segments. Source: After Normark, W.R. Carlson, P.R., 2003. Giant submarine canyons: is size any clue to their importance in the rock record? In: Chan, M.A., Archer, A.W. (Eds.), Extreme Depositional Environments: Mega End Members in Geologic Time. Geological Society of America, Boulder, Colorado. Special Paper 370, pp. 175 190, Geological Society of America.

• Continuously observed global ocean topography using radar altimeter (Fig. A.11B). • Measured sea surface heights with unprecedented accuracy of 3.3 cm. • Monitored effects of currents on global climate change and produced the first global views of seasonal changes of currents. • Monitored large-scale ocean features such as Rossby and Kelvin waves and studied such phenomena as El Nino˜, La Nina˚, and the Pacific Decadal Oscillation. • Mapped basin-wide current variations and provided global data to validate models of ocean circulation. This is of relevance in this book. • Mapped year-to-year changes in heat stored in the upper ocean. Produced the most accurate global maps of tides ever. This is of relevance in this book (see Shanmugam, 2012a for field examples). • Improved our knowledge of Earth’s gravity field. Following the Topex/Poseidon mission, Jason-1 (launched in 2001) and Jason-2 (launched in 2008) satellites have been measuring ocean surface topography. Courtesy: NASA/JPL/CALTECH/ OST. http://sealevel.jpl.nasa.gov/missions/topex/ (accessed August 10, 2011). Newtonian rheology Fluids with no inherent strength. These fluids, such as water, will begin to deform the moment shear stress is applied, and the deformation is linear. Nonuniform flow Spatial changes in velocity at a moment in time. Normal grading Upward decrease in average grain size from the basal contact to the upper contact within a single depositional unit composed of a single rock type. It should not contain any floating mudstone clasts or

466

Concepts, glossary, and methodology

FIGURE A.14 Location map of major submarine canyons in the world’s oceans. Source: Modified after Normark, W.R. Carlson, P.R., 2003. Giant submarine canyons: is size any clue to their importance in the rock record? In: Chan, M.A., Archer, A.W. (Eds.), Extreme Depositional Environments: Mega End Members in Geologic Time. Geological Society of America, Boulder, Colorado. Special Paper 370, pp. 175 190, Geological Society of America. Shaded relief base map: http:// www.ngdc.noaa.gov/mgg/topo/img/globeco3.gif (accessed 26.07.11).

outsized quartz granules. In turbidity currents, waning flows deposit successively finer and finer sediment, resulting in a normal grading (see “Waning flow”). Outcrop A natural exposure of the bedrock without soil capping (e.g., along river-cut subaerial canyon walls or submarine canyon walls) or an artificial exposure of the bedrock due to excavation for roads, tunnels, or quarries. Pelagite or Pelagic sediment is a fine-grained sediment that accumulates as the result of the settling of particles from suspension to the seafloor, commonly far away from the land. These particles consist primarily of either the microscopic, calcareous, or siliceous shells of phytoplankton or zooplankton; clay-size siliciclastic sediment; meteoric dust and variable amounts of volcanic ash also occur within pelagic sediments. Based upon the composition of the ooze, there are three main types of pelagic sediments: siliceous oozes, calcareous oozes, and red clays (Rothwell, 2004). Planar clast fabric Alignment of long axis of clasts parallel to bedding (i.e., horizontal). This fabric implies laminar flow at the time of deposition. Plastic rheology Some naturally occurring materials with strength will not deform until yield stress has been exceeded; once the yield stress is exceeded, the deformation is linear. Such materials with strength are considered to be Bingham-plastic. Johnson (1970) favored a Bingham-plastic rheological model for debris flows. Primary basal glide plane (or de´collement) The basal slip surface along which major displacement occurs. Process sedimentology (aptly “depositional process sedimentology”), a subdiscipline of physical sedimentology, is concerned with the detailed bed-by-bed description of siliciclastic sedimentary rocks for establishing the

Concepts, glossary, and methodology

467

FIGURE A.15 Comparison of cross sections of submarine canyons near shelf edge. Source: After Normark, W.R. Carlson, P.R., 2003. Giant submarine canyons: is size any clue to their importance in the rock record? In: Chan, M.A., Archer, A.W. (Eds.), Extreme Depositional Environments: Mega End Members in Geologic Time. Geological Society of America, Boulder, Colorado. Special Paper 370, pp. 175 190, Geological Society of America. See also Carlson and Karl (1988). link between the deposit and the physics of the depositional process. It is the foundation for reconstructing ancient depositional environments and for understanding sandstone reservoir potential (Fig. A.12). 1. To practice process sedimentology, a certain level of basic knowledge in geology, physics, chemistry, mathematics, zoology, and botany is required. 2. Undergraduate students must have had basic courses in sedimentology and stratigraphy, and a geology field camp. 3. According to Brush (1965, p. 23), a combined knowledge of basic physics, soil mechanics, and fluid mechanics is essential for interpreting the mechanics of various fluid-sediment-gravity processes. These three disciplines were a part of my curriculum. Projected clasts Upward projection of mudstone clasts above the bedding surface of host rock (e.g., sand). This feature implies freezing from a laminar flow at the time of deposition. Reynolds Number It is named after Osborne Reynolds (1842 1912), who proposed it in 1883. This is the most important dimensionless number in fluid dynamics. The Reynolds number (Re) is the ratio of inertial forces to viscous forces and is used for determining whether a flow is laminar (Re , 500) or turbulent (Re . 2000). Rheology relationship between applied shear stress and rate of shear strain in fluids. Root mean square (RMS) amplitudes RMS amplitude attribute is used commonly in the petroleum industry for distinguishing and delineating areas that are sensitive to the sand deposition.

468

Concepts, glossary, and methodology

TABLE A.2 Area of submarine canyons (from several sources). Number (Fig. A.14)

Canyon

Area (km2)

1

Zhemchug

11,350

2

Bering

30,800

3

Navarin

14,600

4

Monterey

2380

5

La Jolla

33

6

Horizon Channel

No data

7

Swatch of No Ground

9000

8

Swatch

1700

9

Amazon

2250

10

Zaire (Congo)

4470

11

Laurentian Fan Valley

No data

See cross sections of these canyons in Fig. A.15. Compiled from Normark, W.R., Carlson, P.R., 2003. Giant submarine canyons: Is size any clue to their importance in the rock record? In Chan, M.A., Archer, A.W. (Eds.), Extreme Depositional Environments: Mega End Members in Geologic Time. Geological Society of America, Boulder, Colorado. Special Paper 370, pp. 175 190.

TABLE A.3 Dimensions of selected modern submarine canyons. Modern canyon

Length [mi (km)]

1. Bering, Bering Sea

929 (1495)

42 (7.9)

6036 (1829)

2. Great Bahama, North Atlantic Ocean

140 (225)

300 (60.0)

14,060 (4285)

3. Zaire (Congo), South Atlantic Ocean

138 (222)

51 (9.5)

4023 (1219)

99 (159)

106 (20.9)

7042 (2134)

292 (470)

138 (26.2)

6035 (1829)

6. Hudson, North Atlantic Ocean

58 (93)

117 (21.9)

4023 (1219)

7. Hydrographer, North Atlantic Ocean

30 (50)

200 (37.5)

3016 (914)

8. Rhone, Mediterranean Sea

17 (28)

287 (53.9)

2013 (610)

9. La Jolla, Pacific Ocean

9 (14)

203 (38.1)

1007 (305)

10. Halawai, Pacific Ocean

7 (11)

478 (89.8)

1007 (305)

4. Pribilof, Bering Sea 5. Monterey, Pacific Ocean

Gradient [ft./mi (m/km)]

Wall relief [ft. (m)]

After Shepard, F.P., Dill, R.F., 1966. Submarine Canyons and Other Sea Valleys. Rand McNally & Co., Chicago, p. 381; Shepard, F.P., Emery, K.O., 1973. Congo submarine canyon and fan valley. AAPG Bull. 57, 1679 1691; and Carlson and Karl (1988).

Concepts, glossary, and methodology

469

FIGURE A.16 Satellite image showing the Great Bahama Canyon, NE of the Great Bahama Bank, which is an isolated carbonate platform in the North Atlantic Ocean. The Tongue of the Ocean and the Providence Channel merge to form the canyon. Moderate resolution imaging spectroradiometer (MODIS) from the Terra satellite on May 18, 2001. Source: Jacques Descloitres, MODIS Rapid Response Team, NASA/GSFC, NASA’s Visible Earth. Website link or uniform resource locator (URL): https://visibleearth.nasa.gov/data/ev254/ev25475_Bahamas.A2001138.1550.1km.jpg (downloaded November 21, 2004).

RMS amplitudes calculated for a defined stratigraphic window, may reveal the internal architecture of that window. Within a stratigraphic window, process of calculation of RMS amplitude is a combination of three steps: (1) squaring of all the amplitude values (thus making all the values positive), (2) calculating the average of the squared amplitude values, and (3) obtaining the square root of that number.

RMS maps were made between two successive proportional slices, thicknesses of which depend upon the thickness and nature of the whole interval for which the slices were made. Rock The term is used for (1) an aggregate of one or more minerals (e.g., sandstone), (2) a body of undifferentiated mineral matter (e.g., obsidian), and (3) solid organic matter (e.g., coal). Scarp A relatively straight, cliff-like face or slope of considerable linear extent, breaking the continuity of the land by failure or faulting. Scarp is an abbreviated form of the term escarpment. Scientific drilling at sea Scientific drilling at sea, which comprise cores recovered from the Deep Sea Drilling Project (DSDP: 1966 83), ODP (1993 2003), Integrated Ocean Drilling Program (IODP: 2003 13), and International Ocean Discovery Program (IODP: 2013 Present). Each ODP offers a great wealth of core data for studying deep-water sedimentation, among other domains. For example, ODP (2007) reported that during its operation, the drilling program recovered 222,430 m of core from 1797 holes. All DSDP/ODP/OPDP core

470

Concepts, glossary, and methodology

FIGURE A.17 Type 1 shelf-incising, river-associated Swatch-No-Ground Canyon adjacent to the GangesBrahmaputra River, Bay of Bengal. Note that I would reclassify this canyon as Type 1B. It is also isolated with nearest canyon about 100 km away. Source: Compiled from Harris, P. T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86, with permission from Elsevier.

FIGURE A.18

Types 2 and 3 canyons on the slope of the Gulf of Lion, northern Mediterranean Sea. They are spaced less than 10 km apart from each other. Source: Compiled from Harris, P. T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86, with permission from Elsevier.

Concepts, glossary, and methodology

471

FIGURE A.19 Submarine canyon statistics (number, percentage, minimum, maximum, mean, and standard deviation) for geomorphic properties of main canyon thalweg length (km), spacing (km) for active versus passive plate margins. Source: After Harris, P. T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86, with permission from Elsevier.

photographs are available online free and can be downloaded from the ODP web page: http://www-odp. tamu.edu/. Home: Texas A&M University, College Station, Texas. Secondary glide plane Internal slip surface within the rock unit along which minor displacement occurs. Sediment-gravity flows or sediment flows They represent sediment gravity flows. They are classified into four types based on sediment-support mechanisms: (1) turbidity current with turbulence, (2) fluidized sediment flow with upward-moving intergranular flow, (3) grain flow with grain interaction (i.e., dispersive pressure), and (4) debris flow with matrix strength (Middleton and Hampton, 1973). Although all turbidity currents are turbulent in state, not all turbulent flows are turbidity currents. For example, subaerial river currents are turbulent, but they are not turbidity currents. River currents are fluid-gravity flows in which fluid is directly driven by gravity. In sediment gravity flows, however, the interstitial fluid is driven by the grains moving downslope under the influence of gravity. Thus turbidity currents cannot operate without their entrained sediment, whereas river currents can do so. River currents are subaerial flows, whereas turbidity currents are subaqueous flows. Sediment flux Represents (1) a flowing sediment-water mixture and (2) transfer of sediment. Sediment deformation It refers to a change in the bulk shape of the aggregate of sediment (Maltman, 1994a,b). It is concerned with deformation early in the burial history. Physical processes involved are (Collinson, 1994): (1) partial loss of strength and density inversion (e.g., flame structures), (2) progressive loading of cohesive sediment (e.g., mud diapirs), (3) partial loss of strength and applied shear (e.g., slump folds), (4) liquefactioninduced upward escape of pore water (e.g., dish and pillar structures) and sediment-water mixture (e.g., sand boil and sediment injection), (5) synsedimentary faults (e.g., extensional and contractional types), (6) sediment shrinkage (e.g., subaerial desiccation cracks and subaqueous syneresis cracks), (7) sediment wetting (e.g., buckling on steep slopes of eolian dunes), and (8) compaction (e.g., reduction in the inclination of dipping surfaces). See entry “Soft-sediment deformation structures” (SSDS) and 12 classifications of SSDS (Shanmugam, 2017a). The phrases “sediment deformation” and “soft-sediment deformation” mean one and the same, although both phrases are in use. Sedimentology Scientific study of sediments (unconsolidated) and sedimentary rocks (consolidated) in terms of their description, classification, origin, and diagenesis. It is concerned with physical, chemical, and biological processes and products. This book deals with physical sedimentology and its branch, process sedimentology.

472

Concepts, glossary, and methodology

FIGURE A.20 Submarine canyon statistics (number, percentage, minimum, maximum, mean, and standard deviation) for geomorphic properties of slope (degrees), depth range (m), canyon dendricity (km), and sinuosity for active versus passive plate margins. Source: After Harris, P. T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86, with permission from Elsevier.

Seismite Seilacher (1969) introduced the generic term “seismite” for deformed beds by earthquakes. However, there are no reliable criteria for recognizing paleoseismicity (Shanmugam, 2016c). Shear strength The maximum shear stress that the soil can withstand (Duncan and Wright, 2005). Slope stability analyses A common method for calculating the slope stability is called the “limit equilibrium analyses,” which refer to the principle in which a slope is stable if the resisting forces exceed the driving forces. See Duncan and Wright (2005). Soft-sediment deformation structures Allen (1984, II, p. 343) provided an accurate account of soft-sediment deformation in terms of physics. The two factors that control the origin of SSDS are prelithification deformation and liquidization. Allen (1984) recognized the following structures as SSDS: • convolute lamination • load casts • heavy mineral sags • passively deformed beds • dish structures • folds and sand mounds • sheet slumps • imbricate structure • deformed cross-bedding

However, a recent inventory suggests a plethora of at least 120 different types of SSDS have been recognized in strata ranging in age from Paleoproterozoic to the present time (Shanmugam, 2017a).

Concepts, glossary, and methodology

473

Soil mechanics Study of the properties and behaviors of soils. Unlike fluid mechanics and solid mechanics that deal with end members, soils consist of a heterogeneous mixture of fluids (usually air and water), particles (usually clay, silt, sand, and gravel), organic matter, and gases. Submarine canyon A steep-sided valley that incises into the continental shelf and slope. V-shaped profile of submarine canyons is common (Fig. A.13), although U-shaped profiles have also been observed. Canyons serve as major conduits for sediment transport from land and the shelf to the deep-sea environment worldwide (Fig. A.14). Smaller erosional features on the continental slope are commonly termed gullies in modern environments; however, there are no standardized criteria to distinguish canyons from gullies in the rock record. Similarly, the distinction between submarine canyons and submarine erosional channels is not straightforward. Thus alternative terms such as gullies, channels, troughs, trenches, fault valleys, and sea valleys are in use for submarine canyons in the published literature. Normark and Carlson (2003) compared submarine canyons and their cross sections near the shelf edge and reported that the Zhemchug Canyon from the North American Margin of the Bering Sea has the largest cross section (Fig. A.15). Zhemchug Canyon has a volume of 5800 km3 (Carlson and Karl, 1988). The Bering Canyon has the largest area of all canyons studied (Table A.2).

The importance of mass movements in shaping large submarine canyons in the Beringian continental margin has been discussed by Carlson et al. (1991). Dimensions of selected modern submarine canyons are listed in Table A.3. The Great Bahama Canyon has the world’s highest wall relief of 14,060 ft. (4285 m) (Fig. A.16). A comprehensive study of submarine canyons worldwide (Figs. A.17 A.20) was carried out by Harris and Whiteway (2011). According to Harris and Whiteway (2011), canyons exhibit an impressive array of statistics from their length and spacing (Fig. A.19) to their slope, depth range, dendricity, and sinuosity (Fig. A.20). Active continental margins contain 44.2% of all canyons (2586) and passive margins contain 38.4% (2244). Canyons are steeper, shorter, more dendritic, and more closely spaced on active than on passive continental margins (Fig. A.21). River-associated, shelf-incising canyons are more numerous on active continental margins (n 5 119) than on passive margins (n 5 34). They are most common on the western margins of South and North America where they comprise 11.7% and 8.6% of canyons, respectively. A variety of deposits, such as slumps, debrites, tidalites, and hemipelagites, can accumulate within submarine canyons. Submarine fan The term “submarine fans” refers loosely to deposits of variable shapes and sizes in deep-marine environments. The principal elements of submarine fans are canyons, channels, and lobes (see a critical review by Shanmugam, 2016a). Total stress The total stress is the sum of all forces, including those transmitted through particle contacts and those transmitted through water pressures, divided by the total area (Duncan and Wright, 2005). Tropical cyclone It is a meteorologic phenomenon characterized by a closed circulation system around a center of low pressure, driven by heat energy released as moist air drawn in over warm ocean waters rises and condenses. Structurally, it is a large, rotating system of clouds, wind, and thunderstorms (Shanmugam, 2008a). The name underscores their origin in the tropics and their cyclonic nature. Worldwide, formation of tropical cyclones peaks in late summer months when water temperatures are warmest. In the Bay of Bengal, tropical cyclone activity has double peaks, one in April and May before the onset of the monsoon and another in October and November just after. Cyclone is a broader category that includes both storms and hurricanes as members. Cyclones in the Northern Hemisphere represent closed counterclockwise circulation. They are classified based on maximum sustained wind velocity as follows: • tropical depression: 37 61 km h21 • tropical storm: 62 119 km h21 • tropical hurricane (Atlantic Ocean): .119 km h21 • tropical typhoon (Pacific or Indian Ocean): .119 km h21

FIGURE A.21 Schematic 3D diagram contrasting the geomorphic attributes characteristic of submarine canyons occurring on passive and active continental margin types. Source: After Harris, P. T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86, with permission from Elsevier.

Concepts, glossary, and methodology

475

The Saffir Simpson Hurricane Scale • • • • •

Category Category Category Category Category

1: 120 153 km h21 2: 154 177 km h21 3: 178 209 km h21 4: 210 249 km h21 5: .249 km h21

Tsunami Oceanographic phenomena that are characterized by a water wave or series of waves with long wavelengths and long periods. They are caused by an impulsive vertical displacement of the body of water by earthquakes, “landslides,” volcanic explosions, or extraterrestrial (meteorite) impacts. The link between tsunamis and sediment flux in the world’s oceans involves four stages (Shanmugam, 2006b): (1) triggering stage, (2) tsunami stage, (3) transformation stage, and (4) depositional stage. During the triggering stage, earthquakes, volcanic explosions, undersea “landslides,” and meteorite impacts can trigger displacement of the sea surface, causing tsunami waves. During the tsunami stage, tsunami waves carry energy traveling through the water, but these waves do not move the water. The incoming wave is depleted in entrained sediment. This stage is one of energy transfer, and it does not involve sediment transport. During the transformation stage, the incoming tsunami waves tend to erode and incorporate sediment into waves near the coast. This sediment entrainment process transforms sediment-depleted waves into outgoing mass-transport processes and sediment flows. During the depositional stage, deposition from slides, slumps, debris flows, and turbidity currents would occur. Turbidite Deposit of a turbidity current with normal grading. Turbidite myths (Shanmugam, 2002a) During the past 60 years, the turbidite paradigm has promoted many myths related to deep-water turbidite deposition. John E. Sanders (1965), a pioneering process sedimentologist, first uncovered many of these turbidite myths. This paper provides a reality check by undoing 10 of these turbidite myths. Myth No. 1: Turbidity currents are nonturbulent flows with multiple sediment-support mechanisms. Reality: turbidity currents are turbulent flows in which turbulence is the principal sediment-support mechanism. Myth No. 2: Turbidites are deposits of debris flows, grain flows, fluidized flows, and turbidity currents. Reality: Turbidites are the exclusive deposits of turbidity currents. Myth No. 3: Turbidity currents are high-velocity flows and therefore, they elude documentation. Reality: Turbidity currents operate under a wide range of velocity conditions. Myth No. 4: HDTCs are true turbidity currents. Reality: Ph. H. Kuenen (1950a) introduced the concept of “turbidity currents of high density” based on experimental debris flows, not turbidity currents. HDTCs are sandy debris flows. Myth No. 5: Slurry flows are HDTCs. Reality: Slurry flows are debris flows. Myth No. 6: Flute structures are indicative of turbidite deposition. Reality: Flute structures are indicative only of flow erosion, not deposition. Myth No. 7: Normal grading is a product of multiple depositional events. Reality: Normal grading is the product of a single depositional event. Myth No. 8: Crossbedding is a product of turbidity currents. Reality: Cross-bedding is a product of traction deposition from bottom currents. Myth No. 9: Turbidite facies models are useful tools for interpreting deposits of turbidity currents. Reality: A reexamination of the Annot Sandstone in SE France, which served as the basis for developing the first turbidite facies model, suggests a complex depositional origin by plastic flows and bottom currents. Myth No. 10: Turbidite facies can be interpreted using seismic facies and geometries. Reality: Individual turbidity-current depositional events, commonly centimeters to decimeters in thickness, cannot be resolved in seismic data. All turbidite myths promote falsehood and should be abandoned. Turbulent flow Irregular fluid motion with macroscopic mixing across fluid layers. For Newtonian fluids, the criterion for initiation of turbulence is the Reynolds Number (Re), which is greater than 2000. Turbidity currents are invariably turbulent in state. Undrained condition A condition under which there is no flow of water into or out of soil in the length of time that the soil is subjected to some change in load (Duncan and Wright, 2005). Uniformitarianism The present is the key to the past (Fig. A.22).

476

Concepts, glossary, and methodology

FIGURE A.22 Diagram illustrating the doctrine of uniformitarianiism. The outer red circle represents critical knowledge sources of modern systems. The inner black circle represents the past (i.e, the ancient rock record). A comprehensive knowledge of modern systems is imperative in interpreting (i.e., vertical blue arrow) the ancient rock record.

Unsteady flow Temporal changes in velocity through a fixed point in space. Waning flow Unsteady flow in which velocity becomes slower and slower at a fixed point through time. As a result, waning flows would deposit successively finer and finer sediment, resulting in a normal grading (Kneller, 1995). Waxing flow Unsteady flow in which velocity becomes faster and faster at a fixed point through time.

A P P E N D I X

B Video of flume experiments on Sandy debris flows

You Tube URL site https://youtu.be/uMO7jffZwK0 This video can be seen at https://youtu.be/uMO7jffZwK0 Introduction Details of the experiments are given in Section 3.4.4.1. Additional details are published in Shanmugam (2000a) and Marr et al. (2001). Length of video: 19 minutes Location of experiments: St. Anthony Falls Laboratory (SAFL), University of Minnesota, Minneapolis, Minnesota Period: 1996 98 Funding Institutions: Mobil Oil Company, Dallas, Texas, and Office of Naval Research Mobil Scientist: G. Shanmugam Project Director: Prof. Gary Parker Student researchers: Jeff Marr and Peter Harff Video production: Jeff Marr

B.1 Composition of slurries used in experiments Sediment slurries were composed of silica sand (120 µm size), clay (bentonite or kaolinite), coal slag (same bulk density as silica sand: 2.6 g cm23), and water (Fig. 3.17). Coal slag of 500 µm size (coarse sand) was used as a tracer material to establish flow behavior and depositional pattern of coarse-grained grains in comparison to very fine-grained sandy matrix. Sandy debris flows were generated with bentonite clay content as low as 0.5% by weight or with kaolinite clay as low as 5% by weight. Sandy debris flows were also generated using medium-grained sand (300 µm size) with bentonite clay content as low as 1.5% by weight or kaolinite clay as low as 5% by weight.

477

478

Video of flume experiments on Sandy debris flows

B.2 Video content Each experimental run is shown twice, which allows the viewer to observe physical attributes of the flow in slow motion. YouTube: 19-minute video starts here: https://youtu.be/uMO7jffZwK0 Running Time in Minutes 00:51: 00.00

Front view of St. Anthony Falls Laboratory

00.36

Title of experiments

00.51

Researchers: Jeff Marr, Peter Harf, G. Shanmugam, and Gary Parker

01.24

Objectives

02.07

Definitions of flow types (Phase 1 and 2)

02.22

Four types of gravity flows

02.25

Turbidity currents

03.43

Weak debris flow

04.25

Moderate debris flow

05.51

Strong debris flow

B.3 Deposits 05.51

Normal grading

07.35

Weak to moderate Type 1

08.45

Weak to moderate Type 2

09.40

Weak to moderate Type 3

Strong flow deposits 10.40

Folded front

11.30

Detached head

12.15

Detached tension blocks (Note isolated blocks behind the flow head)

12.38

Sand volcanoes, water escape

13.45

Detached blocks

14.30

Collapse of snout

16.07

Hydroplaning of head

18.07

Hydroplaning of head

19.00

End of video

Bibliography ´ balos, B., Elorza, J., 2011. Latest Cretaceous cone-in-cone structures and soft-sediment deformation (BasqueA Cantabrian Basin, north Spain): a record of deep-marine paleoseismicity? GSA Bull. 123, 427 438. Ablay, G., Hu¨rlimann, M., 2000. Evolution of the north flank of Tenerife by recurrent giant landslides. J. Volcanol. Geotherm. Res. 103, 135 159. Abreu, V., Sullivan, M., Pirmez, C., Mohrig, D., 2003. Lateral accretion packages (LAPs): an important reservoir element in deep water sinuous channels. Mar. Pet. Geol. 20, 631 648. Acha, E., 2008. An overview of physical and ecological processes in the Rı´o de la Plata Estuary. Cont. Shelf Res. 28 (13), 1579 1588. Ackermann, R.V., Schlische, R.W., Olsen, P.E., 1995. Synsedimentary collapse of portions of the lower Blomidon Formation (Late Triassic), Fundy rift basin, Nova Scotia. Can. J. Earth Sci. 32, 1965 1976. Agnon, A., Migowski, C., Marco, S., 2006. Intraclast breccias in laminated sequences reviewed: recorders of paleoearthquakes. In: Enzel, Y., Agnon, A., Stein, M. (Eds.), Frontiers in Dead Sea Paleoenvironmental Research. GSA Special Paper, 401, pp. 195 214. Ager, D.V., 1974. Storm deposits in the Jurassic of the Moroccan High Atlas. Palaeogeogr. Palaeoclimatol. Palaeoecol. 15, 83 93. Akers, C., Schedl, A.D., Mundy, L., 2012. What caused the landslides in Valles Marineris, Mars? In: 43rd Lunar and Planetary Science Conference, The Woodlands, TX, 19 23 March 2012. ,http://search.aol.com/aol/ search?q 5 What%20caused%20the%20landslides%20in%20Valles%20Marineris%2C%20Mars% 3F&s_it 5 keyword_rollover&ie 5 UTF-8&VR 5 3430. (accessed 27.12.14.). Akhurst, M.A., 1991. Aspects of late Quaternary sedimentation in the Faeroe-Shetland Channel, northwest U.K. continental margin. British Geological Survey Technical Report. WB/91/2. Akhmediev, N., Pelinovsky, E., 2010. Editorial Introductory remarks on “discussion & debate: Rogue waves towards a unifying concept? Eur. Phys. J. Special Top 185, 1 4. Available from: https://doi.org/10.1140/ epjst/e2010-01233-0. Alfaro, P., Moretti, M., Sofia, J.M., 1997. Soft-sediment deformation structures induced by earthquakes (seismites) in Pliocene lacustrine deposits (Guadix-Baza Basin, Central Betic Cordillera). Eclogae Geol. Helv. 90, 531 540. Alfaro, P., Delgado, J., Este´vez, A., Molina, J.M., Moretti, M., Soria, J.M., 2002. Liquefaction and fluidization structures in Messinian storm deposits (Bajo Segura Basin, Betic Cordillera, southern Spain). Int. J. Earth Sci. (Geologische Rundsch.) 91, 505 513. Alfaro, P., Moretti, M., Owen, G., (Eds.), 2016. The environmental significance of soft-sediment deformation. Sediment. Geol., 344, iii iv, 1 408. Allen, J.R.L., 1965. Late Quaternary Niger Delta, and adjacent areas: sedimentary environments and lithofacies. AAPG Bull. 49, 547 600. Allen, J.R.L., 1970. Physical processes of sedimentation. Elsevier, New York, p. 239. Allen, J.R.L., 1982. Mud drapes in sand-wave deposits: a physical model with application to the Folkstone Beds (early Cretaceous, Southeast England). Proc. R. Soc. London, Ser. A 306, 291 345. Allen, J.R.L., 1984. Sedimentary Structures, Their Character and Physical Basis, Unabridged One-Volume. Elsevier, Amsterdam, I, p. 593, and II, p. 663. Allen, J.R.L., 1985. Loose-boundary hydraulics and fluid mechanics: selected advances since 1961. In: Brenchley, P.J., Williams, P.J. (Eds.), Sedimentology: Recent Developments and Applied Aspects. Published for the Geological Society by Blackwell Scientific Publications, Oxford, pp. 7 28. Allen, S.E., Durrieu de Madron, X., 2009. A review of the role of submarine canyons in deep-ocean exchange with the shelf. Ocean. Sci. 5, 607 620. Allison, M.A., Sheremet, A., Goni, M.A., Stone, G.W., 2005. Storm layer deposition on the Mississippi Atchafalaya subaqueous delta generated by Hurricane Lili in 2002. Cont. Shelf Res. 25, 2213 2232.

479

480

Bibliography

Almagor, G., Wiseman, G., 1982. Submarine slumping and mass movements on the slope of Israel. In: Saxov, S., Nieuwenhuis, J.K. (Eds.), Marine Slides and Other Mass Movements. Plenum Press, New York and London, pp. 95 128. Alsop, G.I., Marco, S., 2013. Seismogenic slump folds formed by gravity-driven tectonics down a negligible subaqueous slope. Tectonophysics 605, 48 69. Alsop, G. I., Weinberger, R., Marco, S. & Levi, T. 2019. Folding during soft-sediment deformation. In: Bond, C. E. & Lebit, H. D. (eds), Folding and fracturing of rocks: 50 years of research since the seminal text book of J. G. Ramsay. Geological Society, London, Special Publications, 487, 81 104. https://doi.org/10.1144/SP487.1. Alsop, G.I., Marco, S., Weinberger, R., Levi, T., 2016. Sedimentary and structural controls on seismogenic slumping within Mass Transport Deposits from the Dead Sea Basin. Sediment. Geol. 344, 71 90. Alvarado, A., 2006. Updates on MMS (Minerals Management Service) regulatory issues for offshore operators. ,http://www.southerngas.org/EVENTS/documents/SGAOGO06Alvarado.pdf. (accessed 13.03.07.). Alvarado-Bustos, R., 2011. Mixing in the Continental Slope: Study Case Gulf of Cadiz (Ph.D. thesis). University of Liverpool, Liverpool, 138 p. Alves, J., Carton, X., Ambar, I., 2011. Hydrological structure, circulation and water mass transport in the Gulf of Cadiz. Int. J. Geosci. 2, 432 456. Ambar, I., Howe, M.R., 1979. Observations of the Mediterranean outflow I. Mixing in the Mediterranean outflow. Deep-Sea Res. 26A, 535 554. Ambar, I., Serra, N., Brogueira, M.J., Cabecadas, G., Abrantes, F., Freitas, P., et al., 2002. Physical, chemical and sedimentological aspects of the Mediterranean outflow off Iberia. Deep-Sea Res. II 49, 4163 4177. Available from: https://doi.org/10.1016/S0967-0645(02)00148-0. Amos, A.F., Gordon, A.L., Schneider, E.D., 1971. Water masses and circulation patterns in the region of the BlakeBahama Outer Ridge. Deep-Sea Res. 18, 145 165. Amy, L.A., Talling, P.J., Peakall, J., Wynn, R.B., Arzola Thynne, R.G., 2005. Bed geometry used to test recognition criteria of turbidites and (sandy) debrites. Sediment. Geol. 179, 163 174. Andersen, J.C.Ø., Rasmussen, H., Nielsen, T.F.D., Rønsbo, J.G., 1998. The Triple Group and the Platinova gold and palladium reefs in the Skaergaard intrusion: stratigraphic and petrographic relations. Econ. Geol. 93, 488 509. Anderson, J.G., 1906. Solifluction, a component of subaerial denudation. J. Geol. 14, 91 112. Anderson, S.A., Riemer, M.F., 1995. Collapse of saturated soil due to reduction in confinement. J. Geotech. Eng. 121 (2), 216 220. Anderson, S.A., Sitar, N., 1995. Analysis of rainfall-induced debris flows. J. Geotech. Eng. 121 (7), 544 552. Andreetta, R., 2009. Sedimentological Analysis of Deep Water, Upslopemigrating Cross-Bedded Deposits in a Distally Steepened Carbonate Ramp (Menorca, Balearic Islands, Spain) (Ph.D. thesis). Universita` degli studi di Ferrara, Italy, 194 p. Andresen, A., Bjerrum, L., 1967. Slides in subaqueous slopes in loose sand and silt. In: Richards, A.F. (Ed.), Marine Geotechnique. University of Illinois Press, Urbana, pp. 221 239. Apel, J.R., 1987. Principles of Ocean Physics, International Geophysics 38. Academic Press, London, 63 p. Apel, J.R., 2000. Solitons near Gibraltar: views from the European remote sensing satellites. Report GOA 2000 1, Global Ocean Associates, Silver Spring, MD. Apel, J.R., 2002. Oceanic internal waves and solitons. In: Jackson, C.R., Apel, J.R. (Eds.), An Atlas of Oceanic Internal Solitary-Like Waves and Their Properties (May 2002), by Global Ocean Associates, prepared for Office of Naval Research Code 322PO, pp. 1 40 ,http://www.internalwaveatlas.com/Atlas_PDF/ IWAtlas_Pg001_Introduction.PDF. (accessed 14.05.12.). Apel, J.R., Ostrovsky, L.A., Stepanyants, Y.A., Lynch, J.F., 2006. Internal solitons in the ocean. Woods Hole Oceanographic Institution Technical Report WHOI-2006-04, Woods Hole Oceanographic Institution, Woods Hole, MA, 02543, 109 p. Archer, A.W., 1998. Hierarchy of controls on cyclic rhythmite deposition: carboniferous basins of eastern and mid-continental U.S.A. In: Alexander, C.R., Davis, R.A., Henry, V.J. (Eds.), Tidalites: Processes and Products, 61. SEPM Special Publication, pp. 59 68. Armi, L., Farmer, D.M., 1988. The flow of Mediterranean water through the Strait of Gibraltar. Prog. Oceanogr. 21, 1 105. Arnau, P., Liquete, C., Canals, M., 2004. River mouth plume events and their dispersal in the northwestern Mediterranean Sea. Oceanography 17 (3), 23 31.

Bibliography

481

Arnott, R.W.C., Hand, B.M., 1989. Bedforms, primary structures and grain fabric in the presence of suspended sediment rain. J. Sediment. Petrol. 59, 1062 1069. Arrhenius, G., 1963. Pelagic sediments. In: Hill, M.N. (Ed.), The Sea, vol 3. Wiley, New York, pp. 655 727. Arribas, J., Johnsson, M.J., Critelli, S. (Eds.), 2007. Sedimentary provenance and petrogenesis: perspectives from petrography and geochemistry. GSA Special Paper 379. Assireu, A.T., Alcaˆntara, E., Novo, E.M.L.M., Roland, F., Pacheco, F.S., Stech, J.L., et al., 2011. Hydro-physical processes at the plunge point: an analysis using satellite and in situ data. Hydrol. Earth Syst. Sci. 15 (12), 3689 3700. Baas, J.H., Kesteren, W.V., Postma, G., 2004. Deposits of depletive high-density turbidity currents: a flume analogue of bed geometry, structure and texture. Sedimentology 51, 1053 1088. Bachmann, G.H., Aref, M.A.M., 2005. A seismite in Triassic gypsum deposits (Grabfeld Formation, Ladinian), southwestern Germany. Sediment. Geol. 180, 75 89. Ba´denas, B., Aurell, M., 2001. Proximal-distal facies relationships and sedimentary processes in a stormdominated carbonate ramp (Kimmeridgian, northern Iberian Basin). Sediment. Geol. 139, 319 340. Ba´denas, B., Pomar, L., Aurell, M., Morsilli, B., 2012. A facies model for internalites (internal wave deposits) on a gently sloping carbonate ramp (Upper Jurassic, Ricla, NE Spain). Sediment. Geol. Available from: https:// doi.org/10.1016/j.sedgeo.2012.05.020. Bagnold, R.A., 1954. Experiments on a gravity free dispersion of large solid spheres in a Newtonian fluid under shear. Proc. R. Soc. Lond. Ser. A 225, 49 63. Bagnold, R.A., 1956. The flow of cohesionless grains in fluids. Phil. Trans. R. Soc. Lond. Ser. A. Math. Phys. Sci. 249, 235 297. Bagnold, R.A., 1962. Auto-suspension of transported sediment. Proc. R. Soc. Lond. Ser. A 265, 315 319. Balasubramanian, T., Ajmal Khan, S., 2002. Estuaries of India. Environmental Information System Centre, Centre of Advanced Study in marine biology, Annamalai University, Parangipettai, sponsored by the Ministry of Environment & Forests, Government of India, 195. New Delhi: ENVIS Publication Series: 1/2002. Baltzer, A., 1875. “Uber einen neuerlichen felssturz am Rossberg, negst einigen allgemeinen Bermerkungen uber derartige Erscheinungen in den Alpen,” (On a recent rockfall on Rossberg with a few observations on these phenomena in the Alps). Neues Jahrb. fur Mineralogie, Geologie und Palaeontologie 914 924. Banerjee, I., 1989. Tidal sand sheet origin of the transgressive Basal Colorado Sandstone (Albian): a subsurface study of the Cessford Field, Southern Alberta. Bull. Can. Petrol. Geol. 37, 1 17. Barnard, P.L., Hanes, D.M., Rubin, D.M., Kvitek, R.G., 2006. Giant sand waves at the mouth of San Francisco Bay. EOS Trans. Am. Geophys. Union 87 (29), 285, 289. Barnard, P.L., Hansen, J.E., Erikson, L.H., 2012. Synthesis study of an erosion hot spot, Ocean Beach, California. J. Coast. Res. 28 (4), 903 922. Barnes, N.E., Normark, W.R., 1985. Diagnostic parameters for comparing modern submarine fans and ancient turbidite systems. In: Bouma, A.H., Normark, W.R., Barnes, N.E. (Eds.), Submarine Fans and Related Turbidite Systems. Springer-Verlag, New York, pp. 13 14. and wall chart. Baringer, M.O., Price, J.F., 1997. Mixing and spreading of the Mediterranean outflow. J. Phys. Oceanogr. 27, 1654 1677. Baringer, M.O., Price, J.F., 1999. A review of the physical oceanography of the Mediterranean outflow. Mar. Geol. 155, 63 82. Barrett, J.R., 1965. Subsurface currents off Cape Hatteras. Deep-Sea Res. 12, 173 184. ˇ ´ n, G., Herber, B., et al., 2009/2010. Barton, R., Bird, K., Herna´ndez, J.G., Grajales-Nishimura, J.M., Murillo-Muneto High impact reservoirs. Oilfield Rev. 21 (4), 14 29. Basilici, G., 1997. Sedimentary facies in an extensional and deep-lacustrine depositional system: the Pliocene Tiberino Basin, Central Italy. Sediment. Geol. 109, 73 94. Basilone, L., Lena, G., Gasparo-Morticelli, M., 2014. Synsedimentary-tectonic, soft-sediment deformation and volcanism in the rifted Tethyan margin from the Upper Triassic Middle Jurassic deep-water carbonates in Central Sicily. Sediment. Geol. 308, 63 79. Basilone, L., Sulli, A., Gasparo Morticelli, M., 2016. The relationships between soft sediment deformation structures and synsedimentary extensional tectonics in Upper Triassic deep-water carbonate succession (Southern Tethyan rifted continental margin, Central Sicily). Sediment. Geol. 344, 310 322. Bates, C.C., 1953. Rational theory of delta formation. AAPG Bull. 37, 2119 2162. Bates, R.L., Jackson, J.A., 1980. Glossary of Geology, second ed. American Geological Institute, Falls Church, VA, p. 751.

482

Bibliography

Bea, R.G., Bernard, H.A., Arnold, P., Doyle, E.H., 1975. Soil movements and forces developed by wave-induced slides in the Mississippi Delta. J. Petroleum Technol. 27, 500 514. Bea, R.G., Wright, S.G., Sicar, P., Niedoroda, A.W., 1983. Wave induced slides in South Pass Block 70, Mississippi delta. J. Geotech. Eng. 109, 619 644. Beard, D.C., Weyl, P.K., 1973. Influence of texture on porosity and permeability of unconsolidated sand. AAPG Bull. 57, 349 369. Beaulieu, S., Baldwin, R., 1998. Temporal variability in currents and the benthic boundary layer at an abyssal station off central California. Deep-Sea Res. II 45, 587 615. Beaubouef, R.T., Abreu, V., 2010. MTCs of the Brazos-Trinity slope system; thoughts on the sequence stratigraphy of MTCs and their possible roles in shaping hydrocarbon traps. In: Mosher, D., et al., (Eds.), Submarine Mass Movements and Their Consequences, 28. Advances in Natural and Technological Hazards Research, pp. 475 490. Beck, C., 2009. Lake sediments as Late Quaternary palaeoseismic archives: examples in north-western Alps and clues for earthquake-origin assessment of sedimentary disturbances. Earth Sci. Rev. 96, 327 344. Beckers, A., Beck, C., Hubert-Ferrari, A., Tripsanas, E., Crouzet, C., Sakellariou, D., et al., 2016. Influence of bottom currents on the sedimentary processes at the western tip of the Gulf of Corinth, Greece. Mar. Geol. 378, 312 332. Behrmann, J.H., Peter, B. Flemings, Ce´dric, M. John, the IODP Expedition 308 Scientists, 2006. Rapid Sedimentation, Overpressure, and Focused Fluid Flow, Gulf of Mexico Continental Margin. Scientific Drilling 1 17. Available from: https://doi.org/10.2204/iodp.sd.3.03.2006. September 3. Beicher, R.J., 2000. Physics for Scientists and Engineers. Saunders College, Orlando, FL. Bein, A., Weiler, Y., 1976. The Cretaceous Talme Yafe Formation: a contour current shaped sedimentary prism of calcareous detritus at the continental margin of the Arabian craton. Sedimentology 23, 511 532. Belderson, R.H., Johnson, M.A., Kenyon, N.H., 1982. Bedforms. In: Stride, A.H. (Ed.), Offshore Tidal Sands. Chapman & Hall, London, pp. 27 57. Berggren, W.A., Hollister, C.D., 1977. Plate tectonics and paleocirculation commotion in the ocean. Tectonophysics 38, 11 48. Berra, F., Felletti, F., 2011. Syndepositional tectonics recorded by soft-sediment deformation and liquefaction structures (continental Lower Permian sediments, Southern Alps, Northern Italy): stratigraphic significance. Sediment. Geol. 235, 249 263. Beverage, J.P., Culbertson, J.K., 1964. Hyperconcentration of suspended sediment. Proc. Am. Soc. Civ. Eng. J. Hydraul. Div. 90, 117 128. Bhattacharya, H.N., Bandyopadhyay, S., 1998. Seismites in a Proterozoic tidal succession, Singhbhum, Bihar, India. Sediment. Geol. 119, 239 252. Bhattacharya, H.N., Mukherjee, A., 2020. A. Soft-sediment deformation structures in a Permo-carboniferous glacio-marine setting, Talchir Formation, Dudhi Nala, India. J Earth Syst Sci. 129 (87). Available from: https:// doi.org/10.1007/s12040-020-1357-5. Bhattacharya, J., MacEachern, J., 2009. Hyperpycnal rivers and prodeltaic shelves in the cretaceous seaway of North America. J. Sediment. Res. 79, 184 209. Bhakuni, S.S., Luirei, K., Devi, M., 2012. Soft-sediment deformation structures (seismites) in Middle Siwalik sediments of Arunachal Pradesh, NE Himalaya. Himal. Geol. 33, 139 145. Bi, N.S., Yang, Z.S., Wang, H.J., Hu, B.Q., Ji, Y.J., 2010. Sediment dispersion pattern off the present Huanghe (Yellow River) subdelta and its dynamic mechanism during normal river discharge period. Estuarine Coast. Shelf Sci. 86 (3), 352 362. Boebel, O., Rae, C.D., Garzoli, S., Lutjeharms, J., Richardson, P., Rossby, T., et al., 1998. Float experiment studies interocean exchanges at the tip of Africa. EOS 79 (1), 7 8. Boggs Jr., S., 2001. Principles of Sedimentology and Stratigraphy, third ed. Prentice Hall, Upper Saddle River, NJ, p. 726. Bolt, B.A., Horn, W.L., Macdonald, G.A., Scott, R.F., 1975. Geological Hazards. Springer-Verlag, New York, p. 328. Bond, C. E. & Lebit, H. D. (eds) 2020. Folding and Fracturing of Rocks: 50 Years of Research since the Seminal Text Book of J. G. Ramsay. Geological Society, London, Special Publications, 487. Booth, J.S., O’Leary, D.W., Popenoe, P., Danforth, W.W., 1993. U.S. Atlantic continental slope landslides: their distribution, general attributes, and implications. In: Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), Submarine

Bibliography

483

Landslides: Selected Studies in the U.S. Exclusive Economic Zone, 2002. U.S. Geological Survey Bulletin, pp. 14 22. Bornhold, B.D., Fine, I.V., Rabinovich, A.B., Thomson, R.E., Kulikov, E.A., 2003. The Grand Banks landslidegenerated tsunami of November 18, 1929: analysis and numerical modeling. Geophys. Res. Abstr. 5, 01775. Bouma, A.H., 1962. Sedimentology of Some Flysch Deposits: A Graphic Approach to Facies Interpretation. Elsevier, Amsterdam, p. 168. Bouma, A.H., 1975. Sedimentary structures of Philippine Sea and sea of Japan sediments. DSDP Leg. 31. Available from: https://doi.org/10.2973/dsdp.proc.31.114.1975. Bouma, A.H., Brouwer, A. (Eds.), 1964. Turbidites: Developments in Sedimentology 3. Elsevier, Amsterdam. Bouma, A.H., Hollister, C.D., 1973. Deep ocean basin sedimentation. In: Middleton, G.V., Bouma, A.H. (Eds.), Turbidites and Deep-Water Sedimentation, SEPM Pacific Section Short Course, Anaheim, CA, pp. 79 118. Bouma, A.H., Hampton, M.A., Orlando, R.C., 1977. Sand waves and other bedforms in Lower Cook Inlet, Alaska, Marine Geotechnol., 2. pp. 291 308. Bouma, A.H., Coleman, J.M., 1985. Mississippi fan, Leg 96 program and principal results. DSDP Leg 96 shipboard Scientists In: Bouma, A.H., Normark, W.R., Barnes, N.E. (Eds.), Submarine Fans and Related Turbidite Systems. Springer-Verlag, New York, pp. 247 252. Bouma, A.H., Normark, W.R., Barnes, N.E. (Eds.), 1985. Submarine Fans and Related Turbidite Systems. SpringerVerlag, New York. Bouma, A.H., DeVries, M.B., Stone, C.G., 1997. Reinterpretation of depositional processes in a classic flysch sequence (Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma, Discussion. AAPG Bull. 81, 470 472. Bouma, A.H., Stelting, C.E., Coleman, J.M., 1985b. Mississippi fan, Gulf of Mexico. In: Bouma, A.H., Normark, W.R., Barnes, N.E. (Eds.), Submarine Fans and Related Turbidite Systems. Springer-Verlag, New York, pp. 143 150. Bourgeois, J., Hansen, T.A., Wiberg, P.L., Kauffman, E.G., 1988. A tsunami deposit at the Cretaceous-Tertiary boundary in Texas. Science 241, 567 570. Bornhold, B.D., Fine, I.V., Rabinovich, A.B., Thomson, R.E., Kulikov, E.A., 2003. The Grand Banks landslidegenerated tsunami of November 18, 1929: analysis and numerical modeling. Geophys. Res. Abstr. 5, 01775. Boyd, R., Ruming, K., Goodwin, I., Sandstrom, M., Schro¨der-Adams, C., 2008. Highstand transport of coastal sand to the deep ocean: a case study from Fraser Island, southeast Australia. Geology 36, 15 18. Boyd, R.K., Leckie, D.A., 2004. Comparison of the Cretaceous Paddy Estuary, Canada and the Modern Hervey Bay, Australia insights from analog modeling. Annual Meeting of American Association of Petroleum Geologists, Dallas, TX, April 18_21, 2004. http://www.searchanddiscovery.net/documents/abstracts/ annual2004/Dallas/Boyd02.htm. (accessed 18.04.09.). Boyer, S.E., Elliott, D., 1982. Thrust systems. Am. Assoc. Petrol. Geol. Bull. 66, 1196 1230. Brabb, E.E., 1991. The world landslide problem. Episodes 14 (i), 52 61. Brackenridge, R.E., 2014. Contourite Sands in the Gulf of Cadiz: Characterisation, Controls and Wider Implications for Hydrocarbon Exploration (Ph.D. dissertation). Heriot-Watt University, Edinburgh, 265 p. Brackenridge, R.E., Stow, D.A.V., Herna´ndez-Molina, F.J., Jnes, C., Mena, A., Alejo, I., et al., 2018. Textural characteristics and facies of sand-rich contourite depositional systems. Sedimentology 65 (7), 2223 2252. Available from: https://doi.org/10.1111/sed.12463. Brandt, P., Rubino, A., Fischer, J., 2002. Large-amplitude internal solitary waves in the North Equatorial Countercurrent. J. Phys. Oceanogr. 32, 1567 1573. Brawlower, T.J., Paull, C.K., Leckie, R.M., 1998. The Cretaceous-Tertiary boundary cocktail: Chicxulub impact triggers margin collapse and extensive sediment gravity flows. Geology 26, 331 334. Breien, H., Pagliardi, M., De Balsio, F., Elverhoi, A., Issler, D., Hoeg, K., 2006. Experimental studies of debris flows suabaqueous vs. subaerial: International Centre for Geohazards. ,http://www.geohazards.no/conferences/Breien.pdf. (accessed 01.10.09.). Breien, H., De Blasio, F.V., Elverhøi, A., Nystuen, J.P., Harbitz, C.B., 2010. Transport mechanisms of sand in deepmarine environments—insights based on laboratory experiments. J. Sediment. Res. 80, 975 990. Brennecke, W., 1921. Die oceanographis-chen Arbeiten derdeutschen antarktisschen Expedition 1911 1912. Arch dt. Seewarte 39 (1), 1 216. Briggs, G., Cline, L.M., 1967. Paleocurrents and source areas of Late Paleozoic sediments of the Ouachita Mountains, Southeastern Oklahoma. J. Sediment. Petrol. 37, 985 1000.

484

Bibliography

Brink, K.H., 2016. Cross-shelf exchange. Annu. Rev. Mar. Sci. 8, 59 78. Broecker, W.S., 1991. The great ocean conveyor. Oceanography 4, 79 89. Broecker, W.S., Sanyal, A., Takahashi, T., 2000. The origin of Bahamian Whitings revisited. Geophys. Res. Lett. 27 (22), 3759 3760. Bro¨nnimann, C.S., 2011. Effect of Groundwater on Landslide Triggering (Ph.D. dissertation). E´cole Polytechnique Fe´de´rale de Lausanne, Lausanne, 239 p. Brooke, C.M., Trimble, T.J., Mackay, T.A., 1995. Mounded shallow gas sands from the Quaternary of the North Sea: analogues for the formation of sand mounds in deep water Tertiary sediments? In: Hartley, A.J., Prosser, D.J. (Eds.), Characterisation of Deep Marine Clastic Systems, 94. Geological Society, Special Publications, London, pp. 95 101. Bruhn, R., Surlyk, F., 2004. Sand-grade density flow evolution on a shelf-edge slope basin-floor complex in the Upper Jurassic Olympen Formation, East Greenland. Petrol. Geosci. 10, 81 92. Brunner, C.A., Biscaye, P.E., 1997. Storm-driven transport of foraminifers from the shelf to the upper slope, southern Middle Atlantic Bight. Cont. Shelf Res. 17 (5), 491 508. Bruno, M., Va´zquez, A., Go´mez-Enri, J., Vargas, J.M., Garcı´a-Lafuente, J.M., Ruiz-Can˜avate, A., et al., 2006. Observations of internal waves and associated mixing phenomena in the Portimao Canyon area. Deep. Sea Res. Part. II Topical Stud. Oceanogr. 53 (11 13), 1219 1240. Brunsden, D., 1979. Mass movements. In: Embleton, C., Thornes, J. (Eds.), Processes in Geomorphology. Arnold, London, pp. 130 186. Brunsden, D., Prior, D.B. (Eds.), 1984. Slope Instability. John Wiley & Sons, Chicchester. Brush Jr., L.M., 1965. Experimental work on primary sedimentary structures. In: Middleton, G.V. (Ed.), Primary Sedimentary Structures and Their Hydrodynamic Interpretation, 12. SEPM Special Publication, pp. 17 24. Bryant, G., Cushman, R., Nick, K., Miall, A., 2016. Paleohydrologic controls on soft sediment deformation in the Navajo Sandstone. Sediment. Geol. 344, 205 221. Bryden, H.L., Beal, L.M., Duncan, L.M., 2005. Structure and transport of the Agulhas current and its temporal variability. J. Oceanogr. 61, 479 492. Buatois, L.A., Saccavino, L.L., Zavala, C., 2011. Ichnologic signatures of hyperpycnal flow deposits in Cretaceous river-dominated deltas, Austral Basin, southern Argentina. In: Slatt, R.M., Zavala, C. (Eds.), Sediment Transfer From Shelf to Deep Water—Revisiting the Delivery System, 61. AAPG Studies in Geology, pp. 153 170. Bugge, T., Befring, S., Belderson, R.H., Eidvin, T., Jansen, E., Kenyon, N.H., et al., 1987. A giant three-stage submarine slide off Norway. Geo-Mar. Lett. 7, 191 198. Bulfinch, D.L., Ledbetter, M.T., 1983/1984. Deep western boundary undercurrent Delineated by Sediment Texture at Base of North American Continental Rise. Geo-Mar. Lett. 3, 31 36. Burden, C.A., 2000. Sediment transport in the Mississippi Canyon: the role of currents and storm events on optical variability. December 1999. ,http://www-ocean.tamu.edu/Bpdgroup/abstracts/burden/bur_ab_01.html. (accessed 16.09.07.). Burgess, P.M., Hovius, N., 1998. Rates of delta progradation during highstands: consequences for timing of deposition in deep-marine systems. J. Geol. Soc. (Lond.) 155, 217 222. Busby, C.J., Yip, G., Blikra, L., Renne, P., 2002. Coastal landsliding and catastrophic sedimentation triggered by Cretaceous-Tertiary bolide impact: a Pacific margin example? Geology 30, 687 690. Buss, E., Heim, A., 1881. Der Bergsturz von Elm. Worster, Zurich, p. 133. Butman, B., Twichell, D.C., Rona, P.A., Tucholke, B.E., Middleton, T.J., Robb, J.M., 2006. Sea floor topography and backscatter intensity of the Hudson Canyon Region Offshore of New York and New Jersey. U.S. Geological Survey Open-File Report 2004-1441. Cacchione, D.A., Wiberg, P.L., Lynch, J., Irish, J., Traykovski, P., 1999. Estimates of suspended-sediment flux and bedform activity on the inner portion of the Eel continental shelf. Mar. Geol. 154, 83 97. Cacchione, D.A., Pratson, L.F., Ogston, A.S., 2002. The shaping of continental slopes by internal tides. Science 296, 724 727. Cairns, J.L., 1980. Variability in the Gulf of Cadiz: internal waves and globs. J. Phys. Oceanogr. 10, 579 595. Callot, P., Odonne, F., Sempere, T., 2008. Liquification and soft-sediment deformation in a limestone megabreccia: the Ayabacas giant collapse. Cretaceous, South. Peru.Sediment. Geol. 212, 49 69.

Bibliography

485

Calvo, J.P., Rodrı´guez-Pascua, M., Martin-Velazquez, S., Jimenez, S., De Vicente, G., 1998. Microdeformation of lacustrine laminite sequences from Late Miocene formations of SE Spain: an interpretation of loop bedding. Sedimentology 45, 279 292. Camerlenghi, A., Urgeles, R., Fantoni, I., 2010. A database on submarine landslides of the Mediterranean. In: Mosher, D.C., et al., (Eds.), Submarine Mass Movements and Their Consequences, 28. Advances in Natural and Technological Hazards Research, pp. 503 513. Campbell, C.S., 1989. Self-lubrication for long runout landslides. J. Geol. 97, 653 665. Campbell, K.A., Nesbitt, E.A., Bourgeois, J., 2006. Signatures of storms, oceanic fl oods and forearc tectonism in marine shelf strata of the Quinault Formation (Pliocene), Washington, USA. Sedimentology 53, 945 969. Cannon, G.A., 1978. Circulation in the Strait of Juan De Fuca: some recent oceanographic observations. NOAA Technical Report ERL 399-PMEL 29, 55 p. Cannon, S.H., Kirkham, R.M., Parise, M., 2001. Wildfire-related debris-flow initiation processes, Storm King Mountain, Colorado. Geomorphology 171 188. Carey, S.N., Schneider, J.-L., 2011. Chapter 7. Volcaniclastic processes and deposits in the Deep-Sea. In: Hueneke, H., Mulder, T. (Eds.), Deep-Sea Sediments, 63. Elsevier, Amsterdam, pp. 457 515. Developments in Sedimentology. Carlson, P.R., Molnia, B.F., 1977. Submarine faults and slides on the continental shelf, northern Gulf of Alaska. In: Richards, A.F. (Ed.), Marine Geotechnology, Marine Slope Stability 2. Crane, Russak & Company, Inc., New York, pp. 275 290. Carlson, P.R., Karl, H.A., 1988. Development of large submarine canyons in the Bering Sea indicated by morphologic, seismic, and sedimentologic characteristics. GSA Bulletin 100, 1594 1615. Carlson, P.R., Karl, H.A., Edwards, B.D., et al., 1991. Mass sediment failure and transport features revealed by acoustic techniques, Beringian margin, Bering Sea, AK. Marine Geotechniques 10, 33 51. Carter, R.M., Lindqvist, J.K., 1975. Sealers Bay submarine fan complex, Oligocene, southern New Zealand. Sedimentology 22, 465 483. Carter, R.M., 1975a. Adiscussion and classification of subaqueous mass-transport with particular application to grain flow, slurry flow, and fluxoturbidites. Earth-Sci. Rev. 11, 145 177. Carter, R.M., 1975b. Mass-emplaced sand-fingers at Mararoa construction site, southern New Zealand. Sedimentology 22, 275 288. Carter, R.M., 2007. The role of intermediate-depth currents in continental shelf-slope accretion: Canterbury Drifts, SW Pacific Ocean. In: Viana, A.R., Rebesco, M. (Eds.), Economic and Palaeoceanographic Significance of Contourite Deposits, vol. 276. Geological Society of London Special Publication, pp. 129 154. Cartigny, M.J.B., Ventra, D., Postma, G., van den Berg, J.H., 2014. Morphodynamics and sedimentary structures of bedforms under supercritical-flow conditions: new insights from flume experiments. Sedimentology 61, 712 748. Carvajal, C.R., Steel, R.J., 2006. Thick turbidite successions from supply-dominated shelves during sea-level highstand. Geology 34, 665 668. Carver, R.E., 1971. Procedures in Sedimentary Petrology. Wiley-Interscience, New York, p. 653. Catuneanu, O., 2006. Principles of Sequence Stratigraphy. Elsevier, Amsterdam, p. 386. Charlet, F., Fagel, N., De Batist, M., Hauregard, F., Minnebo, B., Meischner, D., et al., 2005. Sedimentary dynamics on isolated highs in Lake Baikal: evidence from detailed high-resolution geophysical data and sediment cores. Glob. Planet. Change 46, 125 144. Chen, J., Lee, H.S., 2013. Soft-sediment deformation structures in Cambrian siliciclastic and carbonate storm deposits (Shandong Province, China): differential liquefaction and fluidization triggered by storm-wave loading. Sediment. Geol. 288, 81 94. Chen, J.Y., Shen, H.T., Yun, C.X., 1998. Processes of Dynamics and Geomorphology of the Changjiang Estuary. Shanghai Scientific and Technical Publishers, Shanghai, 453 pp. (in Chinese with English Abstract). Chen, S., Steel, R.J., Dixon, J.F., Osman, A., 2014. Facies and architecture of a tide-dominated segment of the Late Pliocene Orinoco Delta (Morne L’Enfer Formation) SW Trinidad. Mar. Petrol. Geol. 57, 208 232. Chen, S., Steel, R.J., Kvale, E.P., Olariu, C., Zhang, J., 2017. Unusually well preserved tidal signals and tidal constituents on Pliocene Paleo-Orinoco Delta, Trinidad. AAPG Datapages/Search and Discovery Article #90291, AAPG Annual Convention and Exhibition, Houston, TX, 2 5 April 2017.

486

Bibliography

Chen, S.N., Geyer, W.R., Hsu, T.J., 2013. A numerical investigation of the dynamics and structure of Hyperpycnal River plumes on sloping continental shelves. J. Geophys. Res. Ocean. 118 (5), 2702 2718. Chen, J.T., Han, Z.Z., Zhang, X.L., Fan, A.P., Yang, R.C., 2010. Early diagenetic deformation structures of the Furongian ribbon rocks in Shandong Province of China—a new perspective of the genesis of limestone conglomerates. Sci. China Earth Sci. 53, 241 252. Chen, Z., Pan, J., Jiang, Y., Lin, H., 2017. Far-reaching transport of Pearl River plume water by upwelling jet in the northeastern South China Sea. J. Mar. Syst. 173, 60 69. Che´rubin, L., Serra, N., Ambar, I., 2003. Low-frequency variability of the Mediterranean Undercurrent downstream of Portimao Canyon. J. Geophys. Res. 108, 3058. Available from: https://doi.org/10.1029/ 2001JC001229. Che´rubin, L., Carton, X., Dritschel, D.G., 2007. Vortex dipole formation by baroclinic instability of boundary currents. J. Phys. Oceanogr. 37, 1661 1677. Chiarella, D., Moretti, M., Longhitano, S.G., Muto, F., 2016. Deformed cross-stratified deposits in the Early Pleistocene tidally-dominated Catanzaro strait-fill succession, Calabrian Arc (Southern Italy): triggering mechanisms and environmental significance. Sediment. Geol. 344, 277 289. Chien, H., Kao, C.-C., Chuang, L.Z.H., 2002. On the characteristics of observed coastal freak waves. Coastal Eng. J 44, 301 319. Chiggiato, J., Bergamasco, A., Borghini, M., Falcieri, F.M., Falco, P., Langone, L.M., et al., 2016. Dense-water bottom currents in the Southern Adriatic Sea in spring 2012. Mar. Geol. 375, 134 145. Chillarige, A.V., Robertson, P.K., Morgenstern, N.R., Christian, H.A., 1997. Evaluation of the in situ state of Fraser River sand. Can. Geotech. J. 34, 510 519. Chikita, K., 1989. A field study on turbidity currents initiated from spring runoffs. Water Resour. Res. 25, 257 271. Chu, V.W., 2014. Greenland ice sheet hydrology: a review. Prog. Phys. Geogr. 38 (1), 19 54. Chu, J.-H., Sampson, C.R., Levine, A.S., Fukada, E., 2002. The joint typhoon warning center tropical cyclone besttracks, 1945 2000, NRL Reference Number: NRL/MR/7540-02-16. August 2002. ,http://metocph.nmci.navy. mil/jtwc/best_tracks/TC_bt_report.html. (accessed 10.11.07.). CIMAS (The Cooperative Institute for Marine and Atmospheric Studies), 2011. Ocean surface currents. ,http:// oceancurrents.rsmas.miami.edu/indian/indian.html. (accessed 20.03.11.). CIMAS (The Cooperative Institute for Marine and Atmospheric Studies), 2015. Ocean surface currents. ,http:// oceancurrents.rsmas.miami.edu/. (accessed 12.07.15.). Claeys, P., Kiessling, W., Alvarez, W., 2002. Distribution of Chicxulub ejecta at the Cretaceous-Tertiary boundary. In: Koeberl, C., MacLeod, K.G. (Eds.), Catastrophic Events and Mass Extinctions: Impacts and Beyond. Geological Society of America Special Paper 356, pp. 55 68. Clague, J.J., Stead, D. (Eds.), 2012. Landslides: Types, Mechanisms, and Modeling. Cambridge University Press, Cambridge. Clare, M.A., Hughes-Clarke, J.E., Talling, P.J., Cartigny, M.J.B., Pratomo, D.G., 2016. Preconditioning and triggering of offshore slope failures and turbidity currents revealed by most detailed monitoring yet at a fjord-head delta. Earth Planet. Sci. Lett. 450, 208 220. Clark, J.D., Stanbrook, D.A., 2001. Formation of large-scale shear structures during the deposition of high-density turbidity currents, Gre`s d’Annot Formation, SE France. In: McCaffrey, W.D., Kneller, B.C., Peakall, J. (Eds.), Particulate Gravity Currents, 31. IAS Special Publication, pp. 219 232. Cleary, P.W., Campbell, C.S., 1993. Self-lubrication for long runout landslides’ examination by computer simulation. J. Geophys. Res. 98 (B12), 21911 21924. Clifton, H.E., (Ed.), 1988. Sedimentologic Consequences of Convulsive Geologic Events, Geological Society of America Special Paper 229, 157 p. Cloud Jr., P.E., 1962. Environment of Calcium Carbonate Deposition West of Andros Island Bahamas. Geological Survey Professional Paper, 350. United States Government Printing Office, Washington, DC, 138 pp. OK. Coates, D.R., 1977. Landslide prospectives. In: Coates, D.R. (Ed.), Landslides. Geological Society of America, Boulder, CO, pp. 3 38. Cochonat, P., Cadet, J.-P., Lallemant, S.J., Mazzotti, S., Nouze, H., Fouchet, C., et al., 2002. Slope instabilities and gravity processes in fluid migration and tectonically active environment in the eastern Nankai accretionary wedge (KAIKO-Tokai’96 cruise). Mar. Geol. 187 (1-2 SU), 193 202.

Bibliography

487

Coleman Jr., J.L., 1997. Reinterpretation of depositional processes in a classic flysch sequence(Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma, Discussion. AAPG Bull. 81, 466 469. Coleman, J.M., Garrison, L.E., 1977. Geological aspects of marine slope stability, northwestern Gulf of Mexico. In: Richards, A.F. (Ed.), Marine Geotechnology, Marine Slope Stability, vol. 2. Crane, Russak & Company, New York, pp. 9 44. Coleman, J.M., Prior, D.B., 1982. Deltaic environments. In: Scholle, P.A., Spearing, D. (Eds.), Sandstone Depositional Environments, 31. AAPG Memoir, pp. 139 178. Coleman, J.M., Prior, D.B., 1988. Mass wasting on continental margins. Annu. Rev. Earth Planet. Sci. 16, 101 119. Collins, G.S., Melosh, H.J., 2003. Acoustic fluidization and the extraordinary mobility of sturzstroms. J. Geophys. Res. 108, 2473. Available from: https://doi.org/10.1029/2003JB002465. Collins, D.S., Johnson, H.D., Allison, P.A., Guilpain, P., Damit, A.R., 2017. Coupled ‘storm-flood’ depositional model: application to the Miocene Modern Baram Delta Province, north west Borneo. Sedimentology 64 (5), 1203 1235. Collinson, J.D., 1994. Sedimentary deformational structures. In: Maltman, A. (Ed.), The Geological Deformation of Sediments. Chapman & Hall, London, pp. 95 125. Collinson, J.D., Thompson, D.B., 1982. Sedimentary Structures. George Allen & Unwin, London, p. 194. Collot, J.-Y., Lewis, K., Lamarche, G., Lallemand, S., 2001. The giant Ruatoria debris avalanche on the northern Hikurangi margin, New Zealand: result of oblique seamount subduction. J. Geophys. Res. 106 (B9), 19,271 19,297. Cook, H.E., 1979. Ancient continental slope sequences and their value in understanding modern slope development. In: Doyle, L.J., Pilkey, O.H. (Eds.), Geology of Continental Slopes, 27. SEPM Special Publication, pp. 287 305.5. Cook, H.E., Enos, P. (Eds.), 1977. Deep-water carbonate environments. SEPM Special Publication 25, p. 335. Cooper, C., Forristall, G.Z., Joyce, T.M., 1990. Velocity and hydrographic structure of two Gulf of Mexico warmcore rings. J. Geophys. Res. 95 (C2), 1663 1679. Cossey, S. P. J. 2011. Mass-transport deposits in the Ppper Paleocene Chicontepec Formation, Mexico. In Shipp et al. (eds.), Mass-Transport Deposits in Deepwater Settings. SEPM (Society for Sedimentary Geology) Special Publication 96, 269 277. Costa, J.E., Williams, G.P., 1984. Debris-flow dynamics (videotape). U.S. Geological Survey Open File Report OF 84-0606. Costard, F., Forget, F., Mangold, N., Peulvast, J.P., 2002. Formation of recent Martian debris flows by melting of near-surface ground ice at high obliquity. Science 295, 110 113. Coussot, P., Meunier, M., 1996. Recognition, classification and mechanical description of debris flows. Earth-Sci. Rev. 40, 209 227. Covault, J.A., Normark, W.R., Romans, B.W., Graham, S.A., 2007. Highstand fans in the California borderland: the overlooked deep-water depositional systems. Geology 35, 783 786. Cowan, D.S., von Huene, R., Aubouin, J., 1982. Origin of ‘vein structure’ in slope sediments on the inner slope of the Middle America Trench off Guatemala. Initial Rep. Deep Sea Drill. Proj. 67, 645 650. Available from: https://doi.org/10.2973/dsdp.proc.67. Cowan, E.A., Cai, J., Powell, R.D., Seramur, K.C., Spurgeon, V.L., 1998. Modern tidal rhythmites deposited in a deep-water estuary. Geo-Mar. Lett. 18, 40 48. Crease, J., 1965. The flow of Norwegian sea water through the Faeroe bank channel. Deep-Sea Res. 12, 143 150. Criado-Aldeanueva, F., Garcı´a-Lafuente, J., Vargas, J.M., Rı´o, J.D., Va´zquez, A., Reul, A., et al., 2006. Distribution and circulation of water masses in the Gulf of Cadiz from in situ observations. Deep-Sea Res. II 53, 1144 1160. Crookshanks, S., Gilbert, R., 2008. Continuous, diurnally fluctuating turbidity currents in Kluane Lake, Yukon Territory. Can. J. Earth Sci. 45, 1123 1138. Crozier, M.J., 1973. Techniques for the morphometric analysis of landslipe. Zeit. Geomorph. Dyn. 17, 78 101. Cruden, D.M., 1991. A simple definition of a landslide. Bull. Int. Assoc. Eng. Geol. 43, 27 28. Cruden, D.M., 2003. The first classification of landslides? Environ. Eng. Geosci. 9, 197 200. Cruden, D.M., Hungr, O., 1986. The debris of the Frank slide and theories of rockslide-avalanche mobility. Can. J. Earth Sci. 23, 425 432. Cruden, D.M., Varnes, D.J., 1996. Landslides types and processes. In: Turner, A.K., Schuster, R.L. (Eds.), Landslides Investigation and Mitigation, 247. Transportation Research Board National Research Council, Washington, DC, pp. 36 75. Special Report.

488

Bibliography

Cuffey, K.M., Paterson, W.S.B., 2010. The Physics of Glaciers, fourth ed. Academic Press, 704 pp. Cutroneo, L., Ferretti, G., Scafidi, D., Ardizzone, G.D., Vagge, G., Capello, M., 2017. Current observations from a looking down vertical V-ADCP: interaction with winds and tide? The case of Giglio Island (Tyrrhenian Sea, Italy). Oceanologia 59 (2), 139 152. D’Agostino, A.E., Jordan, D.W., 1997. Reinterpretation of depositional processes in a classic flysch sequence (Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma: discussion. Am. Assoc. Petrol. Geol. Bull. 473 475. Dade, W.B., Huppert, H.E., 1998. Long-runout rockfalls. Geology 26 (9), 803 806. Dadson, S.J., Hovius, N., Pegg, S., Brian Dade, W., Horng, M.J., Chen, H., 2005. Hyperpycnal river flow from an active mountain belt. J. Geophys. Res. Atmos. 110, F04016. Dallimore, C.J., Imberger, J., Hodges, B.R., 2004. Modeling a plunging underflow. J. Hydraulic Eng. 130 (11), 1068 1076. Dalrymple, R.W., 1992. Tidal depositional systems. In: Walker, R.G., James, N.P. (Eds.), Facies Models: Response to Sea Level Change [GEOtext 1]. Geological Association of Canada, pp. 195 218. Dalrymple, R.W., Narbonne, G.M., 1996. Continental slope sedimentation in the Sheepbed Formation (Neoproterozoic, Windermere Supergroup), Mackenzie Mountains, N.W.T. Can. J. Earth Sci. 33, 848 862. Dalrymple, R.W., John Knight, R., Lambiase, J.J., 1978. Bedforms and their hydraulic stability relationships in a tidal environment, Bay of Fundy, Canada. Nature 275, 100 104. Dalrymple, R.W., Knight, R.J., Zaitlin, B.A., Middleton, G.V., 1990. Dynamics and facies model of a macrotidal sand-bar complex, Cobequid Bay Salmon River estuary, (Bay of Fundy). Sedimentology 37, 577 612. Dalrymple, R.W., Zaitlin, B.A., Boyd, R., 1992. Estuarine facies models: conceptual basis and stratigraphic implications. J. Sediment. Petrol. 62, 1130 1146. Daly, R.A., 1936. Origin of submarine “canyons”. Am. J. Sci. 5 (31), 401 420. Damuth, J.E., Fairbridge, R.W., 1970. Equatorial Atlantic deepsea sands and ice age aridity in tropical South America. Geol. Soc. Am. Bull. 81, 585 601. Damuth, J.E., Embley, R.W., 1981. Mass-transport processes on Amazon Cone: Western Equatorial Atlantic. AAPG Bull. 65, 629 643. Damuth, J.E., Olson, H.C., 2001. Neogene-Quaternary contourite and related deposition on the West Shetland Slope and Faeroe-Shetland Channel revealed by high-resolution seismic studies. Mar. Geophys. Res. 22, 363 398. Damuth, J.E., Flood, R.D., Kowsmann, R.O., Gorini, M.A., Belderson, R.H., Gorini, M.A., 1988. Anatomy and growth-pattern of Amazon deep-sea fan revealed by long-range side-scan sonar (GLORIA) and highresolution seismic studies. AAPG Bull. 72, 885 911. Dan, G., Sultan, N., Savoye, B., 2007. The 1979 Nice harbour catastrophe revisited, trigger mechanism inferred from geotechnical measurements and numerical modeling. Mar. Geol. 245, 40 64. Dasgupta, P., 2003. Sediment-gravity flow the conceptual problems. Earth-Sci. Rev. 62, 265 281. Dasgupta, P., 2008. Experimental decipherment of the soft-sediment deformation observed in the upper part of the Talchir Formation (Lower Permian), Jharia Basin, India. Sedimentary Geology 205, 100 110. Dasgupta, P., Chatterjee, A., 2019. Formation of water-escape structure during shock-induced fluidization: The role of permeability contrast. a. Journal of Structural Geology 124, 1 7. Davies, G.R., 1997. Interpretites: the new universal, non-genetic, non-descriptive classification of sedimentary products and environments. Can. Soc. Petrol. Geol. Reserv. 24, 22 23. Davies, T.R.H., 1982. Spreading of rock avalanche debris by mechanical fluidization. Rock. Mech. Rock Eng. 15, 9 24. De Blasio, F.V., 2011. Landslides in Valles Marineris (Mars): a possible role of basal lubrication by sub-surface ice. Planet. Space Sci. 59 (13), 1384 1392. De Blasio, F.V., Elverhøi, A., 2008. A model for frictional melt production beneath large rock avalanches. J. Geophys. Res. Earth Surf. 113, F02014. Available from: https://doi.org/10.1029/2007JF000867. De Blasio, F.V., Elverhøi, A., Engvik, L.E., Issler, D., Gauer, P., Harbitz, C., 2006. Understanding the high mobility of subaqueous debris flows. Nor. J. Geol. 86, 275 284. de Castro, S., Hernandez-Molina, F.J., Rodriguez-Tovar, F.J., Llave, E., Ng, Z.L., Nishida, N., Mena, A., et al., 2020. Contourites and bottom current reworked sands: Bed facies model and implications. Marine Geology 428, 106 267.

Bibliography

489

de Jong, M.P.C., Battjes, J.A., 2004. Low-frequency sea waves generated by atmospheric convection cells. J. Geophys. Res. Ocean. 109, C01011. de Weger, W., Javier Herna´ndez-Molina, F.J., Flecker, R., Sierro, F.J., Chiarella, D., Krijgsman, W., et al., 2020. Late Miocene contourite channel system reveals intermittent overflow behavior: Geology, v. 48, p. XXX XXX. https://doi.org/10.1130/G47944.1. Denamiel, C., Budgell, W.P., Toumi, R., 2013. The Congo River plume: impact of the forcing on the far-field and near-field dynamics. J. Geophys. Res. Ocean. 118, 964 989. Available from: https://doi.org/10.1002/ jgrc.20062. Dengler, A.T., Wilde, P., 1987. Turbidity currents on steep slopes: application of an avalanche-type numeric model for ocean thermal energy conversion design. Ocean Eng. 14, 409 433. Dengler, A.T., Wilde, P., Noda, E.K., Normark, W.R., 1984. Turbidity currents generated by Hurricane Iwa. GeoMar. Lett. 4, 5 11. Deptuck, M.E., Piper, D.J.W., Savoye, B., Gervais, A., 2008. Dimensions and architecture of late Pleistocene submarine lobes off the northern margin of East Corsica. Sedimentology 55 (4), 869 898. Dierssen, H.M., Zimmerman, R.C., Burdige, D.J., 2009. Optics and remote sensing of Bahamian carbonate sediment whitings and potential relationship to wind-driven Langmuir circulation. Biogeosciences 6, 487 500. Dill, R.F., Marshall, N.F., Reimnitz, E., 1975. In situ submersible observations of sediment transport and erosive features in Rio Balsas Submarine Canyon, Mexico (Abs), Geol. Soc. Am., 7. pp. 1052 1053. Dillon, W.P., Zimmerman, H.P., 1970. Erosion by biological activity in two new england submarine canyons. J. Sediment. Petrol. 40, 542 547. Dikau, R., Cavallin, A., Ja¨ger, S., 1996. Databases and GIS for landslide research in Europe. Geomorphology 15, 227 239. Dingle, R.V., 1977. The anatomy of a large submarine slump on a sheared continental margin (SE Africa). J. Geol. Soc. Lond. 134, 293 310. Dingle, R.V., 1980. Large allochtonous sediment masses and their role in the construction of the continental slope and rise off southwestern Africa. Mar. Geol. 37, 333 354. Dingler, J.R., Anima, R.J., 1989. Subaqueous grain flows at the head of Carmel submarine canyon, California. J. Sediment. Petrol. 59, 280 286. Dixit, J.G., 1982. Resuspension Potential of Deposited Kaolinite Beds (M.S. thesis). University of Florida, Gainesville, FL. Dott Jr., R.H., 1963. Dynamics of subaqueous gravity depositional processes. AAPG Bull. 47, 104 128. Douillet, G.A., Taisne, B., Mu¨ller, S.K., Kueppers, U., Dingwell, D.B., 2015. Syn-eruptive, soft-sediment deformation of deposits from dilute pyroclastic density current: triggers from granular shear, dynamic pore pressure, ballistic impacts and shock waves. Solid. Earth 6, 553 572. Dowdeswell, J.A., O’Cofaigh, C. (Eds.), 2002. Glacier-Influenced Sedimentation on High-Latitude Continental Margins, 203. Geological Society, London, p. 373. Special Publications. Dowdeswell, J.A., Elverhoi, A., Spielhagen, R., 1998. Glacimarine sedimentary processes and facies on the Polar North Atlantic margins. Quat. Sci. Rev. 17 (1 3), 243 272. Doyle, L.J., Pilkey, O.H. (Eds.), 1979. Geology of Continental Slopes, 27. SEPM Special Publication, p. 374. Draper, L., 1964. “Freak” ocean waves. Oceanus 10, 13 15. Driscoll, N.W., Weissel, J.K., Goff, J.A., 2000. Potential for large-scale submarine slope failure and tsunami generation along the U.S. mid-Atlantic coast. Geology 28, 407 410. Droz, L., Rigaut, F., Cochonat, P., Tofani, R., 1996. Morphology and recent evolution of the Zaire turbidite system (Gulf of Guinea). GSA Bull. 108, 253 269. Druzhininaa, O., Bitinasb, A., Molodkovc, A., Kolesnika, T., 2017. Palaeoseismic deformations in the Eastern Baltic region (Kaliningrad District of Russia). Estonian J. Earth Sci. 66 (3), 119 129. Duarte, C.S.L., Viana, A.R., 2007. Santos Drift System: stratigraphic organization and implications for late Cenozoic palaeocirculation in the Santos Basin, SW Atlantic Ocean. In: Viana, A.R., Rebesco, M. (Eds.), Economic and Palegeographic Significance of Contourite Deposits, 276. Geological Society, London, pp. 171 198. Special Publications. Dubois, N., Mitchell, N.C., 2012. Large-scale sediment redistribution on the equatorial Pacific seafloor. Deep. Sea Res. 69, 51 61. PT I.

490

Bibliography

Duda, J.J., Warrick, J.A., Magirl, C.S., 2011. Coastal and lower Elwha River, Washington, prior to dam removal— history, status, and defining characteristics. In: Duda, J.J., Warrick, J.A., Magirl, C.S. (Eds.), Coastal Habitats of the Elwha River: Biological and Physical Patterns and Processes Prior to Dam Removal. U.S. Geological Sand Survey Scientific Investigations Report 2011-5120, Chapter 1. Duke, W.L.R., Arnott, R.W.C., Cheel, R.J., 1991. Shelf sandstones and hummocky-cross stratification: new insights on a stormy debate. Geology 19, 625 628. Duncan, J.M., Wright, S.G., 2005. Soil Strength and Slope Stability:. John Wiley & Sons, Inc, Hoboken, NJ, p. 297. Dunham, J., Saller, A.H., 2014. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands: discussion. AAPG Bull. 98, 851 857. Duranti, D., Hurst, A., Bell, C., Groves, S., Hanson, R., 2002. Injected and remobilized Eocene sandstones from the Alba Field, UKCS: core and wireline log characteristics. Petrol. Geosci. 8, 99 107. Dykstra, M., 2012. Deep-water tidal sedimentology. In: Davis Jr., R.A., Dalrymple, R.W. (Eds.), Principles of Tidal Sedimentology. Springer, Berlin, pp. 371 396. ˝ Dysthe, K., Krogstad, H.E., Muller, P., 2008. Oceanic rogue waves. Annu. Rev. Fluid Mech. 40, 287 310. Dzulynski, S., Sanders, J.E., 1962. Current marks on firm mud bottoms. Conn. Acad. Arts Sci. Trans. 42, 57 96. Dzulynski, S., Ksiazkiewicz, M., Kuenen, Ph. H., 1959. Turbidites in flysch of the Polish Carpathian Mountains. GSA Bull. 70, 1089 1118. Easterbrook, D.J., 1999. Surface Processes and Landforms, second ed. Prentice Hall (Pearson Education Company), Upper Saddle River, NJ, p. 546. Eberli, G., Betzler, C., 2019. Characteristics of modern carbonate contourite drifts. Sedimentology 66, 1163 1191. Eckel, E.B., 1958. Introduction. In: Eckel, E.B. (Ed.), Landslide and Engineering Practice, 29. National Research Council, Highway Research Board Special Report, pp. 1 5. Edgers, L., Karlsrud, K., 1982. Soil flows generated by submarine slides—case studies and consequences. In: Chryssostomidains, C., Connor, J.J. (Eds.), Proceedings of the Third International Conference on the Behavior of Offshore Structures. Hemisphere, Bristol, PA, pp. 425 437. El Taki, H., Pratt, B.R., 2012. Syndepositional tectonic activity in an epicontinental basin revealed by deformation of subaqueous carbonate laminites and evaporites: seismites in Red River strata (Upper Ordovician) of southern Saskatchewan, Canada. Bull. Can. Petrol. Geol. 60, 37 58. Elverhøi, A., Norem, H., Anderson, E.S., Dowdeswell, J.A., Fossen, I., Haflidason, H., et al., 1997. On the origin and flow behavior of submarine slides on deep-sea fans along the Norwegian-Barents Sea continental margin. Geo-Mar. Lett. 17, 119 125. Elverhøi, A., De Blasio, F., Butt, F.A., Issler, D., Harbitz, C.B., Engvik, L., et al., 2002. Submarine mass-wasting on glacially-influenced continental slopes: processes and dynamics. In: Dowdeswell, J.A., O’Cofaigh, C. (Eds.), Glacier-Influenced Sedimentation on High-Latitude Continental Margins, 203. Geological Society, London, pp. 73 87. Special Publications. E-MARSHAL (Earth’s continental MARgins: aSsessing the geoHAzard from submarine Landslides), 2013. IGCP511 legacy. ,http://www.igcp585.org/. (accessed 27.12.14.). Embley, R.W., 1976. New evidence for occurrence of debris flow deposits in the deep sea. Geology 4, 371 374. Embley, R.W., 1980. The role of mass transport in the distribution and character of deep-ocean sediments with special reference to the North Atlantic. Mar. Geol. 38, 23 50. Embley, R.W., 1982. Anatomy of some Atlantic margin sediment slides and some comments on ages and mechanisms. In: Saxov, S., Nieuwenhuis, J.K. (Eds.), Marine Slides and Other Mass Movements. Plenum, New York, pp. 189 214. Embley, R.W., Jacobi, R.D., 1977. Distributions and morphology of large sediment slides and slumps on Atlantic continental margins. Mar. Geotechnol. 2, 205 228. Enos, P., 1977. Flow regimes in debris flow. Sedimentology 24, 133 142. Enzel, Y., Kadan, G., EyaL, Y., 2000. Holocene earthquakes inferred from a fan-delta sequence in the Dead Sea Graben. Quat. Res. 53, 34 48. Erismann, T.H., 1979. Mechanisms of large landslides. Rock. Mech. 12, 15 46. Etienne, S., Mulder, T., Bez, M., Desaubliaux, G., Kwasniewski, A., Parize, O., et al., 2012. Multiple scale characterization of sand-rich distal lobe deposit variability, examples from the Annot Sandstones Formation, Eocene Oligocene, SE France. Sediment. Geol. 273 274, 1 18.

Bibliography

491

Ettensohn, F.R., Zhang, C., Gao, L., Lierman, R.T., 2011. Soft-sediment deformation in epicontinental carbonates as evidence of paleoseismicity with evidence for a possible new seismogenic indicator: accordion folds. Sediment. Geol. 235, 222 233. Expedition 323 Scientists, 2011. Site U1345. In: Takahashi, K., Ravelo, A.C., Alvarez Zarikian, C.A., the Expedition 323 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 323. Integrated Ocean Drilling Program Management International, Inc., Tokyo, pp. 1 79. ,https://doi.org/10.2204/iodp. proc.323.109.2011.. Expedition 333 Scientists, 2011. Integrated Ocean Drilling Program Expedition 333 Preliminary Report NanTroSEIZE Stage 2: Subduction Inputs 2 and Heat Flow. Texas A&M University, College Station, TX, ,https://doi.org/10.2204/iodp.pr.333.2011.. Expedition 308 Scientists, 2006. Site U1322. In: Flemings, P.B., Behrmann, J.H., John, C.M., the Expedition 308 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, Volume 308. Texas A&M University, College Station, Texas, , https://doi.org/10.2204/iodp.proc.308.106.2006 . . Expedition 333 Scientists, 2012a. Site C0012. In: Henry, P., Kanamatsu, T., Moe, K., the Expedition 333 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 333. A&M University, College Station, TX, ,https://doi.org/10.2204/iodp.proc.333.105.2012.. Expedition 333 Scientists, 2012b. Site C0018. In: Henry, P., Kanamatsu, T., Moe, K., the Expedition 333 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 333. Texas A&M University, College Station, TX, ,https://doi.org/10.2204/iodp.proc.333.103.2012.. Expedition 339 Scientists, 2012. Mediterranean outflow: environmental significance of the Mediterranean Outflow water and its global implications. IODP Prel. Rept. 339. Available from: https://doi.org/10.2204/iodp. pr.339.2012. Expedition 344 Scientists, 2013. Mid-slope Site U1380. In: Harris, R.N., Sakaguchi, A., Petronotis, K., the Expedition 344 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 344. Texas A&M University, College Station, TX, ,https://doi.org/10.2204/iodp.proc.344.106.2013.. Expedition 349 Scientists, 2015a. Site U1431. In: Li, C.-F., Lin, J., Kulhanek, D.K., the Expedition 349 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 349. Texas A&M University, College Station, TX, ,https://doi.org/10.14379/iodp.proc.349.103.2015.. Expedition 349 Scientists, 2015b. Site U1434. In: Li, C.-F., Lin, J., Kulhanek, D.K., the Expedition 349 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 349. Texas A&M University, College Station, TX, ,https://doi.org/10.14379/iodp.proc.349.106.2015.. Expedition 351 Scientists, 2015. Site U1438. In: Arculus, R.J., Ishizuka, O., Bogus, K., the Expedition 351 Scientists (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 351. Texas A&M University, College Station, TX, ,https://doi.org/10.14379/iodp.proc.351.103.2015.. Ewing, M., Ettrium, S.L., Ewing, J.L., Le Pichon, X., 1971. Sediment transport and distribution in the Argentine Basin: Part 3, Nepheloid layer and process of sedimentation. In: Ahrens, L.A., Press, F., Runcorn, S.K., Urey, H.C. (Eds.), Physics and Chemistry of the Earth. Pergamon Press, London, pp. 55 77. Ezquerro, L., Moretti, M., Liesa, C.L., Luzo´n, A., Pueyo, E.L., Simo´n, J.L., 2016. Controls on space-time distribution of soft-sediment deformation structures: applying palaeomagnetic dating to approach the apparent recurrence period of paleoseisms at the Concud Fault (eastern Spain). Sediment. Geol. 344, 91 111. Fairbridge, R.W., 1980. The estuary: its definition and geodynamic cycle. In: Olausson, E., Cato, I. (Eds.), Chemistry and Geochemistry of Estuaries. Wiley, New York, pp. 1 35. Fallgatter, C., Kneller, B., Paim, P.S.G., Milana, J.P., et al., 2017. Transformation, partitioning and flow–deposit interactions during the run-out of megaflows. Sedimentology 64, 359 387. Famakinwa, S.B., Shanmugam, G., Hodgkinson, R.J., Blundell, L.C., 1996. Deep-water slump and debris flow dominated reservoirs of the Zafiro Field area, offshore Equatorial Guinea. Offshore West Africa Conference and Exhibition, 5 7 November, Libreville, pp. 1 14. Fan, H., Huang, H., Zeng, T.Q., Wang, K., 2006. River mouth bar formation, riverbed aggradation and channel migration in the modern Huanghe (Yellow) River delta, China. Geomorphology 74, 124 136. Fang, G., Yang, J., 1985. A two-dimensional numerical model of the tidal motions in the Bohai Sea. Chin. J. Oceanol. Limnol. 3 (2), 135 152. Farre, J.A., McGregor, B.A., Ryan, W.B.F., Robb, J.M., 1983. Breaching the shelf-break: passage from youthful to mature phase in submarine canyon evolution. In: Stanley, D.J., Moore, G.T. (Eds.), The Shelf-Break: Critical

492

Bibliography

Interface on Continental Margins, 33. Society of Economic Paleontologists and Mineralogists Special Publication, pp. 25 39. Fauge`res, J.-C., Stow, D.A.V., 1993. Bottom-current-controlled sedimentation: a synthesis of the contourite problem. Sediment. Geol. 82, 287 297. Fauge`res, J.-C., Stow, D.A.V., 2008. Contourite drifts: nature, evolution and controls. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites. Elsevier, Amsterdam, Developments in Sedimentology 60, pp. 259 288. Chapter 14. Fauge`res, J.-C., Mulder, T., 2011. Contour currents and contourite drifts. In: Huneke, H., Mulder, T. (Eds.), DeepSea Sediments, Developments in Sedimentology 63, pp. 149 214. Fauge`res, J.-C., Gonthier, E., Stow, D.A.V., 1984. Contourite drift moulded by deep Mediterranean outflow. Geology 12, 296 300. Farmer, D., Armi, L., 1999. The generation and trapping of internal solitary waves over topography. Science 283, 188 190. Federiuk, J., Allen, J.S., 1996. Model studies of near-inertial waves in flow over the Oregon continental shelf. J. Phys. Oceanogr. 26, 2053 2075. Feeley, K., 2007, Triggering mechanisms of submarine landslides. Research Report, Department of Civil and Environmental Engineering, Northeastern University, Boston, MA, 02115, 45 p. Feldhausen, P.H., Stanley, D.J., Knight, R.J., Maldonado, A., 1981. Homogenization of gravity-emplaced muds and unifites: models from the Hellenic Trench. In: Wezel, F.C. (Ed.), Sedimentary Basins of Mediterranean Margins. Tecnoprint, Bologna, pp. 203 226. Feng, Z.Z., 2017a. A successful symposium of “multi-origin of soft-sediment deformation structures and seismites”. J. Palaeogeogr. 6 (1), 1 6. Feng, Z.Z., 2017b. Preface of the Chinese version of the seismite problem. J. Palaeogeogr. 6 (1), 7 11. Feng, Z.Z., 2017c. Preface: a brief review on 7 papers from the special issue of “The environmental significance of soft-sediment deformation” of the Sedimentary Geology 344 (2016). J. Palaeogeogr. 6 (4), 243 250. Feng, Z.-Z., 2019. Words of the Editor-in-Chief—some ideas about the comments and discussions of hyperpycnal flows and hyperpycnites. J. Palaeogeogr. 8 (3), 301 305. Feng, Z.Z., Bao, Z.D., Zheng, X.J., Wang, Y., 2016. Researches of soft-sediment deformation structures and seismites in China—a brief review. J. Palaeogeogr. 5 (4), 311 317. Fichman, M., 2013. Raindrop Imprints and Their Use in the Retrodeformation of Carboniferous Trace Fossils (M. S. thesis). University of Connecticut Graduate School, Storrs, p. 89. Fichman, M.E., Crespi, J.M., Getty, P.R., Bush, A.M., 2015. Retrodeformation of Carboniferous trace fossils from the Narragansett Basin, United States, using raindrop imprints and bedding-cleavage intersection lineation as strain markers. Palaios 30, 574 588. Fisher, R.V., 1971. Features of coarse-grained, high-concentration fluids and their deposits. J. Sediment. Petrol. 41, 916 927. Fisher, R.V., 1983. Flow transformations in sediment gravity flows. Geology 11, 273 274. Fisher, R.V., 1995. Decoupling of pyroclastic currents: hazards assessments. J. Volcanol. Geotherm. Res. 66, 257 263. Fisher, R.V., Mattinson, J.M., 1968. Wheeler gorge turbidite conglomerate series California—inverse grading. J. Sediment. Petrol. 38, 1013 1023. Flint, R.F., Sanders, J.E., Rodges, J., 1960. Diamictite: a substitute term for symmictite. Geol. Soc. Am. Bull. 71, 1809 1810. Flood, R.D., Hollister, C.D., 1974. Current-controlled topography on the continental margin off the eastern United States, in Burke, C.A. In: Drake, C.L. (Ed.), The Geology of Continental Margins. Springer-Verlag, New York, pp. 197 205. Flood, R.D., Manley, P., Kowsmann, R.O., Appi, C.J., Pirmez, C., 1991. Seismic facies and Late Quaternary growth of Amazon submarine fan. In: Weimer, P., Link, M.H. (Eds.), Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems. Springer-Verlag, New York, pp. 415 433. Flood, R.D., Piper, D.J.W., Klaus, A., Burns, S.J., Busch, W.H., Cisowski, S.M., et al., 1995. Proceedings of the ocean drilling program. Ocean drilling program, College Station, Texas, p. 1233. Initial Reports 155. Flores, G., 1955. Discussion of a paper by F. Beneo: Les resultants des etudes pour la recherche´ pe´trolifeˆre en Sicilie: Proceedings of the 4th World Petroleum Congress, Sec. 1, pp. 121 122.

Bibliography

493

Folk, R.L., 1968. Petrology of Sedimentary Rocks. Hemphill, Austin, TX, p. 170. Fodor, R.V., Thiede, J., 1977. Volcanic breccia from DSDP Site 357: implications for the composition and origin of the Rio Grande Rise. In: Supko, P.R., Perch-Nielsen, K., et al.,Initial Reports of the Deep Sea Drilling Project, vol. 39. United States Government Printing Office, Washington, DC, pp. 537 543. Fonnesu, M., Palermo, D., Galbiati, M., Marchesini, M., Bonamini, E., Bendias, D., 2020. A new world-class deepwater play-type, deposited by the syndepositional interaction of turbidity flows and bottom currents: the giant Eocene Coral Field in northern Mozambique. Mar. Petrol. Geol. 111, 179 201. Ford, C., Bryant, G., Nick, K.E., 2016. Architectural evidence of dune collapse in the Navajo Sandstone, Zion National Park, Utah. Sediment. Geol. 344, 222 236. Forel, F.A., 1885. Les ravins sous-lacustres des fleuves glaciaires. Comptes Rendus de. l’Acade´mie des. Sci. Paris. 101 (16), 725 728. Forel, F.A., 1892. Le Feman, Monographie Limnologique, vol. 1, Geographie, Hydrographie, Geologie, Climatologie, Hydrologie. F. Rouge, Lausanne, p. 543. Foreman, M.G.G., Callendar, W., MacFadyen, A., Hickey, B.M., Thomson, R.E., di Lorenzo, E., 2008. Modeling the generation of the Juan de Fuca Eddy. J. Geophys. Res. Ocean. 113 (C3), CiteID C03006. Forristall, G.Z., Hamilton, R.C., Vardone, V.J., 1977. Continental shelf currents in tropical storm delia: observations and theory. J. Phys. Oceanogr. 87, 532 546. Fortuin, A.R., Dabrio, C.J., 2008. Evidence for Late Messinian seismites, Nijar Basin, south-east Spain. Sedimentology 55, 1595 1622. Fossen, H., 2010. Deformation bands formed during soft-sediment deformation: observations from SE Utah. Mar. Petrol. Geol. 27, 215 222. Fossati, M., Piedra-Cueva, I., 2013. A 3D hydrodynamic numerical model of the Rı´o de la Plata and Montevideos coastal zone. Appl. Math. Model. 37, 1310 1332. Fossati, M., Cayocca, F., Piedra-Cueva, I., 2014. Fine sediment dynamics in the Rı´o de la Plata. Adv. Geosci. 39, 75 80. ˇ Framinan, M.B., Brown, O.B., 1996. Study of the Rı´o de la Plata turbidity front, part I: spatial and temporal distribution. Cont. Shelf Res. 16, 1259 1282. Frey-Martinez, J., Cartwright, J., Hall, B., 2005. 3D seismic interpretation of slump complexes: examples from the continental margin of Israel. Basin Res. 17, 83 108. Friedman, G.M., Sanders, J.E., Kopaska-Merkel, D.C., 1992. Principles of Sedimentary Deposits: Stratigraphy and Sedimentology. McMillan Publishing Company, New York, p. 717. Frohlich, C., Hornbach, M.J., Taylor, F.W., Shen, C.-C., Moala, A., Morton, A.E., et al., 2009. Huge erratic boulders in Tonga deposited by a prehistoric tsunami. Geology 37, 131 134. Fruth, L.S., 1965. The 1929 Grand Banks Turbidite and the Sediments of the Sohm Abyssal Plain (M.Sc. thesis). Columbia University, New York, p. 258. Fudol, Y.A.H., Zhao, Y., Liu, H., Zhou, S., Li, Y., Li, X., 2019. Origin and reservoir properties of deep-water gravity flow sediments in the Upper Triassic Ch6 Ch7 members of the Yanchang Formation in the Jinghe Oilfield, the Southern Ordos Basin, China. Energy Explor. Exploit. 37 (4), 1227 1252. Fuhrmann, A., Kane, I.A., Clare, M.A., Ferguson, R.A., Schomacker, E., Bonamini, E., Contreras, F.A., et al., 2020. Hybrid turbidite-drift channel complexes: An integrated multiscale model:. An integrated multiscale model: Geology 48, 562 568. Available from: https://doi.org/10.1130/G47179.1. Galay, V., 1987. Erosion and Sedimentation in the Nepal Himalaya, 11. Kefford Press, Singapore, p. 10. Galaziy, G.I. (Ed.), 1993. Baikal Atlas. Federal Agency for Geodesy and Cartography, Moscow. Galloway, W.E., 1976. Sediments and stratigraphic framework of the Copper River fan-delta, Alaska. J. Sediment. Petrol. 46, 726 737. Galloway, W.E., 1989. Genetic stratigraphic sequences in basin analysis, I. Architecture and genesis of floodingsurface bounded depositional units. AAPG Bull. 73, 125 142. Gamboa, D., Alves, T., Cartwright, J., Terrinha, P., 2010. MTD distribution on a ‘passive’ continental margin: the Espı´rito Santo Basin (SE Brazil) during the Palaeogene. Mar. Petrol. Geol. 27, 1311 1324. Gangadhara Rao, L.V., Shree Ram, P., 2005. Upper Ocean Physical Processes in the Tropical Indian Ocean: A Monograph Prepared Under CSIR Emeritus Scientist Scheme:. National Institute of Oceanography, Regional Centre, Visakhapatnam, 68 p. ,http://drs.nio.org/drs/bitstream/2264/93/1/UOPP.pdf. (accessed 31.05.10.).

494

Bibliography

Gani, R., 2004. From turbid to lucid: a straightforward approach to sediment gravity flows and their deposits. Sediment. Rec. 3, 4 8. Gao, Z., Eriksson, K.A., 1991. Internal-tide deposits in an Ordovician submarine channel: previously unrecognized facies? Geology 19, 734 737. Gao, Z., Eriksson, K.A., He, Y., Luo, S., Guo, J., 1998. Deep-Water Traction Current Deposits—A Study of Internal Tides, Internal Waves, Contour Currents and Their Deposits. Science Press, Beijing and New York, Utrecht, Tokyo, 128 pp. Gao, W., Li, G., Wang, X., Sun, T., Liu, Y., Cao, L., 2014. Sedimentary characteristics of the hyperpycnal flow in the Modern Yellow River Delta. Indian J. Geo-Mar. Sci. 43 (8), 1438 1448. Gao, S., Wang, D., Yang, Y., Zhou, L., Zhao, Y., Gao, W., et al., 2015. Holocene sedimentary systems on a broad continental shelf with abundant river input: process product relationships. In: Clift, P.D., Harff, J., Wu, J., Qui, Y. (Eds.), River-Dominated Shelf Sediments of East Asian Seas., 429. Geological Society, London, pp. 223 259. Special Publications. Gao, Y., Jiang, Z., Best, J.L., Zhang, J., et al., 2020. Soft-sediment deformation structures as indicators of tectonovolcanic activity during evolution of a lacustrine basin: A case study from the Upper Triassic Ordos Basin, China. Marine and Petroleum Geology 115, 104250. Available from: https://doi.org/10.1016/j. marpetgeo.2020.104250. Garcı´a, M., Herna´ndez-Molina, F.J., Llave, E., Stow, D.A.V., Leo´n, R., Ferna´ndez-Puga, M.C., et al., 2009. Contourite erosive features caused by the Mediterranean Outflow Water in the Gulf of Ca´diz: quaternary tectonic and oceanographic implications. Mar. Geol. 257, 24 40. Gardner, G.H.F., Gardner, L.W., Gregory, A.R., 1974. Formation velocity and density the diagnostic basics for stratigraphic traps. Geophysics 39, 770 780. Gardner, J.V., Bohannon, R.G., Field, M.E., Masson, D.G., 1996. The morphology, processes, and evolution of Monterey Fan: a revisit. In: Gardner, J.V., Field, M.E., Twichell, D.C. (Eds.), Geology of the United States’ Sea Floor: The View From GLORIA. Cambridge University Press, New York, pp. 193 220. Gardner, J.V., Mayer, L.A., Hughes Clarke, J.E., 2000. Morphology and processes in Lake Tahoe (California Nevada). GSA Bull. 112, 736 746. Gardner, J.V., Dartnell, P., Christopher Stone, J., Mayer, L.A., Hughes Clark, J.E., 2002. Bathymetry and selected perspective views Offshore Greater Los Angeles, California. U.S. Geological Survey, Water-Resources Investigations Report 02 4126. Garrett, C., 2003. Internal tides and ocean mixing. Science 301, 1858 1859. Garrett, C., Kunze, E., 2007. Internal tide generation in the deep ocean. Annu. Rev. Fluid Mech. 39, 57 87. Garrison, R.E., Schreiber, B.C., Bernoulli, D., Fabricius, F.H., Kidd, R.B., Melieres, F., et al., 1978. Sedimentary petrology and structures of Messinian evaporitic sediments in the Mediterranean Sea. Initial Reports of the Deep Sea Drilling Project, Leg 42. United States Government Printing Office, Washington, DC, pp. 571 611, ,https://doi.org/10.2973/dsdp.proc.42-1.123.1978.. Gaudin, M., Berne´, S., Jouanneau, J.-M., Palanques, A., Puig, P., Mulder, T., et al., 2006. Massive sand beds deposited by densewater cascading in the Bourcart canyon head, Gulf of Lions (northwestern Mediterranean Sea). Mar. Geol. 234, 111 128. Gavrilov, Y.O., 2017. Reflection of seismic paleoevents in Mesozoic Cenozoic terrigenous sequences of the Northern Caucasus. Lithology Miner. Resour. 52, 1 19. Gawthorpe, R.L., Clemmey, H., 1985. Geometry of submarine slides in the Bowland Basin (Dinantian) and their relation to debris flows. J. Geol. Soc. Lond. 142, 555 565. Gee, M.J.R., Masson, D.G., Watts, A.B., Allen, P.A., 1999. The Saharan debris flow: an insight into the mechanics of long-runout submarine debris flows. Sedimentology 46, 317 335. Gee, M.J.R., Gawthorpe, R.L., Friedmann, S.J., 2006. Triggering and evolution of a giant submarine landslide, offshore Angola, revealed by 3D seismic stratigraphy and geomorphology. J. Sediment. Res. 76, 9 19. Gee, M.J.R., Uy, H.S., Warren, J., Morley, C.K., Lambiase, J.J., 2007. The Brunei slide: a giant submarine landslide on the North West Borneo Margin revealed by 3D seismic data. Mar. Geol. 246, 9 23. Geertsema, M., Schwab, J.W., Blais-Stevens, A., Sakals, E., 2009. Landslides impacting linear infrastructure in west central British Columbia. Nat. Hazards 48, 59 72. Geist, E.L., 2005. Local Tsunami Hazards in the Pacific Northwest From Cascadia Subduction Zone Earthquakes: U.S. Geological Survey Professional Paper 1661-B, 17 p.

Bibliography

495

Gelfenbaum, G., Jaffe, B., 2003. Erosion and sedimentation from the 17 July, 1998 Papua New Guinea Tsunami. Pure Appl. Geophys. 160, 1969 1999. GEOMAR, 2012. 6th International Symposium on Submarine Mass Movements and Their Consequences, 23 25 September 2013, GEOMAR, Kiel. ,http://www.igcp585.org/6th-issmmtc/2nd-circular. (accessed 01.02.13.). Gervais, A., Mulder, T., Savoye, B., Gonthier, E., 2006. Sediment distribution and evolution of sedimentary processes in a small sandy turbidite system (Golo system, Mediterranean Sea): implications for various geometries based on core framework. Geo-Mar. Lett. 26, 373 395. Geyer, W.R., Hill, P., Milligan, T.G., Traykovski, P.A., 2000. The structure of the Eel River plume during flood. Cont. Shelf Res. 20 (16), 2067 2093. Ghibaudo, G., 1992. Subaqueous sediment gravity flow deposits: practical criteria for their field description and classification. Sedimentology 39, 423 454. Gihm, Y.S., Hwang, I.G., 2016. Lacustrine hyperpycnal flow deposits after explosive volcanic eruptions, Cretaceous Beolkeum Member, Wido Island, Korea. Geosci. J. 20 (2), 157 166. Gill, A.E., 1982. Atmosphere-Ocean Dynamics. International Geophysics Series, vol. 30. Academic Press, An Imprint of Elsevier, San Diego, CA, 662 p. Gindre-Chanu, L., Warren, J.K., Puigdefabregas, C., ISharp, I.R., Peacock, D.C.P., Swart, R., et al., 2014. Diagenetic evolution of Aptian evaporites in the Namibe Basin (south-west Angola). Sedimentology 62, 204 233. Glade, T., Anderson, M.G., Crozier, M.J. (Eds.), 2005. Landslide Hazard and Risk. John Wiley & Sons, West Sussex. Gladkov, A.S., Lobova, E.U., Deev, E.V., Korzhenkov, A.M., Mazeika, J.V., Abdieva, S.V., et al., 2016. Earthquakeinduced soft-sediment deformation structures in Late Pleistocene lacustrine deposits of Issyk-Kul Lake (Kyrgyzstan). Sediment. Geol. 344, 112 122. Gladstone, C., Pritchard, D., 2010. Patterns of deposition from experimental turbidity currents with reversing buoyancy. Sedimentology 57, 53 84. Glossary of Coastal Terminology, 1998. Plunging wave. Department of Ecology Publication. Department of Ecology, State of Washington, pp. 98 105. Goguel, J., 1978. Scale-dependent rockslide mechanisms, with emphasis on the role of pore fluid vaporization. In: Voight, B. (Ed.), Rockslides and Avalanches, 1, Natural Phenomena. Elsevier, Amsterdam, pp. 693 705. Goldfinger, C., Grijalva, K., Bu¨rgmann, R., Morey, A.E., Johnson, J.E., Hans Nelson, C., et al., 2008. Late Holocene Rupture of the Northern San Andreas Fault and Possible Stress Linkage to the Cascadia Subduction Zone. Bull. Seismol. Soc. Am. 98, 861 889. Go´mez-Enri, J., Va´zquez, A., Bruno, M., Mariscal, L., Villares, P., 2007. Characterization of internal waves in the Strait of Gibraltar, using SAR and in-situ measurements. In: Proceedings of the ‘Envisat Symposium 2007’, Montreux, 23 27 April 2007: ESA SP-636, July 2007. ,http://earth.esa.int/workshops/envisatsymposium/ proceedings/posters/4P14/464335go.pdf. (accessed 07.04.14.). Gong-Yiming, 1988. Tempestite, seismite, and tsunamite: a discussion of seceral sedimentological terms. Geol. Rev. 34, 481 482. Gonthier, E.G., Fauge`res, J.-C., Stow, D.A.V., 1984. Contourite facies of the Faro Drift, Gulf of Cadiz. In: Stow, D. A.V., Piper, D.J.W. (Eds.), Fine-Grained Sediments: Deep-Water Processes and Facies, 15. Geological Society of London Special Publication, pp. 275 292. ˇ Gonzalez-Silvera, A., Santamaria-del-Angel, E., Milla´n-Nu´nez, R., 2006. Spatial and temporal variability of the Brazil Malvinas Confluence and the La Plata Plume as seen by SeaWiFS and AVHRR imagery. J. Geophys. Res. 111, C06010. Goodbred Jr., S.L., 2003. Response of the Ganges dispersal system to climate change: a source-to-sink view since the last interstade. Sediment. Geol. 162, 83 104. Gordon, A.L., 1985. Indian-Atlantic transfer of thermocline water at the Agulhas Retroflection. Science 227, 1030 1033. Gordon, A.L., 1986. Is there a global scale ocean circulation? EOS Trans. Am. Geophys. Union 67, 109 110. Gordon, A.L., 2005. Oceanography of the Indonesian Seas and their throughflow. Oceanography 18, 14 27. Gordon, A.L., 2001. Bottom water formation, Encyclopedia of Ocean Sciences, second ed. Elsevier, pp. 334 340. Gordon, A.L., 2013. Bottom water formation. In: Steele, J.H., Turekian, K.K., Thorpe, S.A. (Eds.), Encyclopedia of Ocean Sciences, second ed. Academic, San Diego, CA, pp. 415 421. ,https://doi.org/10.1016/B978-0-12409548-9.04019-7..

496

Bibliography

Gordon, A.L., 2019. Bottom water formation. Encyclopedia of Ocean Sciences, High-Latitude Systems and Cryosphere. Elsevier, pp. 120 125. Gordon, A.L., Sprintall, J., Van Aken, H.M., Susanto, D., Wijffels, S., Molcard, R., et al., 2010. The Indonesian throughflow during 2004 2006 as observed by the INSTANT program. Dyn. Atmos. Ocean. 50, 115 128. Gore, R.H., 1992. The Gulf of Mexico. Pineapple Press, Inc., Sarasota, FL, p. 384. Goren, L., Aharonov, E., 2007. Long runout landslides: the role of frictional heating and hydraulic diffusivity. Geophys. Res. Lett. 34, L07301, doi:10.1029/2006GL028895, 2007. Gradmann, S., Beaumont, C., Ings, S.J., 2012. Coupled fluid flow and sediment deformation in margin-scale salt-tectonic systems: 1. Development and application of simple, single-lithology models. Tectonics 31, C4010. Graham, S.A., Dickinson, W.R., Ingersoll, R.V., 1975. Himalayan-Bengal model for flysch dispersal in the Appalachian-Ouachita system. Geol. Soc. Am. Bull. 86, 273 286. Grajales-Nishimura, Cedillo-Pardo, E., Rosales-Domı´nguez, C., Mora´n-Zenteno, D., Alvarez, W., Claeys, P., et al., 2000. Chicxulub impact: the origin of reservoir and seal facies in the southeastern Mexico oil fields. Geology 28, 307 310. Gray, W.M., 1979. Hurricanes: their formation, structure and likely role in the tropical circulation. In: Shaw, D.B. (Ed.), Meteorology Over Tropical Oceans. Royal Meteorological Society, pp. 155 218. Greb, S.F., Archer, A.W., 2007. Soft-sediment deformation produced by tides in a meizoseismic area, Turnagain Arm, Alaska. Geology 35, 435 438. Greene, M., Power, M., Youd, T.L., 1994. Earthquake Basics Brief No. 1: Liquefaction: What It Is and What To Do About It. Earthquake Engineering Research Institute (EERI) Publication, Oakland, CA, p. 8. Greene, H.G., Murai, L.Y., Watts, P., Maher, N.A., Fisher, M.A., Paull, C.E., et al., 2006. Submarine landslides in the Santa Barbara Channel as potential tsunami sources. Nat. Hazards Earth Syst. Sci. 6, 63 88. Gregory, M.R., 1969. Sedimentary features and penecontemporaneous slumping in the Waitemata Group, Whangaparaoa Peninsula, North Auckland, New Zealand. N. Z. J. Geol. Geophys. 12, 248 282. Available from: https://doi.org/10.1080/00288306.1969.10420236. Griffiths, J.C., 1960. Aspects of Measurement in the Geosciences, 29. Mineral Industries. College of Mineral Industries, The Pennsylvania State University, No. 4, 4 8. Griswold, J.P., Iverson, R.M., 2008. Mobility Statistics and Automated Hazard Mapping for Debris Flows and Rock Avalanches. U.S. Geological Survey Scientific Investigations Report 2007-5276, Reston, VI, p. 68. Grotzinger, J., Jordan, T.H., Press, F., Siever, R., 2007. Understanding Earth, fifth ed. W. H. Freeman & Company, New York, p. 579. Gruszka, B., Fard, A.M., van Loon, A.J., 2016. A fluctuating ice front over an esker near Ryssjo¨n (S Sweden) as a cause of a giant load cast. Sediment. Geol. 344, 47 56. Guest, J.E., 1971. Geology of the farside crater Tsiolkovsky. In: Fielder, G. (Ed.), Geology and Physics of the Moon. Elsevier, Amsterdam, pp. 93 103. Guo, L.C., van der Wegen, M., Roelvink, J.A., He, Q., 2014. The role of river flow and tidal asymmetry on 1-D estuarine morphodynamics. J. Geophys. Res. Earth Surf. 119, 2315 2334. Guo, L.C., van der Wegen, M., Jay, D.A., Matte, P., Wang, Z.B., Roelvink, D., et al., 2015. River-tide dynamics: exploration of nonstationary and nonlinear tidal behavior in the Yangtze River estuary. J. Geophys. Res. Ocean. 120, 3499 3521. Gupta, S.K., 2006. Basin architecture and petroleum system of Krishna Godavari Basin, east coast of India. Lead. Edge 25, 830 837. Gutenberg, B.A., 1939. Tsunamis and earthquakes. Seismol. Soc. Am. Bull. 29, 517 526. Gyory, J., Beal, L.M., Bischof, B., Mariano, A.J., Ryan, E.H., 2004. The Agulhas Current. Ocean Surface Currents. ,http://oceancurrents.rsmas.miami.edu/atlantic/agulhas.html.. Gyory, J. Mariano, A.J., Ryan, E.H., 2005. “The Gulf Stream.” Ocean Surface Currents. ,http://oceancurrents. rsmas.miami.edu/atlantic/gulf-stream.html.. Hacker, D.B., Biek, R.F., Rowley, P.D., 2014. Catastrophic emplacement of the gigantic Markagunt gravity slide, southwest Utah (USA): implications for hazards associated with sector collapse of volcanic fields. Geology 42 (11), 943 946. Haflidason, H.L., Reidar Sejrup, H.P., Forsberg, C.F., Bryn, P., 2005. The dating and morphometry of the storegga slide. Mar. Petrol. Geol. 22, 123 136.

Bibliography

497

Hallworth, M.A., Huppert, H.E., 1998. Abrupt transitions in high-concentration, particle-driven gravity currents. Phys. Fluids 10, 1083 1087. Hamilton, P., 2004. Hydrodynamic modeling of the Columbia River Estuary. Final Report on the Simulation Work Unit of the Columbia River Estuary Data Development Program. Science Applications, Inc., Raleigh, NC, p. 433. Hamilton, P.B., Strom, K.B., Hoyal, D., 2015. Hydraulic and sediment transport properties of autogenic avulsion cycles on submarine fans with supercritical distributaries. J. Geophys. Res. Earth Surf. 120 (7), 1369 1389. Hampton, M.A., 1972. The role of subaqueous debris flows in generating turbidity currents. J. Sediment. Petrol. 42, 775 793. Hampton, M.A., 1975. Competence of fine-grained debris flows. J. Sediment. Petrol. 45, 834 844. Hampton, M.A., Lee, H.J., Locat, J., 1996. Submarine landslides. Rev. Geophys. 34, 33 59. Hanselman, D.H., Ryan, W.B.F., 1983. Atlantic 3800-meter radioactive waste disposal site survey: sedimentary, micromorphologic and geophysical analysis. Report EPA 520/1-83-017. U. S. Environmental Protection Agency, Washington, DC, 43 p. Hansen, M.J., 1984. Strategies for classification of landslides. In: Brunsden, D., Prior, D.B. (Eds.), Slope Instability. John Wiley & Sons Ltd, Chichester, pp. 1 25. Hanzawa, H., Kishida, T., 1981. Fundamental considerations of undrained strength characteristics of alluvial marine clays. Soils Found., Jpn Soc. Soil. Mech. Found. Eng. 21 (1), 39 50. Haq, B.U., Hardenbol, J., Vail, P.R., 1988. Mesozoic and Cenozoic chronostratigraphy and cycles of sea-level change. In: Wilgus, C.K., Hastings, B.S., St Kendall, C.G.C., Posamentier, H.W., Ross, C.A., Van Wagoner, J.C. (Eds.), Sea-Level Changes: An Integrated Approach, vol. 42. Society of Economic Paleontologists and Mineralogists, Special Publication, pp. 71 108. Harland, W.B., Herod, K.N., Krinsley, D.H., 1966. The definition and identification of tills and tillites. Earth-Sci. Rev. 2, 225 256. Harmelin-Vivien, M.L., Laboute, P., 1986. Catastrophic impact of hurricanes on atoll outer slopes in the Tuamotu (FP). Coral Reefs 5, 55 62. Harms, J.C., Fahnestock, R.K., 1965. Stratification, bed forms, and flow phenomena (with an example from the Rio Grande). In: Middleton, G.V. (Ed.), Primary Sedimentary Structures and Their Hydrodynamic Interpretation, 12. SEPM Special Publication, pp. 84 115. Harms, J.C., Southard, J.B., Spearing, D.R., Walker, R.G., 1975. Depositional environments as interpreted from primary sedimentary structures and stratification sequences. SEPM Short Course No. 2, Dallas, TX, p. 161. Harms, J.C., Southard, J.B., Walker, R.G., 1982. Structures and sequences in castic rocks. SEPM Short Course No. 9, Calgary, pp. 8 51. Harris, P.T., Whiteway, T., 2011. Global distribution of large submarine canyons: geomorphic differences between active and passive continental margins. Mar. Geol. 285, 69 86. Harrison, J.V., Falcon, N.L., 1938. An ancient landslip at Saidmarreh in southwestern Iran. J. Geol. 46, 296 309. Harrison, K.P., Grimm, R.E., 2003. Rheological constraints on martian landslides. Icarus 163, 347 362. Hart, B., 2008. Stratigraphically significant attributes. Lead. Edge 27, 320 324. Hatayama, T., 2004. Transformation of the Indonesian throughflow water by vertical mixing and its relation to tidally generated internal waves. J. Oceanogr. 60, 569 585. Haver, S., 2003. Freak wave event at Draupner Jacket January 1995. ,http://folk.uio.no/karstent/seminarV05/ Haver2004.pdf. (accessed 28.06.11.). Haver, S., 2004. A possible freak wave event measured at the Draupner Jacket January 1 1995. In: Proceedings of Rogue Waves, IFREMER, Brest, pp. 20 22. Havil, J., 2014. The Irrationals: A Story of the Numbers You Can’t Count On. Princeton University Press, Princeton, NJ, 312 p. Hawati, P., Nugroho Sugianto, D., Anggoro, S., Wirasatriya, A., Widada, S., 2017. Waves induce sediment transport at coastal region of Timbulsloko Demak. IOP Conference Series. Earth Environ. Sci. 55 (1), 012048. Hay, A., Burling, R.W., Murray, J.W., 1982. Remote acoustic detection of a turbidity current surge. Science 217, 833 835. Hayakawa, S.I. 1968. Use the Right Word, Reader’s digest association, incorporation, New York, 726. Hayter, E.J., Mathew, R., Hallden, J., Garland, E., Salerno, H., Svirsky, S.C., 2006. Evaluation of the State-of-theArt Contaminated Sediment Transport and Fate Modeling System. EPA/600/R-06/108 September 2006.

498

Bibliography

National Exposure Research Laboratory, Office of Research and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC, p. 140. He, B.Z., Qiao, X.F., 2015. Advances and overview of the study on paleo-earthquake events: a review of seismites. Acta Geol. Sin. (Engl. Ed.) 89, 1702 1746. He, Y.-B., Luo, J.-X., Li, X.-D., Gao, Z.-Z., Wen, Z., 2011. Evidence of internal-wave and internal-tide deposits in the Xujiajuan Formation of the Xiangshan Group, Ningxia, China. Geo-Mar. Lett. 31, 509 523. Heath, R.A., 1975. Surface oscillations of lake Wakatipu, New Zealand. N. Z. J. Mar. Freshw. Res. 9 (2), 223 238. Available from: https://doi.org/10.1080/00288330.1975.9515560. Hecker, C.K.B., Golubev, V., 1991. Hydrothermal vents in Lake Baikal. Nature 350, 281. Heezen, B.C., 1956. Corrientes de turbidez del Rio Magdelena. Soc. Geografica de. Columbia Bull. 51 52, 135 143. Heezen, B.C., 1959. Dynamic processes of abyssal sedimentation; erosion, transportation, and re-deposition on the deep-sea floor. Geophys. J. R. Astron. Soc. 2, 142 163. Heezen, B.C., Ewing, M., 1952. Turbidity currents and submarine slumps, and the 1929 Grand Banks earthquake. Am. J. Sci. 250, 849 873. Heezen, B.C., Hollister, C.D., 1963. Evidence of deep-sea bottom currents from abyssal sediments. Abstracts of papers, International Association of Physical Oceanography, 13th General Assembly, International Union Geodesy and Geophysics 6, p. 111. Heezen, B.C., Hollister, C.D., 1964. Deep sea current evidence from abyssal sediments. Mar. Geol. 1, 141 174. Heezen, B.C., Hollister, C.D., 1971. The Face of the Deep. Oxford University Press, New York, p. 659. Heezen, B.C., Ericson, D.B., Ewing, M., 1954. Further evidence for a turbidity current following the 1929 Grand Banks earthquake. Deep. Sea Res. 1, 193 202. Heezen, B.C., Tharp, M., Ewing, M., 1959. The floors of the oceans: I. The North Atlantic. Geological Society of America Special Paper 65, 122 p. Heezen, B.C., Hollister, C.D., Ruddiman, W.F., 1966. Shaping of the continental rise by deep geostrophic contour currents. Science 152, 502 508. Heezen, B.C., Menzies, R.J., Schneider, E.D., Ewing, W.M., Granelli, N.C., 1964. Congo submarine canyon. AAPG Bull. 48, 1126 1149. Heezen, B.C., Fischer, A.G., Boyce, R.E., Bukry, D., Douglas, R.G., Garrison, R.E., et al., 1971. Site 44, Leg VI, Deep Sea Drilling Project. In: Fischer, A.G. (Ed.), Initial Reports of the Deep Sea Drilling Project, vol. 6. United States Government Printing Office, Washington, DC, pp. 17 39. Heim, A., 1882. Der Bergsturz von Elm. Zeit. der Deutsch. Geolog. Gesell. 34, 74 115. Heim, A., 1932. Landslides and Human Lives (Bergsturz und Menschenleben). Bi-Tech Publishers, Vancouver, BC, p. 196. Heinio¨, P., Davies, R.J., 2009. Trails of depressions and sediment waves along submarine channels on the continental margin of Espirito Santo Basin, Brazil. Geol. Soc. Am. Bull. 121, 698 711. Helwig, J., 1970. Slump folds and early structures, northeastern Newfoundland Appalachians. J. Geol. 78, 172 187. Hempton, M.R., Dewey, J.F., 1983. Earthquake-induced deformational structures in young lacustrine sediments East-Anatolian Fault, south-east Turkey. Tectonophysics 98, T7 T14. Henkel, D.J., 1970. The role of waves in causing submarine landslide. Geotechnique 20, 75 80. Henrich, R., Hu¨neke, H., 2011. Chapter 5. Hemipelagic advection and periplatform sedimentation. In: Hueneke, H., Mulder, T. (Eds.), Deep-Sea Sediments, 63. Elsevier, Amsterdam, pp. 353 396. Developments in Sedimentology. Henstock, T.J., McNeill, L.C., Tappin, D.R., 2006. Seafloor morphology of the Sumatran subduction zone: surface rupture during megathrust earthquakes? Geology 34, 485 488. Available from: https://doi.org/10.1130/22426.1. Henze, C., 2015. Antarctic Bottom Water in CMIP5 Models: Characteristics, Formation, Evolution (Ph.D. thesis). University of East Anglia, School of Environmental Sciences, Norwich. 226 p. Herdendorf, C.E., 1990. Distribution of the world’s large lakes. In: Tilzer, M.M., Serruya, C. (Eds.), Large Lakes: Ecological Structure and Function. Springer-Verlag, New York, pp. 3 38. Heritier, F.E., Lossel, P., Wathne, E., 1979. Frigg field—large submarine-fan trap in lower eocene rocks of the North Sea Viking Graben. AAPG Bull. 63, 1999 2020. Herna´ndez-Molina, F.J., Llave, E., Somoza, L., Ferna´ndez-Pug, M.C., Maestro, A., Leo´n, R., et al., 2003. Looking for clues to paleoceanographic imprints: a diagnosis of the Gulf of Cadiz contourite depositional systems. Geology 31 (1), 19 22.

Bibliography

499

Herna´ndez-Molina, F.J., Llave, E., Stow, D.A.V., Garcı´a, M., Somoza, L., Va´zquez, J.T., et al., 2006. The contourite depositional system of the Gulf of Cadiz: a sedimentary model related to the bottom current activity of the Mediterranean Outflow Water and the continental margin characteristics. Deep-Sea Res. II 53, 1420 1463. Available from: https://doi.org/10.1016/j.dsr2.2006.04.016. Herna´ndez-Molina, F.J., Maldonado, A., Stow, D.A.V., 2008. Abyssal plain contourites. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites. Elsevier, Amsterdam, pp. 345 378. Developments in Sedimentology 60 Chapter 18. Herna´ndez-Molina, F.J., Serra, N., Stow, D.A.V., Ercilla, G., Llave, E., Van Rooij, D., 2011. Along-slope oceanographic processes and sedimentary products around Iberia. Geo-Mar. Lett. 31, 315 341. Herna´ndez-Molina, F.J., Stow, D.A.V., Alvarez-Zarikian, C., 2013. IODP Expedition 339 in the Gulf of Cadiz and off West Iberia: decoding the environmental significance of the Mediterranean outflow water and its global influence. Sci. Drill. 16, 1 11. Herna´ndez-Molina, F.J., Llave, E., Preu, B., Ercilla, G., Fontan, A., Bruno, M., et al., 2014. Contourite processes associated with the Mediterranean Outflow Water after its exit from the Strait of Gibraltar: global and conceptual implication. Geology 42 (3), 227 230. Herrington, P.M., Pederstad, K., Dickson, J.A.D., 1991. Sedimentology and diagenesis of resedimented and rhythmically bedded chalks from the Eldfisk Field, North Sea Central Graben. AAPG Bull. 75, 1661 1674. Hesse, R., Schacht, U., 2011. Early diagenesis of deep-sea sediments. In: Huneke, H., Mulder, T. (Eds.), Deep-Sea Sediments, Developments in Sedimentology 63, pp. 559 713. Hibsch, C., Alvarado, A., Yepes, H., Perez, V.H., 1997. Holocene liquefaction and soft-sediment deformation in Quito (Ecuador): a paleoseismic history recorded in lacustrine sediments. J. Geodyn. 24, 259 280. Hickey, H., 2013. Tracking sediments’ fate in largest-ever dam removal. UW News (U. Washington), Seattle, WA, 7 March 2013. With technical information on sediment plumes from two UW oceanographers: Charles Nittrouer and Andrea Ogston. ,http://www.washington.edu/news/2013/03/07/tracking-sediments-fate-inlargest-ever-dam-removal/. (accessed 03.03.19.). Higgs, R., 2009. Enigmatic Brushy Canyon Formation, Delaware Basin, USA: Poor Outcrop Analog for Deep-Sea Turbidite Reservoirs. AAPG Search and Discovery Article #51097 (2015) Adapted from oral presentation given at AAPG International Conference & Exhibition, Rio de Janeiro, 15 18 November 2009. Higgs, R., 2010. ‘Hummocky cross-stratification-like structures in deep-sea turbidites: Upper Cretaceous Basque basins (Western Pyrenees, France)’ by Mulder et al. (2009), Sedimentology, 56, 997 1015: discussion. Sedimentology . Available from: https://doi.org/10.1111/j.1365-3091.2010.01163.x. Highland, L.M., Bobrowsky, P., 2008. The Landslide Handbook—A Guide to Understanding Landslides, 1325. U. S. Geological Survey Circular, Reston, VI, 129 p. Hilbert-Wolf, H.L., Roberts, E.M., Simpson, E.L., 2016. New sedimentary structures in seismites from SW Tanzania: evaluating gas- vs. water-escape mechanisms of soft deformation. Sediment. Geol. 344, 253 262. Hiscott, R.N., 1979. Clastic sills and dikes associated with deep-water sandstones, Tourelle Formation, Ordovician, Quebec. J. Sediment. Petrol. 49, 1 10. Hiscott, R.N., Middleton, G.V., 1979. Depositional mechanics of thick-bedded sandstones at the base of a submarine slope, Tourelle Formation (Lower Ordovician), Quebec, Canada. In: Doyle, L.J., Pilkey, O.H. (Eds.), Geology of Continental Slopes, 27. SEPM Special Publication, pp. 307 326. Hiscott, R.N., Aksu, A.E., 1994. Submarine debris flows and continental slope evolution in front of quaternary ice sheets, Baffin Bay, Canadian Arctic. AAPG Bull. 78, 445 460. Hiscott, R.N., Pickering, K.T., Bouma, A.H., Hand, B.M., Kneller, B.C., Postma, G., et al., 1997. Basin-floor fans in the North Sea, Sequence stratigraphic models vs. sedimentary facies, Discussion. AAPG Bull. 81, 662 665. Hizzett, J.L., Hughes Clarke, J.E., Sumner, E.J., Cartigny, M.J.B., Talling, P.J., Clare, M.A., 2018. Which triggers produce the most erosive, frequent, and longest runout turbidity currents on deltas? Geophys. Res. Lett. 45 (2), 855 863. Hollister, C.D., 1967. Sediment Distribution and Deep Circulation in the Western North Atlantic (Unpublished Ph.D. dissertation). Columbia University, New York, p. 467. Hollister, C.D., Heezen, B.C., 1972. Geological effects of ocean bottom currents: Western North Atlantic. In: Gordon, A.L. (Ed.), Studies in Physical Oceanography, vol. 2. Gordon and Breach, New York, pp. 37 66. Hollister, C.D., Johnson, D.A., Lonsdale, P.F., 1974. Current-controlled abyssal sedimentation: Samoan Passage, equatorial west Pacific. J. Geol. 82, 275 300.

500

Bibliography

Holsapple, K.A., 1993. The scaling of impact processes in planetary sciences. Annu. Rev. Earth Planet. Sci. 21, 333 373. Holzer, T.L., Youd, T.L., 2007. Liquefaction, ground oscillation, and soil deformation at the Wildlife Array, California. GSA Bull. 97, 961 976. Hong, Y., Adler, R., Huffman, G., 2007. Use of satellite remote sensing data in the mapping of global landslide susceptibility. J. Nat. Hazards 43, 245 256. Hooper, J.R., Suhayda, J.N., 2005. Hurricane induced seafloor failures in the Mississippi Delta. In: 2005 Offshore Hurricane Readiness and Recovery Conference, American Petroleum Institute, Houston, TX, 26 27 July 2005. ,http://www.mms.gov/tarprojects/559/AC-2_Soils-Geotechnical_Issues-Hooper.pdf. (accessed 31.05.10.). Hopkins, J., Lucas, M., Dufau, C., Sutton, M., Lauret, O., 2013. Detection and variability of the Congo River plume from satellite derived sea surface temperature, salinity, ocean colour and sea level. EGU General Assembly 2013, 7 12 April 2013, Vienna, Austria, id. EGU2013-3135. Hornbach, M.J., Lavier, L.L., Ruppel, C.D., 2007. Triggering mechanism and tsunamogenic potential of the Cape Fear Slide complex, U.S. Atlantic margin. Geochem. Geophy. Geosyst. 8 (12). Available from: https://doi.org/ 10.1029/2007GC001722. Hotchkiss, F.S., Wunsch, C., 1982. Internal waves in Hudson Canyon with possible geological implications. Deep. Sea Res. 29, 415 442. Hoitink, A.J.F., Jay, D.A., 2016. Tidal river dynamics: implications for deltas. Rev. Geophys. 54, 240 272. Hori, K.I., Zhao, Q., Wang, P., 2002. Architecture and evolution of the tide-dominated Changjiang (Yangtze) River delta, China. Sediment. Geol. 146, 249 264. Horowitz, D., 1982. Geometry and origin of large scale deformation structures in some ancient wind blown sand deposits. Sedimentology 29, 155 180. Available from: https://doi.org/10.1111/j.1365-3091.1982. tb01717.x. Houghton, P., Davis, C., McCaffrey, W., Barker, S., 2009. Hybrid sediment gravity flow deposits classification, origin and significance. Mar. Petrol. Geol. 26, 1900 1918. Howard, K.A., 1973. Avalanche mode of motion: implications from lunar examples. Science 180, 1052 1055. Howe, E., 1909. Landslides in the San Juan Mts. Colorado: including consideration of their causes and their classification. U.S. Geological Survey Professional Paper 67, p. 58. Hovland, M., Judd, A.G., 1988. Seabed Pockmarks and Seepages. Graham and Trotman, London, p. 293. Hovland, M., Svensen, H., Forsberg, C.F., Johansen, H., Fichler, C., Fossa˚, J.H., et al., 2005. Complex pockmarks with carbonate-ridges off mid-Norway: products of sediment degassing. Mar. Geol. 218, 191 206. Howe, J.A., 1995. Sedimentary processes and variations in slope-current activity during the last GlacialInterglacial episode on the Hebrides Slope, northern Rockall Trough, North Atlantic Ocean. Sediment. Geol. 96, 201 230. Howe, J.A., 2008. Methods for contourite research. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 19 33. Developments in Sedimentology. Howe, J.A., Humphrey, J.D., 1995. Photographic Evidence for Slope Current Activity on the Hebrides slope, North-East Atlantic Ocean. Scott. J. Geol. 30, 107 115. Howe, J.A., Stoker, M.S., Stow, D.A.V., 1994. Late Cenozoic sediment drift complex, northeast Rockall Trough, North Atlantic. Paleoceanography 9, 989 999. Howe, J.A., Pudsey, C.J., Cunningham, A.P., 1997. Pliocene-Holocene contourite deposition under the Antarctic circumpolar current, western Falkland trough, South Atlantic Ocean. Mar. Geol. 138 (1-2), 27 50. Howe, J.A., Stoker, M.S., Woolfe, K.J., 2001. Deep-marine seabed erosion and gravel lags in the Northwestern Rockall Trough, North Atlantic Ocean, J. Geol. Soc. Lond., 158. pp. 427 438. Hoyal, D.C.H., Demko, T., Postma, G., Wellner, R.W., Pederson, K., Abreu, V., et al., 2014. Evolution, architecture and stratigraphy of Froude supercritical submarine fans. In: AAPG Annual Convention and Exhibition, 6 9 April, Datapages/Search and Discovery Article #90189. Hsu¨, K.J., 1964. Cross-laminated sequence in graded bed sequence. J. Sediment. Petrol. 34, 379 388. Hsu¨, K.J., 1974. Me´langes and their distinction from olistostromes. In: Dott Jr., R.H., Shaver, R.H. (Eds.), Modern and Ancient Geosynclinal Sedimentation, 19. SEPM. Special Publication, pp. 321 333. Hsu¨, K.J., 1975. Catastrophic debris streams (sturzstroms) generated by rock falls. GSA Bull. 86, 129 140. Hsu¨, K.J., 1989. Physical Principles of Sedimentology. Springer-Verlag, New York, p. 233. Hsu¨, K.J., 2004. Physics of Sedimentology, second ed. Springer, Berlin, 240 p.

Bibliography

501

Hsu¨, K.J., 2008. Personal reminiscences on the history of contourites. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 11 17. Developments in Sedimentology. Hsu¨, K.J., Cita, M.B., Ryan, W.B.F., 1973. The origin of the Mediterranean evaporite. In: Ryan, W.B., Hsu¨, K.J. (Eds.), Initial Reports of the Deep Sea Drilling Project, vol. 13, Part 2. United States Government Printing Office, Washington, DC, pp. 1203 1231. ,https://doi.org/10.2973/dsdp.proc.13.143.1973.. Hubbard, D.K., 1992. Hurricane-induced sediment transport in open shelf tropical systems—an example from St. Croix, U.S. Virgin Islands. J. Sediment. Petrol. 62, 946 960. Hu, C.H., Ji, Z.W., Wang, T., 1998. Dynamic characteristics of sea currents and sediment dispersion in the Yellow River Estuary. Int. J. Sediment. Res. 13 (2), 20 30. Hubert, J.F., 1964. Textural evidence for deposition of many western North Atlantic deep-sea sands by oceanbottom currents rather than turbidity currents. J. Geol. 72, 757 785. Hubert-Ferrari, A., El-Ouahabi, M., Garcia-Moreno, D., Avsar, U., Altinok, S., Schmidt, S., et al., 2017. Earthquake imprints on a lacustrine deltaic system: the Ku¨rk Delta along the East Anatolian Fault (Turkey). Sedimentology 64 (5), 1322 1353. Hu¨eneke, H., Stow, D.A.V., 2008. Identification of ancient contourites: problems and palaeoceanographic significance. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites. Elsevier, Amsterdam, pp. 323 346. Developments in Sedimentology 60, Chapter 17. Hu¨neke, H., Henrich, R., 2011. Chapter 4 - Pelagic sedimentation in modern and ancient. In: Hueneke, H., Mulder, T. (Eds.), Deep-Sea Sediments, 63. Elsevier, Amsterdam, pp. 215 351. Developments in Sedimentology. Hu¨eneke, H., Mulder, T. (Eds.), 2011. Deep-Sea Sediments, 63. Elsevier, Amsterdam, p. 750. Developments in Sedimentology. Hu¨neke, H., Hernandez-Molina, F.J., Rodrı´guez-Tovar, F.J., Llave, E., Chiarella, D., Mena, A., et al 2020. Diagnostic criteria using microfacies for calcareous contourites, turbidites and pelagites in the Eocene-Miocene slope succession, southern Cyprus. Sedimentology (In press). Hungr, O., Evans, S.G., 2004. Entrainment of debris in rock avalanches: an analysis of a long run-out mechanism. GSA Bull. 116, 1240 1252. Hungr, O., Evans, S.G., Bovis, M., Hutchinson, J.N., 2001. Review of the classification of landslides of the flow type. Environ. Eng. Geosci. 7, 221 238. Hunt, D., Tucker, M.E., 1992. Stranded parasequences and the forced regressive wedge systems tract: deposition during base-level fall. Sediment. Geol. 81, 1 9. Hunt, D., Tucker, M.E., 1995. Stranded parasequences and the forced regressive wedge systems tract: deposition during base-level fall-reply. Sediment. Geol. 95, 147 160. Hurricane Alley, 2002. Bay of Bengal tropical cyclones. ,http://www.hurricanealley.net/bobhst01.htm. (accessed 10.11.07.). Hurst, A., Cartwright, J., 2007. Sand injectites: implications for hydrocarbon exploration. AAPG Mem. 87, 288. Hurst, A., Cartwright, J.A., Duranti, D., Huuse, M., Nelson, M., 2005. Sand injectites: an emerging global play in deep-water clastic environments. In: Dore´, A.G., Vining, B.A. (Eds.), Petroleum Geology: North-West Europe and Global Perspectives Proceedings of the 6th Petroleum Geology Conference. Petroleum Geology Conferences Ltd. Geological Society, London, pp. 133 144. Hurst, A., Scott, A., Vigorito, M., 2011. Physical characteristics of sand injectites. Earth-Sci. Rev. 106, 215 246. Hussain, M., Warren, J.K., 1989. Nodular and enterolithic gypsum: the “Sabkha-Tization” of Salt Flat playa, west Texas. Sediment. Geol. 63, 13 24. Huthnance, J., Alvarado, R., Ambar, I., Hobbs, R., Krahmann, G., Quentel, E., Silva, P., 2008. Gulf of Cadiz oceanography for comparison with seismic imaging. Geophys. Res, Abstr. 10. EGU2008-A-03530, 2008, SRef- ID: 1607e7962/gra/EGU2008-A-03530. Huuse, M., Duranti, D., Steinsland, N., Guargena, C.G., Prat, P., Holm, K., et al., 2004. Seismic characteristics of large-scale sandstone intrusions in the Paleogene of the South Viking Graben, UK and Norwegian North Sea. In: Davies, R.J., Cartwright, J.A., Stewart, S.A., Lappin, M., Underhill, J.R. (Eds.), 3D Seismic Technology: Application to the Exploration of Sedimentary Basins, 29. Geological Society of London, Memoir, pp. 263 278. ,https://doi.org/10.1144/GSL.EM.2004.029.01.2.. Huuse, M., Jackson, C., Cartwright, J.A., Hurst, A., 2009. Large-scale sand injectites in the North Sea: seismic and event stratigraphy and implications for hydrocarbon exploration. In: AAPG Annual Convention, Denver, CO, 7 10 June 2009. Search and Discovery Article #40481.

502

Bibliography

Huuse, M., Jackson, C.A.L., Van Rensbergen, P., Davies, R.J., Flemings, P.B., Dixon, R.J., 2010. Subsurface sediment remobilization and fluid flow in sedimentary basins: an overview. Basin Res. 22, 342 360. Hutchinson, J.N., 1968. Mass movement. In: Fairbridge, R.W. (Ed.), The Encyclopedia of Geomorphology. Reinhold Book Corp., New York, pp. 688 696. Hutton, J., 1788. Theory of the earth. R. Soc. Edinb. Trans. 1, 209 304. Hyder, P., Jeans, D.R.G., Cauquil, E., Nerzic, R., 2005. Observations and predictability of internal solitons in the northern Andaman Sea. Appl. Ocean. Res. 27, 1 11. ILC (International Landslide Center), 2010. Database at the University of Durham. ,http://www.landslidecentre. org/database.htm. (accessed 29.06.10.). Ilstad, T., Marr, J.G., Elverhoi, A., Harbitz, C.B., 2004. Laboratory studies of subaqueous debris flows by measurements of pore-fluid pressure and total stress. Mar. Geol. 213 (1 4), 403 414. Imran, J., Syvitski, J.P.M., 2000. Impact of extreme river events on the coastal ocean. Oceanography 13, 85 92. Imran, J., Harff, P., Parker, G., 2001. A numerical model of submarine debris flow with graphical user interface. Comput. Geosci. 27 (6), 717 729. Inglis, I., Verstralen, I., Mousset, E., Salim, A., Vially, R., 1981. Etude se´dimentologique des Gre`s d’Annot (Re´gion de Colmars-les-Alpes et du Col de la cayolle). ENSPM, Rueil Malmaison 169. Inman, D.L., Nordstrom, C.E., Flick, R.E., 1976. Currents in submarine canyons: an air-sea-land interaction. Annu. Rev. Fluid Mech. 8, 275 310. Iqbaluddin, G., 1978. Occurrence of seismite in the Precambrian rocks of Rajasthan, India. Curr. Sci. 47, 230 231. Ito, M., 2002. Kuroshio current-influenced sandy contourites from the Plio-Pleistocene Kazusa forearc basin, Boso Peninsula, Japan. In: Stow, D.A.V., Pudsey, C.J., Howe, J.A., Fauge`res, J.-C., Viana, A.R. (Eds.), Deep-Water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics, 22. Geological Society Memoirs, London, pp. 421 432. Ito, M., Ishimoto, S., Ito, K., Kotake, N., 2016. Geometry and lithofacies of coarse-grained injectites and extrudites in a late Pliocene trench-slope basin on the southern Boso Peninsula, Japan. Sediment. Geol. 344, 336 349. Iverson, R.M., 1997. The physics of debris flows. Rev. Geophys. 35, 245 296. Iverson, R.M., 2000. Landslide triggering by rain infiltration. Water Resour. Res. 36 (7), 1897 1910. Iverson, R.M., Vallance, R.W., 2001. New views of granular mass flows. Geology 29, 115 118. Iverson, R.M., Reid, M.E., LaHusen, R.G., 1997. Debris-flow mobilization from landslides. Annu. Rev. Earth Planet. Sci. 25, 85 138. Jacobi, R.D., 1976. Sediment slides on the northwestern continental margin of Africa. Mar. Geol. 22, 157 173. Jackson, B.A., 2004. Seismic evidence for gas hydrates in the north Makassar Basin, Indonesia. Petrol. Geosci. 10, 227 238. Available from: https://doi.org/10.1144/1354-079303-601. Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 Scientists, 2014a. Expedition 341 Summary. In: Jaeger, J. M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 Scientists, (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 341, pp. 1 130. Jaeger, J.M., Gulick, S.P.S., LeVay, L.J., the Expedition 341 Scientists, 2014b. Site U1418. In: Jaeger, J.M., Gulick, S. P.S., LeVay, L.J., the Expedition 341, (Eds.), Proceedings of the Integrated Ocean Drilling Program, vol. 341, ,https://doi.org/10.2204/iodp.proc.341.104.2014.. Jagadeesan, L., Jyothibabu, R., Anjusha, A., Mohan, A.P., Madhu, N.V., Muraleedharan, K.R., et al., 2013. Ocean currents structuring the mesozooplankton in the Gulf of Mannar and the Palk Bay, southeast coast of India. Prog. Oceanogr. 110, 27 48. Jakob, M., Hungr, O. (Eds.), 2005. Debris-Flow Hazards and Related Phenomena. Praxis-Springer, Berlin; Heidelberg. Jaimes, B., Shay, L.K., 2010. Near-inertial wave wake of Hurricanes Katrina and Rita over mesoscale oceanic eddies. J. Phys. Oceanogr. 40, 1320 1337. Jansen, E., Befring, S., Bugge, T., Eidvin, T., Holtedahl, H., Petter Sejrup, H., 1987. Large submarine slides on the Norwegian continental margin: sediments, transport and timing. Mar. Geol. 78, 77 107. Jayne, S.R., St. Laurent, L.C., Gille, S.T., 2004. Connections between ocean bottom topography and earth’s climate. Oceanography 17, 65 74. Jennette, D.C., Garfield, T.R., Mohrig, D.C., Cayley, G.T., 2000. The interaction of shelf accommodation, sediment supply and sea level in controlling the facies, architecture and sequence stacking patterns of the Tay and Forties/Sele Basin-floor fans, Central North Sea. In: Weimer, P., Slatt, R.M., Coleman, J., Rossen, N.C., Nelson, H., Bouma, A.H., Styzen, M.J.,

Bibliography

503

Lawrence, D.T. (Eds.), Deep-Water Reservoirs of the World: GCSSEPM 20th Annual Bob F. Perkins Research Conference, 3 6 December 2000, Houston, TX, pp. 402 421. Jia, J., Liu, Z., Miao, C., Fang, S., Zhou, R., Meng, Q., et al., 2014. Depositional model and evolution for a deepwater sublacustrine fan system from the syn-rift Lower Cretaceous Nantun Formation of the Tanan Depression (Tamtsag Basin, Mongolia). Mar. Petrol. Geol. 57, 264 282. Jiang, J., Zhong, N., Li, Y., Xu, H., Yang, H., Peng, X., 2016. Soft sediment deformation structures in the Lixian lacustrine sediments, Eastern Tibetan Plateau and implications for postglacial seismic activity. Sediment. Geol. 344, 123 134. Johns, E., Wilson, W.D., Molinari, R.L., 1999. Direct observations of velocity and transport in the passages between the Intra-Americas Sea and the Atlantic Ocean, 1984 1996. J. Geophys. Res. 104, 25805 25820. p. Johnson, A.M., 1970. Physical Processes in Geology. Freeman, Cooper and Co., San Francisco, CA, p. 577. Johnson, A.M., 1984. Debris flow. In: Prior, D.B. (Ed.), Slope Instability. John Wiley, New York, pp. 257 361. Johnson, D.A., Damuth, J.E., 1979. Deep thermohaline flow and current controlled sedimentation in the Amirante Passage: western Indian Ocean. Mar. Geol. 33, 1 44. Johnson, T.C., Carlson, T.W., Evans, J.E., 1980. Contourites in Lake Superior. Geology 8, 437 441. Johnson, K.S., Paull, C.K., Barry, J.P., Chavez, F.P., 2001. A decadal record of underflows from a coastal river into the deep sea. Geology 29, 1019 1022. Jolly, R.J.H., Lonergan, L., 2002. Mechanisms and control on the formation of sand intrusions. J. Geol. Soc. Lond. 159, 605 617. Jones, A.P., Omoto, K., 2000. Towards establishing criteria for identifying trigger mechanisms for soft-sediment deformation: a case study of Late Pleistocene lacustrine sands and clays, Onikobe and Nakayamadaira Basins, northeastern Japan. Sedimentology 47, 1211 1226. Jones, M.E., Preston, R.M.F. (Eds.), 1987. Deformation of Sediments and Sedimentary Rocks. Geological Society of London, Special Publication. Jones, I., Christensen, R., Haynes, J., Faragher, J., Novianti, I., Morris, H., et al., 2005. Multi-disciplinary geoscience: the ‘Brenda’ North Sea development. CSEG (Canadian Society of Exploration Geophysicists) RECORDER, January, pp. 39 43. Jonk, R., Hurst, A., Duranti, D., Parnell, J., Mazzini, A., Fallick, A.E., 2005. Origin and timing of sand injection, petroleum migration, and diagenesis in Tertiary reservoirs, south Viking Graben, North Sea. AAPG Bull. 89, 329 357. Kahle, C.E., 2002. Seismogenic deformation structures in microbialites and mudstones, Silurian Lockport Dolomite, northwestern Ohio, USA. J. Sediment. Res. 72 (1), 201 216. Kang, H., Paik, I., Lee, H., Lee, J., Chun, J., 2010. Soft-sediment deformation structures in Cretaceous non-marine deposits of southeastern Gyeongsang Basin, Korea: occurrences and origin. Isl. Arc 19 (4), 628 646. Kao, S.J., Dai, M., Selvaraj, K., Zhai, W., Cai, P., Chen, S.N., et al., 2010. Cyclone-driven deep sea injection of freshwater and heat by hyperpycnal flow in the subtropics. Geophys. Res. Lett. 37 (21), CiteID L21702. Karcz, I., Shanmugam, G., 1974. Decrease in Scour Rate of Fresh Deposited Muds. Proc. American Society of Civil Engineers (ASCE). J. Hydraulics Div. 100 (HY11), 1735 1738. Karl, H.A., Cacchione, D.A., Carlson, P.R., 1986. Internal-wave currents as a mechanism to account for large sand waves in Navarinsky Canyon Head, Bering Sea. J. Sediment. Petrol. 56, 706 714. Kassem, A., Imran, J., 2001. Simulation of turbid underflows generated by the plunging of a river. J. Geol. 29 (7), 655 658. Kastens, K.A., Cita, M.B., 1981. Tsunami induced sediment transport in the Abyssal Mediterranean Sea. GSA Bull. 89, 591 604. Kastens, K.A., Cita, M.B., 1981. Tsunami induced sediment transport in the Abyssal Mediterranean Sea. Geol. Soc. Am. Bull 89, 591 604. Kearey, P., Klepeis, K.A., Vine, F.J., 2009. Global Tectonics, third ed. Wiley-Blackwell, p. 496. Keen, T.R., Glenn, S.,M., 2002. Predicting bed scour on the continental shelf during Hurricane Andrew. J. Waterw. Port. Coast. Ocean. Eng. 128, 249 257. Keevil, G.M., Peakall, J., Best, J.L., Amos, K.J., 2006. Flow structure in sinuous submarine channels: velocity and turbulence structure of an experimental submarine channel. Mar. Geol. 229, 241 257. Keller, G.H., Shepard, F.P., 1978. Currents and sedimentary processes in submarine canyons off the northeast United States. In: Stanley, D.J., Kelling, G.K. (Eds.), Sedimentation in Submarine Canyons, Fans, and Trenches. Dowden, Hutchinson & Ross, Inc, Stroudsburg, PA, pp. 15 32.

504

Bibliography

Kelling, G., Walton, E.K., 1957. Load-cast structures: their relationship to upper surface structures and their mode of formation. Geol. Mag. 94, 481 490. Kennedy, W.J., 1987. Sedimentology of Late Cretaceous-Palaeocene Chalk reservoirs, North Sea Central Graben. In: Brooks, J., Glennie, K.W. (Eds.), Petroleum Geology of North West Europe. Graham & Trotman, London, pp. 469 481. Kennett, J.P., Fackler-Adams, B.N., 2000. Relationship of clathrate instability to sediment deformation in the Upper Neogene of California. Geology 28 (3), 215 218. Kent, P.E., 1966. The transport mechanism in catastrophic rock falls. J. Geol. 74, 79 83. Kenyon, N.H., 1987. Mass-wasting features on the continental slope of northwest Europe. Mar. Geol. 74, 57 77. Kenyon, N.H., Belderson, R.H., 1973. Bedforms of the Mediterranean undercurrent observed with side scan sonar. Sediment. Geol. 9, 77 100. Kenyon, N.H., Klaucke, I., Millington, J., Ivanov, M.K., 2002a. Sandy submarine canyon-mouth lobes on the western margin of Corsica and Sardinia, Mediterranean Sea. Mar. Geol. 184, 69 84. Kenyon, N.H., Akhmetzhanov, A.M., Twichell, D.C., 2002b. Sand wave fields beneath the Loop Current, Gulf of Mexico: reworking of fan sands. Mar. Geol. 192, 297 307. Kerr, B.C., Scholl, D.W., Klemperer, S.L., 2005. Seismic stratigraphy of Detroit Seamount, Hawaiian-Emperor seamount chain: post-hot-spot shield-building volcanism and deposition of the Meiji drift. Geochem. Geophys. Geosyst. 6 (7). Available from: https://doi.org/10.1029/2004GC000705. Khan, S.M., Imran, J., Bradford, S., Syvitski, J.P.M., 2005. Numerical modeling of hyperpycnal plume. Mar. Geol. 222 223, 193 211. Khanna, D.R., Yadav, P.R., 2008. Biology of Coelenterata. Discovery Publishing, 393 pp. Khripounoff, A., Vangriesheim, A., Babonneau, N., Crassous, P., Dennielou, B., Savoye, B., 2003. Direct observation of intense turbidity current activity in the Zaire submarine valley at 4000 m water depth. Mar. Geol. 194, 151 158. Khripounoff, A., Crassous, P., Lo Bue, N., Dennielou, B., Silva Jacinto, R., 2012. Different types of sediment gravity flows detected in the Var submarine canyon (northwestern Mediterranean Sea). Prog. Oceanogr. 106, 138 153. Khvorova, I.V., 1978. Terrigenous clastic formations of oceans and certain seas. Lithologiya i Polenzye Iskopaemye 3, 3 24. Kim, S.B., Chough, S.K., Chun, S.S., 1995. Bouldery deposits in the lowermost part of the Cretaceous Kyokpori Formation, SW Korea: cohesionless debris flows and debris falls on a steep-gradient delta slope. Sediment. Geol. 98, 97 119. Kim, B.-C., Yu, K.-M., Chun, H.-Y., Choi, S.-J., Kim, Y.-B., 1997. The southeastern margin of the Cretaceous Yongdong Basin, Korea: a lacustrine fan-delta system. Geosci. J. 1, 61 74. Kim, D.C., Lee, G.S., Lee, G.H., Park, S.C., 2008. Sediment echo types and acoustic characteristics of gas-related acoustic anomalies in Jinhae Bay, southern Korea. Geosci. J. 12, 47 61. Available from: https://doi.org/ 10.1007/s12303-008-0007-8. Kimura, G., Koga, K., Fujioka, K., 1989. Deformed soft-sediments at the junction between the Mariana and Yap trenches. J. Struct. Geol. 11, 463 472. King, L.H., MacLean, B., 1970. Pockmarks on the Scotian Shelf. GSA Bull. 81, 3141 3148. King, E.L., Haflidason, H., Sejrup, H.P., Lovlie, R., 1998. Glacigenic debris flows on the North Sea Trough Mouth Fan during ice stream maxima. Mar. Geol. 152 (1 3), 217. Kirkland, D.W., Anderson, R.Y., 1970. Microfolding in the Castile and Todilto Evaporites, Texas and New Mexico. GSA Bull. 81, 3259 3282. Kirschbaum, D.B., Adler, R., Hong, Y., Lerner-Lam, A., 2009. Evaluation of a preliminary satellite-based landslide hazard algorithm using global landslide inventories. Nat. Hazards Earth Syst. Sci. 9, 673 686. Kjeldsen, S.P., 1984. Dangerous wave groups. Nor. Marit. Res. 2, 4 16. Klaucke, I., Masson, D.G., Kenyon, N.H., Gardner, J.V., 2004. Sedimentary processes of the lower Monterey Fan channel and channel-mouth lobe. Mar. Geol. 206, 181 198. Klaus, A., Ledbetter, M.T., 1988. Deep-sea sedimentary processes in the Argentine Basin revealed by high-reso lution seismic records (3.5 kHz echograms). Deep-Sea Res. 40, 899 917. Klein, G.D., 1966. Dispersal and petrology of sandstones of Stanley-Jackfork boundary, Ouachita foldbelt, Arkansas and Oklahoma. AAPG Bull. 50, 308 326. Klein, G.D., 1970. Depositional and dispersal dynamics of intertidal sand bars. J. Sediment. Petrol. 40, 1095 1127.

Bibliography

505

Klein, G.D., 1971. A sedimentary model for determining paleotidal range. Geol. Soc. Am. Bull. 82, 2585 2592. Klein, G.D., 1975. Resedimented pelagic carbonate and volcaniclastic sediments and sedimentary structures in Leg 30 DSDP cores from the western equatorial Pacific. Geology 3, 39 42. Klein, G.D., 1998. Clastic tidalites—a partial retrospective view. In: Alexander, C.R., Davis, R.A., Henry, V.J. (Eds.), Tidalites: Processes and Products, 61. SEPM Special Publication, pp. 5 14. Klein, G.D., Melo, U., Della Favera, J.C., 1972. Subaqueous gravity processes on the front of Cretaceous deltas, Recoˆncavo basin, Brazil. GSA Bull. 83, 1469 1492. Klein, G.D., Kobayashi, K., White, S.M., 1980. Introduction and Explanatory Notes. Deep. Sea Drill. Proj. Leg. 58, 3 18. Available from: https://doi.org/10.2973/dsdp.proc.58.101.1980. Knauss, J., 1965. A technique for measuring deep ocean currents close to the bottom with an unattached current meter and some preliminary results. J. Mar. Res. 23, 237 245. Kneller, B., 1995. Beyond the turbidite paradigm: physical models for deposition of turbidites and their implications for reservoir prediction. In: Hartley, A.J., Prosser, D.J. (Eds.), Characterization of Deep-Marine Clastic Systems. Geological Society Special Publication No. 94, pp. 31 49. Kneller, B., Buckee, C., 2000. The structure and fluid mechanics of turbidity currents: a review of some recent studies and their geologic implications. Sedimentology 47 (Suppl. 1), 62 94. Kneller, B., Martinsen, O.J., McCaffrey, W. (Eds.), 2009. External Controls on Deep-water Depositional Systems, vol. 92. SEPM Special Publication, p. 402. Knight, J., 2003. Geomorphic and sedimentary evidence for patterns of late Midlandian ice retreat in the Tempo Valley, north-central Ireland. Ir. Geogr. 36, 127 144. Koch, S.P., Barker, J.W., Vermersch, J.A., 1991. The Gulf of Mexico Loop Current and deepwater drilling. J. Petrol. Technol. 43, 1046 1050; 1118 1119. Kohut, J.T., Glenn, S.M., Paduan, J.D., 2006. Inner shelf response to Tropical Storm Floyd. J. Geophys. Res. 111, C09S91. Available from: https://doi.org/10.1029/2003JC002173. ,http://www.agu.org/pubs/crossref/2006/ 2003JC002173.shtml. (accessed 16.09.07.). Komar, P.D., 1971. Hydraulic jumps in turbidity currents. Geol. Soc. Am. Bull. 82, 1477 1488. Komar, P.D., 1976. Beach Processes and Sedimentation. Prentice-Hall, Inc., Englewood Cliffs, NJ, p. 429. Koning, H.L., 1982. On an explanation of marine flow slides in sand. In: Saxov, S., Nieuwenhuis, J.K. (Eds.), Marine Slides and Other Mass Movements. Plenum Press, New York and London, pp. 83 94. Kopf, A.J., Stegmann, S., Garziglia, S., Sultan, N., Henry, P., Dennielou, B., et al., 2016. Soft sediment deformation in the shallow submarine slope off Nice (France) as a result of a variably charged Pliocene aquifer and mass wasting processes. Sediment. Geol. 344, 290 309. Koppejan, A.W., van Wamelan, B.M., Weinberg, L.J.H., 1948. Coastal flow slides in the Dutch province of Zeeland. In: Proceedings of the 2nd International Conference on Soil Mechanics and Foundation, Rotterdam, 5, pp. 89 96. Korznikov, K.A., 2014. Mud volcanoes of Sakhalin Island within the regional system of nature protection areas. ВЕСТН. МОСК. УН-ТA. СЕР. 5. ГЕОГРAФИЯ, 3, pp. 34 39. Kostic, S., Parker, K., 2003. Progradational sand-mud deltas in lakes and reservoirs. Part 2. Experiment and numerical simulation. J. Hydraulic Res. 41 (2), 141 152. Kostic, S., Parker, G., 2007. Conditions under which a supercritical turbidity current traverses an abrupt transition to vanishing bed slope without a hydraulic jump. J. Fluid Mech. 586, 119 145. Kostic, S., Parker, G., Marr, J.G., 2002. Role of turbidity currents in setting the foreset slope of clinoforms prograding into standing fresh water. J. Sediment. Res. 72 (3), 353 362. Kottke, B., Schwenk, T., Breitzke, M., Wiedicke, M., Kudrass, H.R., Spiess, V., 2003. Acoustic facies and depositional processes in the upper submarine canyon Swatch of No Ground (Bay of Bengal). Deep-Sea Res. II 50, 979 1001. Krumbein, W.C., 1934. Size frequency distribution of sediments. J. Sediment. Petrol. 4, 65 77. Krumbein, W.C., Sloss, L.L., 1963. Stratigraphy and Sedimentation, second ed. W.H. Freeman and Company, San Francisco, CA, p. 660. Krynine, P.D., 1948. The megascopic study and field classification of sedimentary rocks. J. Geol. 56, 130 165. Kudrass, H.R., Michels, K.H., Wiedicke, M., 1998. Cyclones and tides as feeders of a submarine canyon off Bangladesh. Geology 26, 715 718. Kuehl, S.A., Hariu, T.M., Moore, W.S., 1989. Shelf sedimentation off the Ganges-Brahmaputra river system: evidence for sediment bypassing to the Bengal Fan. Geology 17, 1132 1135.

506

Bibliography

Kuenen, Ph.H., 1950a. Turbidity currents of high density. In: 18th International Geological Congress 1948, London, Reports, pt. 8, pp. 44 52. Kuenen, Ph.H., 1950b. Slump structures of Waitamata Beds around Auckland. Trans. R. Soc. N. Z. 78, 467 475. Kuenen, Ph.H., 1951. Properties of turbidity currents of high density. In: Hough, J.L. (Ed.), Turbidity Currents and the Transportation of Coarse Sediments to Deep Water. A Symposium, 2. SEPM Special Publication, pp. 14 33. Kuenen, Ph.H., 1953. Significant features of graded bedding. AAPG Bull. 37, 1044 1066. Kuenen, Ph.H., 1957. Sole markings of graded greywacke beds. J. Geol. 65, 231 258. Kuenen, Ph.H., 1964. Deep-sea sands and ancient turbidites. In: Bouma, A.H., Brouwer, A. (Eds.), Turbidites, Developments in Sedimentology, 3. Elsevier, Amsterdam, pp. 3 33. Kuenen, Ph. H., 1966. Experimental turbidite lamination in a circular flume. J. Geol. 74, 523 545. Kuenen, Ph. H., 1967. Emplacement of flysch-type sand beds. Sedimentology 9, 203 243. Kuenen, Ph. H., Migliorini, C.I., 1950. Turbidity currents as a cause of graded bedding. J. Geol. 58, 91 127. Kuhn, T.S., 1970. The Structure of Scientific Revolutions. University of Chicago Press, Chicago, IL, p. 172. Kuhn, T.S., 1996. The Structure of Scientific Revolutions, third ed. The University of Chicago Press, Chicago, IL, p. 212. Kundu, A., Eriksson, P.G., Matin, A., 2015. Soft-sediment deformation structures in the sub-Himalayan Middle Siwalik Subgroup, Lish River section, India. Episodes 38 (3), 197 207. Kunze, K., Rosenfeld, L.K., Carter, G.S., Gregg, M.C., 2002. Internal waves in Monterey submarine canyon. J. Phys. Oceanogr. 32, 1890 1913. Kvenvolden, K.A., 1993. Gas hydrates: geological perspective and global change. Rev. Geophys. 31, 173 187. Labaume, P., Mutti, E., Seguret, M., 1987. Megaturbidites: a depositional model from the Eocene of the SWPyrenean foreland basin, Spain. Geo-Mar. Lett. 7, 91 101. Laberg, J.S., Vorren, T.O., Dowdeswell, J.A., Kenyon, N.H., Taylor, J., 2000. The Andøya Slide and the Andøya Canyon, north-eastern Norwegian-Greenland Sea. Mar. Geol. 162, 259 275. Laberg, J.S., Camerlenghi, A., 2008. The significance of contourites for submarine slopoe stability. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 537 556. Developments in Sedimentology. Laberg, J.S., Kawamura, K., Amundsen, H., Baeten, N., Forwick, M., Rydningen, T.A., et al., 2014. A submarine landslide complex affecting the Jan Mayen Ridge, Norwegian Greenland Sea: slide-scar morphology and processes of sediment evacuation. Geo-Mar. Lett. 34, 51 58. Available from: https://doi.org/10.1007/s00367-0130345-z. Ladd, G.E., 1935. Landslides, subsidences and rockfalls. Am. Railw. Eng. Assoc. Bull. 37, 1091 1162. LaFond, E.C., 1962. Internal waves. In: Hill, M.N. (Ed.), The Sea., vol. 1. Wiley-Interscience, New York, pp. 731 751. Lafuente, J.G., Sarhan, T., Vargas, M., Vargas, J.M., Plaza, F., 1999. Tidal motions and tidally induced fluxes through La Linea submarine canyon, western Alboran Sea. J. Geophys. Res. 104, 3109 3119. Laine, E.P., 1978, Geological Effects of the Gulf Stream in the North American Basin (Unpublished Ph.D. dissertation). Woods Hole Oceanographic Institution/Massachusetts Institute of Technology joint program in oceanography and ocean engineering, Woods Hole, MA, 164 p. Laine, E.P., Hollister, C.D., 1981. Geological effects of the Gulf Stream system on the northern Bermuda Rise. Mar. Geol. 39, 277 310. Lamarche, G., Mountjoy, J., Bull, S., Hubble, T., Krastel, S., Lane, E., et al. (Eds.), 2018. Springer, 621 p. 2018: 8th International Symposium on Submarine Mass Movements and Their Consequences, Victoria, Canada, 7 9 May, 2018, Springer, 621 p. Lamb, M.P., Mohrig, D., 2009. Do hyperpycnal-flow deposits record river-flood dynamics? Geology 37 (12), 1067 1070. Lamb, M.P., McElroy, B., Kopriva, B., Shaw, J., Mohrig, D., 2010. Linking river-flood dynamics to hyperpycnalplume deposits: experiments, theory, and geological implications. GSA Bull. 122, 1389 1400. Lamb, M.P., Nittrouer, J.A., Mohrig, D., Shaw, J., 2012. Backwater and river plume controls on scour upstream of river mouths: implications for fluvio-deltaic morphodynamics. J. Geophys. Res. Earth Surf. 117, F01002. Lambe, T.W., Whitman, R.V., 1969. Soil Mechanics. John Wiley & Sons, New York, 553 pp. Lambert, A.M., Kelts, K.R., Marshall, N.F., 1976. Measurements of density underflows from Walensee, Switzerland. Sedimentology 23, 87 105.

Bibliography

507

Lancaster, J.T., Spittler, T.E., Short, W.R., 2015. Alluvial fan flooding hazards: an engineering geologic approach to preliminary assessment. California Geological Survey Special Report 227, California Department of Conservation, Sacramento, CA, 46 p. Lane, E.W., 1957. A study of the shape of channels formed by natural streams flowing in erodible material. U.S. Army Corps of Engineers, Missouri River Division, Omaha, NE. revised Lewis, T., Lamoureux, S.F., Normandeau, A., Dugan, H.A., 2018. Hyperpycnal flows control the persistence and flushing of hypoxic highconductivity bottom water in a High Arctic lake. Arct. Sci. 4, 26 41. Larcombe, P., Carter, R.M., 2004. Cyclone pumping, sediment partitioning and the development of the Great Barrier Reef shelf system: a review. Quat. Sci. Rev. 23, 107 135. Larsen, P.-H., 1986. Soft sediment deformation structures in Silurian turbidites from North Greenland. Bull. Geol. Soc. Den. 35, 19 23. Lario, J., Luque, L., Zazo, C., Goy, J.L., Spencer, C., Cabero, A., et al., 2010. Tsunami vs. storm surge deposits: a review of the sedimentological and geomorphological records of extreme wave events (EWE) during the Holocene in the Gulf of Cadiz, Spain. Z. fu¨r Geomorphologie 54 (Suppl. 3), 301 316. Stuttgart, Juli 2010. Lastras, G., Canals, M., Urgeles, R., Amblas, D., Ivanov, M., Droz, L., et al., 2007. A walk down the Cap de Creus canyon, Northwestern Mediterranean Sea: recent processes inferred from morphology and sediment bedforms. Mar. Geol. 246, 176 192. Lathrop, E., 2015. Sediment Composition in the Gulf of Ca´diz Contourites During the Pleistocene (Undergraduate Research Thesis). The Ohio State University, Columbus, OH, 61 p. LaViolette, P.E., Lacombe, H., 1988. Tidal-induced pulses in the flow through the Strait of Gibraltar. In: Minas, H. J., Nival, P. (Eds.), Ocdanographie pdlagique mdditerrandenne: Oceanologica Acta, pp 13 27. Lavrenov, I., 1998. The wave energy concentration at the Agulhas current of South Africa. Nat. Hazards 17, 117 127. Lawton, T.F., Shipley, K.W., Aschoff, J.L., Giles, K.A., Vega, F.J., 2005. Basinward transport of Chicxulub ejecta by tsunami-induced backflow, La Popa basin, northeastern Mexico, and its implications for distribution of impact-related deposits flanking the Gulf of Mexico. Geology 33, 81 84. Leat, P.T., Tate, A.J., Tappin, D.R., Day, S.J., Owen, M.J., 2010. Growth and mass wasting of volcanic centers in the northern South Sandwich arc, South Atlantic, revealed by new multibeam mapping. Mar. Geol. 275, 110 126. Leclair, S., Arnott, R.W.C., 2005. Parallel lamination formed by high-density turbidity currents. J. Sediment. Res. 75, 1 5. Lee, H.J., 2005. Undersea landslides: extent and significance in the Pacific Ocean, an update. Nat. Hazards Earth Syst. Sci. 5, 877 892. Lee, H.J., 2009. Timing of occurrence of large submarine landslides on the Atlantic Ocean margin. Mar. Geol. 264, 53 64. Lee, H.J., Schwab, W.C., Booth, J.S., 1993. Submarine landslides: an introduction. In: Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone, 2002. U.S. Geological Survey Bulletin, pp. 1 13. Lee, H.J., Kayen, R.E., Edwards, B.D., Field, M.E., Gardner, J.V., Schwab, W.C., et al., 1996. Ground-truth studies of west coast and Gulf of Mexico submarine fans. In: Gardner, J.V., Field, M.E., Twichell, D.C. (Eds.), Geology of the United States’ Sea Floor, The View From GLORIA. Cambridge University Press, New York, pp. 221 222. Lee, H.J., Syvitski, J.P.M., Parker, G., Orange, D., Locat, J., Hutton, E.W.H., et al., 2002. Distinguishing sediment waves from slope failure deposits: field examples, including the ‘Humboldt slide’, and modeling results. Mar. Geol. 192, 79 104. Lee, H., Ryan, H., Kayen, R.E., Haeussler, P.J., Dartnell, P., Hampton, M.A., 2006. Varieties of submarine failure morphologies of seismically-induced landslides in Alaskan fjords. Nor. J. Geol. 86, 221 230. Lee, I.-H., Wang, Y.-H., Liu, J.T., Chuang, W.-S., Xu, J., 2009a. Internal tidal currents in the Gaoping (Kaoping) Submarine Canyon. J. Mar. Syst. 76, 397 404. Lee, I.H., Lien, R.C., Liu, J.T., Chuang, W.S., 2009b. Turbulent mixing and internal tides in Gaoping (Kaoping) Submarine Canyon, Taiwan. J. Mar. Syst. 76, 383 396. Lee, S.H., Jung, W., Bahk, J., Gardner, J.M., 2011. Depositional Style of Mass-flow Lobes in Ulleung Basin (East Sea, Korea) Interpreted from MR1 Sonar Image, Chirp Profiles, and Piston Cores. American Geophysical Union. Fall Meeting 2011, abstract #OS13B-1520.

508

Bibliography

Lee, S.H., Stow, D.A.V., 2007. Laterally contiguous, concave-up basal shear surfaces of submarine landslide deposits (Miocene), southern Cyprus: differential movement of sub-blocks within a single submarine landslide lobe. Geosci. J. 11, 315 321. Legou, I., Savoye, B., Pirmez, C., Droz, L., 2008. Channel-mouth lobe complex of the recent Amazon fan: the missing piece. Mar. Geol 252, 62 77. Legros, F., 2002. Can disperssive pressure cause inverse grading in grain flows? J. Sediment. Res. 72, 166 170. Legros, F., 2003. Can disperssive pressure cause inverse grading in grain flows?—reply. J. Sediment. Res. 73, 335. Le Heron, D.P., Etiene, J.L., 2005. A complex subglacial clastic dyke swarm, So´lheimajo¨kull, Southern Iceland. Sediment. Geol. 181, 25 37. Le Heron, D.P., Sutcliffe, O.E., Whittington, R.J., Craig, J., 2005. The origins of glacially related softsediment deformation structures in Upper Ordovician glaciogenic rocks: implication for ice-sheet dynamics. Palaeogeogr. Palaeoclimatol. Palaeoecol. 218, 75 103. Leo´n, R., Somoza, L., Medialdea, T., Javier Gonza´lez, F., Gimenez-Moreno, C.J., Pe´rez-Lo´pez, R., 2014. Pockmarks on either side of the Strait of Gibraltar: formation from overpressured shallow contourite gas reservoirs and internal wave action during the last glacial sea-level lowstand? Geo-Mar. Lett. 34, 131 151. Le Roux, J.P., Nielsen, S.N., Kemnitz, H., Henriquez, A., 2008. A Pliocene mega-tsunami deposit and associated features in the Ranquil Formation, southern Chile. Sediment. Geol. 203, 164 180. Levi, T., Weinberger, R., Aifa, T., Eyal, Y., Marco, S., 2006. Earth-quake induced clastic dikes detected by anisotropy of magnetic susceptibility. Geology 34, 69 72. Lewis, K.B., 1971. Slumping on a continental slope inclined at 1 4 . Sedimentology 16, 97 110. Lewis, K., Collot, J.-Y., 2001. Giant submarine avalanche: was this “deep impact” New Zealand style? Water Atmos. 9, 26 27. Lewis, J., Clinch, S., Meyer, D., Richards, M., Skirius, C., Stokes, R., et al., 2007. Exploration and appraisal challenges in the Gulf of Mexico deep-water Wilcox: Part 1 - Exploration overview, reservoir quality, and seismic imaging. In: Kennan, L., Pindell, J., Rosen, N.C. (Eds.), The Paleogene of the Gulf of Mexico and Caribbean Basins: Processes, Events, and Petroleum Systems. 27th Annual GCSSEPM Foundation Bob F. Perkins Research Conference, Houston, TX, pp. 398 414. ,http://www.bergan.com/Downloads/LewsEtal.pdf. (accessed 31.05.10.). Lewis, T., Lamoureux, S.F., Normandeau, A., Dugan, H.A., 2018. Hyperpycnal flows control the persistence and flushing of hypoxic high-conductivity bottom water in a High Arctic lake. Arct. Sci. 4, 26 41. Li, T., 1989. Landslides—extent and economic significance in China. In: Brabb, E.E., Harrod, B.L. (Eds.), Landslides—Extent and Economic Significance. A. A. Balkema Publishers, Rotterdam, pp. 271 287. Li, G.X., Tang, Z.S., Yue, S.H., Zhuang, K.L., Wei, H.L., 2001. Sedimentation in the shear front off the Yellow River mouth. Cont. Shelf Res. 21 (6 7), 607 625. Li, S., Du, Y., Zhang, Z., Wu, J., 2008. Earthquake-related softsediment deformation structures in Palaeogene on the continental shelf of the East China Sea. Front. Earth Sci. 2 (2), 177 186. ,https://doi.org/10.1007/s11707008-0026-9. (in Chinese with English abstract). Li, X., Zhao, Y., Liu, H., Wang, Y., Wei, L., Liao, J., et al., 2010. The characteristics of the sandy debris flow of the Triassic Yanchang Formation and its exploration significance in the Ordos Basin, China. AAPG Search and Discovery Abstracts. ,http://www.searchanddiscovery.com/abstracts/pdf/2010/intl/abstracts/ndx_li02. pdf. (accessed 31.10.10.). Li, X., Chen, Q., Liu, H., Wan, Y., Wei, L., Liao, J., et al., 2011. Features of sandy debris flows of the Yanchang Formation in the Ordos Basin and its oil and gas exploration significance. Acta Geol. Sin. (Engl. Ed.) 85 (5), 187 1202. Available from: https://doi.org/10.1111/j.1755-6724.2011.00550.x. Lien, R.-C., Ma, B., Lee, C.M., et al., 2015. The Kuroshio and Luzon Undercurrent east of Luzon Island. Oceanography 28 (4), 54 63. Available from: https://doi.org/10.5670/oceanog.2015.81. Liesa, C.L., Rodrı´guez-Lo´pez, J.P., Ezquerro, L., Alfaro, P., Rodrı´guez-Pascua, M.A., Lafuente, P., et al., 2016. Facies control on seismites in an alluvial aeolian system: the Pliocene dune field of the Teruel half-graben basin (eastern Spain). Sediment. Geol. 344, 237 252. Lindberg, B., Laberg, J.S., Vorren, T.O., 2004. The Nyk Slide - morphology, progression, and age of a partly buried submarine slide offshore northern Norway. Mar. Geol. 213, 277 289. Link, M.H., Osborne, R.H., 1978. Lacustrine facies in the Pliocene Ridge Basin, California. In: Matter, A., Tucker, M.E. (Eds.), Modern and Ancient Lake Sediments, 2. IAS Special Publication, pp. 167 187.

Bibliography

509

Lipman, P.W., Normark, W.R., Moore, J.G., Wilson, J.B., Gutmacher, C.E., 1988. The Giant Submarine Alika Debris Slide, Mauna Loa, Hawaii. J. Geophys. Res. 93 (B5), 4279 4299. Lisitsyn, A.P., 1986. Avalanche sedimentation in seas and oceans. Part V: Special mechanisms of transport of sedimentary matter and formation of sedimentary bodies of the second global level. Gravitites, their classes and series. Lithologiya i Polenzye Iskopaemye 4, 3 28. Liu, K.B., Sun, S., Jiang, X., 1992. Environmental change in the Yangtze River Delta since 12,000 Years B.P. Quat. Res. 38, 32 45. Liu, J.P., Li, A.C., Xu, K.H., Velozzi, D.M., Yang, Z.S., Milliman, J.D., et al., 2006. Sedimentary features of the Yangtze River-derived along-shelf clinoform deposit in the East China Sea. Cont. Shelf Res. 26, 2141 2156. Liu, J.T., Wang, Y.H., Yang, R.J., Hsu, R.T., Kao, S.J., Lin, H.L., et al., 2012. Cyclone-induced hyperpycnal turbidity currents in a submarine canyon. J. Geophys. Res. Ocean. 117, C04033. Llave, E., Matias, H., Herna´ndez-Molina, F.J., Ercilla, G., Stow, D.A.V., Medialdea, T., 2011. Pliocene Quaternary contourites along the northern Gulf of Cadiz margin: sedimentary stacking pattern and regional distribution. Geo-Mar. Lett. 31, 377 390. Locat, J., Lee, H.J., 2000. Submarine landslides: advances and challenges. In: Proceedings of the 8th International Symposium on Landslides, Cardiff, June. ,http://www.landslides.ggl.ulaval.ca/saguenay/publi/locatleeul. pdf. (accessed 31.06.10.). Locat, J., Lee, H.J., 2002. Submarine landslides: advances and challenges. Can. Geotech. J. 39, 193 212. Locat, J., Mienert, J. (Eds.), 2003. Submarine Mass Movements and Their Conseqenecs. Kluwer Academic Publishers, Dordrecht. Locat, J., Lee, H.J., 2005. Chapter 9: Subaqueous debris flows. In: Jakob, M., Hungr, O. (Eds.), Debris-Flow Hazards and Related Phenomena: Praxis. Springer, Berlin, pp. 203 245. Locat, J., Lee, H.J., Locat, P., Imran, J., 2004. Numerical analysis of the mobility of the Palos Verdes debris avalanche, California, and its implication for the generation of tsunamis. Mar. Geol. 203, 269 280. Locat, J., Lee, H.J., ten Brink, U.S., Twichell, D., Geist, E., Sansoucy, M., 2009. Geomorphology, stability and mobility of the Currituck slide. Mar. Geol. 264, 28 40. Logan, W.E., 1863. Report on the Geology of Canada. John Lovell, Montreal, p. 464. Lombard, A., 1963. Laminites: a structure of flysch-type sediments. J. Sediment. Petrol. 33, 14 22. Loncke, L., Gaullier, V., Bellaiche, G., Mascle, J., 2002. Recent depositional patterns of the Nile deep-sea fan from echo-character mapping. Am. Assoc. Petrol. Geol. Bull. 86, 1165 1186. Longva, O., Janbu, N., Blikra, L.H., Boe, R., 2003. The Finneidfjord slide: seafloor failure and slide dynamics. In: Locat, J., Mienert, J. (Eds.), Submarine Mass Movements and Their Consequences. Kluwer Academic Publishers, Dordrecht, pp. 531 538. Lonsdale, P., Malfait, B., 1974. Abyssal dunes of foraminiferal sand on the Carnegie Ridge. GSA Bull. 85, 1697 1712. Lonsdale, P., Nornaark, W.R., Newman, W.A., 1972. Sedimentation and erosion on Horizon Guyot. Geol. Soc. Am. Bull. 83, 289 316. Lovell, J.P.B., Stow, D.A.V., 1981. Identification of ancient sandy contourites. Geology 9, 347 349. Lowe, D.R., 1975. Water escape structures in coarse grained sediments. Sedimentology 22, 157 204. Lowe, D.R., 1976a. Subaqueous liquefied and fluidized sediment flows and their deposits. Sedimentology 23, 285 308. Lowe, D.R., 1976b. Grain flow and grain flow deposits. J. Sediment. Petrol. 46, 188 199. Lowe, D.R., 1979. Sediment gravity flows: their classification, and some problems of application to natural flows and deposits. In: Doyle, L.J., Pilkey, O.H. (Eds.), Geology of Continental Slopes, 27. SEPM Special Publication, pp. 75 82. Lowe, D.R., 1982. Sediment gravity flows: II. Depositional models with special reference to the deposits of highdensity turbidity currents. J. Sediment. Petrol. 52, 279 297. Lowe, D.R., 1997. Reinterpretation of depositional processes in a classic flysch sequence (Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma, Discussion. AAPG Bull. 81, 460 465. Lowe, D.R., Guy, M., 2000. Slurry-flow deposits in the Britannia formation (Lower Cretaceous), North Sea: a new perspective on the turbidity current and debris flow problem, Sedimentology, 47. pp. 31 70. Lu, S., Tong, C.F., Lee, D.Y., Zheng, J.H., Shen, J., Zhang, W., et al., 2015. Propagation of tidal waves up in Yangtze Estuary during the dry season. J. Geophys. Res. Ocean. 120, 6445 6473. Lucchitta, B.K., 1979. Landslides in Valles Marineris, Mars. J. Geophys. Res. 84, 8097 8113.

510

Bibliography

Lucchitta, B.K., 1987. Valles Marineris, Mars: Wet debris flows and ground ice. Icarus 72, 411 429. Lu¨dmann, T., Paulat, M., Betzler, C., et al., 2016. Carbonate mounds in the Santaren Channel, Bahamas: a currentdominated periplatform depositional regime. Mar. Geol. 376, 69 85. Lunina, O.V., Gladkov, A.S., 2016. Soft-sediment deformation structures induced by strong earthquakes in southern Siberia and their paleoseismic significance. Sediment. Geol. 344, 5 19. Luo, Z., Zhu, J., Wu, H., Li, X., 2017. Dynamics of the sediment plume over the Yangtze Bank in the Yellow and East China Seas. J. Geophys. Res. Ocean. 122, 10,073 10,090. ,https://doi.org/10.1002/2017JC013215.. Lykousis, V., Sakellariou, D., Locat, J., (Eds.), 2007. Submarine Mass Movements and Their Consequences. Advances in Natural and Technological Hazards Research 27, 424 p. Ma, Y.X., 2009. Continental Shelf Sediment Transport and Depositional Processes on An Energetic, Active Margin: The Waiapu River Shelf, New Zealand (Ph.D. dissertation). The College of William and Mary, Williamsburg, VI, 206 pp. See changes. Macdonald, D.I.M., Moncrieff, A.C.M., Butterworth, P.J., 1993. Giant slide deposits from a Mesozoic fore-arc basin, Alexander Island, Antarctica. Geology 21, 1047 1050. Madrussani, G., Rossi, G., Rebesco, M., Picotti, S., Urgeles, R., Llopart, J., 2018. Sediment properties in submarine mass-transport deposits using seismic and rock-physics off NW Barents Sea. Mar. Geol. 402, 264 278. Magalha˜es, J., da Silva, J.C., Grimshaw, R., Griffiths, S., New, A., 2010. Internal solitary waves off the western Iberian coast and in the Gulf of Cadiz: main hotspots and some generation mechanisms. In: SEASAR 2010: The 3rd International Workshop on Advances in SAR Oceanography From Envisat, ERS and ESA Third Party Missions. 25 29 January 2010, Rome. Magalha˜es, V.H., Pinheiro, L.M., Ivanov, M.K., Kozlova, E., Blinova, V., Kolganova, J., et al., 2012. Formation processes of methane-derived authigenic carbonates from the Gulf of Cadiz. Sediment. Geol. 243 244, 155 168. Major, J.J., 1998. Pebble orientation on large, experimental debris-flow deposits. Sediment. Geol. 117, 151 164. Maier, K.L., Gales, J.A., Paull, C.K., Rosenberger, K., Talling, P.J., Simmons, S.M., Gwiazda, R., McGann, M., Cartigny, M.J.B., Lundsten, E., Anderson, K., Clare, M.A., Xu, J., Parsons, D., Barry, J.P., Wolfson-Schwehr, M., Nieminski, N.M., Sumner, E.J., et al., 2019. Linking Direct Measurements of Turbidity Currents to Submarine Canyon-Floor Deposits. Front. Earth Sci. 7, 144. Available from: https://doi.org/10.3389/feart.2019.00144. Major, J.J., Iverson, R.M., 1999. Debris-flow deposition: effects of pore-fluid pressure and friction concentrated at flow margins. Geol. Soc. Am. Bull. 111 (10), 1424 1434. Malin, M.C., 1992. Mass movements on Venus: preliminary results from Magellan cycle I observations. J. Geophys. Res. 97 (E10), 16,337 16,352. Malinverno, A., Ryan, W.B.F., Auffret, G., Pautot, G., 1988. Sonar images of the path of recent failure events on the continental margin off Nice, France. In: Clifton, H.E. (Ed.), Sedimentologic Consequences of Convulsive Geologic Events. Geological Society of America, pp. 59 75. Special Paper 229. Malkawi, A.I.H., Alawneh, A.S., 2000. Paleoearthquake features as indicators of potential earthquake activities in the Karameh Dam Site. Nat. Hazards 22, 1 16. Maltman, A.J., 1987. Microstructures in deformed sediments, Denbigh Moors, North Wales. Geol. J. 22, 87 94. Maltman, A., 1994a. Deformation structures preserved in the rocks. In: Maltman, A. (Ed.), The Geological Deformation of Sediments. Chapman & Hall, London, pp. 261 307. Maltman, A., 1994b. Introduction and overview. In: Maltman, A. (Ed.), The Geological Deformation of Sediments. Chapman & Hall, London, pp. 1 35. Maltman, A. (Ed.), 1994c. The Geological Deformation of Sediments. Chapman & Hall, London. Maltman, A.J., Hubbard, B., Hambrey, M.J. (Eds.), 2000. Deformation of Glacial Materials. Geological Society of London, Special Publication 176. Manca, F., Capelli, G., La Vigna, F., Mazza, R., Pascarella, A., 2014. Wind-induced salt-wedge intrusion in the Tiber River mouth (Rome-Central Italy). Environ. Earth Sci. 72 (4), 1083 1095. Manica, R., 2012. Sediment gravity flows: study based on experimental simulations, hydrodynamics—natural water bodies, In: Schulz, H.E., Andre´, Simo˜es, L.A., Lobosco, R.J. (Eds.), IntechOpen, ,https://doi.org/ 10.5772/28794.. Available from: ,https://www.intechopen.com/books/hydrodynamics-natural-water-bodies/sediment-gravity-flows-study-based-on-experimental-simulations.. Mankoff, K.D., Straneo, F., Cenedese, C., Das, S.B., Richards, C.G., Singh, H., 2016. Structure and dynamics of a subglacial discharge plume in a Greenlandic fjord. J. Geophys. Res. Ocean. 121, 8670 8688.

Bibliography

511

Marco, S., Katz, O., Dray, Y., 2014. Historical sand injections on the Mediterranean shore of Israel: evidence for liquefaction hazard, Nat. Hazards, 74. pp. 1449 1459. Marr, J.G., Harff, P.A., Shanmugam, G., Parker, G., 2001. Experiments on subaqueous sandy gravity flows: the role of clay and water content in flow dynamics and depositional structures. GSA Bull. 113, 1377 1386. Marshall, N.F., 1978. Large storm-induced sediment slump reopens an unknown Scripps submarine canyon tributary. In: Stanley, D.J., Kelling, G. (Eds.), Sedimentation in Submarine Canyons, Fans, and Trenches. Hutchinson and Ross, Stroudsburg, PA, pp. 73 84. Martin, B.D., Emery, K.O., 1967. Geology of Monterey Canyon, California. AAPG Bull. 51, 2281 2304. Martın-Chivelet, J., Fregenal-Martinez, M.A., Chaco´n, B., 2008. Traction structures in contourites. In: Rebesco, M., ˜ Camerlenghi, A. (Eds.), Contourites. Developments in Sedimentology, 60. pp. 159 182. Martı´n-Chivelet, J., Palma, R.M., Lo´pez-Go´mez, J., Kietzmann, D.A., 2011. Earthquake induced soft-sediment deformation structures in Upper Jurassic open-marine microbialites (Neuque´n Basin, Argentina). Sediment. Geol. 235, 210 221. Martinsen, O.J., 1989. Styles of soft-sediment deformation on a Namurian (Carboniferous) delta slope, Western Irish Namurian Basin, Ireland. In: Whateley, K.G., Pickering, K.T. (Eds.), Deltas: Sites and Traps for Fossil Fuels. Geological Society Special Publication No. 41, pp. 167 177. Martinsen, O.J., 1994. Mass movements. In: Maltman, A. (Ed.), The Geological Deformation of Sediments. Chapman & Hall, London, pp. 127 165. Mascarenhas, A., 2004. Oceanographic validity of buffer zones for the east coast of India: a hydrometeorological perspective. Curr. Sci. 86 (3), 399 406. Massari, F., Guibaudo, G., D’Alessandro, T., Davaud, E.J., 2001. Water-upwelling pipes and soft-sedimentdeformation structures in lower Pleistocene calcarenites (Salento, southern Italy). GSA Bull. 113, 545 560. Maslin, E., 2015. Scientific Drilling. Offshore Engineer, OE, ,http://www.oedigital.com/geoscience/item/10495scientific-drilling. (accessed 29.10.15.). Maslin, M., Owen, M., Day, S., Long, D., 2004. Linking continental-slope failures and climate change: testing the clathrate gun hypothesis. Geology 32, 53 56. Maslin, M., Vilela, C., Mikkelsen, N., Grootes, P., 2005. Causes of catastrophic sediment failures of the Amazon Fan. Quat. Sci. Rev. 24, 2180 2193. Masson, D.G., van Niel, B., Weaver, P.P.E., 1997. Flow processes and sediment deformation in the Canary Debris Flow on the NW African Continental Rise. Sediment. Geol. 110, 163 179. Masson, D.G., Canals, M., Alonso, B., Urgeles, R., Huhnerbach, V., 1998. The Canary Debris flow: source area morphology and failure mechanisms. Sedimentology 45, 411 432. Masson, D.G., Howe, J.A., Stoker, M.S., 2002a. Bottom-current sediment waves, sediment drifts and contourites in the northern Rockall Trough. Mar. Geol. 192, 215 237. Masson, D.G., Watts, A.B., Gee, M.J.R., Urgeles, R., Mitchell, N.C., Le Bas, T.P., et al., 2002b. Slope failures on the flanks of the western Canary Islands. Earth-Sci. Rev. 57, 1 35. Masson, D.G., Harbitz, C.B., Wynn, R.B., Pedersen, G., Løvholt, F., 2006. Submarine landslides: processes, triggers, and hazard prevention. R. Soc. Lond. Trans. Ser. A 364 (1845), 2009 2039. Masunaga, E., Hommaa, H., Yamazaki, H., Fringer, O.B., Nagai, T., Kitade, Y., et al., 2015. Mixing and sediment resuspension associated with internal bores in a shallow bay. Cont. Shelf Res. 110, 85 99. Matano, R.P., Palma, E.D., Piola, A.R., 2010. The influence of the Brazil and Malvinas Currents on the southwestern Atlantic shelf circulation. Ocean. Sci. 6, 983 995. Matsumoto, D., Naruse, H., Fujino, S., Surphawajruksakul, A., Jarupongsakul, T., Sakakura, N., et al., 2008. Truncated flame structures within a deposit of the Indian Ocean Tsunami: evidence of syn-sedimentary deformation. Sedimentology 55, 1559 1570. Matsuyama, M., Ohta, S., Hibiya, T., Yamada, H., 1993. Strong tidal currents observed near the bottom in the Suruga Trough, Central Japan. J. Oceanogr. 49, 683 696. Mattox, T.N., Mangan, M.T., 1997. Littoral hydrovolcanic explosions: a case study of lava seawater interaction at Kilauea Volcano. J. Volcanol. Geotherm. Res. 75, 1 17. Maxworthy, T., 1979. A note on the internal solitary waves produced by tidal flow over a threedimensional ridge. J. Geophys. Res. 84, 338 346. Mayer, L.M., 1994. Surface area control of organic carbon accumulation in continental shelf sediments, Geochimica et Cosmochimica Acta, 58, 1271 1284.

512

Bibliography

Mazumder, R. (Ed.), 2016. Sediment Provenance: Influences on Compositional Change From Source to Sink. Elsevier, 614 p. Mazumder, R., Van Loon, A.J., Arima, M., 2006. Soft-sediment deformation structures in the Earth’s oldest seismites. Sediment. Geol. 186, 19 26. Mazumder, R., Van Loon, A.J., Malviya, V.P., Arima, M., Ogawa, Y., 2016. Soft-sediment deformation structures in the Mio-Pliocene Misaki Formation within alternating deep sea clays and volcanic ashes (Miura Peninsula, Japan). Sediment. Geol. 344, 323 335. MBARI (Monterey Bay Aquarium Research Institute), 2003. Monterey Canyon turbidity flow: caught in the act. . .again!, 17 19 December 2002. ,http://www.mbari.org/benthic/Turbidity%20Event%202002/canyon% 20events.html. (accessed 16.09.07.). McAdoo, B.G., Pratson, L.F., Orange, D.L., 2000. Submarine landslide geomorphology, US continental slope. Mar. Geol. 169, 103 136. McCaffrey, W.D., Eggenhuisen, J., Haughton, P., Butler, R., Barker, S., Del Pino Sanchez, S.A., et al., 2008. Impact of multiple scales of remobilization on deep marine clastic reservoir architecture: the Aptian Britannia Sandstone Formation, North Sea. In: Schofield, K., Rosen, N.C., Pfeiffer, D., Johnson, S. (Eds.), Answering the Challenges of Production from Deep-Water Reservoirs: Analogues and Case Histories to Aid a New Generation: 28th Annual Gulf Coast Section SEPM Foundation Bob F. Perkins Research Conference, Houston, TX, 7 9 December 2008, Program and Abstracts. McCave, I.,N., Jones, P.N., 1988. Deposition of ungraded muds from high-density non turbulent turbidity currents. Nature 333, 250 252. ISSN 0016-7649. McConnel, R.G., Brock, R.W., 1904. The great landslide at Frank, Alberta. Department of Interior, Annual Report 1902 1903, Sess. Paper 25, pp. 4 17. McEwen, A.S., 1989. Mobility of rock avalanches: evidence from Valles Marineris, Mars. Geology 17, 1111 1114. McGowen, J.H., Garner, L.E., Wilkinson, B.H., 1977. The Gulf Shoreline of Texas: processes, characteristics, and factors in use. The University of Texas at Austin, Bureau of Economic Geology. Geol. Circ. 77, 27. McGregor, B.A., Rothwell, R.G., Kenyon, N.H., Twichell, D.C., 1993. Salt tectonics and slope failure in an area of salt domes in the northwestern Gulf of Mexico. In: Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone, 2002. U.S. Geological Survey Bulletin, pp. 92 96. McHugh, C.M.G., Damuth, J.E., Mountain, G.S., 2002. Cenozoic mass-transport facies and their correlation with relative sea-level change, New Jersey continental margin. Mar. Geol. 184, 295 334. McKee, E.D., Weir, G.W., 1953. Terminolgy for stratification and cross-stratification in sedimentary rocks. Geol. Soc. Am. Bull. 64, 385. McLean, C.E., Brown, D.J., Rawcliffe, H.J., 2016. Extensive soft-sediment deformation and peperite formation at the base of a rhyolite lava: Owyhee Mountains, SW Idaho, USA. Bull. Volcanol. 78, 42. Available from: https://doi.org/10.1007/s00445-016-1035-2. McMurtry, G.M., Fryer, G.J., Tappin, D.R., Wilkinson, I.P., Williams, M., Fietzke, J., et al., 2004. Megatsunami deposits on Kohala volcano, Hawaii, from flank collapse of Mauna Loa. Geology 32, 741 744. McPherson, J.G., Shanmugam, G., Moiola, R.J., 1987. Fan-deltas and braid deltas: varieties of coarse-grained deltas. Geol. Soc. Am. Bull. 99, 331 340. McPherson, J.G., Shanmugam, G., Moiola, R.J., 1988. Fan deltas and braid deltas: varieties of coarse-grained deltas: discussion and reply, reply. GSA Bull. 100, 1309 1310. Mearns, D.L., Hine, A.C., Riggs, S.R., 1988. Comparison of sonographs taken before and after Hurricane Diana; Onslow Bay, North Carolina. Geology 16, 267 270. Meckel III, L.D., 2009. Reservoir characteristics and classification of sand-prone submarine mass-transport deposits. In: Shipp, et al., (Eds.), Mass-Transport Deposits in Deepwater Settings. SEPM Special Publication No. 96, pp. 423 454. Meckel, T., 2010. Classifying and characterizing sand-prone submarine mass-transport deposits. AAPG Annual Convention and Exhibition, New Orleans, LA, 11 14 April 2010, Search and Discovery Article #50270 (2010), Posted 16 July 2010. ,http://www.searchanddiscovery.com/documents/2010/50270meckel/ndx_meckel. pdf. (accessed 28.11.10.). Mellere, D., Plink-Bj˝orklund, P., Steel, R., 2002. Anatomy of shelf deltas at the edge of a prograding Eocene shelf margin, Spitsbergen. Sedimentology 49, 1181 1206.

Bibliography

513

Melosh, H.J., 1979. Acoustic fluidization: a new geologic process? J. Geophys. Res. 84, 7513 7520. Melosh, H.J., 1989. Impact Cratering: S Geologic Process. Research Supported by NASA. Oxford Monographs on Geology and Geophysics, 11. Oxford University Press, New York, p. 253. Meng, Q., Qu, H., Hu, J., 2007. Triassic deep-marine sedimentation in the western Qinling and Songpan terrane. Sci. China Ser. D-Earth Sci. 50 (Suppl. II), 246 263. Meng, X., Peng, Z., Yang, H., Allman, S., 2013. Hurricane Irene’s impacts on the aftershock sequence of the 2011 Mw5.8 Virginia earthquake. In: American Geophysical Union, Fall Meeting 2013, Abstract# S51B-2369. Menzies, J., van der Meer, J.J.M., Ravier, E., 2016. A kinematic unifying theory of microstructures in subglacial tills. Sediment. Geol. 344, 57 70. Menard, H.W., 1952. Deep ripple marks in the sea. J. Sediment. Petrol. 22, 3 9. Meshram, D.C., Sangode, S.J., Gujar, A.R., Ambre, N.V., Dhongle, P., 2011. Occurrence of soft sediment deformation at Dive Agar beach, west coast of India: possible record of the Indian Ocean tsunami (2004). Nat. Hazards 57, 385 393. Me`tivier, F., Lajeunesse, E., Cacas, M.-C., 2005. Submarine canyons in the bathtub. J. Sediment. Res. 75, 6 11. Metz, J.M., 2010. A Study of the Record of Ancient Sedimentary Rocks on Mars Using MER, HIRISE and CRISM Images (Ph.D. dissertation). California Institute of Technology, Pasadena, CA, p. 250. Metz, M., Grotzinger, J., Okubo, C., Milliken, R., 2010. Thin-skinned deformation of sedimentary rocks in Valles Marineris, Mars. J. Geophys. Res. 115, E11004. Available from: https://doi.org/10.1029/2010JE003593. 201. Meyer, D., Zarra, L., Yun, J., 2007. From BAHA to Jack, evolution of the Lower Tertiary Wilcox trend in the deepwater Gulf of Mexico. Sediment. Rec. 5, 4 9. Miall, A.D., 1984. Principles of Sedimentary Basin Analysis. Springer-Verlag, New York, 490 p. Miall, A.D., 1985. Architectural-element analysis: a new method of facies analysis applied to fluvial deposits. Earth-Sci. Rev. 22, 261 308. Miall, A.D., 1996. The Geology of Fluvial Deposits. Springer, Berlin, p. 582. Miall, A.D., 2014. The emptiness of the stratigraphic record: a preliminary evaluation of missing time in the Mesaverde Group, Book Cliffs, Utah, U.S.A. J. Sediment. Res. 84, 457 469. Michalik, J., 1997. Tsunamites in a storm-dominated Anisian carbonate ramp (Vysoka´ Formation, Male´ Karpaty Mts., western Carpathians). Geol. Carpath. 48, 221 229. Michels, K.H., Kuhn, G., Hillenbrand, C.D., Diekmann, B., Futterer, D.K., Grobe, H., et al., 2002. The southern Weddell Sea: combined contourite-turbidite sedimentation at the southeastern margin of the Weddell Gyre. In: Stow, D.A.V., Pudsey, C.J., Howe, J.A., Fauge`res, J.-C., Viana, A.R. (Eds.), Deep-Water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics, vol. 22. Geol. Soc., London, pp. 305 324. Michels, K.H., Suckow, A., Breitzke, M., Kudrass, H.R., Kottke, B.V., 2003. Sediment transport in the shelf canyon “Swatch of No Ground” (Bay of Bengal). Deep Sea Res. Part II: Top. Stud. Oceanogr. 50, 1003 1022. Middleton, G.V. (Ed.), 1965. Primary Sedimentary Structures and Their Hydrodynamic Interpretation, 12. SEPM Special Publication, p. 265. Middleton, G.V., 1966. Experiments on density and turbidity currents. I. Motion of the head. Can. J. Earth Sci. 3, 523 546. Middleton, G.V., 1967. Experiments on density and turbidity currents: III. Deposition of sediment. Can. J. Earth Sci. 4, 475 505. Middleton, G.V., 1970. Experimental studies related to problems of flysch sedimentation. In: Lajoie, J. (Ed.), Flysch Sedimentology in North America. Geological Association of Canada, pp. 253 272. Special Paper No. 7. Middleton, G.V., 1973. Johannes Walther’s Law of the correlation of fades. GSA Bulletin 84, 979 988. Middleton, G.V., 1993. Sediment deposition from turbidity currents. Annu. Rev. Earth Planet. Sci. 21, 89 114. Middleton, G.V., Bouma, A.H. (Eds.), 1973. Turbidites and Deep-Water Sedimentation. SEPM Pacific Section Short Course, Anaheim, CA. Middleton, G.V., Hampton, M.A., 1973. Sediment gravity flows: mechanics of flow and deposition. In: Middleton, G.V., Bouma, A.H. (Eds.), Turbidites and Deep-Water Sedimentation. SEPM Pacific Section Short Course, Anaheim, CA, pp. 1 38. Middleton, G.V., Southard, J.B., 1977. Mechanics of sediment movement. Short Course No. 3, Eastern section of SEPM, Binghamton, NY, p. 10.2. Middleton, G.V., Wilcock, P.R., 1994. Mechanics in the Earth and Environmental Sciences. Cambridge University Press, Cambridge, p. 459.

514

Bibliography

Mienert, J., Vanneste, M., Bunz, S., Andreassen, K., Haflidason, H., Sejrup, H.P., 2005. Ocean warming and gas hydrate stability on the mid-Norwegian margin at the Storegga Slide. Mar. Petrol. Geol. 22, 233 244. Migeon, S., 2000. Dunes ge´antes et leve´es se´dimentaires en domainemarin profond: Approche morphologique, sismique et se´dimentologique (Ph.D. thesis). Universite´ Bordeaux 1, Talence, 288 pp. Migeon, S., Ducassou, E., Le Gonidec, Y., Rouillard, P., Mascle, J., Revel-Rolland, M., 2010. Lobe construction and sand/mud segregation by turbidity currents and debris flows on the western Nile deep-sea fan (Eastern Mediterranean). Sediment. Geol. 229, 124 143. Mikhailova, M.V., Bellotti, P., Valeri, P., Tortora, P., 1998. The Tiber river delta and the hydrological and morphological features of its formation. Water Resour. 25 (5), 572 582. Translated from Vodnye Resursy, Voi 25, No. 5, 1998, pp. 620-630. Mikhailov, V.N., Kravtsova, V.I., Isupova, M.V., 2015. Impact of reservoirs on the hydrological regime and morphology of the lower reaches and delta of the Zambezi River (Mozambique). Water Resour. 42 (2), 170 185. Milia, A., Torrente, M.M., Zuppetta, A., 2003. Offshore debris avalanches at Somma Vesuvius volcano (Italy): implications for hazard evaluation. J. Geol. Soc. Lond. 160, 309 317. Milia, A., Torrente, M.M., Giordano, F., 2006. Chapter 4 Gravitational instability of submarine volcanoes offshore Campi Flegrei (Naples Bay, Italy). Dev. Volcanol. 9, 69 83. Miller, D.J., 1960. The Alaska earthquake of July 10, 1958: giant wave in Lituya Bay. Seismol. Soc. Am. Bull. 50, 253 266. Milliff, R.F., Morzel, J., Chelton, D.B., Freilich, M.H., 2004. Wind stress curl and wind stress divergence biases from rain effects on QSCAT surface wind retrievals. J. Atmos. Ocean. Technol. 21 (8), 1216 1231. Milliman, J.D., 2001. River inputs. In: Steele, J.H., Thorpe, S.A., Turekian, K.K. (Eds.), Encyclopedia of Ocean Sciences. Academic Press, pp. 2419 2427. Milliman, J.D., Meade, R.H., 1983. World-wide delivery of river sedment to the oceans. J. Geol. 91, 1 21. Milliman, J.D., Shen, H., Yang, Z., Meade, R.H., 1985. Transport and deposition of river sediment in the Changjiang estuary and adjacent continental shelf. Cont. Shelf Res. 4, 37 45. Milliman, J.D., Lin, S.W., Kao, S.J., Liu, J.P., Liu, C.S., Chiu, J.K., et al., 2007. Short-term changes in seafloor character due to flood-derived hyperpycnal discharge: typhoon Mindule, Taiwan, July 2004. Geology 35 (9), 779 782. Miramontes, E., Garreau, P., Caillaud, M., Jouet, G., Pellen, R., Herna´ndez-Molina, F.J., et al., 2019. Contourite distribution and bottom currents in the NW Mediterranean Sea: coupling seafloor geomorphology and hydrodynamic modelling. Geomorphology 333, 43 60. Available from: https://doi.org/10.1016/j.geomorph.2019.02.030. MMS (Minerals Management Service), 2005. Oil and gas production in the Gulf of Mexico continues to stabilize, MMS Issues Damage Assessment and Review of Hurricane Ivan, News Release 3223, 2 February 2005. MMS, U.S. Department of the Interior. ,http://www.mms.gov/ooc/press/2005/press0202.htm. (accessed 31.05.10.). MMS (Minerals Management Service), 2006. Impact assessment of offshore facilities from Hurricanes Katrina and Rita, News Release 3418, 19 January 2006. MMS, U.S. Department of the Interior. ,http://www.gomr.mms. gov/homepg/whatsnew/newsreal/2006/060119.pdf. (accessed 31.05.10.). Mitchell, D.A., Teague, W.J., Jarosz, E., Wang, D.W., 2005a. Observed currents over the outer continental shelf during Hurricane Ivan. Geophys. Res. Lett. 32, L11610. Available from: https://doi.org/10.1029/ 2005GL023014. Mitchell, D.A., Teague, W.A., Jarosz, E., Wang, D.W., Hulbert, M.S., 2005b. Current measurements under Hurricane Ivan. ,http://www.agu.org/meetings/sm05/sm05-sessions/sm05_OS34A.html. (accessed 16.09.07.). Mitchell, J.K., Holdgate, G.R., Wallace, M.W., Gallagher, S.J., 2007. Marine geology of the Quaternary Bass Canyon system, southeast Australia: a cool-water carbonate system. Mar. Geol. 237, 71 96. Mitchum Jr., R.M., 1977. Seismic stratigraphy and global changes of sea level. Part 11: glossary of terms used in seismic stratigraphy. In: Payton, C.E. (Ed.), Seismic Stratigraphy Applications to Hydrocarbon Exploration, vol. 26. A.A.P.G. Memoir, pp. 205 212. Mitchum, R., Vail, P., Thompson, S., 1977. Seismic stratigraphy and global changes in sea level, part 2: the depositional sequence as the basic unit for stratigraphic analysis. In: Payton, C. (Ed.), Seismic Stratigraphy: Application to Hydrocarbon Exploration, 26. AAPG Memoir, pp. 53 62.

Bibliography

515

Mitsuzawa, K., Momma, H., Fukasawa, M., Hotta, H., 1993. Observation of deep sea current and change of bottom shapes in the Suruga Trough: OCEANS apos 93. Eng. Harmony Oceanapos, Proc. 3 (18 21), III149 III154. Available from: https://doi.org/10.1109/OCEANS.1993.326176. ,http://ieeexplore.ieee.org/ Xplore/login.jsp?url 5 /iel2/1081/7739/00326176.pdf?isnumber 5 7739&prod 5 CNF&arnumber 5 326176&arSt 5 III149&ared 5 III154 1 vol.3&arAuthor 5 Mitsuzawa%2C 1 K.%3B 1 Momma%2C 1 H.%3B 1 Fukasawa%2C 1 M. %3B 1 Hotta%2C 1 H. (accessed 16.09.07.). Miyamoto, H., Dohm, J.M., Baker, V.R., Beyer, R.A., Bourke, M., 2004. Dynamics of unusual debris flows on Martian sand dunes. Geophys. Res. Lett. 31, L13701. Available from: https://doi.org/10.1029/2004GL020313. Miramontes, E., Penven, P., et al., 2019. The influence of bottom currents on the Zambezi Valley morphology (Mozambique Channel, SW Indian Ocean): in situ current observations and hydrodynamic modelling. Mar. Geol. 410, 42 55. Available from: https://doi.org/10.1016/j.margeo.2019.01.002. Moberly, R., Jenkyns, H.C., 1981. 10. Cretaceous volcanogenic sediments of the Nauru Basin, Deep Sea Drilling Project Leg 61, Initial Reports of the Deep Sea Drilling Project, vol. 61. United States Government Printing Office, Washington, DC, pp. 533 548, ,https://doi.org/10.2973/dsdp.proc.61.110.198.. Moernaut, J., De Batist, M., 2011. Frontal emplacement and mobility of sublacustrine landslides: results from morphometric and seismostratigraphic analysis. Mar. Geol. 285, 29 45. Mohrig, D., Whipple, K.X., Hondzo, M., Ellis, C., Parker, G., 1998. Hydroplaning of subaqueous debris flows. GSA Bull. 110, 387 394. Mohindra, R., Thakur, V.C., 1998. Historic large earthquakeinduced soft sediment deformation features in the Sub-Himalayan Doon Valley. Geol. Mag. 135, 269 281. Available from: https://doi.org/10.1017/ S0016756898008413. Moiola, R.J., Shanmugam, G., 1984. Submarine fan sedimentation, Ouachita Mountains, Arkansas and Oklahoma. Trans. Gulf Coast. Assoc. Geol. Soc. 34, 175 182. Montgomery, D.R., Som, S.M., Jackson, M.P.A., Schreiber, B.C., Gillepsie, A.R., 2009. Continental-scale salt tectonics on Mars and the origin of Valles Marineris and associated outflow channels. Geol. Soc. Am. Bull. 121, 117 133. Moore, D.G., 1961. Submarine slumps. J. Sediment. Petrol. 31, 343 371. Moore, D., 1966. Deltaic sedimentation. Earth-Sci. Rev. 1, 87 104. Moore, D.G., 1977. Submarine slides. In: Voight, I.B. (Ed.), Rock Slides and Avalanches, 1. Elsevier, Amsterdam, pp. 563 604. Natural Phenomena. Moore, D.G., Curray, J.R., Emmel, F.J., 1976. Large submarine slide (olistostrome) associated with Sunda arc subduction zone, northeast Indian Ocean. Mar. Geol. 21, 211 226. Moore, J.G., Clague, D.A., Holcomb, R.T., Lipman, P.W., Normark, W.R., Torresan, M.E., 1989. Prodigious submarine landslides on the Hawaiian Ridge. J. Geophys. Res. 94, 17465 17484. Moore, J.G., Normark, W.R., Holcomb, R.T., 1994. Giant Hawaiian landslides. Annu. Rev. Earth Planet. Sci. 22, 119 144. Moraes, M.A.S., Maciel, W.B., Braga, M.S.S., Viana, A.R., 2007. Bottom-current reworked Palaeocene-Eocene deepwater reservoirs of the Campos Basin, Brazil. In: Viana, A.R., Rebesco, M. (Eds.), Economic and Palaeoceanographic Significance of Contourite Deposits, 276. Geological Society of London Special Publication., pp. 81 94. Morales de Luna, T., Ferna´ndez Nieto, E.D., Castro Dı´az, M.J., 2017. Derivation of a multilayer approach to model suspended sediment transport: application to hyperpycnal and hypopycnal plumes. Commun. Comput. Phys. 22 (5), 1439 1485. Moretti, M., Owen, G., Tropeano, M., 2011. Soft-sediment deformation induced by sinkhole activity in shallow marine environments: a fossil example in the Apulian Foreland (Southern Italy). Sediment. Geol. 235, 331 342. Morgenstern, N.R., 1967. Submarine slumping and the initiation of turbidity currents. In: Richards, A.F. (Ed.), Marine Geotecnnique. University of Illinois Press, Urbana, pp. 189 220. Morisawa, M., 1968. Streams Their Dynamics and Morphology. McGraw-Hill Book Company, New York, p. 175. Moritz, H.R., 2004. The effects of infragravity energy and storm-induced current on short waves beyond the surf zone. In: 8th International Workshop on Wave Hindcasting and Forecasting: North Shore, Oahu, HI, 14 19 November 2004. ,http://www.waveworkshop.org/8thWaves/Papers/O2.pdf. (accessed 16.09.07.).

516

Bibliography

Morner, N.A., 1991. Intense earthquakes and seismotectonics as a function of glacial isostasy. Tectonophysics 117, 139 153. Morozov, E.G., Trulsen, K., Velarde, M.G., Vlasenko, V.I., 2002. Internal tides in the Strait of Gibraltar. J. Phys. Oceanogr. 32, 3193 3206. Morton, R.A., 1993. Attributes and origins of ancient submarine slides and filled embayments: examples from the Gulf Coast Basin. AAPG Bull. 77, 1064 1081. Moscardelli, L., Wood, L., 2007. New classification system for mass transport complexes in offshore Trinidad. Basin Res. Available from: https://doi.org/10.1111/j.1365-2117.2007.00340.x. Moscardelli, L., Wood, L., Mann, P., 2006. Mass-transport complexes and associated processes in the offshore area of Trinidad and Venezuela. Am. Assoc. Petrol. Geol. Bull. 90, 1059 1088. Mosher, D.C., Piper, D.J.W., Campbell, D.C., Jenner, K.A., 2004. Near-surface geology and sedimentfailure geohazards of the central Scotian slope. AAPG Bull. 88, 703 723. Mosher, D.C., Shipp, R.C., Moscardelli, L., Chaytor, J.D., Baxter, C.D.P., Lee, H.J., et al., 2010. Submarine mass movements and their consequences, Adv. Nat. Technol. Hazards Res., 28. p. 786. Mulder, T., 2006. Book review of “Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs” (Handbook of Petroleum Exploration and Production, vol. 5), by G. Shanmugam, 2006. Elsevier B.V. J. Sediment. Res. ,http://www.sepm.org/jsr/book_revs/2006_revs/br_shanmugam.pdf. (accessed 30.05.11.). Mulder, T., 2011. Chapter 2-Gravity processes and deposits on continental slope, rise and abyssal plains. In: Hueneke, H., Mulder, T. (Eds.), Deep-Sea Sediments, 63. Elsevier, Amsterdam, pp. 25 148. Developments in Sedimentology. Mulder, T., Syvitski, J.P.M., 1995. Turbidity currents generated at river mouths during exceptional discharges to the world oceans. J. Geol. 103, 285 299. Mulder, T., Cochonat, P., 1996. Classification of offshore mass movements. J. Sediment. Res. 66, 43 57. Mulder, T., Alexander, J., 2001. The physical character of subaqueous sedimentary density flows and their deposits. Sedimentology 48, 269 299. Mulder, T., Migeon, S., Savoye, B., Faugeres, J.-C., 2001. Inversely graded turbidite sequences in the deep Mediterranean. a record of deposits from flood-generated turbidity currents? Geo-Mar. Lett. 21, 86 93. Mulder, T., Migeon, S., Savoye, B., Faugeres, J.-C., 2002. Reply to discussion by Shanmugam on. Mulder et al. (2001, Geo-Marine Letters, 21, 86 93) Inversely graded turbidite sequences in the deep Mediterranean. A record of deposits from flood-generated turbidity currents? Geo-Mar. Lett. 22, 112 120. Mulder, T., Syvitski, J.P.M., Migeon, S., Fauge`res, J.-C., Savoye, B., 2003. Marine hyperpycnal flows: initiation, behavior and related deposits. A review. Mar. Petrol. Geol. 20, 861 882. Mulder, T., Razin, P., Fauge`res, J.-C., 2009a. Hummocky cross-stratification-like structures in deep-sea turbidites: upper cretaceous basque basins (Western Pyrenees, France). Sedimentology 56, 997 1015. Mulder, T., Zaragosi, S., Razin, P., Grelaud, C., Lanfumey, V., Bavoil, F., 2009b. A new conceptual model for the deposition process of homogenite: application to a cretaceous megaturbidite of the western Pyrenees (Basque region, SW France). Sediment. Geol. 222, 263 273. Mulder, T., Philippe, R., Fauge`res, J.-C., Ge´rard, J., 2010. Reply to the discussion by Roger Higgs on ‘Hummocky crossstratification-like structures in deep-sea turbidites: Upper Cretaceous Basque basins (Western Pyrenees, France)’ by Mulder et al., Sedimentology, 56, 997 1015. Sedimentology. ,https://doi.org/10.1111/j.13653091.2010.01162.. Mulder, T., Philippe, R., Fauge`res, J.-C., Ge´rard, J., 2011. Reply to the discussion by Roger Higgs on ‘Hummocky cross stratification-like structures in deep-sea turbidites: Upper Cretaceous Basque basins (Western Pyrenees, France)’ by Mulder et al., Sedimentology, 56: 997 1015, Sedimentology 58, 671 577. Mulder, T., Hassan, R., Ducassou, E., Zaragosi, S., Gonthier, E., Hanquiez, V., et al., 2013. Contourites in the Gulf of Cadiz: a cautionary note on potentially ambiguous indicators of bottom current velocity. Geo-Mar. Lett. 33, 357 367. Mulder, T., et al., 2019. Contour current imprints and contourite drifts in the Bahamian archipelago. Sedimentology 66 (4), 1192 1221. Mullins, H.T., Neumann, A.C., Wilber, R.J., Hine, A.C., Chinburg, S.J., 1980. Carbonate sediment drifts in the northern Straits of Florida. AAPG Bull. 64, 1701 1717. Mullins, H.T., Gardulski, A.F., Wise Jr., S.W., Applegate, J., 1987. Implications for seismic stratigraphic succession and Loop Current/Gulf Stream circulation. GSA Bull. 98, 702 713.

Bibliography

517

Mu¨nchow, A., 1998. Tidal currents in a topographically complex channel. Cont. Shelf Res. 18, 561 584. Munk, W.H., Wunsch, C., 1998. Abyssal recipes II: energetics of tidal and wind mixing. Deep-Sea Res. 45, 1977 2010. Murchison, R.I., 1827. Supplementary remarks on the oolitic series in the counties of Sutherland and Ross, and in the Hebrides. Trans. Geol. Soc. 2, 353. Murphy, M.A., Schlanger, S.O., 1962. Sedimentary structures in llhas sao Sebastiao Formations (Cretaceous), Reconcavo Basin, Brazil. AAPG Bull. 46, 457 477. Murray, S.P., 1970. Bottom currents near the coast during Hurricane Camille. J. Geophys. Res. 75, 4579 4582. Murray, J., Renard, A.F., 1891. Report on deep-sea deposits based on specimens collected uring he voyage of H. M.S. Challenger in the years 1872 1876. Government Printer, Challenger Reports, London. Muto, T., Steel, R.J., 1997. The Middle Jurassic Oseberg delta, northern North Sea; a sedimentological and sequence stratigraphic interpretation. AAPG Bull. 81, 1070 1086. Mutti, E., 1977. Distinctive thin-bedded turbidite facies and related depositional environments in the Eocene Hecho Group (southcentral Pyrenees, Spain). Sedimentology 24, 107 131. Mutti, E., G. Davoli, and R. Tinterri. 1994. Flood-related gravity-flow deposits in fluvial and fluvio-deltaic depositional systems and their sequencestratigraphic implications. In: Posamentier, H.W. and Mutti, E. (ed.) Second High-Resolution Sequence Stratigraphy Conference (tremp, abstract book), 137 143. Mutti, E., 1992. Turbidite Sandstones. Agip Special publication, Milan, p. 275. Mutti, E., 2009. The future of field-based stratigraphic and sedimentologic studies from a personal perspective. J. Mediterr. Earth Sci. 1, 89 90. Mutti, E., 2019. Thin-bedded plumites: an overlooked deep-water deposit. J. Mediterr. Earth Sci. 11. in press. Mutti, E. 2020. The beauty of rocks: A journey through the Pyrenees with Emiliano Mutti (90 episodes of videos). Accessed June 21, 2020. https://www.eni.com/en-IT/operations/emiliano-mutti.html. Mutti, E., Normark, W.R., 1987. Comparing examples of modern and ancient turbidite systems: problems and concepts. In: Leggett, J.R., Zuffa, G.G. (Eds.), Marine Clastic Sedimentology: Concepts and Case Studies. Graham and Trotman, London, pp. 1 37. Mutti, E., Ricci Lucchi, F., 1972. Turbidites of the northern Apennines: introduction to facies analysis (English translation by Nilsen, T.H., 1978). Int. Geol. Rev. 20, 125 166. Mutti, E., Carminatti, M., 2011. Deep-water sands in the Brazilian offshore basins. AAPG Search and Discovery Article 30219. ,http://www.searchanddiscovery.com/documents/2012/30219mutti/ndx_mutti.pdf. (accessed 29.10.15.). Mutti, E., Ricci Lucchi, F., Segure, T.,M., Zanzucchi, G., 1984a. Seismoturbidites: a new group of resedimented deposits. In: Cita, M.B., Ricci Lucchi, F. (Eds.), Seismicity and Sedimentation. Elsevier Scientific Publication, Amsterdam, pp. 103 116. Mutti, E., Allen, G.P., Rosell, J., 1984b. Sigmoidal cross stratification and sigmoidal bars: depositional features diagnostic of tidal sandstones (abstract). In: International Association of Sedimentologists, 5th European Regional Meeting of Sedimentology, Abstracts, pp. 312 313. Mutti, E., Davoli, G., Tinterri, R., Zavala, C., 1996. The importance of fluvio-deltaic systems dominated by catastrophic flooding in tectonically active basins. Mem. di Scienze Geologiche 48, 223 291. Mutti, E., Tinterri, R., Remacha, E., Mavilla, N., Angella, S., Fava, L., 1999. An Introduction to the Analysis of Ancient Turbidite Basins From an Outcrop Perspective. American Association of Petroleum Geologists Continuing Education Course, Tulsa, OK, Note Series No. 39, 61 p. Mutti, E., Tinterri, R., Benvelli, G., di Biase, D., Gavanna, G., 2003. Deltaic, mixed and turbidite sedimentation of ancient foreland basins. Mar. Petrol. Geol. 20, 733 755. Mutti, E., Bernoulli, D., Ricci Lucchi, F., Tineterri, R., 2009. Turbidites and turbidity currents from Alpine ‘flysch’ to the exploration of continental margins. Sedimentology 56, 267 318. Nadim, F., Kjekstad, O., Peduzzi, P., Herold, C., Jaedicke, C., 2006. Global landslide and avalanche hotspots. Landslides 3, 159 173. Nanfito, A.F., 2008. Timing of Diagenesis and Deformation of Carboniferous Gypsum & Anhydrites in Spitsbergen (B.Sc. thesis). Department of Geology and Geography, University of Nebraska, Omaha, NE, p. 62. Nardin, T.R., Hein, F.J., Gorsline, D.S., Edwards, B.D., 1979. A review of mass movement processes, sediment and acoustic characteristics, and contrasts in slope and base-of-slope systems versus canyon-fan-basin floor

518

Bibliography

systems. In: Doyle, L.J., Pilkey, O.H. (Eds.), Geology of Continental Slopes. Society of Economic Paleontologists and Mineralogists, pp. 61 73. Special Publication 27. NASA (National Aeronautics and Space Administration), 2006a, Shelf water during fair-weather on 5 April 2000. ,http://veimages.gsfc.nasa.gov/825/S2000096172719_md.jpg. (accessed 10.11.07.). NASA (National Aeronautics and Space Administration), 2006b, Shelf water stirred by Hurricane Floyd on 16 September 1999. ,http://veimages.gsfc.nasa.gov/1127/S19992591701393.jpg. (accessed 10.11.07.). NASA (National Aeronautics and Space Administration), 2017. Earth observatory. ,https://earthobservatory. nasa.gov/. (accessed 08.10.17.). NASA (National Aeronautics and Space Administration), 2019. Earth observatory. ,https://earthobservatory. nasa.gov/. (accessed 30.01.18.). Natland, M.L., 1967. New classification of water-laid clastic sediments. AAPG Bull. 51, 476. Nelson, C.H., Meyer, A.W., Thor, D., Larsen, M., 1986. Crater Lake, Oregon: a restricted basin with base-of-slope aprons of nonchannelized turbidites. Geology 14, 238 241. Nelson, C.H., Twichell, D.C., Schwab, W.C., Lee, H.J., Kenyon, N.H., 1992. Upper Pleistocene turbidite sand beds and chaotic silt beds in the channelized, distal, outer-fan lobes of the Mississippi fan. Geology 20, 693 696. Nelson, C.H., Baraza, J., Maldonado, A., 1993. Mediterranean undercurrent sandy contourites, Gulf of Cadiz, Spain. Sediment. Geol. 82, 103 131. Nelson, H., Escutia, C., Damuth, J.E., Twichell Jr., D.C., 2009. Interplay of mass-transport and turbidite-system deposits in different active tectonic and passive continental margin settings: external and local controlling factors. In: Shipp, et al., (Eds.), Mass-Transport Deposits in Deepwater Settings. SEPM Special Publication No. 96, pp. 39 68. Nefzaoui, E., Skurtys, O., 2012. Impact of a liquid drop on a granular medium: inertia, viscosity and surface tension effects on the drop deformation. Exp. Therm. Fluid Sci. 41, 43 50. Nemec, W., 1990. Aspects of sediment movement on steep delta slopes. In: Colella, A., Prior, D.B. (Eds.), CoarseGrained Deltas. International Association of Sedimentologists, pp. 29 73. Special Publication No. 10. Newman, M. St., Leeder, M.L., Woodruff, A.H.W., Hatton, I.R., 1993. The geology of the Gryphon Field. In: Parker, J.D. (Ed.), Proceedings of the 4th Conference of the Geological Society, Petroleum Geology of Northwest Europe, London, pp. 123 133. Neumann, G., Pierson Jr., W.J., 1966. Principles of Physical Oceanography. Prentice-Hall, Englewood Cliffs, NJ, p. 545. Neumann, A.C., Ball, M.M., 1970. Submersible observations in the Straits of Florida: geology and bottom currents. Geol. Soc. Am. Bull. 81, 2861 2874. Neumeier, U., 1998. Tidal dunes and sand waves in deep outer-shelf environments, Bajocian, SE Jura, France. J. Sediment. Res. 68, 507 514. Newton, C.S., Shipp, R.C., Mosher, D.C., Wach, G.D., 2004. Importance of mass transport complexes in the Quaternary development of the Nile Fan, Egypt. In: Offshore Technology Conference, 3 6 May 2004, Houston, TX. ,http://www.onepetro.org/mslib/servlet/onepetropreview?id 5 OTC-16742-MS&soc 5 OTC. (accessed 11.06.10.). NGDC (National Geophysical Data Center), 2006. Tsunami data at NGDC. ,http://www.ngdc.noaa.gov/seg/ hazard/tsu.shtml. (accessed 10.11.07.). NGDC (National Geophysical Data Center) Tsunami Event Database, http://www.ngdc.noaa.gov/nndc/struts/results? bt_0=&st_0=&type_8=EXACT&query_8=None 1 Selected&op_14=eq&v_14=&type_15=EXACT&query_15=None 1 Selected&type_7=Like&query_7=&st_1=&bt_2=&st_2=&bt_1=&bt_10=&st_10=&ge_9=&le_9=&bt_3=&st_3=&type_19=EXACT&query_19=83&op_17=eq&v_17=&type_18=EXACT&query_18=None 1 Selected&bt_20=&st_20=&bt_13=&st_13=&bt_16=&st_16=&bt_6=&st_6=&bt_11=&st_11=&d=7&t=101650&s=7. (accessed 10.11.07.). Ni, L.T., Zhong, J.H., Shao, Z.F., Li, Y., Mao, C., Liu, S., 2015. Characteristics, genesis, and sedimentary environm of duplex-like structures in the Jurassic sediments of Western Qaidam Basin, China. J. Earth Sci. 26, 677 689. Nichols, R.J., 1995. The liquefaction and remobilization of sandy sediments. In: Hartley, A.J., Prosser, D.J. (Eds.), Characterisation of Deep Marine Clastic Systems, 94. Geological Society of London Special Publication, pp. 63 76. Nielsen, T., Knutz, P.C., Kuijpers, A., 2008. Seismic expression of contourite depositional systems. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 301 321. Developments in Sedimentology. Nilsen, T.H., Abbott, P.L., 1979. Introduction. In: Nilsen, T.H., Arthur, M.A. (Eds.), Upper Cretaceous Deep-Sea Fan Deposits, San Diego, 11. Geological Society of America Annual Meeting, San Diego, CA, pp. 137 166. Fieldtrip.

Bibliography

519

Nio, S.-D., Yang, C.-S., 1991. Diagnostic attributes of clastic tidal deposits: a review. In: Smith, D.G., Zaitlin, B.A., Reinson, G.E., Rahmani, R.A. (Eds.), Clastic Tidal Sedimentology. Canadian Society of Petroleum Geologists, Calgary, pp. 3 27. Nixon, L.D., Shepard, N.K., Bohannon, C.M., Montgomery, T.M., Kazanis, E.G., Gravois, M.P., 2009. Deepwater Gulf of Mexico 2009: Interim Report of 2008 Highlights. OCS Report MMS 2009-016: U.S. Department of the Interior, Minerals Management Service (MMS), New Orleans, LA, Gulf of Mexico OCS Region, 87 pp. NOAA (National Oceanic and Atmospheric Administration), 2005. NOAA News Online, Story 2365. ,http:// www.noaanews.noaa.gov/stories2005/s2365.htm. (accessed 15.06.05.). NOAA (National Oceanic and Atmospheric Administration) Fisheries Glossary, 2006. River plume. NOAA Technical Memorandum NMFS-F/SPO-69. Revised Edition, U.S. Department of Commerce, 71 p. NOAA-USGS Debris Flow Task Force, 2005. NOAA-USGS debris-flow warning system—final report. U.S. Geological Survey Circular 1283, 47 p. Nodine, M.C., Wright, S.G., Gilbert, R.B., Ward, E.G., 2006. Mudflows and mudslides during Hurricane Ivan. In: Offshore Technology Conference, Houston, TX, 1 4 May 2006., OTC Paper # 18328. Nogueira, A.C.R., Riccomini, C., Sial, A.N., Moura, C.A.V., Fairchild, T.R., 2003. Soft-sediment deformation at the base of the Neoproterozoic Puga cap carbonate (southwestern Amazon Craton, Brazil): conbrmation of rapid icehouse-greenhouse transition in snowball Earth. Geology 31, 613 616. Norem, H., Locat, J., Schieldrop, B., 1990. An approach to the physics and the modeling of submarine flowslides. Mar. Geotechnol. 9, 93 111. Normark, W.R., 1989. Observed parameters for turbidity-current flow in channels, Reserve Fan, Lake Superior. J. Sediment. Petrol. 59, 423 431. Normark, W.R., 1991. Turbidite elements and the obsolescence of the suprafan concept. G. di Geologia 53 (2), 1 10. ser 3a. Normark, W.R., Carlson, P.R., 2003. Giant submarine canyons: is size any clue to their importance in the rock record? In: Chan, M.A., Archer, A.W. (Eds.), Extreme Depositional Environments: Mega End Members in Geologic Time:. Geological Society of America, Boulder, CO, pp. 175 190. Special Paper 370. Normark, W.R., Reid, J.A., 2003. Extensive deposits on the Pacific Plate from Late Pleistocene North American Glacial Lake Outbursts. J. Geol. 111, 617 637. Normark, W.R., Hess, G.R., Stow, D.A.V., Bowen, A.J., 1980. Sediment waves on the Monterey fan levee: a preliminary physical interpretation. Mar. Geol. 37, 1 18. Normark, W.R., Gutmacher, C.E., Chase, T.E., Wilde, P., 1985. Monterey Fan, Pacific Ocean. In: Bouma, A.H., Normark, W.R., Barnes, N.E. (Eds.), Submarine Fans and Related Turbidite Systems. Springer-Verlag, New York, pp. 79 86. Normark, W.R., Moore, J.G., Torresan, M.E., 1993. Giant volcano-related landslides and the development of the Hawaiian Islands. In: Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone, 2002. U.S. Geological Survey Bulletin, pp. 184 196. Normark, W.R., Damuth, J.E., The Leg 155 Sedimentology Group, 1997. Sedimentary facies and associated depositional elements of the Amazon Fan. In: Flood, R.D., Piper, D.J.W., Klaus, A., Peterson, J.C. (Eds.), Proceedings of the Ocean Drilling Program Scientific Results, pp. 611 651. Normark, W.R., McGann, M., Sliter, R., 2004. Age of Palos Verdes submarine debris avalanche, southern California. Mar. Geol. 203, 247 259. Norris, R.D., Firth, J.V., 2002. Mass wasting of Atlantic continental margins following the Chicxulub impact event. In: Koeberl, C., MacLeod, K.G. (Eds.), Catastrophic Events and Mass Extinctions: Impacts and Beyond. Geological Society of America Special Paper 356, pp. 79 95. Nowlin Jr., W.D., 1972. Winter circulation patterns and property distributions. In: Capurro, L.R.A., Reid, J.L. (Eds.), Contributions on the Physical Oceanography of the Gulf of Mexico. Texas A&M University Oceanographic Studies 2. Gulf Publishing Co, Houston, TX, pp. 3 51. Nowlin Jr., W.D., Hubert, J.M., 1972. Contrasting summer circulation patterns for the eastern Gulf. In: Capurro, L. R.A., Reid, J.L. (Eds.), Contributions on the Physical Oceanography of the Gulf of Mexico. Texas A&M University Oceanographic Studies 2. Gulf Publishing Co, Houston, TX, pp. 119 137. Oakeshott, J.M. D’Ivry, 1989. Aspects of Depositional Mechanisms of High Concentration Sediment Gravity Flows (Unpublished Ph.D. thesis). University of Keele, Keele, p. 502.

520

Bibliography

Oard, M., Hergenrather, J., Klevberg, P., 2006. Flood transported quartzites: Part 3—failure of uniformitarian interpretations. J. Creation 20, 78 86. Obermeier, S.F., 1989. The New Madrid Earthquakes: an engineering-geologic interpretation of relict liquefaction features. U.S. Geological Survey Professional Paper, 1336-B. Obermeier, S.F., 1996. Use of liquefaction-induced features for paleoseismic analysis—an overview of how seismic liquefaction features can be distinguished from other features and how their regional distribution and properties of source sediment can be used to infer the location and strength of Holocene paleoearthquakes. Eng. Geol. 44, 1 76. Obermeier, S.F., 1998. seismic liquefaction features: examples from paleoseismic investigations in the continental United States. U.S. Geological Survey, Open-File Report 98-488. ,http://pubs.usgs.gov/of/1998/of98-488/. (accessed 31.05.10.). Obermeier, S.F., Pond, E.C., Olson, S.M., Green, R.A., 2002. Paleoliquefaction studies in continental settings. In: Ettensohn, F.R., Rast, N., Brett, C.E. (Eds.), Ancient Seismites. GSA Special Papers, 359, 13 27. Obermeier, S.F., Olson, S.M., Russell, A.G., 2005. Field occurrences of liquefaction-induced features: a primer for engineering geologic analysis of paleoseismic shaking. Eng. Geol. 76 (2005), 209 234. Occhietti, S., 1973. Les structures et deformations engendrres par les glaciers essal de mise au point. I. Drformations et structures glacitectoniques. Rev. Grogr. Montr. 27, 365 380. Ocean Motion (NASA), 2011. Wind-driven surface currents: gyres. Adapted from DataStreme Ocean and used with permission of the American Meteorological Society. ,http://oceanmotion.org/html/background/winddriven-surface.htm. (accessed 20.03.11.). Odonne, F., Callot, P., Debroas, E., Sempere, T., Hoareau, G., Maillard, A., 2011. Soft-sediment deformation from submarine sliding: favourable conditions and triggering mechanisms in examples from the Eocene Sobrarbe delta (Ainsa, Spanish Pyrenees) and the mid-Cretaceous Ayabacas Formation (Andes of Peru). Sediment. Geol. 235, 234 248. Ogata, K., Festa, A., Pini, G.A. (Eds.), 2019. Submarine Landslides: Subaqueous Mass Transport Deposits From Outcrops to Seismic Profiles. American Geophysical Union. Wiley, 384 P. Ogston, A.S., Cacchione, D.A., Sternberg, R.W., Kineke, G.C., 2000. Observations of storm and river flood-driven sediment transport on the northern California continental shelf. Cont. Shelf Res. 20, 2141 2162. Olabode, S., 2014. Soft sediment deformation structures in the Maastrichtian Ajali Formation, Western Flank of Anambra Basin, Southern Nigeria. J. Afr. Earth Sci. 89, 16 30. O’Leary, D.W., 1993. Submarine mass movement, a formative process of passive continental margins: the Munson-Nygren landslide complex and the southeast New England complex. In: Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone, 2002. U.S. Geological Survey Bulletin, pp. 23 39. Oliveira, C.M.M., Hodgson, D.M., Flint, S.S., 2011. Distribution of soft-sediment deformation structures in clinoform successions of the Permian Ecca Group, Karoo Basin, South Africa. Sediment. Geol. 235, 314 330. Olson, H.C., Damuth, J.E., 2010. Character, distribution and timing of latest quaternary mass-transport deposits in Texas Louisiana intraslope basins based on high-resolution (3.5 kHz) seismic facies and piston cores. In: Mosher, D.C., et al., (Eds.), Submarine Mass Movements and Their Consequences, 28. Advances in Natural and Technological Hazards Research, pp. 607 617. Onorato, M.R., Perucca, L., Coronato, A., Rabassa, J., Lo´pez, R., 2016. Seismically-induced soft-sediment deformation structures associated with the Magallanes-Fagnano Fault System (Isla Grande de Tierra del Fuego, Argentina). Sediment. Geol. 344, 135 144. Opreanu, G., 2003 2004. Porosity density and other physical properties of deep-sea sediments from the Black Sea. GEO-ECO-MARINA 9-10/2003-2004. Ortner, H., Kilian, S., 2016. Sediment creep on slopes in pelagic limestones: Upper Jurassic of Northern Calcareous Alps, Austria. Sediment. Geol. 344, 350 363. Oxford Dictionaries, 2018. Estuary. ,https://en.oxforddictionaries.com/definition/estuary.. Ozcelik, M., 2016. Evaluation of soft sediment deformation structures along the Fethiye Burdur Fault Zone, SW Turkey. J. Earth Syst. Sci. 125, 343 358. Palanques, A., Durieu de Madron, X., Puig, P., Fabres, J., Guille´n, J., Calafat, A.M., et al., 2006. Suspended sediment fluxes and transport processes in the Gulf of Lions submarine canyons. The role of storms and dense water cascading. Mar. Geol. 234, 43 61.

Bibliography

521

Palladino, G., Prosser, G., Bentivenga, M., Alsop, G.I., 2019. Mass transport deposits overprinted by contractional tectonics: a case study from the southern Apennines of Italy. Geol. Mag. 156 (5), 849 873. Pan, S., Liu, H., Zavala, C., Liu, C., Liang, S., Zhang, Q., et al., 2017. Sublacustrine hyperpycnal channel-fan system in a large depression basin: a case study of Nen 1 Member, Cretaceous Nenjiang Formation in the Songliao Basin, NE China. Petrol. Explor. Dev. 44 (6), 911 922. Pandey, D.K., Clift, P.D., Kulhanek, D.K., The Expedition 355 Scientists, 2016a. Expedition 355 summary. In: Proceedings of the International Ocean Discovery Program, vol. 355. ,https://doi.org/10.14379/iodp.proc.355.101.2016.. Pandey, D.K., Clift, P.D., Kulhanek, D.K., The Expedition 355 Scientists, 2016b. Site U1457. In: Proceedings of the International Ocean Discovery Program, vol. 355. ,https://doi.org/10.14379/iodp.proc.355.104.2016.. Parchure, T.M., 1980. Effect of Bed Shear Stress on the Erosional Characteristics of Kaolinite (M.S. thesis). University of Florida, Gainesville, FL. Parker, G., Toniolo, H., 2007. Note on the analysis of plunging of density flows. J. Hydraulic Eng. 133, 690 694. Parker, G., Fukushima, Y., Pantin, H.M., 1986. Self accelerating turbidity currents. J. Fluid Mech. 171, 145 181. Parsons, J.D., Schweller, W.J., Stelting, C.W., Southard, J.B., Lyons, W.J., Grotzinger, J.P., 2003. A preliminary experimental study of turbidite fan deposits—reply. J. Sediment. Res. 73, 839 841. Parsons, J.D., Whipple, K.X., Simoni, A., 2001a. Experimental study of the grain-flow, fluid-mud transition in debris flows. J. Geol. 109, 427 447. Parsons, J.D., Bush, J.W.M., Syvitski, J.P.M., 2001b. Hyperpycnal plume formation from riverine outfl ows with small sediment concentrations. Sedimentology 48, 465 478. Parsons, J.D., Friedrichs, C.T., Trykovski, P.A., Mohrig, D., Imran, J., Syvitski, J.P.M., et al., 2007. The mechanics of marine sediment gravity flows. In: Nittrouer, C. (Ed.), Continental Margin Sedimentation: Transport to Sequence, 37. Blackwell Publishing Ltd, Oxford, pp. 275 338. IAS Special Publication. Parthasarathy, A., Nagarajan, R., Panchapakesan, V., 2013. Engineering Geology. Wiley India, New Delhi, 532 p. Passchier, S., 2000. Soft-sediment deformation features in core from CRP-2/2A, Victoria Land Basin, Antarctica. Terra Antarctica 7 (3), 401 412. Pati, J.K., Reimold, W.U., Koeberl, C., Pati, P., 2008. The Dhala structure, Bundelkhand craton, Central India— eroded remnant of a large Paleoproterozoic impact structure. Meteorit. Planet Sci. 43 (8), 1383 1398. Paull, C.K., Buelow, W.J., Ussler III, W., Borowski, W.S., 1996. Increased continental-margin slumping frequency during sea-level lowstands above gas hydrate bearing sediments. Geology 24, 143 146. Paull, C.K., Matsumoto, R., Wallace, P.J., Dillon, W.P., 2000. Scientific results. Proceedings of Ocean Drilling Program, College Station, TX, vol. 164, p. 459. Paull, C.K., Ussler III, W., Greene, H.G., Keaten, R., Mitts, P., Barry, J., 2003. Caught in the act: the 20 December 2001 gravity flow event in Monterey Canyon. Geo-Mar. Lett. 22, 227 232. Paull, C.K., Mitts, P., Ussler III, W., Keaten, R., Greene, H.G., 2005. Trail of sand in upper Monterey Canyon: offshore California. Geol. Soc. Am. Bull. 117, 1134 1145. Paull, C.K., Talling, P.J., Maier, K.L., Parsons, D., Xu, J., Caress, D.W., et al., 2018. Powerful turbidity currents driven by dense basal layers. Nat. Commun. 9 (1), 4114. Available from: https://doi.org/10.1038/s41467-018-06254-6. Paull, C.K., Ussler, I.I.I.W., Borowski, W.S., Spiess, F.N., 1995. Methane-rich plumes on the Carolina continental rise: associations with gas hydrates. Geology 23 (1), 89 92. Pavec, M., Carton, X., Swaters, G., 2005. Baroclinic instability of frontal geostrophic currents over a slope. J. Phys. Oceanogr. 35, 911 918. Payenberg, T.H.D., Lang, S.C., Deller, K., Ceglar, N., Reilly, M.R.W., Cotton, T., et al., 2005. Reducing critical uncertainties for deepwater stratigraphic architecture in the Officer Basin - the significance of outcrop analogues in the Flinders Ranges. Unpublished Report ASP-RARG-2005-01, Reservoir Analogues Research Group, Australian School of Petroleum, The University of Adelaide, March 2005, 82 p. Payton, C. (Ed.), 1977. Seismic stratigraphy: applications to hydrocarbon exploration. AAPG Memoir 26, 516 p. PBS (Public Broadcasting Service), 2014. Nova “Killer landslides”, a TV documentary, broadcast on November 19, 2014 in the U.S. ,http://www.pbs.org/wgbh/nova/earth/killer-landslide.html. (accessed 27.12.14.). Peakall, J., Amos, K.J., Keevil, G.M., Bradbury, P.W., Gupta, S., 2007. Flow processes and sedimentation in submarine channel bends. Mar. Petrol. Geol. 24, 470 486. Peliz, A., Marchchesiello, P., Santos, A.M.P., Dubert, J., Teles-Machado, A., Marta-Almeida, M., 2009. Surface circulation in the Gulf of Cadiz: 2. Inflow outflow coupling and the Gulf of Cadiz slope current. J. Geophys. Res. Ocean. 114, C03011.

522

Bibliography

Pequegnat, W.E., 1972. A deep bottom-current on the Mississippi Cone. In: Capurro, L.R.A., Reid, J.L. (Eds.), Contribution on the Physical Oceanography of the Gulf of Mexico. Texas A&M University Oceanographic Studies, 2. Gulf Publishing, Houston, TX, pp. 65 87. Perch-Nielsen, K., Supko, P.R., Boersma, A., Carlson, R.L., Dinkelman, M.G., Fodor, R.V., et al. (Eds.), Initial Reports of the Deep Sea Drilling Project, Leg 39. United States Government Printing Office, Washington, DC, pp. 231 327, ,https://doi.org/10.2973/dsdp.proc.39.106.1977.. Pe´rez, L.F., Herna´ndez-Molina, F.J., Esteban, F.D., Tassone, A., Piola, A.R., Maldonado, A., et al., 2015. Erosional and depositional contourite features at the transition between the western Scotia Sea and southern South Atlantic Ocean: links with regional water-mass circulation since the Middle Miocene. Geo-Mar. Lett. 35, 271 288. Peterson, G.L., 1966. Structural interpretation of Sandstone dikes, Northwest Sacramento Valley, California. GSA Bull. 77, 833 842. Petley, D.N., 2010. An analysis of fatal landslides in the Asia-Pacific region for 2006 to 2008. ,http://daveslandslideblog.blogspot.com/2010/02/analysis-of-fatal-landslides-in-asia.html. (accessed 19.11.10.). Petley, D., 2012. Global patterns of loss of life from landslides. Geology 40 (10), 927 930. Available from: https:// doi.org/10.1130/G33217.1. Petruncio, E.T., Rosenfeld, L.K., Paduan, J.D., 1998. Observations of the internal tide in Monterey Canyon. J. Phys. Oceanogr. 28, 1873 1903. Petter, A.L., Steel, R.J., 2006. Hyperpycnal flow variability and slope organization on an Eocene shelf margin, central basin, Spitsbergen. AAPG Bull. 90, 1451 1472. Petters, S.W., 1984. An ancient submarine canyon in the Oligocene-Miocene of the western Niger Delta. Sedimentology 31, 805 810. Pettijohn, F.J., 1957. Sedimentary Rocks, second ed. Harper, New York, p. 718. Pettijohn, F.J., 1975. Sedimentary Rocks, third ed. Harper & Row, Publishers, New York, 628 p. Pettijohn, F.J., Potter, P.E., 1964. Atlas and Glossary of Primary Sedimentary Structures. Springer, Berlin, p. 770. Phillips, C.J., Davies, T.R.H., 1991. Determining rheological parameters of debris flow material. Geomorphology 4, 101 110. Pickering, K.T., Hilton, V., 1998. Turbidite Systems of Southeast France. Vallis Press, London, p. 229. Pickering, K.T., Hiscott, R.N., 2015. Deep Marine Systems: Processes, Deposits, Environments, Tectonics and Sedimentation. American Geophysical Union. Wiley, p. 696. Pickering, K.T., Hiscott, R.N., Hein, F.J., 1989. Deep-Marine Environments. Unwin Hyman, London, p. 416. Pickering, K.T., Hiscott, R.N., Kenyon, N.H., Ricci Lucchi, F., Smith, R.D.A. (Eds.), 1995. Atlas of Deep Water Environments: Architectural Style in Turbidite Systems. Chapman and Hall, London, 333 p. Pickering, K.T., Hodgson, D.M., Platzman, E., Clark, J.D., Stephens, C., 2001. A new type of bedform produced by backfilling processes in a submarine channel, late Miocene, Tabernas-Sorbas basin, SE Spain. J. Sediment. Res. 71, 692 704. Pierce, L.E.R., 2012. Poverty Shelf, New Zealand From the Holocene to Present: Stratigraphic Development and Event Layer Preservation in Response to Sediment Supply, Tectonics and Climate (Unpublished Ph.D. dissertation). The College of William & Mary, Williamsburg, VA, 274 p. Pierson, T., 1981. Dominant particle support mechanisms in debris flows at Mt Thomas, New Zealand, and implications for flow mobility. Sedimentology 28 (1), 49 60. Pierson, T., 1985. Initiation and flow behavior of the 1980 Pine Creek and Muddy River lahars, Mount St. Helens, Washington. Geol. Soc. Am. Bull. 96, 1056 1069. Pierson, T., 1990. Perturbation and melting of snow and ice by the 13 November 1985 eruption of Nevado del Ruiz, Colombia, and consequent mobilization, flow and deposition of lahars. J. Volcanol. Geotherm. Res. 41 (1 4), 17 66. Pierson, T.C., Costa, J.E., 1987. Arheologic classification of subaerial sediment-water flows. In: Costa, J.E., Wieczorek, G.F. (Eds.), Debris Flows/Avalanches: Process, Recognition, and Mitigation, VII. Geological Society of America Reviews in Engineering Geology, pp. 1 12. Pinardi, N., Masetti, E., 2000. Variability of the large scale general circulation of the Mediterranean Sea from observations and modelling: a review. Palaeogeogr. Palaeoclimatol. Palaeoecol. 158, 153 173. Available from: https://doi.org/10.1016/S0031-0182(00)00048-1. Pinheiro, L.M., Ivanov, M.K., Sautkin, A., Akhmanov, G., Magalhaes, V.H., Volkonskaya, A., et al., 2003. Mud volcanism in the Gulf of Cadiz: results from the TTR-10 cruise. Mar. Geol. 195, 131 151.

Bibliography

523

Piper, D.J.W., 1970. A Silurian deep sea fan deposit in western Ireland and its bearing on the nature of turbidity current. J. Geol. 78, 509 522. Piper, D.J.W., 1975. Appendix II. Deformation of stiff and semilithified cores from Legs 18 and 28. Deep Sea Drilling Project 28, 977 979. ,https://doi.org/10.2973/dsdp.proc.28.app2.1975.. Piper, D.J.W., 1978. Turbidite muds and silts in deep-sea fans and abyssal plains. In: Stanley, D.J., Kelling, G. (Eds.), Sedimentation in Submarine Fans, Canyons, and Trenches. Hutchinson and Ross, Stroudsburg, PA, pp. 163 176. Piper, D.J.W., Brisco, C.D., 1975. Deep-water continental-margin sedimentation, DSDP Leg 28, Antarctica. In: Hayes, D.E., et al., (Eds.), Initial Reports of the Deep Sea Drilling Project. U.S. Government Printing Office, Washington, DC, pp. 727 755. Piper, D.J.W., Aksu, A.E., 1987. The source and origin of the 1929 Grand Banks turbidity current inferred from sediment budgets. Geo-Mar. Lett. 7, 177 182. Piper, D.J.W., Shor, A.N., Hughes Clarke, J.E., 1988. The 1929 ‘Grand Banks’ earthquake, slump, and turbidity current. In: Clifton, H.E. (Ed.), Sedimentologic Consequences of Convulsive Geologic Events. Geological Society of America Special Paper 229, pp. 77 92. Piper, D.J.W., Mosher, D.C., Campbell, D.C., 2012. Controls on the distribution of major types of submarine landslides. In: Clague, J.J., Stead, D. (Eds.), Landslides: Types, Mechanismsand Modeling. Cambridge University Press, Cambridge, UK, pp. 95 107. Piper, D.J.W., Pirmez, C., Manley, P.L., Long, D., Flood, R.D., Normark, W.R., et al., 1997. Mass-transport deposits of the Amazon Fan. In: Flood, R.D., Piper, D.J.W., Klaus, A., Peterson, L.C. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results 155, pp. 109 143. Piper, D.J.W., Cochonat, P., Morrison, M.L., 1999. The sequence of events around the epicenter of the 1929 Grand Banks earthquake: initiation of debris flows and turbidity current inferred from sidescan sonar. Sedimentology 46, 79 97. Pirmez, C., Imran, J., 2003. Reconstruction of turbidity currents in Amazon Channel. Mar. Petrol. Geol 20, 823 849. Plaziat, J.C., Purser, B.H., Philobbos, E., 1990. Seismic deformation structures (seismites) in the syn-rift sediments of the NW Red Sea (Egypt). Bull. de. la. Socie´te´ ge´ologique de Fr. 6 (3), 419 434. Plink-Bjo¨rklund, P., Steel, R., 2004. Initiation of turbidity currents: outcrop evidence for Eocene hyperpycnal flow turbidites. Sediment. Geol. 165, 29 52. Plint, A.G., Nummedal, D., 2000. The falling stage systems tract: recognition and importance in sequence stratigraphic analysis. In: Hunt, D., Gawthorpe, R.L. (Eds.), Sedimentary Response to Forced Regression, vol. 172. Geological Society of London Special Publication, pp. 1 17. Pluenneke, J.L., 1976. Comparative Analysis of Debrites, Turbidites, and Contourites (Unpublished M.S. thesis). Texas A&M University, College Station, TX, p. 93. Pochat, S., Van den Driessc, J., 2007. Impact of synsedimentary metre-scale normal fault scarps on sediment gravity flow dynamics: an example from the Gre`s d’Annot Formation, SE France. Sediment. Geol. 202, 796 820. Pomar, L., Morsilli, M., Hallock, P., Ba´denas, B., 2012. Internal waves, an under-explored source of turbulence events in the sedimentary record. Earth-Sci. Rev. 111, 56 81. Popenoe, P., 1993. Summary geologic report for the South Atlantic Outer Continental Shelf (OCS) Planning Area. U.S. Geological Survey Open-File Report 84 476, 27 p. Popenoe, P., Schmuck, E.A., Dillon, W.P., 1993. The cape fear landslide; slope failure associatedwith salt diapirism and gas hydrate decomposition. In: Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone, 2002. U.S. Geological Survey Bulletin, pp. 40 53. Posamentier, H.W., Erskine, R.D., 1991. Seismic expression and recognition criteria of ancient submarine fans. In: Weimer, P., Link, M.H. (Eds.), Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems. Springer-Verlag, New York, pp. 197 222. Posamentier, H.W., Allen, G.P., 1999. Siliciclastic sequence stratigraphy: concepts and applications, SEPM Concepts in Sedimentology and Paleontology no. 7, 210 p. Posamentier, H.W., Kolla, V., 2003. Seismic geomorphology and stratigraphy of depositional elements in deepwater settings. J. Sediment. Res. 73, 367 388. Posamentier, H.W., Allen, G.P., James, D.P., Tesson, M., 1992. Forced regressions in a sequence-stratigraphic framework: concepts, examples, and exploration significance. AAPG Bull. 76, 1687 1709.

524

Bibliography

Postma, G., 1986. Classification of sediment gravity-flow deposits based on flow conditions during sedimentation. Geology 14, 291 294. Postma, G., Cartigny, M.J.B., 2014. Supercritical and subcritical turbidity currents and their deposits a synthesis. Geology 42, 987 990. Postma, G., Nemec, W., Kleinspehn, K.L., 1988. Large floating clasts in turbidites: a mechanism for their emplacement. Sediment. Geol. 58, 47 61. Postma, G., Kleverlaan, K., Cartigny, M.J.B., 2014. Recognition of cyclic steps in sandy and gravelly turbidite sequences, and consequences for the Bouma facies model. Sedimentology . Available from: https://doi.org/ 10.1111/sed.12135. Potter, P.E., Pettijohn, F.J., 1977. Paleocurrents and Basin Analysis. Springer, 425 p. Powell, R.D., Laird, M.G., Naish, T.R., Fieldling, C.R., Krissek, L.A., Van Der Meer, J.J.M., 2001. Depositional environments for strata cored in CRP-3 (Cape Roberts Project), Victoria Land Basin, Antarctica: palaeoglaciological and palaeoclimatological inferences. Terra Antarctica 8, 207 216. Prandtl, L., 1925. Bericht fiber Untersuchungen zur ausgebildeten Turbulenz. ZAMM 5, 136 139. Prandtl, L., 1926. Bericht fiber neuere Turbulenzforschung. Hydraulische Probleme. VDI-Verlag, Berlin, pp. 1 13. Prater, M.D., Sanford, T.B., 1994. A meddy off Cape St. Vincent. Part 1: description. J. Phys. Oceanogr. 24 (7), 1572 1586. Price, J.F., Baringer, M.O., Lueck, R.G., Johnson, G.C., Ambar, I., Parrilla, G., et al., 1993. Mediterranean outflow mixing dynamics. Science 259, 1277 1282. Prins, M.A., Bouwer, L.M., Beets, C.J., Troelstra, S.R., Weltje, G.J., Kruk, R.W., et al., 2002. Ocean circulation and iceberg discharge in the glacial North Atlantic: inferences from unmixing of sediment size distributions. Geology 30, 555 558. Prave, A.R., Duke, W.L., 1990. Small-scale hummocky cross-stratification in turbidites; a form of antidune stratification. Sedimentology 37, 531 539. Prior, D.B., Coleman, J.M., 1979. Submarine landslides—geometry and nomenclature. Z. Geomorphol. N. F. 23, 415 426. Prior, D.B., Coleman, J.M., 1984. Submarine slope instability. In: Brunsden, D., Prior, D.B. (Eds.), Slope Instability. John Wiley & Sons Ltd, Chichester, pp. 419 455. Prior, D.B., Hooper, J.R., 1999. Sea floor engineering geomorphology: recent achievements and future directions. Geomorphology 31, 411 439. Prior, D.B., Bornhold, B.D., 1990. The underwater development of Holocene fan deltas. In: Colella, A., Prior, D.B. (Eds.), Coarse-Grained Deltas. Blackwell Scientific, Oxford, International Association of Sedimentologists Special Publication 10: 7590. Prior, D.B., Wiseman, W.J., Bryant, W.R., 1981. Submarine chutes on the slopes of fjord deltas. Nature 290, 326 328. Prior, D.B., Bornhold, B.D., Coleman, J.M., Bryant, W.R., 1982. The morphology of a submarine slide, Kitimat Arm, British Columbia. Geology 10, 588 592. Prior, D.B., Bornhold, B.D., Johns, M.W., 1984. Depositional characteristics of a submarine debris flow. J. Geol. 92, 707 727. Prior, D.B., Bornhold, B.D., Wiseman Jr., W.J., Lowe, D.R., 1987. Turbidity current in activity in a British Columbia fjord. Science 237, 1330 1333. Prior, D.B., Suhayda, J.N., Lu, N.-Z., Bornhold, B.D., Keller, G.H., Wiseman, W.J., et al., 1989. Storm wave reactivation of a submarine landslide. Nature 341, 47 50. Prytkov, A.S., Vasilenko, N.F., Ershov, V.V., Russ, J., 2014. Simulation of the 2011 South Sakhalin mud volcano eruption based on the GPS data. Russian J. Pac. Geol. 8, 224 231. Pudsey, C.J., Howe, J.A., 1998. Quaternary history of the Antarctic Circumpolar Current: evidence from the Scotia Sea. Mar. Geol. 148, 83 112. Puig, P., Ogston, A.S., Mullenbach, B.L., Nittrouer, C.A., Sternberg, R.W., 2003. Shelf-to-canyon sedimenttransport processes on the Eel continental margin (northern California). Mar. Geol. 193, 129 149. Puig, P., Palanques, A., Martı´n, J., 2014. Contemporary sediment-transport processes in submarine canyons. Annu. Rev. Mar. Sci. 6, 53 77. Purdy, E.G., 1963. Recent calcium carbonate facies of the Great Bahama Bank: 2. Sedimentary facies, J. Geol., 71. pp. 472 497.

Bibliography

525

Purkey, S.G., Smethie, W.M., Gebbie, G., Gordon, A.L., Sonnerup, R.E., Warner, M.J., et al., 2018. A synoptic view of the ventilation and circulation of Antarctic bottom water from chlorofluorocarbons. Annu. Rev. Mar. Sci. 10, 503 527. Purkis, S., Cavalcante, G., Rohtla, L., Oehlert, A.M., Harris, P.M., Swart, P.K., 2017. Hydrodynamic control of whitings on Great Bahama Bank. Geology 45 (10), 939 942. Qiao, L.L., Bao, X.W., Wu, D.X., Wang, X.H., 2008. Numerical study of generation of the tidal shear front off the Yellow River mouth. Cont. Shelf Res. 28 (14), 1782 1790. Quaresma, L.S., Pichon, A., 2013. Modelling the barotropic tide along the West-Iberian margin. J. Mar. Syst. 109 110, S3 S25. Ouyang, C., Zhao, W., An, H., Zhou, S., Wang, D., Xu, Q., et al., 2019. Early identification and dynamic processes of ridge-top rockslides: implications from the Su Village landslide in Suichang County, Zhejiang Province, China. Landslides 16, 799 813. Available from: https://doi.org/10.1007/s10346-018-01128-w. Rahmstorf, S., 2002. Ocean circulation and climate during the past 120,000 years. Nature 419, 207 214. Rahmstorf, S., 2006. Thermohaline ocean circulation. In: Elias, S.A. (Ed.), Encyclopedia of Quaternary Sciences. Elsevier, Amsterdam, pp. 1 10. Ramprasad, T., Dewangan, P., Ramana, M.V., Mazumdar, A., Karisiddaiah, S.M., Ramya, E.R., et al., 2011. Evidence of slumping/sliding in Krishna-Godavari offshore basin due to gas/fluid movements. Mar. Petrol. Geol. 28, 1806 1816. Ramsay, J.G., 1967. Folding and Fracturing of Rocks. McGraw Hill, San Francisco, p. 568. Ramsay, J.G., Graham, R.H., 1970. Strain variation in shear belts. Can. J. Earth Sci. 7, 786 813. Rana, N., Prakash Sati, S., Sundriyal, Y., Juyal, N., 2016. Genesis and implication of soft sediment deformation structures in high energy fluvial deposits of the Alaknanda valley, Garhwal Himalaya, India. Sediment. Geol. 344, 263 276. Rapp, A., 1963. The debris slides of Ulvadal, western Norway: an example of catastrophic slope processes in Scandinavia. Nachrichten der Akademie der Wiss. Gottingen II. Mathematisch-Physikalische Kl. 13, 197 210. Rautman, C.A., Dott, R.H., 1977. Dish structures formed by fluid escape in Jurassic shallow marine sandstones. J. Sediment. Petrol. 47, 101 106. Rawson, M.D., Ryan, W.B.F., 1983, Geologic observations at the 2800-meter radioactive waste disposal site and associated deepwater dumpsite 106 (DWD-106) in the Atlantic Ocean: Report EPA 520/1-83-018, U.S. Environmental Protection Agency, Washington, DC, p. 61. Rebesco, M., Camerlenghi, A. (Eds.), 2008. Contourites, Developments in Sedimentology 60. Elsevier, Amsterdam, 663 p. Rebesco, M., Camerlenghi, A., Volpi, V., Neagu, C., Accettella, D., Lindberg, B., et al., 2007. Interaction of processes and importance of contourites: insights from the detailed morphology of sediment Drift 7, Antarctica. In: Viana, A.R., Rebesco, M. (Eds.), Economic and Palaeoceanographic Significance of Contourite Deposits, vol. 276. Geological Society of London Special Publication, pp. 95 110. Rebesco, M., Camerlenghi, A., Van Loon, A.J., 2008. Contourite research: a field in full development. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 3 10. Developments in Sedimentology. Rebesco, M., Herna´ndez-Molina, F.J., Van Rooij, D., Wa˚hlin, A., 2014. Contourites and associated sediments controlled by deep-water circulation processes: state-of-the-art and future considerations. Mar. Geol. 352, 111 154. Reiche, P., 1937. The Toreva-Block—a distinctive landslide type. J. Geol. 45, 538 548. Reid, J.L., 1969. Preliminary results of measurements of deep currents in the Pacific Ocean. Nature 221, 848. Reilly, M.R.W., Lang, S.C., 2003. A ponded basin floor outcrop analogue: bunkers sandstone, Northern flinders ranges, South Australia. APPEA J. 537 553. Reineck, H.E., Wunderlich, F., 1968. Classification and origin of flaser and lenticular bedding. Sedimentology 11, 99 104. Reineck, H.E., Singh, I.B., 1980. Depositional Sedimentary Environments, second ed. Springer-Verlag, Berlin, p. 549. Richardson, M.J., Wimbush, M., Mayer, L., 1981. Exceptionally strong near-bottom flows on the continental rise of Nova Scotia. Science 213, 887 888. Ridente, D., Foglini, F., Minisini, D., Trincardi, F., Verdicchio, G., 2007. Shelf-edge erosion, sediment failure and inception of Bari Canyon on the Southwestern Adriatic Margin (Central Mediterranean). Mar. Geol. 246, 193 207.

526

Bibliography

Ritchie, A.C., Warrick, J.A., East, A., et al., 2018. Morphodynamic evolution following sediment release from the world’s largest dam removal. Sci. Rep. 8, 13279. ,https://doi.org/10.1038/s41598-018-30817-8-1.. Ritter, D.F., Kochel, R.C., Miller, J.R., 1995. Process Geomorphology, third edition William C. Brown, Dubuque, IA, p. 546. Rivera-Rosario, G.A., Diamessis, P.J., Jenkins, J.T., 2017. Bed failure induced by internal solitary waves. J. Geophys. Res. Ocean. 7, 5468 5485. Rizzini, A., Passega, R., 1964. E´volution de la se´dimentation et orogene`se, Valle´e du Santerno, Apennin Septentrional. In: Bouma, A.H., Brouwer, A. (Eds.), Turbidites, Developments in Sedimentology, 3. Elsevier, Amsterdam, pp. 65 74. Roberts, N.J., Evans, S.G., 2009. Controls on size and occurrence of the largest sub-aerial landslide on Earth: Seymareh (Saidmarreh) landslide, Zagros fold-thrust belt, In: American Geophysical Union, Fall Meeting 2009, abstract #NH41C-1260. Robertson, R., 2010. Tidal currents and mixing at the INSTANT mooring locations. Dyn. Atmos. Ocean. 50, 331 373. Rodine, J.D., Johnson, A.M., 1976. The ability of debris, heavily freighted with coarse clastic material to flow on gentle slopes. Sedimentology 23, 213 234. Rodriguez, A.B., Anderson, J.B., 2004. Contourite origin for shelf and upper slope sand sheet, offshore Antarctica. Sedimetology 51, 699 711. Rodriguez, M., Chamot-Rooke, N., He´bert, H., Fournier, M., Huchon, P., 2013. Owen Ridge deep-water submarine landslides: implications for tsunami hazard along the Oman coast. Nat. Hazards Earth Syst. Sci. 13, 417 424. Available from: https://doi.org/10.5194/nhess-13-417-2013. Available from: www.nat-hazards-earth-syst-sci. net/13/417/2013/. Rodrı´guez-Pascua, M.A., Calvo, J.P., De Vicente, G., Gomez-Gras, D., 2000. Soft-sediment deformation structures interpreted as seismites in lacustrine sediments of the Prebetic Zone, SE Spain, and their potential use as indicators of earthquake magnitudes during the Late Miocene. Sediment. Geol. 135, 117 135. Rodrı´guez-Pascua, M.A., Silva, P.G., Perucha, M.A., Giner- Robles, J.L., Heras, C., Bastida, A.B., et al., 2016. Seismically induced liquefaction structures in La Magdalena archaeological site, the 4th century AD Roman Complutum (Madrid, Spain). Sediment. Geol. 344, 34 46. Rodrı´guez-Tovar, F.J., Herna´ndez-Molina, F.J., 2018a. Ichnological analysis of contourites: past, present and future. Earth-Sci. Rev. 182, 28 41. Rodrı´guez-Tovar, F.J., Herna´ndez-Molina, F.J., 2018b. Reply to comment on “Ichnological analysis of contourites: past, present and future” by Francisco J. Rodrı´guez-Tovar and F. Javier Hernandez-Molina [Earth-Science Reviews, 182 (2018), 28 41]. Earth-Sci. Rev. 184, 50 51. Rossetti, D.F., 1999. Soft-sediment deformation structures in late Albian to Cenomanian deposits, Sao Luis Basin, northern Brazil: evidence for palaeoseismicity. Sedimentology 46, 1065 1081. Rothwell, R.G.. 2004. Deep ocean pelagic oozes. In: Selley, R.C.; Cocks, R.M.; Plimer, I.R., (eds.) Encyclopedia of Geology. Oxford, Elsevier, 2750 p. Rothwell, R.G., Kenyon, N.H., McGregor, B.A., 1991. Sedimentary features of the south Texas continental slope as revealed by side-scan sonar and high-resolution seismic data. AAPG Bull. 75, 298 312. Rouse, C., 1984. Flowslides. In: Brunsden, D., Prior, D.B. (Eds.), Slope Instability. John Wiley & Sons Ltd, Chichester, pp. 491 522. Roveri, M., Bergamasco, A., Marcello Falcieri, F., Gennari, R., Lugli, S., Manzi, V., et al., 2013. The origin of Messinian canyons in the Mediterranean: the role of brine-related dense shelf water cascading currents. EGU General Assembly 2013, 7 12 April 2013, Vienna, id. EGU2013-4016. Rowe, G.T., 1971. Observations on bottom currents and epibenthonic populations in Hatteras submarine canyon. Deep-Sea Res. 18, 569 581. Rowe, G.T., Menzies, R.T., 1968. Deep bottom currents off the coast of North Carolina. Deep-Sea Res. 15, 711 720. Roy, S.K., Banerjee, S., 2016. Soft sediment deformation structures in the Andaman Flysch Group, Andaman Basin: evidence for Paleogene seismic activity in the Island Arc. Beri Sediment. 35, 55 64. Ruban, V., et al., 2010. Rogue waves towards a unifying concept?: discussions and debates. Eur. Phys. J. Spec. Top. 185, 5 15. Available from: https://doi.org/10.1140/epjst/e2010-01234-y. Rudersdorf, A., Hartmann, K., Yu, K., Stauch, G., Reicherter, K., 2015. Seismites as indicators for Holocene seismicity in the northeastern Ejina Basin, Inner Mongolia. In: Landgraf, A., Ku¨bler, S., Hintersberger, E.,

Bibliography

527

Stein, S. (Eds.), Fault Rupture and Earthquake Hazards in Slowly Deforming Regions, 432. Geological Society of London, Special Publication, pp. 213 231. Rudnick, D.L., Jan, S., Lee, C.M., 2015. A new look at circulation in the western North Pacific. Oceanography 28 (4), 16 23. ,https://doi.org/10.5670/oceanog.2015.77.. Ruppel, C.D., Kessler, J.D., 2017. The interaction of climate change and methane hydrates. Rev. Geophys. 55, 126 168. Sakaguchi, A., Kimura, G., Strasser, M., Screaton, E.J., Curewitz, D., Murayama, M., 2011. Episodic seafloor mud brecciation due to great subduction zone earthquakes. Geology 39, 919 922. Saller, A.H., Lin, R., Dunham, J., 2006. Leaves in turbidite sands: the main source of oil and gas in the deep-water Kutei Basin, Indonesia. AAPG Bull. 90, 1585 1608. Saller, A.H., Dunham, J., Lin, R., 2008a. Leaves in turbidite sands. The main source of oil and gas in the deepwater Kutei Basin, Indonesia, reply. AAPG Bull. 92, 139 141. Saller, A.H., Noah, J.T., Ruzuar, A.P., Schneider, R., 2004. Linked lowstand delta to basin-floor fan deposition, offshore Indonesia, an analog for deep-water reservoir systems. AAPG Bulletin 88, 21 46. Saller, A.H., Werner, K., Sugiaman, F., Cebastiant, A., May, R., Glenn, D., et al., 2008b. Characteristics of Pleistocene deep-water fan lobes and their application to an upper Miocene reservoir model, offshore East Kalimantan, Indonesia. AAPG Bull. 92, 919 949. Sammartini, M., Moernaut, J., Anselmetti, F.S., Hilbe, M., Lindhorst, K., Praet, N., et al., 2019. An atlas of masstransport deposits in lakes. In: Ogata, et al., (Eds.), Submarine Landslides: Subaqueous Mass Transport Deposits from Outcrops to Seismic Profiles. AGU Books, Wiley, pp. 201 226. Sanchez-Garrido, J.C., Sannino, G., Liberti, L., 2011. Generation and evolution of internal waves in the Strait of Gibraltar. EAI (Energia, Ambiente, Innov.) 4 5, 74 79. Sa´nchez-Roma´n, A., Criado-Aldeanueva, F., Garcı´a-Lafuente, J., Sa´nchez, J.C., 2008. Vertical structure of tidal currents over Espartel and Camarinal sills, Strait of Gibraltar. J. Mar. Syst. 74, 120 133. Sanders, J.E., 1963. Concepts of fluid mechanics provided by primary sedimentary structures. J. Sediment. Petrol. 33, 173 179. Sanders, J.E., 1965. Primary sedimentary structures formed by turbidity currents and related resedimentation mechanisms. In: Middleton, G.V. (Ed.), Primary Sedimentary Structures and Their Hydrodynamic Interpretation, 12. SEPM Special Publication, pp. 192 219. Sanders, J.E., Friedman, G.M., 1997. History of petroleum exploration in turbidites and related deepwater deposits. Northeastern Geol. Environ. Sci. 19 (1/2), 67 102. Santek, D.A., Winguth, A., 2005. A satellite view of internal waves induced by the Indian Ocean tsunami. Int. J. Remote. Sens. ,http://drake.aos.wisc.edu/_winguth/Papers/Tsunami_Internal_Waves.pdf.. Sapozhnikov, V.V., Ivanova, O.S., Mordasova, N.V., 2011. Identification of local upwelling zone in the Bering Sea usig hydrochemical parameters. Oceanology 51, 247 254. Sarkar, S., Bose, P.K., Eriksson, P.G., 2011. Neoproterozoic tsunamiite: Upper Bhander Sandstone, Central India. Sediment. Geol. 238, 181 190. S˘as˘aran, L., B˘alc, R., S˘as˘aran, E., 2005. Sedimentologic and micropalaeontological study of an Upper Cretaceous (Santonian-Campanian) regressive facies development -from basin to upper slope sediments (Gil˘au Mountains, Plescu¸ta Valley, NW-Romania). Acta Palaeontol. Roman. 5, 441 450. Sassa, K., 2009. Report of the 2008 First World Landslide Forum on 18 21 November 2008 at UNU, Tokyo. Landslides 6, 167 179. Available from: https://doi.org/10.1007/s10346-009-0161-8. Saxov, S., 1982. Marine slides some introductory remarks. In: Saxov, S., Nieuwenhuis, J.K. (Eds.), Marine Slides and Other Mass Movements. Plenum Press, New York and London, pp. 1 10. Saxov, S., Nieuwenhuis, J.K. (Eds.), 1982. Marine Slides and Other Mass Movements. Plenum Press, New York and London. Sayles, R.W., 1931. Bermuda during the Ice Age. Proc. Am. Acad. Arts Sci. 66, 382 467. Schaller, P.J., 1991. Analysis and Implications of Large Martian and Terrestrial Landslide (Ph.D. dissertation). California Institute of Technology, Pasadena, CA, p. 596. Scheidegger, A.E., 1973. On the prediction of the reach and velocity of catastrophic landslides. Rock. Mech. Rock Eng. 5, 231 236. Scheidegger, A.E., 1975. Physical Aspects of Natural Catastrophes. Elsevier Science Publishing, New York, p. 289. Schillereff, D.N., Chiverrell, R.C., Macdonald, N., Hooke, J.M., 2014. Flood stratigraphies in lake sediments: a review. Earth-Sci. Rev. 135, 17 37.

528

Bibliography

Schmidt, M., Dehn, M., 2003. Examining links between climate change and landslide activity using GCMs: case studies from Italy and New Zealand. In: McLaren, S.J., Kniveton, D.R. (Eds.), Advances in Global Change Research, Linking Climate Change to Land Surface Change, 6. Springer, pp. 123 141. Schoemaker, E.M., 1963. Impact mechanics at Meteor Crater, Arizona. In: Middlehurts, B., Kuiper, G.P. (Eds.), The Moon Meteorites and Comets. The University of ChicagoPress, Chicago, pp. 301 336. Schmitz, W.J., 1996. On the World Ocean Circulation: Volume I. Some Global Features/North Atlantic Circulation. Woods Hole Oceanographic Institution Technical Report, Woods Hole, MA, WHOI-96-03, 141 p. Schulte, P., et al., 2010. The Chicxulub Asteroid Impact and Mass Extinction at the Cretaceous-Paleogene Boundary. Science 327 (5970), 1214 1218. Available from: https://doi.org/10.1126/science.1177265. Schut, E.W., Uenzelmann-Neben, G., Gersonde, R., 2002. Seismic evidence for bottom current activity at the Agulhas Ridge. Glob. Planet. Change 34, 185 198. Schuster, R.L., 1983. Engineering aspects of the 1980 Mount St. Helens Eruptions. Bull. Assoc. Eng. Geol. 20, 125 143. Schuster, R.L., Highland, L.M., 2001. Socioeconomic and environmental impacts of landslides in the Western Hemisphere. US Geological Survey Open-File Report 01 0276. ,http://pubs.usgs.gov/of/2001/ofr-01-0276/. (accessed 27.12.14.). Schuster, R.L., Wieczorek, G.F., 2002. Landslide triggers and types. In: Rybar, J., Stemberk, J., Wagner, P. (Eds.), Proceedings of the 1st European Conference on Landslides. A.A. Balkema, Prague, pp. 59 78. 24 26 June 2002. Schwab, W.C., Lee, H.J., Twichell, D.C. (Eds.), 1993. Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone, 2002. U.S. Geological Survey Bulletin, p. 204. Schwab, W.C., Lee, H.J., Twichell, D.C., Locat, J., Hans Nelson, C., McArthur, W.C., et al., 1996. Sediment massflow processes on a depositional lobe, outer Mississippi Fan. J. Sediment. Res. 66, 916 927. Schwalbach, J.R., Edwards, B.D., Gorsline, D.S., 1996. Contemporary channel-levee systems in active borderland basin plains, California Continental Borderland. Sediment. Geol. 104, 53 72. Seibold, E., Hinz, K., 1974. Continental slope construction and destruction, West Africa. In: Burk, C.A., Drake, C. L. (Eds.), The Geology of Continental Margins. Springer-Verlag, New York, pp. 179 196. Seilacher, A., 1969. Fault-graded beds interpreted as seismites. Sedimentology 13, 155 159. Seilacher, A., 1984. Sedimentary structures tentatively attributed to seismic events. Mar. Geol. 55, 1 12. Sepm Strata, 2011. SEPM sequence stratigraphy web. ,http://www.sepmstrata.org/. (accessed 04.04.11.). Sepu´lveda, S.A., Padilla, C., 2008. Rain-induced debris and mudflow triggering factors assessment in the Santiago cordilleran foothills, Central Chile. Nat. Hazards 47, 201 215. Sepu´lveda, H.H., Valle-Levinson, A., Framin˜an, M.B., 2004. Observations of subtidal and tidal flow in the Rı´o de la Plata Estuary. Cont. Shelf Res. 24, 509 525. Sepu´lveda S.A., Rebolledo, S., Lara, M., Padilla, C., 2006. Landslide hazards in Santiago, Chile: an overview. The Geological Society of London 2006 1, IAEG2006 Paper number 105. ,http://www2.cose.isu.edu/Bcrosby/ teach/udec/reading/sepulveda_etal_ls_hazards_in_santiago_iaeg_2006.PDF. (accessed 27.12.14.). Serra, N., 2004. Observations and Numerical Modeling of the Mediterranean Outflow (Ph.D. thesis). Faculdade de Ciencias, Universidade de Lisboa, Portugal, p. 234 (unpublished). Seth, A., Sarkar, S., Bose, P.K., 1990. Synsedimentary seismic activity in an immature passive margin basin (Lower Member of the Katrol Formation, Upper Jurassic, Kutch, India). Sediment. Geol. 68, 279 291. Shakesby, R.A., Doerr, S.H., 2006. Wildfire as a hydrological and geomorphological agent. Earth-Sci. Rev. 74, 269 307. Shanmugam, G., 1978. The Stratigraphy, Sedimentology, and Tectonics of the Middle Ordovician Sevier Shale Basin in East Tennessee (Unpublished Ph.D. dissertation). The University of Tennessee, Knoxville, TN, p. 222. Shanmugam, G., 1984. Rock names challenged. Geotimes 29, 6. Shanmugam, G., 1985a. Significance of coniferous rain forests and related Organic matter in generating commercial quantities of oil, Gippsland basin, Australia: Am. Assoc. Petroleum Geol. Bull. 69, 1241 1254. Shanmugam, G., 1985b. Ophiolite source rocks for Taconic-age flysch: trace-element evidence: discussion. GSA Bull. 96, 1221 1222. Shanmugam, G., 1985c. Types of porosity in sandstones and their significance in interpreting provenance. In: Zuffa, G.G. (Ed.), Provenance of Arenites. D. Reidel Publishing Company, pp. 115 137. Shanmugam, G., 1988. Origin, recognition and importance of erosional unconformities in sedimentary basins. In: Kleinspehn, K.L., Paola, C. (Eds.), New Perspectives in Basin Analysis. Springer-Verlag, New York, pp. 83 108.

Bibliography

529

Shanmugam, G., Higgins, J.B., 1988. Porosity enhancement from chert dissolution beneath Neocomian unconformity: Ivishak Formation, North Slope, Alaska. AAPG Bull. 72, 523 535. Shanmugam, G., 1990a. Comment and Reply on “Diagenetic quartzarenite and destruction of secondary porosity: an example from the Middle Jurassic Brent sandstone of northwest Europe”. Geology 18 (3), 287. Shanmugam, G., 1990b. Deep-marine facies models and the interrelationship of depositional components in time and space. In: Brown, G.C., Gorsline, D.S., Schweller, W.J. (Eds.), Deep-Marine Sedimentation: Depositional Models and Case Histories in Hydrocarbon Exploration & Development, vol. 66. SEPM Pacific Section Short Course, San Francisco, CA, pp. 199 246. Shanmugam, G., 1990c. Porosity prediction in sandstones using erosional unconformities. In: Meshri, I.D., Ortoleva, P.J. (Eds.), Prediction of Reservoir Quality Through Chemical Modelling. AAPG Memoir, pp. 1 23. Shanmugam, G., 1996a. High-density turbidity currents: are they sandy debris flows? J. Sediment. Res. 66, 2 10. Shanmugam, G., 1996b. Perception vs. reality in deep-water exploration. World Oil 217, 37 41. Shanmugam, G., 1997a. The Bouma Sequence and the turbidite mind set. Earth-Sci. Rev. 42, 201 229. Shanmugam, G., 1997b. Deep-water exploration: conceptual models and their uncertainties: NAPE (Nigerian Association of Petroleum Explorationists). Bulletin 12/01, 11 28. Shanmugam, G., 1997c. Slope turbidite packets in a fore-arc basin fill sequence of the Plio-Pleistocene Kakegawa Group: their formation and sea-level changes—discussion. Sediment. Geol. 112, 297 300. Shanmugam, G., 1998. Dimensions and Geometries of the Components of Deep-Water Systems. Mobil Technology Company Special Publication, Dallas, TX, p. 110. Shanmugam, G., Poffenberger, M., Toro Alava, J., 2000. Tide-dominated estuarine facies in the Hollin and Napo (‘T’and ‘U’) formations (Cretaceous), Sacha field, Oriente Basin, Ecuador. AAPG Bulletin 84, 652 682. Shanmugam, G., 2000a. 50 years of the turbidite paradigm (1950s 1990s): deep-water processes and facies models a critical perspective. Mar. Petrol. Geol. 17, 285 342. Shanmugam, G., 2000b. John E. Sanders and the turbidite controversy. In: Friedman, G.M. (Ed.), Conference on the History of Geologic Pioneers. Rensselaer Center of Applied Geology, Troy, NY, pp. 19 20. Shanmugam, G., 2002a. Ten turbidite myths. Earth-Sci. Rev. 58, 311 341. Shanmugam, G., 2002b. Discussion on Mulder et al. (2001, Geo-Marine Letters, 21, 86 93) Inversely graded turbidite sequences in the deep Mediterranean. A record of deposits from flood-generated turbidity currents? GeoMar. Lett. 22, 108 111. Shanmugam, G., 2002c. Deep-Marine Sediments. Access Science McGraw-Hill Encyclopedia of Science & Technology, ,http://accessscience.com/abstract.aspx?id 5 757553&referURL 5 http%3a%2f%2faccessscience. com%2fcontent.aspx%3fid%3d757553. (accessed 20.05.11.). Shanmugam, G., 2002d. Submarine Canyon. AccessScience McGraw-Hill Encyclopedia of Science & Technology, ,http://accessscience.com/abstract.aspx?id 5 664300&referURL 5 http%3a%2f%2faccessscience.com%2fcontent.aspx%3fid%3d664300. (accessed 20.05.11.). Shanmugam, G., 2003a. Deep-marine tidal bottom currents and their reworked sands in modern and ancient submarine canyons. Mar. Petrol. Geol. 20, 471 491. Shanmugam, G., Poffenberger, M., Toro Alava, J., 2000. Tide-dominated estuarine facies in the Hollin and Napo (‘T’and ‘U’) formations (Cretaceous), Sacha field, Oriente Basin, Ecuador. AAPG Bulletin 84, 652 682. Shanmugam, G., 2003b. A preliminary experimental study of turbidite fan deposits: discussion. J. Sediment. Res. 73, 838 839. Shanmugam, G., 2006a. Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs. Elsevier, Amsterdam, p. 476. Shanmugam, G., 2006b. The tsunamite problem. J. Sediment. Res. 76, 718 730. Shanmugam, G., 2007. The obsolescence of deep-water sequence stratigraphy in petroleum geology. Indian J. Petrol. Geol. 16 (1), 1 45. Shanmugam, G., 2008a. The constructive functions of tropical cyclones and tsunamis on deepwater sand deposition during sea level highstand: implications for petroleum exploration. AAPG Bull. 92, 443 471. Shanmugam, G., 2008b. Chapter 5 Deep-water bottom currents and their deposits. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, Developments in Sedimentology, Vol. 60. Elsevier, Amsterdam, pp. 59 81. Shanmugam, G., 2008c. Leaves in turbidite sand: the main source of oil and gas in the deep-water Kutei Basin, Indonesia: discussion. AAPG Bull. 92, 127 137.

530

Bibliography

Shanmugam, G., 2008d. Slides, slumps, debris flows, and turbidity currents. In: Steele, J.H., Turekian, K.K., Thorpe, S.A. (Eds.), Encyclopedia of Ocean Sciences, second ed. Elsevier, ISBN: 978-0-12-374473-9 pp. 447 467. ,http://www.sciencedirect.com/science?_ob 5 ArticleURL&_udi 5 B99B0-4V2BCCJ-FM&_rdoc 5 51&_hierId 5 390954000021&_refWorkId 5 9119&_explode 5 390954000021&_alpha 5 S&_fmt 5 high&_orig 5 na&_docanchor5 &_idxType 5 AR&view 5 c&_ct 5 65&_acct 5 C000050221&_version 5 1&_urlVersion 5 0&_userid 5 10&md55 685a6d9dc8b0582f270ac8385a5f12ba. (accessed 31.05.10.). Shanmugam, G., Poffenberger, M., Toro Alava, J., 2000. Tide-dominated estuarine facies in the Hollin and Napo (‘T’ and ‘U’) formations (Cretaceous), Sacha Field, Oriente Basin, Ecuador. AAPG Bulletin 84, 652 682. Shanmugam, G., 2009a. Slides, slumps, debris flows, and turbidity currents. In: Steele, J.H., Thorpe, S.A., Turekian, K.K. (Eds.), Encyclopedia of Ocean Sciences, second ed. Academic Press (Elsevier), Waltham, MA, pp. 447 467. Shanmugam, G., 2009b. Comment on “Late Holocene Rupture of the Northern San Andreas Fault and Possible Stress Linkage to the Cascadia Subduction Zone” by C. Goldfinger, K. Grijalva, R. Bu¨rgmann, A. E. Morey, J. E. Johnson, C. Hans Nelson, J. Gutie´rrez-Pastor, A. Ericsson, E. Karabanov, J. D. Chaytor, J. Patton, and E. Gra`cia. Bull. Seismol. Soc. Am. 99 (4), 2594 2598. Shanmugam, G., 2009c. Deep-marine process sedimentology vs. deep-marine sequence stratigraphy: implications for petroleum exploration. In: Raman, C.V. (Ed.), International Seminar on Prospects of Marine Sediments as Resource Base for the Growth of Mankind. Andhra University, Visakhapatnam, Abstract v, 9-12.G. Shanmugam, G., 2010a. Sandy-mass-transport deposits (SMTD) in deep-water environments: recognition, geometry, and reservoir quality. American Association of Petroleum Geologists Annual Convention and Exhibition, New Orleans, LO, Abstracts Volume, p. 234. ,http://www.searchanddiscovery.net/documents/2010/ 50291shanmugam/ndx_shanmugam.pdf. (accessed 11.10.10.). Shanmugam, G., 2010b. Discussion: turbidites and turbidity currents from Alpine ‘flysch’ to the exploration of continental margins by Mutti et al. (2009), Sedimentology, 56, 267 318. Sedimentology 57, 932. Available from: https://doi.org/10.1111/j.1365-3091.2009.01132.x. Shanmugam, G., 2011a. Transport mechanisms of sand in deep-marine environments—insights based on laboratory experiments: discussion. J. Sediment. Res. 81, 841. Shanmugam, G., 2011b. Book review of “Deep-Sea Sediments by Hu¨eneke, H., Mulder, T. (Eds.), 2011. Elsevier, Amsterdam, Developments in Sedimentology 63”. Geologos. Shanmugam, G., 2012a. New Perspectives on Deep-Water Sandstones: Origin, Recognition, Initiation, and Reservoir Quality, 9. Elsevier, Handbook of Petroleum Exploration and Production, Amsterdam, p. 524. Shanmugam, G., 2012b. Process-sedimentological challenges in distinguishing paleo-tsunami deposits. In: Kumar, A., Nister, I. (Eds.), Paleo-tsunamis. Natural Hazards, 63, 5 30. Shanmugam, G., 2012c. Discussion of He et al., 2011 (Geo-Marine Letters) Evidence of internal-wave and internaltide deposits in the Middle Ordovician Xujiajuan Formation of the Xiangshan Group, Ningxia, China. GeoMar. Lett. 32, 359 366. Shanmugam, G., 2013a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands. AAPG Bull. 97, 767 811. Shanmugam, G., 2013b. Comment on “Internal waves, an underexplored source of turbulence events in the sedimentary record” by L. Pomar, M. Morsilli, P. Hallock, and B. Ba´denas [Earth-Science Reviews, 111 (2012), 56 81]. Earth-Sci. Rev. 116, 195 205. Shanmugam, G., 2013c. Slides, slumps, debris flows, and turbidity currents. In: Elias, S.A. (Ed.), Reference Module in Earth Systems and Environmental Science. Elsevier Online. Shanmugam, G., 2014a. Modern internal waves and internal tides along oceanic pycnoclines: challenges and implications for ancient deep-marine baroclinic sands: reply. Am. AAPG Bull. 98, 858 879. Shanmugam, G., 2014b. Review of research in internal-wave and internal-tide deposits of China, discussion. J. Palaeogeogr. 3 (4), 332 350. Shanmugam, G., 2015a. The landslide problem. J. Palaeogeogr. 4, 109 166. Shanmugam, G., 2015b. 3D palaeogeographic reconstructions of the Phanerozoic versus sea-level and Sr-ratio variations, discussion. J. Palaeogeogr. 4 (3), 234 243. Shanmugam, G., 2016a. Submarine fans: a critical retrospective (1950 2015). J. Palaeogeogr. 5 (2), 110 184.

Bibliography

531

Shanmugam, G., 2016b. The contourite problem. In: Mazumder, R. (Ed.), Sediment Provenance. Elsevier, pp. 183 254. Chapter 9. Shanmugam, G., 2016c. The seismite problem. J. Palaeogeogr. 5 (4), 318 362. Shanmugam, G., 2016d. Glossary: a supplement to “Submarine fans: a critical retrospective (1950-2015)”. J. Palaeogeogr. 5 (3), 258 277. Shanmugam, G., 2017a. Global case studies of soft-sediment deformation structures (SSDS): definitions, classifications, advances, origins, and problems. J. Palaeogeor. 6 (4), 251 320. Shanmugam, G., 2017b. Contourites: physical oceanography, process sedimentology, and petroleum geology. Petrol. Explor. Dev. 44 (2), 183 216. Shanmugam, G., 2017c. The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India: discussion and liquefaction basics. J. Palaeogeogr. 6 (3), 224 234. Shanmugam, G., 2017d. The fallacy of interpreting SSDS with different types of breccias as seismites amid the multifarious origins of earthquakes: implications. J. Palaeogeogr. 6 (1), 12 44. Shanmugam, G., 2018a. Slides, slumps, debris flows, turbidity currents, and bottom currents: implications. Earth Systems and Environmental Sciences. Elsevier Online Module, 120 p. Shanmugam, G., 2018b. The hyperpycnite problem. J. Palaeogeogr. 7 (3), 197 238. Shanmugam, G., 2018c. A global satellite survey of density plumes at river mouths and at other environments: plume configurations, external controls, and implications for deep-water sedimentation. Petrol. Explor. Dev. 45 (4), 640 661. Shanmugam, G., 2018d. Preface. In: Special Issue dedicated to George Devries Klein by the Journal of the Indian Association of Sedimentologists (JIAS). J. Indian Assoc. Sedimentol. 35 (2), 1 5. Shanmugam, G., 2018e. Bioturbation and trace fossils in deep-water contourites, turbidites, and hyperpycnites: a cautionary note. In: Special Issue dedicated to George Devries Klein by the Journal of the Indian Association of Sedimentologists (JIAS). J. Indian Assoc. Sedimentol. 35 (2), 13 32. Shanmugam, G., 2018f. Comment on “Ichnological analysis of contourites: past, present and future” by Francisco J. Rodrı´guez-Tovar and F. Javier Hernandez-Molina [Earth-Science Reviews, 182 (2018), 28 41]. Earth-Sci. Rev. 184, 46 49. Shanmugam, G., 2019a. Slides, slumps, debris flows, turbidity currents, hyperpycnal flows, and bottom currents. In: third ed. Kirk Cochran, J. (Ed.), Henry Bokuniewicz Patricia Yager Encyclopedia of Ocean Sciences, vol. 4. pp. 228 257. Shanmugam, G., 2019b. Climatic and tectonic controls of lacustrine hyperpycnite origination in the Late Triassic Ordos Basin, central China: implications for unconventional petroleum development: discussion. AAPG Bull. 103 (2), 505 510. Shanmugam, G., 2019c. Reply to discussions by Zavala (2019) and by Van Loon, Hu¨eneke, and Mulder (2019) on Shanmugam, G. (2018, Journal of Palaeogeography, 7 (3): 197 238): the hyperpycnite problem. J. Palaeogeogr. 8 (4), 408 421. Shanmugam, G., 2019d. Global significance of wind forcing on deflecting sediment plumes at river mouths: implications for hyperpycnal flows, sediment transport, and provenance. J. Indian Assoc. Sedimentol. 36 (2), 1 37. Shanmugam, G. 2020a. The turbidite – contourite –tidalite – hybridite problem: Orthodoxy Vs Empirical Evidence behind the “Bouma Sequence”. The Drifters VGT (Virtual Get-Together) Zoom Lecture organized by F. J. Hernandez-Molina, Dept. Earth Sciences, Royal Holloway, University of London (UK), July 27, 2020, Monday, 2.30 PM London (UK) Time. YouTube video of lecture with editorial comments: https://www.youtu.be/ ayxlusmkDCA. Accessed August 17, 2020. Shanmugam, G. 2020b. Deep-water processes and deposits. In: Elsevier’s https://www.sciencedirect.com/science/referenceworks/9780124095489 Reference Module in Earth Systems and Environmental Sciences. https://www.sciencedirect.com/science/article/pii/B9780124095489125412. Shanmugam, G., Benedict, G.L., 1978. Fine-grained carbonate debris flow, Ordovician basin margin, Southern Appalachians. J. Sediment. Petrol. 48, 1233 1240. Shanmugam, G., Walker, K.R., 1978. Tectonic significance of distal turbidites in the Middle Ordovician Blockhouse and lower Sevier formations in east Tennessee. Am. J. Sci. 278, 551 578. Shanmugam, G., Walker, K.R., 1980. Sedimentation, subsidence, and evolution of a foredeep basin in the Middle Ordovician, Southern Appalachians. Am. J. Sci. 280, 479 496.

532

Bibliography

Shanmugam, G., Moiola, R.J., 1982. Eustatic control of turbidites and winnowed turbidites. Geology 10, 231 235. Shanmugam, G., Benedict III, G.L., 1983. Manganese distribution in the carbonate fraction of shallow and deep marine lithofacies, Middle Ordovician, eastern Tennessee. Sediment. Geol. 35, 159 175. Shanmugam, G., Moiola, R.J., 1983. Comment and reply on “Eustatic control of turbidites and winnowed turbidites”-reply. Geology 11, 58 60. Shanmugam, G., Lash, G.G., 1983a. Comment and reply on “Analogous tectonic evolution of the Ordovician foredeeps, southern and central Appalachians”-reply. Geology 11, 492 493. Shanmugam, G., Lash, G.G., 1983b. Comment and reply on “Analogous tectonic evolution of the Ordovician foredeeps, southern and central Appalachians”-reply. Geology 11, 732 733. Shanmugam, G., Moiola, R.J., 1985a. Submarine fan models: problems and solutions. In: Bouma, A.H., Normark, W.R., Barnes, N.E. (Eds.), Submarine Fans and Related Turbidite Systems. Springer- Verlag, New York, pp. 29 34. Shanmugam, G., Moiola, R.J., 1985b. Sedimentary facies of the Nova Scotian upper and middle continental slope, offshore eastern Canada, Discussion. Sedimentology 32, 455. Shanmugam, G., McPherson, J.G., 1987. Sedimentation in the Chile Trench: depositional morphologies, lithofacies, and stratigraphy: discussion and reply. GSA Bull. 99 (4), 598. Shanmugam, G., Alhilali, K.A., 1988. Parameters influencing Porosity in sandstones: a model for sandstone porosity prediction: discussion. AAPG Bull. 72, 852 853. Shanmugam, G., Moiola, R.J., 1988. Submarine fans: characteristics, models, classification, and reservoir potential. Earth-Sci. Rev. 24, 383 428. Shanmugam, G., Clayton, C.A., 1989. Reservoir description of a sand rich submarine fan complex for a steamflood project: Upper Miocene Potter Sandstone, North Midway Sunset Field, California. AAPG Bull. 73, 411. Shanmugam, G., Higgins, J.B., 1990a. Porosity enhancement from chert dissolution beneath Neocomian unconformity: Ivishak formation, North Slope, Alaska: reply, AAPG Bull., 74. pp. 89 90. Shanmugam, G., Higgins, J.B., 1990b. Porosity enhancement from chert dissolution beneath Neocomian unconformity: Ivishak formation, North Slope, Alaska: reply, AAPG Bull., 74. pp. 1752 1753. Shanmugam, G., Moiola, R.J., 1991. Types of submarine fan lobes, models and implications. AAPG Bull. 75, 156 179. Shanmugam, G., Moiola, R.J., 1995. Reinterpretation of depositional processes in a classic flysch sequence (Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma. AAPG Bull. 79, 672 695. Shanmugam, G., Zimbrick, G., 1996. Sandy slump and sandy debris flow facies in the Pliocene and Pleistocene of the Gulf of Mexico: implications for submarine fan models. In: AAPG International Congress and Exhibition, Caracas, Official Program, A45. Shanmugam, G., Moiola, R.J., 1997. Reinterpretation of depositional processes in a classic flysch sequence (Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma: reply. AAPG Bull. 81, 476 491. Shanmugam, G., Damuth, J.E., Moiola, R.J., 1985a. Is the turbidite facies association scheme valid for interpreting ancient submarine fan environments? Geology 13, 234 237. Shanmugam, G., Walker, K.R., Lash, G.G., 1985b. Comment and reply on “Provenance of the Middle Ordovician Blount clastic Wedge, Georgia and Tennessee”, Comment. Geology 13, 889 891. Shanmugam, G., Moiola, R.J., Sales, J.K., 1988a. Duplex-like structures in submarine fan channels, Ouachita Mountains, Arkansas. Geology 16, 229 232. Shanmugam, G., Moiola, R.J., McPherson, J.G., O’Connell, S., 1988b. Comparison of turbidite facies associations in modern passive-margin Mississippi Fan with ancient active-margin fans. Sediment. Geol. 58, 63 77. Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993a. Process sedimentology and reservoir quality of deepmarine bottom-current reworked sands (sandy contourites): an example from the Gulf of Mexico. AAPG Bull. 77, 1241 1259. Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993b. Traction structures in deep-marine bottom-current reworked sands in the Pliocene and Pleistocene, Gulf of Mexico. Geology 21, 929 932. Shanmugam, G., Lehtonen, L.R., Straume, T., Syversten, S.E., Hodgkinson, R.J., Skibeli, M., 1994. Slump and debris flow dominated upper slope facies in the Cretaceous of the Norwegian and Northern North Seas (61 67 N): implications for sand distribution. AAPG Bull. 78, 910 937.

Bibliography

533

Shanmugam, G., Bloch, R.B., Mitchell, S.M., Beamish, G.W.J., Hodgkinson, R.J., Damuth, J.E., et al., 1995a. Basin-floor fans in the North Sea: sequence stratigraphic models vs. sedimentary facies. AAPG Bull. 79, 477 512. Shanmugam, G., Hermance, W.E., Olaifa, J.O., Odior, E.G., 1995b. Slump dominated upper slope reservoir facies, Intra Qua Iboe (Pliocene), Edop Field, Offshore Nigeria (abs.). AAPG Annu. Convention, Off. Program. 88A. Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1995c. Deep-marine bottom-current reworked sand (Pliocene and Pleistocene), Ewing Bank 826 Field, Gulf of Mexico. In: Winn, R.D., Armentrout, J.M. (Eds.), Turbidites and Associted Deep-Water Facies, SEPM Core Workshop No. 20, pp. 25 54. Shanmugam, G., Bloch, R.B., Mitchell, S.M., Damuth, J.E., Beamish, G.W.J., Hodgkinson, R.J., et al., 1996. Slump and debris-flow dominated basin-floor fans in the North Sea: an evaluation of conceptual sequencestratigraphical models based on conventional core data. In: Hesselbo, S.P., Parkinson, D.N. (Eds.), Sequence Stratigraphy in British Geology. Geological Society Special Publication No. 103, pp. 145 175. Shanmugam, G., Bloch, R.B., Damuth, Hodgkinson, R.J., 1997a. Basin-floor fans in the North Sea: sequence stratigraphic models vs. sedimentary facies: reply. AAPG Bull. 81, 666 672. Shanmugam, G., Famakinwa, S.B., Hodgkinson, R.J., et al., 1997b. Deep-marine slump and debris flow dominated reservoirs of the Zafiro Field, offshore Equatorial Guinea, 1997. In: AAPG International Conference and Exhibition, Vienna, 6. Shanmugam, G., Poffenberger, M., Toro Alava, J., 2002. Tide-dominated estuarine facies in the Hollin and Napo (“T” and “U”) Formations (Cretaceous), Sacha Field, Oriente Basin, Ecuador: reply, AAPG Bull., 86. pp. 335 340. Shanmugam, G., Shrivastava, S.K., Das, B., 2009. Sandy debrites and tidalites of Pliocene reservoir sands in upper-slope canyon environments, Offshore Krishna-Godavari Basin (India): implications. J. Sediment. Res. 79, 736 756. Shapiro, S.A., Marsaglia, K.M., Carter, L., 2007. The petrology and provenance of sand in the Bounty Submarine Fan, New Zealand. In: Arribas, J., Johnsson, M.J., Critelli, S. (Eds.), Sedimentary Provenance and Petrogenesis: Perspectives From Petrography and Geochemistry. GSA Special Paper 420, pp. 277 296. Sharp, R.P., 1954. Glacier flows; a review. Geol. Soc. Am. Bull. 65, 821 838. Sharpe, C.F.S., 1938. Landslides and Related Phenomena. Columbia University Press, New York, p. 137. Shepard, F.P., 1932. Landslide modifications of submarine valleys. Trans. Am. Geophys. Union. 13, 226 230. Shepard, F.P., 1954. High-velocity turbidity currents, a discussion. Proc. R. Soc. London. Ser. A, Math. Phys. Sci. 222, 323 326. Shepard, F.P., 1976. Tidal components of currents in submarine canyons. J. Geol. 84, 343 350. Shepard, F.P., 1981. Submarine canyons: multiple causes and long-time persistence. AAPG Bull. 65, 1062 1077. Shepard, F.P., Dill, R.F., 1966. Submarine Canyons and Other Sea Valleys. Rand McNally & Co., Chicago, IL, p. 381. Shepard, F.P., Emery, K.O., 1973. Congo submarine canyon and fan valley. AAPG Bull. 57, 1679 1691. Shepard, F.P., Marshall, N.F., 1978. Currents in submarine canyons and other sea valleys. In: Stanley, D.J., Kelling, G. (Eds.), Sedimentation in Submarine Fans, Canyons, and Trenches. Hutchinson and Ross, Stroudsburg, PA, pp. 3 14. Shepard, F.P., Dill, R.F., Von Rad, U., 1969. Physiography and sedimentary processes of La Jolla Submarine Fan and Fan-Valley, California. AAPG Bull. 53, 39 420. Shepard, F.P., Marshall, N.F., McLoughlin, P.A., Sullivan, G.G., 1979. Currents in submarine canyons and other sea valleys. AAPG Stud. Geol. 8, 173. Sherrod, L., Simpson, E.L., Higgins, R., Miller, K., Morgano, K., Snyder, E., et al., 2016. Subsurface structure of water gas escape features revealed by ground-penetrating radar and electrical resistivity tomography, Glen Canyon National Recreation Area, Lake Powell delta, Utah, USA. Sediment. Geol. 344, 160 174. Shiki, T., Cita, M.B., Gorsline, D.S. (Eds.), 2000. Sedimentary features of seismites, seismo-turbidites and tsunamites. Sediment. Geol., 135, p. 324. Shillington, F.A., Hutchings, L., Probyn, T.A., Waldron, H.N., Peterson, W.T., 1992. Filaments and the Benguela frontal zone: offshore advection or recirculating loops? South. Afr. J. Mar. Sci. 12 (1), 207 218. Shinn, E.A., Steinen, R.P., Lidz, B.H., Swart, P.K., 1989. Whitings, a sedimentologic dilemma. J. Sediment. Petrol. 59, 147 161.

534

Bibliography

Shipboard Scientific Party, 1973a. Introduction, Initial Reports of the Deep Sea Drilling Project, vol. 18. United States Government Printing Office, Washington, DC, pp. 5 8, ,https://doi.org/10.2973/dsdp. proc.18.101.1973.. Shipboard Scientific Party, Speeden, I.G., 1973b. 4. Site 204. In: Burns, R.E., Andrews, J.E. (Eds.), Initial Reports of the Deep Sea Drilling Project, vol. 21. United States Government Printing Office, Washington, DC, pp. 33 56. ,https://doi.org/10.2973/dsdp.proc.21.104.1973.. Shipboard Scientific Party, 1975. 7. Site 296. In: Ingle Jr., J.C., Karig, D.E., Bouma, A.H., et al.,Initial Reports of the Deep Sea Drilling Project, vol. 31. United States Government Printing Office, Washington, DC, pp. 191 274. ,https://doi.org/10.2973/dsdp.proc.31.107.1975.. Shipboard Scientific Party, 1983. 5. Site 516: Rio Grande Rise. In: Barker, P.F., Carlson, R.L., Johnson, D.A., et al., Initial Reports of the Deep Sea Drilling Project, Volume 72. United States Government Printing Office, Washington, p. 155e338. Available from: http://dx.doi.org/10.2973/dsdp.proc.72.105.1983. Shipboard Scientific Party, 1978. 2. Cape basin continental rise—Sites 360 and 361. In: Bolli, H.M., Ryan, W.B.F., et al.,Initial Reports of the Deep Sea Drilling Project, 40. United States Government Printing Office, Washington, DC, pp. 29 182. ,https://doi.org/10.2973/dsdp.proc.40.102.1978.. Shipboard Scientific Party, 1981. 6. Site 451: East Edge of the West Mariana Ridge. In: Kroenke, L., Scott, R., et al., Initial Reports of the Deep Sea Drilling Project, 59 Vol.. United States Government Printing Office, Washington, DC, pp. 405 483. ,https://doi.org/10.2973/dsdp.proc.59.106.1981.. Shipboard Scientific Party, 1988. Sites 712 and 713. In: Backman, J., Duncan, R.A., et al. (Eds.), Proceedings of Ocean Drilling Program, Initial Reports, 115, College Station, TX, pp. 733 845. ,https://doi.org/10.2973/ odp.proc.ir.115.111.1988.. Shipboard Scientific Party, 1989. 9. Site 755. In: Peirce, J., Weissel, J., et al. (Eds.), Proceedings of Ocean Drilling Program, Initial Reports, 121. College Station, TX, pp. 237 258, ,https://doi.org/10.2973/odp.proc. ir.121.109.1989.. Shipboard Scientific Party, 1992. Site 833. In: Collot, J.-Y., Greene, H.G., Stokking, L.B., et al. (Eds.), Proceedings of Ocean Drilling Program, Initial Reports, 134, College Station, TX, pp. 479 557. ,https://doi.org/10.2973/ odp.proc.ir.134.113.1992.. Shipboard Scientific Party, 1993. Site 884. In: Rea, D.K., Basov, I.A., Janecek, T.R., Palmer-Julson, A., et al. (Eds.), Proceedings of Ocean Drilling Program, Initial Reports, 134, College Station, TX, pp. 209 302. ,https://doi. org/10.2973/odp.proc.ir.145.108.1993.. Shipboard Scientific Party, 1994. Site 905. In: Mountain, G.S., Miller, K.G., Blum, P., et al. (Eds.), Proceedings of Ocean Drilling Program, Initial Reports, 134, College Station, TX, pp. 255 308. ,https://doi.org/10.2973/ odp.proc.ir.150.109.1994.. Shipboard Scientific Party, 1995a. Site 934. In: Flood, R.D., Piper, D.J.W., Klaus, A., et al. (Eds.), Proceedings of the ODP Initial Reports 155, College Station, TX, Ocean Drilling Program, pp. 241 271. Shipboard Scientific Party, 1995b. Site 944. In: Flood, R.D., Piper, D.J.W., Klaus, A., et al. (Eds.), Proceedings of the ODP Initial Reports 155, College Station, TX, Ocean Drilling Program, 591 633. Shipboard Scientific Party, 1998. Site 1085. In: Wefer, G., Berger, W.H., Richter, C., et al. (Eds.), Proceedings of Ocean Drilling Program, Initial Reports, 134, College Station, TX, pp. 385 428. ,https://doi.org/10.2973/ odp.proc.ir.175.113.1998.. Shipboard Scientific Party, 2001. Site 1166. In: O’Brien, P.E., Cooper, A.K., Richter, C., et al. (Eds.), Proceedings of Ocean Drilling Program, Initial Reports, 134, College Station, TX, pp. 1 110. ,https://doi.org/10.2973/odp. proc.ir.188.104.2001.. Shipboard Scientific Party, Coleman, R., Hamilton, N., 1974. 7. Site 223. In: Whitmarsh, R.B., Weser, O.E., et al., Initial Reports of the Deep Sea Drilling Project, 23. United States Government Printing Office, Washington, DC, pp. 292 381. ,https://doi.org/10.2973/dsdp.proc.23.107.1974.. Shipley, S., Sarna-Wojcicki, A.M., 1982. Distribution, thickness, and mass of late pleistocene and holocene tephra from major volcanoes in the northwestern United States: a preliminary assessment of hazards from volcanic ejecta to nuclear reactors in the Pacific Northwest. US Geological Survey Miscellaneous Field Studies Map MF1435. Shipboard Scientific Party, 1995d. Site 946. In: Flood, R.D., Piper, D.J.W., Klaus, A., et al. (Eds.), Proceedings of the ODP Initial Reports 155. Ocean Drilling Program, College Station, Texas, pp. 657e688.

Bibliography

535

Shipp, R.C., Weimer, P., Posamentier, H. (Eds.), 2009. Mass-Transport Deposits in Deepwater Settings. SEPM Special Publication No. 96. Shirley, K., 2003. India gas find has major impact. American Association of Petroleum Geologists, Explorer. ,http://www.aapg.org/explorer/2003/01jan/india.cfm. (accessed 18.04.09.). Shoaei, Z., Ghayoumian, J., 1998. Seimareh landslide, the largest complex slide in the world. In: Moore, D., Hungr, O. (Eds.), Proceedings of the Eighth International Congress of the International Association for Engineering Geology and the environment, pp. 1337 1342. Shreve, R.L., 1968. The Blackhawk Landslide. Geological Society of America Special Paper 108, p. 47. Shultz, A.W., 1984. Subaerial debris-flow deposition in the upper Paleozoic Cutler Formation, Western Colorado. J. Sediment. Petrol. 54, 759 772. Simons, D.B., Richardson, E.V., Nordin Jr., C.F., 1965. Sedimentary structures generated by flow in alluvial channels. In: Middleton, G.V. (Ed.), Primary Sedimentary Structures and Their Hydrodynamic Interpretation, 12. SEPM Special Publication, pp. 34 52. Simpson, E.J., 1972. Effects of the lower boundary on the head of a gravity current. J. Fluid Mech. 53, 759 768. Singer, K.N., McKinnon, W.B., Schenk, P.M., Moore, J.M., 2012. Massive ice avalanches on Iapetus mobilized by friction reduction during flash heating. Nat. Geosci. 5, 574 578. Skipper, K., 1971. Antidune cross-stratification in a turbidite sequence, Cloridorme Formation (Middle Ordovician) Gaspe´, Quebec. Sedimentology 17, 51 68. Skempton, A.W., 1960. Terzaghi’s discovery of effective stres. In: Bjerrum, L., Casagrande, A., Peck, R.B., Skempton, A.W. (Eds.), From Theory to Practice in Soil Mechanics. John Wiley, Hoboken, New Jersey, pp. 42 53. Slatt, R.M., Weimer, P., Stone, C.G., 1997. Reinterpretation of depositional processes in a classic flysch sequence (Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma, discussion. AAPG Bull. 81, 449 459. Smit, J., Roep, T.B., Alvarez, W., Montanari, A., Claeys, P., Grajales-Nishimura, J.M., et al., 1996. Coarse-grained, clastic sandstone complex at the K-T boundary around the Gulf of Mexico: deposition by tsunami waves induced by the Chicxulub impact? In: Ryder, G., et al., (Eds.), The Cretaceous-Tertiary Event and Other Catastrophes in Earth History. Geological Society of America Special Paper 307, pp. 151 182. Smith, R.D.A., 1995. Reservoir architecture of syn-rift lacustrine turbidite systems, early Cretaceous, offshore South Gabon. In: Lambiase, J.J. (Ed.), Hydrocarbon Habitat in Rift Basins, Geological Society Special Publication, 80. pp. 197 210. Snedden, J.W., Nummedal, D., Amos, A.F., 1988. Storm- and fair-weather combined flow on the Central Texas continental shelf. J. Sediment. Petrol. 58, 580 595. Snyder, M.E., Waldron, J.W.F., 2016. Unusual soft-sediment deformation structures in the Maritimes Basin, Canada: possible seismic origin, Sediment. Geol., 344. pp. 145 159. So, Y.S., Rhee, C.W., Choi, P.-Y., Kee, W.S., Seo, J.Y., Lee, E.-J., 2013. Distal turbidite fan/lobe succession of The Late Paleozoic Taean Formation, Western Korea. Geosci. J. 17 (1), 9 25. Sobiesiak, M.S., Kneller, B., Alsop, G.I., Milana, J.P., 2016. Internal deformation and kinematic indicators within a tripartite mass transport deposit, NW Argentina. Sediment. Geol. 344, 364 381. Soh, W., 1989. Coarse clast dominant submarine debrite, the Mio-Pliocene Fujukawa Group, Central Japan. In: Taira, A., Mazuda, F. (Eds.), Sedimentary Facies in the Active Plate Margin. Terra Scientific Publishing Company (TERRAPUB), Tokyo, pp. 495 510. Sohn, Y.K., Kim, S.B., Hwang, J.G., Bahk, J.J., Choe, M.Y., Chough, S.K., 1997. Characteristics and depositional processes of large-scale gravelly Gilbert-type foresets in the Miocene Doumsan fan delta, Pohang Basin, SE Korea. J. Sediment. Res. 67, 130 141. Solheim, A. (Ed.), 2006. Submarine mass movements and their consequences: Proceedings of the 2nd International Conference, Oslo, Norway, September 5 7, 2005. Nor. J. Geol. 86, 151 372. Solheim, A., Bryn, P., Sejrup, H.P., Mienert, J., Berg, K., 2005a. Ormen Lange—an integrated study for the safe development of a deep-water gas field within the Storegga Slide Complex, NE Atlantic continental margin; executive summary. Mar. Petrol. Geol. 22, 1 9. Solheim, A., Berg, K., Forsberg, C.F., Bryn, P., 2005b. The Storegga Slide complex: repetitive large scale sliding with similar cause and development. Mar. Petrol. Geol. 22 (2005), 97 107.

536

Bibliography

Soomere, T., 2010. Rogue waves in shallow water. Eur. Phys. J. Spec. Top. 185, 81 96. Available from: https:// doi.org/10.1140/epjst/e2010-01240-1. Sorby, H.C., 1908. On the application of quantitative methods to the study of rocks. Q. J. Geol. Soc. 64, 171 233. Southard, J.B., 1975. Bed configutations. In: Harms, J.C., Southard, J.B., Spearing, D.R., Walker, R.G. (Eds.), Depositional Environments as Interpreted from Primary Sedimentary Structures and Stratification Sequences, Dallas, TX, Lecture Notes for SEPM Short Course No. 2, pp. 5 43. Southard, J.B., Boguchwal, L.A., 1990. Bed configurations in steady unidirectional water flows, 2: synthesis of flume data. J. Sediment. Petrol. 60, 658 679. Southard, J.B., Cacchione, D.A., 1972. Experiments on bottom sediment movement by breaking internal waves. In: Swift, D.J.P., Duane, D.B., Pilkey, O.H. (Eds.), Shelf Sediment Transport: Process and Pattern. Dowden, Hutchinson & Ross, Stroudsburg, PA, pp. 83 97. Southard, J.B., Stanley, D.J., 1976. Shelf-break processes and sedimentation. In: Stanley, D.J., Swift, J.P. (Eds.), Marine Sediment Transport and Environmental Management. John Wiley & Sons, New York, pp. 351 377. Sowers, G., 1979. Introductory Soil Mechanics and Foundations, 4th ed. Geotechnical Engineering, 640. Prentice Hall, New Jersey. Soyinka, O.A., Slatt, R., 2008. Identification and microstratigraphy of hyperpycnites and turbidites in Cretaceous Lewis Shale, Wyoming. Sedimentology 55, 1117 1133. Sridhar, P.N., Ali, M.M., Vethamony, P., Babu, M.T., Ramana, V., Jayakumar, S., 2008a. Seasonal occurrence of unique sediment plume in the Bay of Bengal. Eos Trans. Am. Geophys. Union. 89 (3), 22 23. Sridhar, P.N., Ramana, I., Vali, M.M., Veeranarayana, B., 2008b. Understanding the suspended sediment dynamics in the coastal waters of the Bay of Bengal using high resolution ocean colour data. Curr. Sci. 94 (11), 1499 1502. Srurkel, B.F.F., Ormo¨, J., 1997. Impact-related clastic injections in the marine Ordovician Lockne impact structure, Central Sweden. Sedimentology 44, 793 804. St. Laurent, L., Alford, M.H., Paluszkiewicz, T., 2012. An introduction to the special issue on internal waves. Oceanography 25 (2), 15 19. Available from: https://doi.org/10.5670/oceanog.2012.37. Stanford, J.D., Rohling, E.J., Bacon, S., Holliday, N.P., 2011. A review of the deep and surface currents around Eirik Drift, south of Greenland: comparison of the past with the present. Glob. Planet. Change 79, 244 254. Stanley, D.J., 1963. Vertical petrographic variability in Annot Sandstone turbidites. J. Sediment. Petrol. 33, 783 788. Stanley, D.J., 1975. Submarine canyon and slope sedimentation (Gres D’Annot) in the French Maritime Alps. In: IXth Congress International de Sedimentologie, Nice, Field Guide, p. 129. Stanley, D.J., 1980. The Saint-Antonin Conglomerate in the Maritime Alps: A Model for Coarse Sedimentation on a Submarine Slope: Smithsonian Contributions to the Marine Sciences 5. Smithsonian Institution Press, Washington, DC, 36 pp. Stanley, D.J., 1981. Unifites: structureless muds of gravity flow origin in Mediterranean basins. Geo-Mar. Lett. 1, 77 83. Stanley, D.J., 1993. Model for turbidite-to-contourite continuum and multiple process transport in deep marine settings: example in the rock record. Sediment. Geol. 82, 241 255. Stanley, D.J., Kelling, G. (Eds.), 1978. Sedimentation in Submarine Canyons, Fans, and Trenches. Dowden, Hutchinson and Ross, Inc., Stroudsburg, PA. Stanley, D.J., Moore, G.T. (Eds.), 1983. The Selfbreak: Critical Interface on Continental Margins, No. 33. SEPM Special Publication, p. 467. Stanley, D.J., Palmer, H.D., Dill, R.F., 1978. Coarse sediment transport by mass flow and turbidity current processes and downslope transformations in Annot Sandstone canyon-fan valley systems. In: Stanley, D.J., Kelling, G. (Eds.), Sedimentation in Submarine Canyons, Fans, and Trenches. Hutchinson and Ross, Stroudsburg, PA, pp. 85 115. Stauffer, P.H., 1967. Grain flow deposits and their implications, Santa Ynez Mountains, California. J. Sediment. Petrol. 37, 487 508. Steel, E., Simms, A.R., Warrick, J., Yokoyama, Y., 2016. Highstand shelf fans: the role of buoyancy reversal in the deposition of a new type of shelf sand body. GSA Bull. 128, 1717 1724. Steel, E., Simms, A.R., Steel, R., Olariu, C., 2018. Hyperpycnal delivery of sand to the continental shelf: insights from the Jurassic lajas formation, Neuque´n Basin, Argentina. Sedimentology. Available from: https://doi.org/ 10.1111/sed.12460.

Bibliography

537

Steele, T.H., Barrett, T.R., Worthington, L.V., 1962. Deep currents south of Iceland. Deep-Sea Res. 9, 465 474. Steele, J.H., Turekian, K.K., Thorpe, S.A. (Eds.), 2008. Encyclopedia of Ocean Sciences. second ed. Elsevier, ISBN 978-012-374473-9. ,http://www.sciencedirect.com/science/referenceworks/9780123744739. (accessed 25.07.11.). Sterling, G.H., Strohbeck, G.E., 1975. The Failure of South Pass 70B Platform in Hurricane Camille. J. Petrol. Technol. 263 268. Stevens, M.R., Bossong, C.R., Litke, D.W., Viger, R.J., Rupert, M.G., Char, S.J., 2008. Estimated probability of postwildfire debris-flow occurrence and estimated volume of debris flows from a pre-fire analysis in the Three Lakes Watershed, Grand County, CO. U.S. Geological Survey Scientific Investigations Map 3009, 1 sheet. Stewart, R.H., 2008. Introduction to Physical Oceanography. E-book. Texas A&M University, College Station, TX, 353 p. ,http://oceanworld.tamu.edu/resources/ocng_textbook/PDF_files/book.pdf. (accessed 01.10.12.). Sto¨ffler, D., Grieve, R.A.F., 2003. Towards a unified nomenclature of metamorphic petrology. 11. Impactites. A proposal on behalf of the IUGS Subcommission on the Systematics of Metamorphic Rocks. Web version of 30 April 2003. ,http://www.bgs.ac.uk/scmr/docs/paper_12/scmr_paper_12_1.pdf. (accessed 22.02.05.). Stommel, H., 1958. The abyssal circulation. Deep-Sea Res. 5, 80 82. Stone, G.W., Walker, N.D., Hsu, S.A., Babin, A., Liu, B., Keim, B.D., et al., 2005. Hurricane Ivan’s impact along the Northern Gulf Of Mexico. EOS Trans. Am. Geophys. Union 86 (48), 497 508. Stow, D.A.V., 1979. Distinguishing between fine-grained turbidites and contourites off Nova Scotia. Sedimentology 26, 371 387. Stow, D.A.V., 1985. Deep-sea clastics: where are we and where are we going? In: Brenchly, P.J., Williams, P.J. (Eds.), Sedimentology: Recent Developments and Applied Aspects. Published for the Geological Society by Blackwell Scientific Publications, Oxford, pp. 67 94. Stow, D.A.V., Lovell, J.P.B., 1979. Contourites: their recognition in modern and ancient sediments. Earth-Sci. Rev. 14, 251 291. Stow, D.A.V., Shanmugam, G., 1980. Sequence of structures in fine-grained turbidites: comparison of recent deepsea and ancient flysch sediments. Sediment. Geol. 25, 23 42. Stow, D.A.V., Piper, D.J.W. (Eds.), 1984. Fine-Grained Sediments: Deep-Water Processes and Facies, No. 15. Geological Society Special Publication, p. 659. Stow, D.A.V., Fauge`res, J.-C. (Ed.) 1998. Contourites, turbidites and process interaction. Sediment. Geol. 115, pp. 1 384. Stow, D.A.V., Johansson, M., 2000. Deep-water massive sands: nature, origin and hydrocarbon implications. Mar. Petrol. Geol. 17, 145 174. Stow, D.A.V., Fauge`res, J.-C., 2008. Contourite facies and the facies model. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites. Elsevier, Amsterdam, pp. 223 256. Stow, D and Smillie, Z., 2020. Distinguishing between deep-water sediment facies: Turbidites, contourites and hemipelagites. Geosciences 10: 68. https://doi.org/10.3390/geosciences10020068. Stow, D.A.V., Cremer, M., Droz, L., Normark, W.R., O’Connell, S., Pickering, K.T., et al., 1985. Mississippi fan sedimentary facies, composition, and texture. In: Bouma, A.H., Normark, W.R., Barnes, N.E. (Eds.), Submarine Fans and Related Turbidite Systems. Springer-Verlag, New York, pp. 259 266. Stow, D.A.V., Amano, K., Batson, B.S., Brass, G.W., Corrigan, J., Raman, C.V., et al., 1990. 31. Sediment facies and processes on the distal Bengal Fan, Leg 116. In: Cochran, J.R., Stow, D.A.V., et al. (Eds.), Proceedings of the Ocean Drilling Project, Leg 116, Scientific Results, Ocean Drilling Program, College Station, TX, pp. 377 396. Stow, D.A.V., Reading, H.G., Collinson, J., 1996. Deep seas. In: Reading, H.G. (Ed.), Sedimentary Environments: Processes, Facies, and Stratigraphy. Blackwell Science, Oxford, pp. 395 453. Stow, D.A.V., Fauge`res, J.C., Viana, A., Gonthier, E., 1998. Fossil contourites: a critical review. Sediment. Geol. 115, 3 31. Stow, D.A.V., Pudsey, C.J., Howe, J.A., Faugeres, J.-C., Viana, A.R. (Eds.), 2002. Deep-water contourite systems: modern drifts and ancient series, seismic and sedimentary characteristics. Geol. Soc. Lond. Mem. 22, 464. Stow, D.A.V., Hunter, S., Wilkinson, D., Herna´ndez-Molina, F.J., 2008. Chapter 9 The nature of contourite deposition. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 143 156. Developments in Sedimentology. Stow, D.A.V., Herna´ndez-Molina, F.J., Llave, E., Sayago, M., Dı´az del Rı´o, V., Branson, A., 2009. Bedform-velocity matrix: the estimation of bottom current velocity from bedform observations. Geology 37 (4), 327 330.

538

Bibliography

Stow, D.A.V., Brackenridge, R., Hernandez-Molina, F.J., 2011. Contourite sheet sands: new deepwater exploration target. In: Abstract American Association of Petroleum Geologists Annual Conference, Houston, TX, 2011. Stow, D.A.V., Herna´ndez-Molina, F.J., Llave, E., Bruno, M., Garcı´a, M., Dı´az del Rio, V., et al., 2013. The Cadiz Contourite Channel: sandy contourites, bedforms and dynamic current interaction. Mar. Geol. 343, 99 114. Strangways, W.T.H.F., 1821. Description of strata in the Brook Pulcova, near the village of Great Pulcova in the neigbourhood of St. Petersburg. Trans. Geol. Soc. 5, 382 458. First Series. Strasser, M., Moore, G., Kimura, G., Kopf, A., Underwood, M., Guo, J., et al., 2011. Slumping and mass-transport deposition in the Nankai forearc: evidence from IODP drilling and 3-D reflection seismic data. Geochem. Geophys. Geosyst. 12, 1 24. Strickland, H.E., 1840. On some remarkable dykes of Calcareous Grit, at Ethie in Ross-shire. Trans. Geol. Soc. 5, 599 600. Second Series. Stride, A.H. (Ed.), 1982. Offshore Tidal Sands: Proceses and Deposits. Chapman & Hall, London. Stride, A.H., Tucker, M.J., 1960. Internal waves and waves of sand. Nature 188, 933. Su, D.C., Sun, A.P., 2012. Typical earthquake-induced soft sediment deformation structures in the Mesoproterozoic Wumishan Formation, Yongding River Valley, Beijing, China and interpreted earthquake frequency. J. Palaeogeogr. 1, 71 89. Suhayda, J.N., Prior, D.B., 1978. Explanation of submarine landslide morphology by stability analysis and rheological models. In: Offshore Technology Conference, Houston, TX. 8 11 May 1978. ,http://oai.dtic.mil/oai/ oai?verb 5 getRecord&metadataPrefix 5 html&identifier 5 ADA054927. (accessed 27.12.14.). Sultan, N., Cochonat, P., Canals, M., Cattaneo, A., Dennielou, B., Haflidason, H., et al., 2004. Triggering mechanisms of slope instability processes and sediment failures on continental margins: a geotechnical approach. Mar. Geol. 213, 291 321. Sun, F.N., Yang, R.C., Li, D.Y., 2016. Research progresses on hyperpycnal flow deposits. Acta Sedimentol. Sin. 34 (3), 452 462. Sundborg, A., 1956. The River Klara¨lven: a study of fluvial processes. Geografiska Annaler 38, 197. Seies A. Surlyk, F., 1987. Slope and deep shelf gully sandstones, Upper Jurassic, East Greenland. AAPG Bull. 71, 464 475. Surlyk, F., Hurst, J.M., 1984. The evolution of the early Paleozoic deep-water basin of North Greenland. GSA Bull. 95, 131 154. Surlyk, F., Noe-Nygaard, N., 2001. Sand remobilization and intrusion in the Upper Jurrasic Hareelv Formantion, East Greenland. Bull. Geol. Soc. Den. 48, 169 188. Swallow, J.C., Worthington, L.V., 1961. An observation of a deep counter current in the western North Atlantic. Deep-Sea Res. 8, 1 9. Swanson, R.G., 1981. Sample examination manual. AAPG Methods Explor. Ser. IV, 45. Sylvester, Z., Pirmez, C., Cantelli, A., 2011. A model of submarine channel-levee evolution based on channel trajectories: implications for stratigraphic architecture. Mar. Petrol. Geol. 28, 716 727. Syvitski, J.P.M., Farrow, G.E., 1989. Fjord sedimentation as an analogue for small hydrocarbon-bearing fan deltas. In: Whateley, M.K.G., Pickering, K.T. (Eds.), Deltas: Sites and Traps for Fossil Fuels. The Geological Society, London, pp. 21 44. Tada, R., 1991. Compaction and cementation in siliceous rocks and their possible effect on bedding enhancement. In: Einsele, G. (Ed.), Cycles and Events in Stratigraphy. Springer-Verlag, Berlin, Heidelberg, pp. 480 491. Takahashi, T., 1981. Debris flows. Annu. Rev. Fluid Mech. 13, 57 77. Takahasi, R., Hatori, T., 1961. A summary report of the Chilean tsunami of 1960. In: Takahasi, R. (Ed.), Report on the Chilean Tsunami of May 24, 1960, as Observed Along the Coast of Japan. Committee for Field Investigation of the Chilean Tsunami of 1960. Maruzen Co, Tokyo, pp. 23 34. Takayama, H., Tada, R., Matsui, T., Iturralde-Vinent, M.A., Oji, T., Tajika, E., et al., 2000. Origin of the Pen˜alver Formation in northwestern Cuba and its relation to K/T boundary impact event. Sediment. Geol. 135, 295 320. Talley, L.D., 2013. Closure of the global overturning circulation through the Indian, Pacific, and Southern Oceans: schematics and transports. Oceanography 26 (1), 80 97. Talling, P.J., 2013. Hybrid submarine flows comprising turbidity current and cohesive debris flow: Deposits, theoretical and experimental analyses, and generalized models. Geosphere 9 (3), 460 488. Talling, P.J., 2014. On the triggers, resulting flow types and frequencies of subaqueous sediment density flows in different settings. Mar. Geol. 352, 155 182.

Bibliography

539

Taira, A., Hill, I., Firth, J.V., Berner, U., Brueckmann, W., Byrne, T., et al., 1992. Sediment deformation and hydrogeology of the Nankai Trough accretionary prism: Synthesis of shipboard results of ODP Leg 131. Earth Planet. Sci. Lett. 109, 431 450. Talling, P.J., Wynn, R.B., Masson, D.G., Frenz, M., Cronin, B.T., Schiebel, R., et al., 2007. Onset of submarine debris flow deposition far from original giant landslide. Nature 450, 541 544. Talling, P.J., Masson, D.G., Sumner, F.J., Malgesini, G., 2012. Subaqueous sediment density flows, depositional processes and deposit types. Sedimentology 59, 1937 2003. Tappin, D.R., 2010. Submarine mass failures as tsunami sources: their climate control. Philos. Trans. R. Soc. A 368, 2417 2434. Available from: https://doi.org/10.1098/rsta.2010.0079. Tappin, D.R., Watts, P., McMurtry, G.M., Lafoy, Y., Matsumoto, T., 2001. The Sissano, Papua New Guinea tsunami of July 1998 offshore evidence on the source mechanism. Mar. Geol. 175, 1 23. Taylor, B., Huchon, P., Klaus, A., et al., 2000. Initial reports In: Proceedings of Ocean Drilling Program, 180 (Online). ,http://www-odp.tamu.edu/publications/180_IR/VOLUME/CORES/COR_1108.PDF. (accessed 22.07.04.). Teale, T., Young, J.R., 1987. Isolated olistoliths from the Longobucco Basin Calabria, Southern Italy. In: Leggett, J. K., Zuffa Shaver, G.G. (Eds.), Advances in Marine Clastic Sedimentology. Graham & Trotman, pp. 75 88. Terwindt, J.H.J., 1981. Origin and sequences of sedimentary structures in inshore mesotidal deposits of the North Sea. In: Nio, S.-D., Shuttenhelm, R.T.E., Van Weering, Tj.C.E. (Eds.), Holocene Marine Sedimentation in the North Sea Basin, No. 5. International Association of Sedimentologists Special Publication, pp. 4 26. Terzaghi, K., 1936. The shear resistance of saturated soils. In: Proceedings of the First International Conference on Soil Mechanics and Foundation Engineering, Cambridge, MA, pp. 54 56. Terzaghi, K., 1950. Mechanism of landslides. In: Paige, S. (Ed.), Application of Geology to Engineering Practice (Berkey Volume). Geological Society of America, New York, pp. 83 123. Terzaghi, K., Peck, R.B., Mesri, G., 1996. Soil Mechanics in Engineering Practice, third ed. Wiley, Hoboken, NJ, p. 592. The Learning Channel, 1997. Landslides (videotape), Produced for the Learning Channel (a cable television channel in the U.S.) by the BBC Television, London. Thiede, J., 1977. 10. Sedimentary structures in pelagic and hemipelagic sediments from the Central and Southern Atlantic Ocean. Initial Reports of the Deep Sea Drilling Project, Leg 39. United States Government Printing Office, Washington, DC, pp. 407 421, ,https://doi.org/10.2973/dsdp.proc.39.110.1977.. Thiede, J., Myhre, A.M., 1996. 36. The paleoceanographic history of the North Atlantic Arctic Gateways: synthesis of the Leg 151 drilling results. In: Thiede, J., Myhre, A.M., Firth, J.V., Johnson, G.L., Ruddiman, W.F. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results, 151, pp. 645 658. Thomson, R.E., Miha´Ly, S.F., Kulikov, E.A., 2007. Estuarine versus transient flow regimes in Juan de Fuca Strait. J. Geophys. Res. Ocean. 112 (C9), CiteID C09022. Tilling, R.I., Topinka, L., Swanson, D.A., 1990. Eruptions of Mount St. Helens: Past, Present, and Future. U.S. Geological Survey Special Interest Publication, 56 p. Tinterri, R., Muzzi Magalhaes, P., Tagliaferri, A., Cunha, R.S., 2016. Convolute laminations and load structures in turbidites as indicators of flow reflections and decelerations against bounding slopes. Examples from the Marnoso-arenacea Formation (northern Italy) and Annot Sandstones (south eastern France). Sediment. Geol. 344, 382 407. ¨ zkul, M., 2014. Soft-sediment deformation structures interpreted as seismites in the Kolankaya Topal, S., O Formation, Denizli Basin (SW Turkey). Sci. World J. 2014, 352654. ,https://doi.org/10.1155/2014/3526.. To¨ro˝ , B., Pratt, B.R., 2016. Sedimentary record of seismic events in the Eocene Gran River Formation and its implications for regional tectonics on lake evolution (Bridger Basin, Wyoming). Sediment. Geol. 344, 175 204. Trincardi, F., Cattaneo, A., Correggiari, A., Mongardi, S., Breda, A., Asioli, A., 2003. Submarine slides during relative sea level rise: two examples from the eastern Tyrrhenian margin. In: Locat, J., Mienert, J. (Eds.), Submarine Mass Movements and Their Consequences. Kluwer Academic Publishers, Dordrecht, pp. 469 478. Tripsanas, E.K., Bryant, W.R., Phaneuf, B.A., 2004. Slope-instability processes caused by salt movements in a complex deep-water environment, Bryant Canyon area, northwest Gulf of Mexico. AAPG Bull. 88, 801 823. Tripsanas, E.K., Piper, D.J.W., Jenner, K.A., Bryant, W.R., 2008. Submarine mass-transport facies: new perspectives on flow processes from cores on the eastern North Atlantic margin. Sedimentology 55, 97 136.

540

Bibliography

Truswell, J.F., 1972. Sandstone sheets and related intrusions from Coffee Bay, South Africa. J. Sediment. Petrol. 42, 578 583. Tsuji, Y., 1993. Tide influenced high energy environments and rhodolith-associated carbonate deposition on the outer shelf and slope off the Miyako Islands, southern Ryukyu Island Arc, Japan. Mar. Geol. 113, 255 271. Tucholke, B.E., Embley, R.W., 1984. Cenozoic regional erosion of the abyssal sea floor off South Africa. In: Schlee, J.S.M. (Ed.), Interregional Unconformities and Hydrocarbon Accumulation, 36. AAPG Memoir, pp. 145 164. Tuholke, B.E., Wright, W.R., Hollister, C.D., 1973. Abyssal circulation over the Greater Antilles Outer Ridge. Deep-Sea Res. 20, 973 995. Turner, J.S., 1980. Buoyancy Effects in Fluids. Series: Cambridge Monographs on Mechanics. Cambridge University Press, 412 pp. Turner, F.J., Weiss, L.E., 1963. Structural Analysis of Metamorphic Tectonites. McGraw-Hill, New York, p. 545. Twichell, D.C., Roberts, D.G., 1982. Morphology, distribution, and development of submarine canyons in the United States Atlantic continental slope between Hudson and Baltimore Canyons. Geology 10, 408 412. Twichell, D.C., Schwab, W.C., Nelson, C.H., Kenyon, N.H., Lee, H.J., 1992. Characteristics of a sandy depositional lobe on the outer Mississippi Fan from Sea MARC 1Asidescan sonar images. Geology 20, 689 692. Twichell, D.C., Schwab, W.C., Kenyon, N.H., 1995. Geometry of sandy deposits at the distal edge of the Mississippi Fan, Gulf of Mexico. In: Pickering, K.T., Hiscott, R.N., Kenyon, N.H., Ricci Lucchi, F., Smith, R.D. A. (Eds.), Atlas of Deep Water Environments, Architectural Style in Turbidite Systems. Chapman & Hall, London, pp. 282 286. Twichell, D.C., Chaytor, J.D., ten Brink, U.S., Buczkowski, B., 2009. Morphology of late Quaternary submarine landslides along the U.S. Atlantic continental margin. Mar. Geol. 264, 4 15. Uchupi, E., 1975. Physiography of the Gulf of Mexico and Caribbean Sea. In: Nairn, A.E.M., Stehli, F.G. (Eds.), The Ocean Basins and Margins, vol. 3: The Gulf of Mexico and the Caribbean. Plenum Press, New York, pp. 1 64. Underwood, M.B., Bachman, S.G., 1982. Sedimentary facies associations within subduction complexes. In: Leggett, J.K. (Ed.), Trench-Forearc Geology: Sedimentation and Tectonics on Modern and Ancient Active Plate Margins, 10. Geological Society of London Special Publication, pp. 537 550. Unrug, R., 1977. Ancient deep-sea traction currents deposits in the Lgota Beds (Albian) of the Carpathian flysch. Rocznik Polskiego Towarzystwa Geologicznego 47, 355 370. Urgeles, R., Canals, M., Baraza, J., Alonso, B., Masson, D., 1997. The most recent megalandslides of the Canary Islands, El Golfo debris avalanche and Canary debris flow, west El Hierro Island. J. Geophys. Res. 102 (B9), 20305 20323. Urgeles, R., Camerlenghi, A., Ercilla, G., Anselmetti, F., Bru¨ckmann, W., Canals, M., et al., 2007. Scientific ocean drilling behind the assessment of geohazards from submarine slides: Barcelona, Spain, 25 27 October 2006. EOS Trans. Am. Geophys. Union 88 (17), 192. ,https://doi.org/10.1029/2007EO170009. (accessed 27.12.14.). USACE (U.S. Army Core of Engineers), 2003. Slope stability. Department of the Army, Engineers Manual, EM 1110-2-1902, Washington, DC, 31 October 2003, p. 205. USGS (U.S. Geological Survey), 1994. Geologic features of the sea bottom around a municipal sludge dumpsite near 39 N, 73 W, Offshore New Jersey and New York. U.S. Geological Survey Open-file Report 94-152. ,http://pubs.usgs.gov/of/1994/of94-152/description.html. (accessed 06.07.11.). USGS (U.S. Geological Survey), 2004. Landslids types and processes. USGS Fact Sheet 2004-3072. ,http://pubs. usgs.gov/fs/2004/3072/fs-2004-3072.html. (accessed 15.07.10.). USGS (U.S. Geological Survey), 2010. Catastrophic landslids of the 20th century worldwide. ,http://landslides.usgs.gov/learning/majorls.php. (accessed 14.06.10.). USGS (U.S. Geological Survey), 2011. Magnitude 9.0 near the east coast of Honshu. 11 March 2011. ,http:// earthquake.usgs.gov/earthquakes/recenteqsww/Quakes/usc0001xgp.php. (accessed 12.04.11.). USGS (U.S. Geological Survey), 2014. The 1964 Great Alaska Earthquake and Tsunamis—A modern perspective and enduring legacies. Fact Sheet 2014-3018, March 2014. U.S. Department of the Interior, p. 6. USGS (U.S. Geological Survey), 2018. Moving mountains: Elwha River still changing five years after world’s largest dam-removal project: more than 20 million tons of sediment flushed to the sea. USGS News. 5 September 2018. ,https://www.usgs.gov/news/moving-mountains-elwha-river-still-changing-five-years-after-world-slargest-dam-removal. (accessed 03.03.19.). Vail, P.R., 1987. Seismic-stratigraphy interpretation using sequence stratigraphy. Pt. 1: seismic stratigraphyinterpretation procedures. In: Bally, A.W. (Ed.), Atlas of Seismic Stratigraphy, vol. 1. AAPG Studies in Geology 27, pp. 1 10.

Bibliography

541

Vail, P.R., Mitchum Jr., R.M., Thompson III, S., 1977a. Seismic stratigraphy and global changes of sea level: Part 3. Relative changes of sea level from coastal onlap. In: Payton, C.E. (Ed.), Eismic Stratigraphy-Applications to Hydrocarbon Exploration, 26. AAPG Memoir, pp. 63 81. Vail, P.R., Todd, R.G., Sangree, J.B., 1977b. Seismic stratigraphy and global changes of sea level: Part 5. Chronostratigraphic significance of seismic reflections. In: Payton, C.E. (Ed.), Seismic StratigraphyApplications to Hydrocarbon Exploration, 26. AAPG Memoir, pp. 99 116. Vail, P.R., Wornardt, W.W., 1990. Well log-seismic sequence stratigraphy: an integrated tool for the 90’s. In: Sequence Stratigraphy as an Exploration Tool, Concepts and Practices in the Gulf Coast, 11th Annual Research Conference, Gulf Coast SEPM Foundation, pp. 379 388. Vail, P.R., Audemard, F., Bowman, S.A., Eisner, P.N., Perez-Cruz, C., 1991. The stratigraphic signatures of tectonics, eustacy and sedimentology an overview. In: Einsele, G., Ricken, W., Seilacher, A. (Eds.), Cycles and Events in Stratigraphy. Springer-Verlag, Berlin, pp. 618 659. Vallance, J.W., Scott, K.M., 1997. The Oseola mudflow from Mount Rainier: sedimentology and hazard implications of a huge clay-rich debris flow. GSA Bull. 109, 143 163. Van der Lingen, G.J., 1969. The turbidite problem. N. Z. J. Geol. Geophys. 12, 7 50. Van Loon, A. J. (Tom), Hu¨eneke, H., Mulder, T., 2019. The hyperpycnite problem: comment. J. Palaeogeogr. 8 (3), 314 320. Van Loon, A.J., Mazumder, R., De, S., 2017. The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India: reply. J. Palaeogeog 6 (3), 235 241. Van Loon, A.J., Mazumder, R., Shuvabrata, D., 2016a. The response of stromatolites to seismic shocks: tomboliths from the Palaeoproterozoic Chaibasa Formation, E India. J. Palaeogeogr. 5 (4), 381 390. van Maren, D.S., Winterwerp, J.C., Wu, B.S., Zhou, J.J., 2009. Modelling hyperconcentrated flow in the Yellow River. Earth Surf. Process. Landf. 34 (4), 596 612. Van Rooij, D., 2013. Deep-water bottom current dynamics: processes, products & challenges. Marine and River Dune Dynamics, MARID IV, 15 16 April 2013, Bruges, pp. 297 300. Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M., Vail, P.R., Sarg, J.F., Loutit, T.S., et al., 1988. An overview of the fundamentals of sequence stratigraphy and key definitions. In: Wilgus, C.K., Hastings, B.S., Kendall, C. G. St. C., Posamentier, H.W., Ross, C.A., Van Wagoner, J.C. (Eds.), Sea-Level Changes: An Integrated Approach, No. 42. SEPM Special Publication, pp. 39 45. Vanneste, M., De Batist, M., Golmshtok, A., Kremlev, A., Versteeg, W., 2001. Multi-frequency seismic study of gas hydrate-bearing sediments in Lake Baikal, Siberia. Mar. Geol. 172, 1 21. Vanney, J.-R., Stanley, D.J., 1983. Shelfbreak physiography: an overview. In: Stanley, D.J., Moore, G.T. (Eds.), The Shelfbreak: Critical Interface on Continental Margins, No. 33. SEPM Special Publication, pp. 1 24. ˇ Vargas-Yanez, M., Viola, T.S., Jorge, F.P., Rubin, J.P., Garc!ıa-Martı´nez, M.C., 2002. The influence of tidetopography interaction on low-frequency heat and nutrient fluxes. Application to Cape Trafalgar. Cont. Shelf Res. 22, 115 139. Varnes, D.J., 1958. Landslide types and processes. In: Eckel, E.D. (Ed.), Landslide and Engineering Practice, 29. Highway Research Board Special Report, pp. 20 47. Varnes, D.J., 1978. Slope movement types and processes. In: Schuster, R.L., Krizek, R.J. (Eds.), Landslides: Analysis and Control. National Academy of Science, Washington, DC, pp. 11 33. Special Report 176. Varnes, D.J., IAEG (International Association for Engineering Geology), Commission on Landslides and other MassMovements, 1984. Landslide Hazard Zonation: A Review of Principles and Practice. UNESCO Press, Paris, 63 p. Velasco-Villareal, M., Urrutia-Fucugauchi, J., Rebolledo-Vieyra, M., Perez-Cruz, L., 2011. Paleomagnetism of impact breccias from the Chicxulub crater—Implications for ejecta emplacement and hydrothermal processes. Phys. Earth Planet. Inter. 186, 154 171. Venzke, E., Wunderman, R.W., McClelland, L., Simkin, T., Luhr, J.F., Siebert, L., et al., (Eds.), 2002. Global Volcanism, 1968 to the Present. Smithsonian Institution, Global Volcanism Program Digital Information Series, GVP-4 ,http://www.volcano.si.edu/reports/.. Verdicchio, G., Trincardi, F., 2008. Shallow-water contourites. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites. Developments in Sedimentology, vol. 60, pp. 409 433. Verdicchio, G., Trincardi, F., Asioli, A., 2007. Mediterranean bottom-current deposits: an example from the Southwestern Adriatic Margin. In: Viana, A.R., Rebesco, M. (Eds.), Economic and Palaeoceanographic Significance of Contourite Deposits, vol. 276. Geological Society of London, Special Publication, pp. 199 224.

542

Bibliography

Viana, A.R., 2008. Economic relevance of contourites. Chapter 23. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 493 510. Developments in Sedimentology. Viana, A.R., Rebesco, M. (Eds.), 2007. Economic and Palaeoceanographic Significance of Contourite Deposits, 276. Geological Society Special Publication, London, 350 p. Viana, A.R., Faugeres, J.C., Kowsmann, R.O., Lima, J.A.M., Caddah, L.F.G., Rizzo, J.G., 1998. Hydrology, morphology, and sedimentology of the Campos continental margin, offshore Brazil, Sediment. Geol., 115. pp. 133 157. Viana, A.R., Almeida Jr., W., Nunes, M.C.V., Bulho¨es, E.M., 2007. The economic importance of contourites. In: Viana, A.R., Rebesco, M. (Eds.), Economic and Palaeoceanographic Significance of Contourite Deposits, vol. 276. Geological Society of London Special Publication, pp. 1 23. Visser, M.J., 1980. Neap-spring cycles reflected in Holocene subtidal large-scale bedform deposits: a preliminary note. Geology 8, 543 546. Vivas, V., Macsotay, O., Furrer, M., Alvarez, E., 1988. Injectitas cla´sticas asociadas a desplomes en sedimentitas batiales del Creta´ceo Superior de Venezuela nor-oriental. Boletı´n de. la. Soc. Venezolana de Geologı´a 34, 3 33. Voight, B., Faust, C., 1982. Frictional heat and strength loss in some rapid landslides. Geotechnique 32 (1), 43 54. Voight, B., Janda, R.J., Glicken, H., Douglas, P.M., 1983. Nature and mechanics of the Mount St. Helens rockslideavalanche of 18 May, 1980. Geotechnique 33, 243 273. Volkman, C., 1962. Deep current observations in the western North Atlantic. Deep-Sea Res. 9, 493 500. Vrolijk, P.J., Southard, J.B., 1997. Experiments on rapid deposition of sand from high-velocity flows. Geosci. Can. 24, 45 54. Vsemirnova, E., Hobbs, R., Serra, N., Klaeschen, D., Quentel, E., 2009. Estimating internal wave spectra using constrained models of the dynamic ocean. Geophys. Res. Lett. 36, L00D07. Available from: https://doi.org/ 10.1029/2009GL039598. Wagreich, M., 2003. A slope-apron succession filling a piggyback basin: the Tannheim and Losenstein Formations (Aptian - Cenomanian) of the eastern part of the Northern Calcareous Alps (Austria). Mitt. Oster. Geol. Ges. 93, 31 54. Wajsowicz, R.C., Gordon, A.L., Field, A., Susanto, R.D., 2003. Estimating transport in Makassar Strait. Deep-Sea Res. II 50, 2163 2181. Waldron, J.W.F., Gagnon, J.F., 2011. Recognizing soft-sediment structures in deformed rocks of orogens. J. Struct. Geol. 33, 271 279. Wallis, G.B., 1969. One-Dimensional Two-Phase Flow. McGraw-Hill, New York, p. 408. Walker, R.G., 1965. The origin and significance of the internal sedimentary structures of turbidites. Proc. Yorks. Geol. Soc. 35, 1 32. Walker, R.G., 1973. Mopping up the turbidite mess. In: Ginsburg, R.N. (Ed.), Evolving Concepts in Sedimentology. The Johns Hopkins University Press, Baltimore, MD, pp. 1 37. Walker, R.G., 1975. Nested submarine-fan channels in the capistrano formation, San Clemente, California. GSA Bull. 86 (7), 915 924 (1975). Walker, R.G., 1992. Facies, facies models, and modern stratigraphic concepts. In: Walker, R.G., James, N.P. (Eds.), Facies Models: Response to Sea Level Change, GEOtext 1. Geological Association of Canada, pp. 1 14. Walker, K.R., Shanmugam, G., Ruppel, S.C., 1984. A model for carbonate to terrigenous clastic sequences: discussion and reply. GSA Bull. 95, 753. Walker, J.R., Massingill, J.V., 1970. Slump features on the Mississippi Fan, northern Gulf of Mexico. Geol. Soc. Am. Bull. 81, 3101 3108. Walker, N.D., Rouse Jr., L.J., 1993. Satellite Assessment of Mississippi River Discharge Plume Variability. OCS Study MMS 93 0044. U.S. Dept. of the Interior, Minerals Management Service, Gulf of Mexico OCS Regional Office, New Orleans, LA, 50 p. Wang, X.H., Wang, H.J., 2010. Tidal straining effect on the suspended sediment transport in the Huanghe (Yellow River) Estuary, China. Ocean. Dyn. 60, 1273 1283. Wang, H.J., Yang, Z.S., Saito, Y., Liu, J.P., Sun, X.X., 2006. Interannual and seasonal variation of the Huanghe (Yellow River) water discharge over the past 50 years: connections to impacts from ENSO events and dams. Glob. Planet. Change 50 (3), 212 225. Wang, H.J., Yang, Z.S., Li, Y.H., Guo, Z.G., Sun, X.X., Wang, Y., 2007. Dispersion pattern of suspended sediment in the shear frontal zone off the Huanghe (Yellow River) mouth. Cont. Shelf Res. 27 (6), 854 871.

Bibliography

543

Wang, Q., Guo, X.Y., Takeoka, H., 2008. Seasonal variations of the Yellow River plume in the Bohai Sea: a model study. J. Geophys. Res. Ocean. 113, C08046. Wang, H., Bi, N., Wang, Y., Saito, Y., Yang, Z., 2010. Tide-modulated hyperpycnal flows off the Huanghe (Yellow River) Mouth, China. Earth Surf. Process. Landf. 35 (11), 1315 1329. Wang, Y., Wang, Y., Xu, Q., Li, D., Zhuo, H., Zhou, W., 2012. The early-middle Miocene submarine fan system in the Pearl River Mouth Basin, South China Sea. Petrol. Sci. 9, 1 9. Available from: https://doi.org/10.1007/ s12182-012-0175-z. Wang, J.D., Li, S.Z., Santosh, M., Dai, L.M., Suo, Y.H., Yu, S., et al., 2013. Lacustrine turbidites in the Eocene Shahejie Formation, Dongying Sag, Bohai Bay Basin, North China Craton. Geol. J. Available from: https://doi. org/10.1002/gj.2517. Wang, N., Li, G.X., Xu, J.S., Qiao, L.L., Dada, O.A., Zhou, C.Y., 2015. The marine dynamics and changing trend off the modern Yellow River mouth. J. Ocean. Univ. China 14 (3), 433 445. Wang, N., Li, G.X., Qiao, L.L., Shi, J.H., Dong, P., Xu, J.S., et al., 2017. Long-term evolution in the location, propagation, and magnitude of the tidal shear front off the Yellow River mouth. Cont. Shelf Res. 137, 1 12. Warme, J.E., Slater, R.A., Cooper, R.A., 1978. Bioerosion in submarine canyons. In: Stanley, D.J., Kelling, G.K. (Eds.), Sedimentation in Submarine Canyons, Fans, and Trenches. Dowden, Hutchinson & Ross, Inc, Stroudsburg, PA, pp. 65 70. Warne, A.G., Meade, R.H., White, W.A., Aslan, A., Guevara, E.H., Gibeaut, J.C., et al., 2002. Regional controls on geomorphology, hydrology, and ecosystem integrity of the Orinoco Delta, Venezuela. Geomorphology 44, 273 307. Warrick, J.A., Stevens, A.W., Miller, J.M., Gelfenbaum, G., 2011. Coastal Processes of the Elwha River Delta. In: Duda, J.J., Warrick, J.A., Magirl, C.S. (Eds.), Coastal Habitats of the Elwha River: Biological and Physical Patterns and Processes Prior to Dam Removal. Chapter: 1. U.S. Geological Survey Scientific Investigations Report 2011-5120, Chapter 5. Warrick, J.A., Simms, A.R., Ritchie, A., Steel, E., Dartnell, P., Conrad, J.E., et al., 2013. Hyperpycnal plumederived fans in the Santa Barbara Channel, California. Geophys. Res. Lett. 40, 2081 2086. Weathley, D.F., Chan, M.A., Sprinkel, D.A., 2016. Clastic pipe characteristics and distributions throughout the Colorado Plateau: implications for paleoenvironment and paleoseismic controls. Sediment. Geol. 344, 20 33. Weaver, P.P.E., Thomson, J., Hunter, P.M., 1987. Introduction. In: Weaver, P.P.E., Thomson, J. (Eds.), Geology and Geochemistry of Abyssal Plains:, 31. Geological Society of London Special Publication, pp. vii xii. Weber, M.E., Wiedicke, M.H., Kudrass, H.R., Huebscher, C., Erlenkeuser, H., 1997. Active growth of the Bengal Fan during sea-level rise and highstand. Geology 25, 315 318. Weidinger, J.T., Korup, O., 2009. Frictionite as evidence for a large Late Quaternary rockslide near Kanchenjunga, Sikkim Himalayas, India-Implications for extreme events in mountain relief destruction. Geomorphology 103, 57 65. Welbon, A.I.F., Brockbank, P.J., Brunsden, D., Olsen, T.S., 2007. Characterizing and producing from reservoirs in landslides: challenges and opportunities. In: Jolley, S.J., Barr, D., Walsh, J.J., Knipe, R.J. (Eds.), Structurally Complex Reservoirs, 292. Geological Society Special Publication, London, pp. 49 74. Wentworth, C.K., 1922. A scale of grade and class terms for clastic sediments. J. Geol. 30, 377 392. Wessels, M., Bussmann, I., Schloemer, S., Schlu¨ter, M., Bo¨der, V., 2010. Distribution, morphology, and formation of pockmarks in Lake Constance, Germany. Limnol. Oceanogr. 55 (6), 2623 2633. Available from: https://doi. org/10.4319/lo.2010.55.6.2623. Wetzel, A., Werner, F., Stow, D.A.V., 2008. Chapter 11. Bioturbation and biogenic sedimentary structures in contourites. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites, 60. Elsevier, Amsterdam, pp. 183 202. Developments in Sedimentology. Wheeler, A.J., Stadnitskaia, A., 2011. Chapter 6. Benthic Deep Sea carbonates: reefs and seeps. In: Hueneke, H., Mulder, T. (Eds.), Deep-Sea Sediments, 63. Elsevier, Amsterdam, pp. 397 455. Developments in Sedimentology. Whitaker, Mc.D., J.H. (Ed.), 1976. Submarine Canyons and Deep-Sea Fans. Dowden, Hutchinson and Ross, Inc., Stroudsburg, PA. Wikipedia, the free Encyclopedia, 2011. ,http://en.wikipedia.org/wiki/Main_Page. (accessed 05.08.11.). Wilson, R.D., Schieber, J., 2014. Hyperpycnites in the Lower Genesee Group of Central New York, USA: implications for mud transport in epicontinental seas. J. Sediment. Res. 84, 866 874. Wilson, R.D., Schieber, J., 2017. Sediment transport processes and lateral facies gradients across a muddy shelf: examples from the Geneseo Formation of central New York, United States. AAPG Bull. 101, 423 431.

544

Bibliography

Winn Jr., R.D., Dott Jr., R.H., 1979. Deep-water fan-channel conglomerates of Late Cretaceous age, southern Chile. Edimentology 26, 203 228. Winsemann, N.J., Asprion, U., Meyer, T., Schramm, C., 2007. Facies characteristics of Middle Pleistocene (Saalian) icemargin subaqueous fan and delta deposits, glacial Lake Leine, NW Germany. Sediment. Geol. 193, 105 129. Wiseman Jr, W.J., Fan, Y.B., Bornhold, B.D., Keller, G.H., Su, Z.Q., Prior, D.B., et al., 1986. Suspended sediment advection by tidal currents off the Huanghe (Yellow River) delta. Geo-Mar. Lett. 6 (2), 107 113. Wizevich, M.C., Simpson, E.L., Hilbert-Wolf, H.L., Tindall, S.E., 2016. Development of large-scale seismites in Upper Cretaceous fluvial sandstones in a fault-proximal setting. Sedimentology 63 (6), 1719 1738. Woodbury, H.O., Spotts, J.H., Akers, W.H., 1978. Gulf of Mexico continental slope sediments and sedimentation. In: Bouma, A.H., Moore, G.T., Coleman, J.M. (Eds.), Framework, Facies, and Oikl-Trapping Characteristics of the Upper Continental Margin, 7. AAPG Studies in Geology, pp. 117 137. Woodcock, N.H., 1976. Structural style in slump sheets: Ludlow Series, Powys, Wales. J. Geol. Soc. Lond. 132, 399 415. Woodcock, N.H., 1979. Sizes of submarine slides and their significance. J. Struct. Geol. 1, 137 142. Woodcock, N.H., Miller, A.V.M., Woodhouse, C.D., 2014. Chaotic breccia zones on the Pembroke Peninsula, south Wales: evidence for collapse into voids along dilational faults. J. Struct. Geol. 69, 91 107. Worthington, L.V., Volkman, G.H., 1965. The volume transport of the Norwegian Sea overflow in the North Atlantic. Deep-Sea Res. 12, 667 676. Wren, P.A., Leonard, L.A., 2005. Sediment transport on the mid-continental shelf in Onslow Bay, North Carolina during Hurricane Isabel. Estuarine Coast. Shelf Sci. 63, 43 56. Wright, L.D., Nittrouer, C.A., 1995. Dispersal of river sediments in coastal seas: six contrasting cases. Estuaries 18 (3), 494 508. Wright, S.G., Rathje, E.M., 2003. Triggering mechanisms of slope instability and their relationship to earthquakes and tsunamis. Pure Appl. Geophys. 160, 1865 1877. Wright, L.D., Yang, Z.S., Bornhold, B.D., Keller, G.H., Prior, D.B., Wiseman Jr., W.J., 1986. Hyperpycnal plumes and plume fronts over the Huanghe (Yellow River) delta front. Geo-Mar. Lett. 6, 97 105. Wright, L.D., Wiseman, W.J., Bornhold, B.D., Prior, D.B., Suhayda, J.N., Keller, G.H., et al., 1988. Marine dispersal and deposition of Yellow River silts by gravity-driven under flows. Nature 332, 629 632. Wright, L.D., Wiseman, W.J., Yang, Z.S., Bornhold, B.D., Keller, G.H., Prior, D.B., et al., 1990. Processes of marine dispersal and deposition of suspended silts off the modern mouth of the Huanghe (Yellow River). Cont. Shelf Res. 10, 1 40. Wright, L.D., Friedrichs, C., Kim, S., Scully, M., 2001. Effects of ambient currents and waves on gravity-driven sediment transport on continental shelves. Mar. Geol. 175, 25 45. Wu, J., Ren, J., Liu, H., Qiu, C., Cui, Y., Zhang, Q., 2016. Trapping and escaping processes of Yangtze Riverderived sediments to the East China Sea. In: Clift, P.D., Harff, J., Wu, J., Qui, Y. (Eds.), River-Dominated Shelf Sediments of East Asian Seas, 429. Geological Society, London, pp. 153 169. Special Publications. Wunsch, C.I., 1969. Progressive internal waves on slopes. J. Fluid Mech. 35, 131 144. Wunsch, C.I., 1996. The Ocean Circulation Inverse Problem. Cambridge University Press, Cambridge, 442 pp. Wunsch, C., 2002. What is the thermohaline circulation? Science 298 (5596), 1179 1181. Available from: https:// doi.org/10.1126/science.1079329. PMID 12424356. Wunsch, C., Gill, A.E., 1976. Observations of equatorially trapped waves in Pacific sea level variations. Deep-Sea Res. Oceanogr. Abstr. 23 (5), 371 390. Wunsch, C., Webb, S., 1979. Climatology of deep ocean internal waves. J. Phys. Oceanogr. 9, 235 243. Wu¨st, G., 1933. Schichtung und Zirkulation des Atlantischen Ozeans. Das Bodenwasser und die Gliederung der Atlantischen Tiefsee. Wiss. Erg. Dt. Atl. Exp. “Mete” (1925e1927) 6(1), 106 pp. In: Olson, B.E. (Ed.), Bottom Water and the Distribution of the Deep Water of the Atlantic, Slessers, M. (translator). US Naval Oceanographic Office, Washington, DC, 145 p. Wust, G., 1950. Block diagramme der Atlantischhen Zirkulation auf Grand der Meteor Ergibribe. Kider Meeresforsch. 7, 24 34. Wust, G., 1963. On the stratification and circulation in the cold water sphere of the Antillean-Caribbean basins. Deep-Sea Res. 10, 165 167. Wynn, R.B., Stow, D.A.V., 2002. Classification and characterisation of deep-water sediment waves. Mar. Geol. 192, 7 22. Wynn, R.B., Masson, D.G., Stow, D.A.V., Weaver, P.P., 2000. The Northwest African slope apron: a modern analogue fordeep-water systems with complex seafloor topography. Mar. Petrol. Geol. 17, 253 265.

Bibliography

545

Xu, J.P., Noble, M.A., Rosenfeld, L.K., 2004. In-situ measurements of velocity structure within turbidity currents. Geophys.l Res. Lett. 31, L09311. https://doi.org/10.1029/2004GL019718, 2004. Xu, J.P., Noble, M.A., Eittreim, S.L., Rosenfeld, L.K., Schwing, F.B., Pilskaln, C.H., 2002. Distribution and transport of suspended particulate matter in Monterey Canyon, California. Mar. Geol. 181, 215 234. Yang, T., Cao, Y., Wang, Y., 2017b. A new discovery of The Early Cretaceous supercritical hyperpycnal flow deposits on Lingshan Island, East China. Acta Geol. Sin. (Engl. Ed.) 91 (2), 749 750. Yao, Z.G., He, R.Y., Ban, X.W., Wu, D.X., 2012. M2 tidal dynamics in Bohai and Yellow Seas: a hybrid data assimilative modeling study. Ocean. Dyn. 62, 753 769. Yang, R., Jin, Z., van Loon, A.J., Han, Z., Fan, A., 2017a. Climatic and tectonic controls of lacustrine hyperpycnite origination in the Late Triassic Ordos Basin, central China: implications for unconventional petroleum development. AAPG Bulletin 101 (1), 95 117. Yamada, Y., Kawamura, K., Ikehara, K., Ogawa, Y., Urgeles, R., Mosher, D., et al., 2012. Submarine Mass Movements and their Consequences 5th International Symposium. In: Advances in Natural and Technological Hazards Research, vol. 31. Springer, 756 p. Yamazaki, T., Yamaoka, M., Shiki, T., 1989. Miocene offshore tractive current-worked conglomeratesd Tsubutegaura, Chita Peninsula, central Japan. In: Taira, A., Masuda, F. (Eds.), Sedimentary Features in the Active Plate Margin. Terra Scientific, Tokyo, Japan, pp. 483 494. Yu, J., Fu, Y., Li, Y., Han, G., Wang, Y., Zhou, D., et al., 2011. Effects of water discharge and sediment load on evolution of modern Yellow River Delta, China, over the period from 1976 to 2009. Biogeosciences 8 (9), 2427 2435. Yu, X., Wang, J., Liang, J., Li, S., Zeng, X., Li, W., 2014. Depositional characteristics and accumulation model of gas hydrates in northern South China Sea. Mar. Petrol. Geol. 56, 74 86. Zavala, C., 2019. The new knowledge is written on sedimentary rocks a comment on Shanmugam’s paper “The hyperpycnite problem”. J. Palaeogeogr. 8 (3), 306 313. Zavala, C., Arcuri, M., 2016. Intrabasinal and extrabasinal turbidites: origin and distinctive characteristics. Sediment. Geol. 337, 36 54. Zavala, C., Pan, S.X., 2018. Hyperpycnal flows and hyperpycnites: origin and distinctive characteristics. Lithol. Reserv. 30 (1), 1 27. Zavala, C., Ponce, J., Drittanti, D., Arcuri, M., Freije, H., Asensio, M., 2006. Ancient lacustrine hyperpycnites: a depositional model from a case study in the rayoso formation (cretaceous) of west-central Argentina. J. Sediment. Res. 76, 41 59. Zenk, W., 1975. On the Mediterranean outflow west of Gibraltar. Meteor. Forscir Ergebuisse (16), 23 34. Zenk, W.O., 1981. Detection of overflow events in the Shag Rocks Passage, Scotia Ridge. Science 213, 1113 1114. Zenk, W., 2008. Abyssal and contour currents. In: Rebesco, M., Camerlenghi, A. (Eds.), Contourites. Elsevier, Amsterdam, pp. 37 57. Zenk, W., Armi, L., 1990. The complex spreading patterns of Mediterranean water off the Portuguese continental slope. Deep-Sea Res. 37, 1805 1823. Zhang, E.F., Savenije, H.H.G., Chen, S.L., Mao, X.H., 2012. An analytical solution for tidal propagation in the Yangtze Estuary, China. Hydrol. Earth Syst. Sci. 16 (9), 3327 3339. Zhang, W., Feng, H., Hoitink, A.J.F., Zhu, Y., Gong, F., Zheng, J., 2017. Tidal impacts on the subtidal flow division at the main bifurcation in the Yangtze River Delta. Estuarine Coast. Shelf Sci. 196, 301 314. Zhao, R., Zhang, Q., Tjugito, H., Cheng, X., 2015. Granular impact cratering by liquid drops: understanding raindrop imprints through an analogy to asteroid strikes. Proc. Natl Acad. Sci. U.S.A. 112 (2), 342 347. Zimmerman, H.B., 1971. Bottom currents on the New England continental rise. J. Geophys. Res. 76, 5865 5876. Zou, C., Wang, L., Li, Y., Tao, S., Hou, L., 2012. Deep-lacustrine transformation of sandy debrites into turbidites, upper Triassic, central China. Sediment. Geol. 265 266, 143 155. Zuffa, G.G. (Ed.), 1985. Provenance of Arenites. D. Reidel Publishing Company, Dordrecht, 408 p.

Author Index Note: Page numbers followed by “f” and “t” refer to figures and tables, respectively.

A

B

´ balos, B., 378t A Abbott P.L., 342 343 Ablay, G., 81 Acha, E., 256 Ackermann, R.V., 378t Agnon, A., 378t Aharonov, E., 79 Ajmal Khan, S., 183t, 219t Akers, C., 79 80 Akhurst, M.A., 326t Aksu, A.E., 28t Alexander, J., 93 95 Alfaro, P., 378t Alhilali, K.A., 264t Allen, J.R.L., 2, 93 95, 128, 133 135, 135f, 247 248, 333, 377 396, 401, 456, 459 461, 472 Allen, S.E., 335 Allison, M.A., 290t Almagor, G., 40t Alsop, G.I., 14 22, 70, 131 Alvarado, A., 289 Alvarado-Bustos, R., 339 340 Alves, J., 338 339 Ambar, I., 337 338, 343 344 Amos, A.F., 326t Anderson, J.B., 314t Anderson, J.G., 40t Anderson, R.Y., 378t Anderson, S.A., 36 38 Andreetta, R., 178 Andresen, A., 40t Apel, J.R., 310, 328 330, 337, 339 340, 364 365 Archer, A.W., 359, 378t Arcuri, M., 200, 216 217, 225, 228, 236, 239, 245 248, 252, 259 Aref, M.A.M., 378t Armi, L., 337 339 Arnau, P., 183t, 194 195, 216 217, 219t Arnott, R.W.C., 138 Arrhenius, G., 429t Arribas, J., 369 370 Assireu, A.T., 224

Baas, J.H., 93 95 Bachman, S.G., 322 Bachmann, G.H., 378t Ba´denas, B., 314t Bagnold, R.A., 14 22, 40t, 79 80, 92 95, 97, 122, 127f, 131, 140, 226 227, 243 Balasubramanian, T., 183t, 219t Baldwin, R., 356 357 Banerjee, I., 164 165, 359 Baringer, M.O., 338, 340 Barnard, P.L., 219t, 239, 260 261 Barrett, J.R., 326t Barton, R., 275t Basilone, L., 378t Bates, C.C., 92, 140 143, 143f, 182, 214, 216, 225, 269, 271, 282, 443 444, 455 456 Bates, R.L., 39, 40t, 48 49 Battjes, J.A., 239 240, 262 Bea, R.G., 275t, 289 Beard, D.C., 461 Beaulieu, S., 356 357 Beckers, A., 183t, 198 Beicher, R.J., 144 Bein, A., 314t Belderson, R.H., 339, 346 Benedict, G.L., 30t, 101f, 245, 314t, 345 346 Berggren, W.A., 302, 336 Berra, F., 378t Betzler, C., 90 92, 310, 370 Beverage, J.P., 95 Bhattacharya, J., 216 217 Biscaye, P.E., 290t Bjerrum, L., 40t Bobrowsky, P., 40t, 45 Boggs Jr., S., 401 Bolt, B.A., 28t Booth, J.S., 37 Bornhold, B.D., 28t, 40t, 52t Bouma, A.H., 2, 92 95, 97, 127f, 136 140, 138f, 143f, 149 150, 161 165, 244, 246 247, 252, 264t, 310, 324 325, 341 343, 377 395, 397t, 442 443, 450 452, 459 461

547

548 Bourgeois, J., 277 Boyd, R., 245 246, 250, 275t, 292, 314t Boyer, S.E., 63 68 Brabb, E.E., 49 50 Brackenridge, R.E., 341, 348, 372 Breien, H., 90, 245 Brennecke, W., 319 320 Briggs, G., 169 170 Brink, K.H., 182 186 Brisco, C.D., 344 345 Broecker, W.S., 240, 261, 313, 328 330 Bro¨nnimann, C.S., 275t, 298 Brown, O.B., 225, 256 Brunner, C.A., 290t Bruno, M., 339 Brunsden, D., 79 Brush Jr., L.M., 50, 467 Bryant, G., 378t Bryden, H.L., 314t, 326t Buatois, L.A., 243 Buckee, C., 93 95 Bugge, T., 28t, 52t, 80 Bulfinch, D.L., 323 324, 326t Burden, C.A., 290t

Author Index

Clayton, C.A., 30t, 314t Cleary, P.W., 79 80 Cline, L.M., 169 170 Cloud Jr., P.E., 261 Coates, D.R., 40t, 45 Cochonat, P., 80 Coleman, J.M., 52t, 80, 244, 275t, 301 302 Coleman Jr., J.L., 171, 264t Collins, D.S., 216 217 Collins, G.S., 78, 81 Collinson, J.D., 50, 377 395, 401 402, 459 461, 471 Collot, J.-Y., 48, 52t, 275t Cook, H.E., 50 Cooper, C., 206 207, 326t, 349 352 Cossey, S.P.J., 378t Costa, J.E., 40t, 92 93, 95, 96f, 98f, 102 105, 225 Costard, F., 81 82 Covault, J.A., 306 307 Cowan, D.S., 397t Cowan, E.A., 245 246 Criado-Aldeanueva, F., 338, 343 344 Cruden, D.M., 48 49, 79 Culbertson, J.K., 95

D C Cacchione, D.A., 290t Cairns, J.L., 337, 339 Camerlenghi, A., 2, 40t, 45, 244 245, 302, 310, 311t Campbell, C.S., 79 80 Cannon, G.A., 183t, 189 Cannon, S.H., 275t, 298 Carlson, P.R., 28t, 251, 465f, 466f, 467f, 468t, 473 Carminatti, M., 324 325, 344 345, 370 Carter, R.M., 40t, 290t, 314t, 342 343 Cartigny, M.J.B., 177 179 Catuneanu, O., 171 Charlet, F., 452 Chen, J., 378t Chen, J.T., 378t Chen, J.Y., 239 Chen, S., 183t, 200 201, 245 246 Chen, S.N., 216 217 Chen, Z., 245 246 Che´rubin, L., 337 Chiarella, D., 378t Chiggiato, J., 314t, 326t Chillarige, A.V., 277 Chu, J.-H., 291t Cita, M.B., 40t, 281 282, 429t Claeys, P., 275t, 277 Clare, M.A., 243, 274 Clark, J.D., 162

Dabrio, C.J., 378t Dade, W.B., 81 D’Agostino, A.E., 102 105, 171, 264t Dallimore, C.J., 224 225 Dalrymple, R.W., 231, 239, 342 343, 359 Damuth, J.E., 14 22, 28t, 150, 152 153, 275t, 326t, 408 413 Dan, G., 275t, 298 Das, B., 116f, 118t, 120f, 121f, 122f, 123f, 124t, 125t Dasgupta, P., 93 95 Davies, G.R., 429t Davies, T.R.H., 96f De Blasio, F.V., 79 80 de Castro, S., 2, 310, 348 de Jong, M.P.C., 239 240, 262 Denamiel, C., 199 Dengler, A.T., 290t Dierssen, H.M., 261 Dill, R.F., 11 12, 14 22, 40t, 100f, 133f, 250 251, 302 304, 356 357, 468t Dillon, W.P., 275t, 302 304 Dingle, R.V., 28t, 50, 52t Dixit, J.G., 37 38 Doerr, S.H., 298 Dott Jr., R.H., 14 22, 40t, 45 48, 50, 81, 92 95, 96f, 123 126, 133 135, 226 228, 245, 248 Douillet, G.A., 378t Dowdeswell, J.A., 300

Author Index

Doyle, L.J., 2 Driscoll, N.W., 28t, 52t, 304 Duarte, C.S.L., 314t, 322, 336 Duncan, J.M., 35 37, 274 276, 449, 453, 472 473, 475 Dunham, J., 50, 314t, 365 366, 370 Durrieu de Madron, X., 335 Dykstra, M., 347 Dysthe, K., 275t, 288 Dzulynski, S., 40t, 377 395, 429t

E Easterbrook, D.J., 50 52 Eberli, G., 90 92, 310, 370 Eckel, E.B., 40t El Taki, H., 378t Elliott, D., 63 68 Elorza, J., 378t Elverhoi, A., 79 Embley, R.W., 9, 14 22, 28t, 52t, 113 114, 302, 304, 336, 408 413 Emery, K.O., 156 158, 468t Enos, P., 50, 96f, 97f, 245, 345 346 Erismann, T.H., 79 Etienne, S., 165 Evans, S.G., 52t, 79 Ewing, M., 28t, 52t, 275t, 326t Ezquerro, L., 378t

F Fahnestock, R.K., 143f Fairbridge, R.W., 232, 275t Falcon, N.L., 28t Fallgatter, C., 454 455 Famakinwa, S.B., 30t Fan, H., 225, 233 Fang, G., 236 237 Farmer, D.M., 337, 339 Farre, J.A., 250 Farrow, G.E., 114 116 Fauge`res, J.-C., 2, 243 245, 314t, 330 331, 337 338, 341 346, 348, 442 443 Faust, C., 79 Feeley, K., 274 Feldhausen, P.H., 429t Felletti, F., 378t Feng, Z.-Z., 215 216 Feng, Z.Z., 377 395, 378t Fichman, M., 378t Fichman, M.E., 378t, 402 403 Fisher, R.V., 40t, 48 50, 123 126, 129f, 163, 243, 345 346, 429t Flint, R.F., 429t Flood, R.D., 153 155

Flores, G., 40t Fodor, R.V., 397t Folk, R.L., 457 459 Fonnesu, M., 2, 310, 372, 373f, 374f, 454 455 Ford, C., 378t Forel, F.A., 1 2, 216, 268 270, 455 456 Foreman, M.G.G., 182 186, 183t, 189 Forristall, G.Z., 290t Fortuin, A.R., 378t Fossati, M., 256 Fossen, H., 378t Framinan, M.B., 225, 256 Frey-Martinez, J., 28t Friedman, G.M., 40t Frohlich, C., 282 Fudol, Y.A.H., 90 92 Fuhrmann, A., 2, 310, 454 455

G Gagnon, J.F., 22, 72f, 401 402 Galay, V., 98f, 226 227 Gangadhara Rao, L.V., 296 Gani, R., 429t Gao, S., 233 234 Gao, W., 231, 233 234, 244 Garcı´a, M., 337 Gardner, J.V., 159, 452 453 Garrison, R.E., 397t Gaudin, M., 40t, 49, 429t Gavrilov, Y.O., 378t Gebbie, G., 145f Gee, M.J.R., 28t, 38, 45, 80 Geertsema, M., 22 23, 45, 81 Geist, E.L., 280 281 Gelfenbaum, G., 378t Geyer, W.R., 183t, 191 Ghayoumian, J., 28t Ghibaudo, G., 93 95 Gihm, Y.S., 216 217 Gill, A.E., 310, 328 330, 363 364, 456 Gindre-Chanu, L., 378t Glade, T., 22 23 Gladkov, A.S., 378t Gladstone, C., 249 250 Goguel, J., 79 Go´mez-Enri, J., 340 Gong-Yiming, 429t Gonthier, E.G., 314t, 323 324, 326t, 338, 348 349 Gonzalez-Silvera, A., 183t, 193, 219t, 256 260 Goodbred Jr., S.L., 296 Gordon, A.L., 2, 144 146, 145f, 322, 328 330 Goren, L., 79 Grajales-Nishimura, 277

549

550

Author Index

Greb, S.F., 378t Greene, H.G., 13, 275t, 300 Gregory, M.R., 378t Grieve, R.A.F., 429t Griffiths, J.C., 462 Grimm, R.E., 80 81 Griswold, J.P., 78, 81 Grotzinger, J., 327 328 Gruszka, B., 378t Guest, J.E., 80 Guo, L.C., 237 239 Guy, M., 40t, 165

H Hacker, D.B., 28t Haflidason, H.L., 28t, 52t Hamilton, P.B., 177 178 Hampton, M.A., 2, 14 23, 38, 40t, 48 50, 78 81, 90, 92 94, 94f, 98f, 99, 101 105, 112 113, 126 127, 133 135, 139f, 140 145, 142f, 165, 176, 182 186, 214, 226 227, 245, 250, 336, 443 444, 453, 471 Hansen, M.J., 40t, 48 Hanzawa, H., 37 38 Haq, B.U., 336 Harland, W.B., 429t Harmelin-Vivien, M.L., 289 292 Harms, J.C., 143f, 459 461 Harris, P.T., 251 252, 356 357, 470f, 471f, 472f, 473 475, 474f Harrison, J.V., 28t Harrison, K.P., 80 81 Haver, S., 288 Havil, J., 327 328 Hawati, P., 239, 246 Hayter, E.J., 37 38 He, B.Z., 128 130 He, Y.-B., 314t Heath, R.A., 203 204 Heezen, B.C., 28t, 52t, 275t, 301 302, 310, 314t, 319 320, 323 324, 333 336, 348, 397t, 449 Heim, A., 79, 81 82 Helwig, J., 22, 50, 71, 378t Henkel, D.J., 289 Henstock, T.J., 275t Herdendorf, C.E., 452 Heritier, F.E., 370 Herna´ndez-Molina, F.J., 310, 314t, 326t, 336 340, 344, 348 Hickey, H., 189 Higgins, J.B., 264t Higgs, R., 178 Highland, L.M., 22 23, 40t, 45 Hilbert-Wolf, H.L., 378t

Hilton, V., 30t Hiscott, R.N., 2, 5, 175 176, 264t Hizzett, J.L., 274 Hoitink, A.J.F., 237 Hollister, C.D., 227, 244 245, 302, 310, 323 325, 326t, 333 336, 343 345, 367 368, 371t, 429t, 450 Holsapple, K.A., 402 403, 405 Hooper, J.R., 289 Hopkins, J., 199 Hori, K.I., 239 Hornbach, M.J., 300 301 Horowitz, D., 378t Houghton, P., 40t, 325 327 Howard, K.A., 80 Howe, J.A., 244, 302, 326t, 328 331, 333f, 336, 342 Howe, M.R., 338 Hoyal, D.C.H., 177 179 Hsu¨, K.J., 40t, 48 49, 79, 81, 138, 310, 324 325, 344 345, 429t Hubbard, D.K., 289, 290t Hubert, J.F., 310, 324 325, 344 345 Hubert, J.M., 206 207, 349 352 Hu¨eneke, H., 5 Humphrey, J.D., 326t Hungr, O., 48 49, 79 Hunter, P.M., 450f Huppert, H.E., 81 Hu¨rlimann, M., 81 Hutton, J., 462 Hwang, I.G., 216 217

I Ilstad, T., 38, 80 Imran, J., 183t, 191, 216 217, 219t, 224, 246 Inglis, I., 165 Inman, D.L., 48 49, 290t Ito, M., 165 166, 324 325, 378t Iverson, R.M., 37 39, 40t, 78 79, 81, 92, 126

J Jackson, B.A., 364 365 Jackson, J.A., 39, 40t, 48 49 Jacobi, R.D., 28t, 52t, 113 114 Jaeger, J.M., 397t, 407 Jaffe, B., 378t Jagadeesan, L., 183t, 219t, 239 Jansen, E., 52t Jay, D.A., 237 Jenkyns, H.C., 397t Jia, J., 176 Johansson, M., 245 Johnson, A.M., 37 38, 50, 98f, 245, 466 Johnson, D.A., 326t

Author Index

Johnson, K.S., 216 217 Johnson, T.C., 334 Jones, A.P., 378t Jones, P.N., 93 95 Jordan, D.W., 102 105, 171, 264t

K Kahle, C.E., 378t Kang, H., 378t Kao, S.J., 205 206 Karcz, I., 37 38 Kassem, A., 216 217 Kastens, K.A., 40t, 281 282, 429t Keller, G.H., 357 359 Kelling, G., 2 Kent, P.E., 79 Kenyon, N.H., 206 207, 339, 349 352 Kerr, B.C., 314t Khan, S.M., 216 217 Khanna, D.R., 261 Khripounoff, A., 93 95 Khvorova, I.V., 429t Kilian, S., 378t King, E.L., 300 Kirkland, D.W., 378t Kirschbaum, D.B., 22 23 Kishida, T., 37 38 Kjeldsen, S.P., 288 Klaucke, I., 159 Klaus, A., 153 155, 314t, 347 348 Klein, G.D., 92, 136, 136f, 182 186, 310, 324 325, 344 345, 359, 429t Kleinspehn, K.L., 127f, 128f Knauss, J., 326t Kneller, B., 2, 92 95, 143f, 476 Koch, S.P., 206 207, 326t Kohut, J.T., 290t Komar, P.D., 177 178, 182 186, 289 292 Koning, H.L., 40t Kopf, A.J., 378t Koppejan, A.W., 40t Korup, O., 79 Korznikov, K.A., 378t Kostic, S., 178, 216 217, 225, 228, 244, 249 250 Kottke, B., 292 Kuenen, Ph. H., 2, 14 22, 40t, 90, 92 95, 98f, 122, 165, 333, 335 336, 377 395, 429t, 475 Kuhn, T.S., 263 268 Kundu, A., 378t Kvenvolden, K.A., 304

L Labaume, P., 429t

Laberg, J.S., 28t, 302 Laboute, P., 289 292 Lacombe, H., 337, 339 Laine, E.P., 367 368, 371t Lamb, M.P., 216 217, 225, 228, 243 244, 249 250 Lambe, T.W., 296 Lancaster, J.T., 13 Larcombe, P., 290t Lario, J., 334 335, 340 Lash, G.G., 264t Lastras, G., 116 117 Lathrop, E., 341 LaViolette, P.E., 337, 339 Lavrenov, I., 288 Lawton, T.F., 277 Le Roux, J.P., 378t Leat, P.T., 277 279 Leclair, S., 138 Ledbetter, M.T., 314t, 323 324, 326t, 347 348 Lee, H.J., 28t, 49 50, 274, 275t, 300, 302 Lee, H.S., 378t Legros, F., 81 Leo´n, R., 337, 340 Leonard, L.A., 290t Lewis, K., 48 Lewis, K.B., 52t Lewis, T., 216 217 Li, G.X., 200, 235 236 Li, S., 131, 378t Li, T., 49 50 Li, X., 176 177 Lien, R.-C., 314t Liesa, C.L., 378t Lindberg, B., 300 Lipman, P.W., 277 279 Lisitsyn, A.P., 429t Liu, J.P., 183t, 219t, 237, 239 Liu, J.T., 225, 253 254 Liu, K.B., 237 239 Locat, J., 28t, 34, 274, 275t, 302 Logan, W.E., 378t Longva, O., 296 Lonsdale, P., 314t, 326t, 344 345, 365 366, 397t Lovell, J.P.B., 310, 331 334 Lowe, D.R., 2, 14 22, 40t, 50, 92 95, 98f, 106f, 108 112, 126 128, 131 132, 138 140, 139f, 149 150, 165, 171, 264t, 377 395 Lu, S., 239 Lucchitta, B.K., 80 Lu¨dmann, T., 314t Lunina, O.V., 378t Luo, Z., 183t, 216 217, 219t, 241 242 Lykousis, V., 34

551

552

Author Index

M MacEachern, J., 216 217 Madrussani, G., 14 22, 88 Magalha˜es, V.H., 340 Major, J.J., 38 Malfait, B., 326t, 344 345 Malin, M.C., 52t, 81 82 Maltman, A., 14 22, 377 395, 471 Maltman, A.J., 50, 71 Manca, F., 197 198 Manica, R., 93 95 Marco, S., 14 22, 70, 131 Marr, J.G., 14 22, 45 47, 63 68, 73, 80, 90, 93 95, 98 99, 105, 107f, 122, 126f, 245, 477 Marshall, N.F., 48 49, 289 Martin, B.D., 156 158 Martin-Chivelet, J., 244, 324 325, 337, 344 345 Martinsen, O.J., 79 Maslin, E., 370 Maslin, M., 275t Massari, F., 378t Masson, D.G., 48, 52t, 274 Masunaga, E., 219t, 240 Matano, R.P., 183t, 193, 219t Matsumoto, D., 378t Mattinson, J.M., 243 Mazumder, R., 210 211, 378t McAdoo, B.G., 28t, 54 55, 80 81 McCave, I.N., 93 95 McEwen, A.S., 78, 81 McGowen, J.H., 246 McGregor, B.A., 28t, 52t McLean, C.E., 378t McLoughlin, P.A., 359t McMurtry, G.M., 277 279 McPherson, J.G., 14 22, 183t, 203 204, 219t, 228 229, 239, 264t, 397t, 449 Meade, R.H., 218, 229 230 Mearns, D.L., 290t Meckel, T., 90 92 Meckel III, L.D., 90 92 Melosh, H.J., 78 79, 81, 402 403 Menzies, J., 378t Menzies, R.T., 326t Meshram, D.C., 378t Metz, J.M., 378t Metz, M., 378t Miall, A.D., 336 Michalik, J., 40t Michels, K.H., 314t, 326t Middleton, G.V., 2, 14 22, 40t, 50, 92 95, 94f, 96f, 98f, 126 127, 133 135, 139f, 140 145, 142f, 165, 176,

182 186, 214, 226 227, 243, 245, 250, 336, 443 444, 453, 459 461, 471 Mienert, J., 34, 304 Migeon, S., 259 Migliorini, C.I., 333 Mikhailova, M.V., 183t, 197 198 Milia, A., 40t, 279 280 Milliff, R.F., 182 186, 205 206 Milliman, J.D., 216 218, 229 230, 239 Miramontes, E., 326t Mitsuzawa, K., 290t Miyamoto, H., 81 82 Moberly, R., 397t Mohrig, D., 38, 80, 108, 110f, 216 217, 233 234 Moiola, R.J., 104f, 169 171, 246 247, 249, 264t, 275t, 429t Molnia, B.F., 28t Montgomery, D.R., 11, 14 34, 52t, 80 81 Moore, D., 225 Moore, D.G., 28t, 52t Moore, G.T., 2 Moore, J.G., 28t, 52t, 79, 117, 277 279 Moraes, M.A.S., 370 Morales de Luna, T., 216 217, 225, 234 Moretti, M., 378t Morgenstern, N.R., 40t Moritz, H.R., 290t Morner, N.A., 300 Morozov, E.G., 337 Moscardelli, L., 14 22, 28t Mosher, D.C., 2, 14 22, 34, 274 276, 275t, 298, 300, 333 Mulder, T., 2, 5, 40t, 90 95, 149 150, 182, 200, 214, 216 217, 224, 229 230, 236, 242 244, 247 249, 259, 268 270, 274, 310, 314t, 330 331, 338, 340 343, 348, 370, 429t, 442 443, 456 Mullins, H.T., 90 92, 206 207, 314t, 335, 349, 370, 371t Mu¨nchow, A., 250 Murray, J., 14 22 Murray, S.P., 290t Mutti, E., 2, 14 22, 40t, 63, 73 74, 93 95, 122, 140, 142f, 149 150, 153, 155 156, 216 217, 244, 246 248, 324 325, 344 345, 370, 429t Myhre, A.M., 397t

N Nanfito, A.F., 378t Narbonne, G.M., 342 343 Nardin, T.R., 14 22, 40t, 49, 73, 94, 98 99 Natland, M.L., 40t, 344 345, 429t Nelson, C.H., 155 156 Nemec, W., 40t, 49, 127f, 128f, 301 302 Newton, C.S., 28t

Author Index

Ni, L.T., 63 68 Nilsen T.H., 342 343 Nio, S.-D., 359 Nittrouer C.A., 90, 214 Nodine, M.C., 289 Nogueira, A.C.R., 378t Norem, H., 40t, 80 Normark, W.R., 28t, 52t, 150, 152 153, 159, 251, 277 279, 465f, 466f, 467f, 468t, 473 Nowlin, Jr., W.D., 206 207, 349 352

O Obermeier, S.F., 378t Odonne, F., 378t Ogston, A.S., 216 217 Olabode, S., 378t O’Leary, D.W., 52t Oliveira, C.M.M., 378t Omoto, K., 378t Onorato, M.R., 378t Ortner, H., 378t Ouyang, C., 9 10 Ozcelik, M., 378t ¨ zkul, M., 378t O

P Palanques, A., 49, 253 254, 290t Palladino, G., 14 22, 395 Pan, S., 216 217 Pan, S.X., 216 217 Pandey, D.K., 397t Parchure, T.M., 37 38 Parker, G., 93 95, 178, 227 Parker, K., 216 217, 225, 228, 244 Parsons, J.D., 90, 93 95, 216 217 Passchier, S., 378t Passega, R., 429t Paull, C.K., 156 158, 289 292, 304, 397t Paull C.K., 95 Pavec, M., 337 Peliz, A., 183t, 196, 219t Penven, P., 326t Pequegnat, W.E., 206 207, 310, 334 335, 349 352 Perch-Nielsen, K., 397t Petley, D., 22 23, 275t Petruncio, E.T., 356 357 Petter, A.L., 216 217 Pettijohn, F.J., 210 211, 366 369, 459 461 Phillips, C.J., 96f Pichon, A., 339 Pickering, K.T., 2, 5, 30t, 38, 50 Pierce, L.E.R., 216 217 Pierson, T., 38, 52t, 80

553

Pierson, T.C., 40t, 92 93, 95, 96f, 98f, 225 Pilkey, O.H., 2 Pinheiro, L.M., 340 Piper, D.J.W., 2, 14 22, 28t, 48 49, 153 155, 310, 344 345, 397t, 407 413 Plink-Bjo¨rklund, P., 225 Pluenneke, J.L., 429t Pomar, L., 314t, 429t Popenoe, P., 28t, 275t, 304 Postma, G., 2, 14 22, 40t, 93 95, 108, 122 123, 126f, 127f, 128f, 165, 177 179, 233 234 Potter, P.E., 210 211, 366 369, 459 461 Prandtl, L., 92 Pratt, B.R., 378t Price, J.F., 338, 340, 343 344 Prior, D.B., 40t, 52t, 80, 113, 116 117, 244, 275t, 296, 301 302 Sultan, N., 38, 275t, 302 Pritchard, D., 249 250 Prytkov, A.S., 378t Pudsey, C.J., 333f Puig, P., 216 217, 290t Purdy, E.G., 261 Purkey, S.G., 90, 145 146, 145f, 322 Purkis, S., 261

Q Qiao, L.L., 216 217, 235 236 Qiao, X.F., 128 130 Quaresma, L.S., 339

R Rana, N., 378t Rathje, E.M., 282, 289 Rebesco, M., 2, 92, 244 245, 310, 311t, 322, 331 333, 335 336, 341, 343 344, 349 Reiche, P., 40t, 49 Reid, J.L., 326t Reineck, H.E., 359, 459 461 Renard, A.F., 14 22 Ricci Lucchi, F., 14 22 Richardson, M.J., 323 324, 326t Ridente, D., 117 Riemer, M.F., 36 37 Ritchie, A.C., 189 Ritter, D.F., 81 Rivera-Rosario, G.A., 292 Rizzini, A., 429t Roberts, N.J., 52t Rodine, J.D., 98f Rodriguez, A.B., 314t Rodriguez, M., 28t Rothwell, R.G., 179

554

Author Index

Rouse, C., 40t, 48 Rouse, L.J., Jr., 183t, 191 192, 219t Roveri, M., 250 251 Rowe, G.T., 326t Ruban, V., 288 Rudersdorf, A., 378t Rudnick D.L., 326t

S Sakaguchi, A., 397t Saller, A.H., 30t, 50, 246, 304 307, 314t, 365 366, 370 Sammartini, M., 453 457 Sanchez-Garrido, J.C., 339 Sanders, J.E., 14 22, 40t, 47 48, 50, 92 95, 122, 133 135, 138, 142f, 226 228, 248, 342 343, 377 395, 475 Sarkar, S., 378t Saxov, S., 40t Sayles, R.W., 429t Schaller, P.J., 79 Scheidegger, A.E., 79, 81 Schieber, J., 216 217, 243 244 Schillereff, D.N., 216 217 Schuster, R.L., 22 23, 28t Schwab, W.C., 28t, 155 156 Scott, K.M., 52t Seilacher, A., 378t, 395, 414 416, 429t, 472 Seth, A., 378t Shakesby, R.A., 298 Shanmugam, G., 1 2, 5, 11, 13 35, 28t, 30t, 37 39, 40t, 45 50, 52t, 63 68, 80 81, 90 95, 96f, 97, 97f, 98f, 99, 100f, 101 105, 101f, 102f, 103f, 104f, 105f, 106f, 107f, 109f, 110f, 111f, 113 117, 113f, 114f, 115f, 116f, 118t, 120f, 121f, 122, 122f, 123f, 124f, 124t, 125f, 125t, 126f, 128f, 130 131, 130f, 133 136, 133f, 134f, 135f, 137f, 138, 138f, 139f, 141f, 142f, 143f, 144, 149 151, 155 158, 160 166, 169 172, 175 176, 179, 182 186, 183t, 188, 193, 199, 202, 210 211, 215 218, 219t, 224 227, 227t, 233 234, 237, 239, 243 252, 263, 264t, 268 270, 274, 275t, 277 283, 288 292, 300, 302, 304 307, 310, 314t, 320 321, 324 325, 327 328, 333 336, 342 346, 349 357, 359, 363, 365 366, 370 372, 377 396, 378t, 397t, 408 436, 426f, 429t, 449 454, 450f, 454f, 456 462, 458f, 459f, 460f, 461f, 464f, 465, 471 475 Shapiro, S.A., 369 370 Sharpe, C.F.S., 40t, 48 Shepard, F.P., 11 12, 14 22, 40t, 48 49, 100f, 117, 133f, 250 251, 302 304, 356 359, 468t Marshall, N.F., 245 246, 310, 314t, 333, 335, 356 359, 359t Sherrod, L., 378t Shillington, F.A., 219t, 240

Shinn, E.A., 261 Shoaei, Z., 28t Shree Ram, P., 296 Shreve, R.L., 40t, 48, 79, 81, 113 Shrivastava, S.K., 116f, 118t, 120f, 121f, 122f, 123f, 124t, 125t Shultz, A.W., 98f Simons, D.B., 178, 347 Simpson, E.J., 93 95 Singer, K.N., 52t, 80 81 Singh, I.B., 459 461 Sitar, N., 37 38 Skipper, K., 178 Slatt, R., 216 217 Slatt, R.M., 171, 264t Smethie, W.M., 145f Smit, J., 277 Smith, R.D.A., 453 Snedden, J.W., 289 292, 290t Snyder, M.E., 378t So, Y.S., 176 Sobiesiak, M.S., 378t Solheim, A., 34 Solheim, A., 54 55, 302 Sonnerup, R.E, 145f Soomere, T., 288 Sorby, H.C., 462 Southard, J.B., 40t, 182 186, 245, 320 321, 459 461 Soyinka, O.A., 216 217 St. Laurent, L., 310 Stanbrook, D.A., 162 Stanford J.D., 314t Stanley, D.J., 2, 40t, 163, 165 166, 182 186, 320 321, 324 325, 429t Stauffer, P.H., 133 Steel, E., 200, 216 217, 228, 236, 245, 247, 249 250 Steel, R., 225 Steel, R.J., 216 217 Steele, T.H., 326t Sterling, G.H., 289 Stevens, M.R., 298 Stewart, R.H., 310, 328 330 Sto¨ffler, D., 429t Stommel, H., 310, 313 Stone, C.G., 171 Stone, G.W., 290t Stow, D.A.V., 2, 90 92, 139f, 155, 243 245, 310, 314t, 330 347, 372, 429t Strasser, M., 54 55, 63, 397t Strohbeck, G.E., 289 Su, D.C., 130 131, 378t Suhayda, J.N., 38, 289 Sullivan, G.G., 359t

Author Index

Sun, A.P., 130 131, 378t Sun, F.N., 216 217 Sundborg, A., 48 49 Swallow, J.C., 326t Swanson, R.G., 461 Syvitski, J.P.M., 114 116, 183t, 191, 219t, 224, 229 230, 246

555

Viana, A.R., 90 92, 310, 314t, 322, 326t, 336, 370 Visser, M.J., 164, 245 246, 359 Vivas, V., 429t Voight, B., 79 Volkman, C., 326t Volkman, G.H., 326t Vrolijk, P.J., 40t Vsemirnova, E., 337

T Takahashi, T., 245 Talley, L.D., 237, 310, 313, 328 330 Talling, P.J., 80, 216 217, 226 227, 249, 274 Tappin, D.R., 49 Taylor, B., 459, 460f Teale, T., 63 Terwindt, J.H.J., 164 165, 359 Terzaghi, K., 36 38, 79 Thiede, J., 397t Thompson, D.B., 459 461 Thomson, J., 450f Thomson, R.E., 183t, 189 Tilling, R.I., 28t, 275t Tinterri, R., 378t Toniolo, H., 227 Topal, S., 378t To¨ro, B., 378t Trincardi, F., 275t, 302, 335 336 Tripsanas, E.K., 300 301 Tsuji, Y., 326t Tucholke, B.E., 302, 336 Turner, F.J., 429t Turner, J.S., 248 Twichell, D.C., 54 55, 155 156, 274 276, 300, 408 413

U Underwood, M.B., 322 Urgeles, R., 22 23, 275t

V Vail, P.R., 153, 171 172, 275t Vallance, J.W., 52t Vallance, R.W., 40t Van der Lingen, G.J., 138, 263, 377 395 Van Loon, A.J., 378t Van Loon, A.J. (Tom), 215 216, 268 270 van Maren, D.S., 225, 233 Van Rooij, D., 346 347 Vanneste, M., 452 Vargas-Yanez, M., 337 Varnes, D.J., 14 22, 38, 40t, 45 49, 81 Velasco-Villareal, M., 378t Verdicchio, G., 335 336

W Waldron, J.W.F., 22, 72f, 378t, 401 402 Walker, K.R., 314t Walker, N.D., 183t, 191 192, 219t Walker, R.G., 139f, 348 Wallis, G.B., 98f Wang, H., 183t, 200, 219t, 225, 229 230, 233 236, 239, 269 270 Wang, H.J., 216 217, 231, 236 Wang, J.D., 176 Wang, N., 216 217, 231, 233, 236 Wang, X.H., 216 217, 231 Wang, Y., 176 Warme, J.E., 275t, 302 304 Warne, A.G., 245 246 Warner, M.J., 145f Warrick, J.A., 189, 200, 216 217, 236, 246 247, 249 250 Weathley, D.F., 378t Weaver, P.P.E., 449, 450f Weber, M.E., 296 Weidinger, J.T., 79 Weiler, Y., 314t Weimer, P., 171 Weiss, L.E., 429t Welbon, A.I.F., 83 84 Wetzel, A., 337, 342 Weyl, P.K., 461 Whiteway, T., 251 252, 356 357, 470f, 471f, 472f, 473 475, 474f Whitman, R.V., 296 Wilcock, P.R., 96f Williams, G.P., 102 105 Wilson, R.D., 216 217, 243 244 Wiseman, G., 40t Wiseman, Jr. W.J., 236 237 Wizevich, M.C., 378t Woodbury, H.O., 9, 179 Woodcock, N.H., 50 Worthington, L.V., 326t Wren, P.A., 290t Wright, L.D., 90, 214, 216 218, 224 225, 228, 230, 233 234

556 Wright, S.G., 35 37, 274 276, 282, 289, 449, 453, 472 473, 475 Wu, J., 237 Wunderlich, F., 359 Wunsch, C., 349 Wust, G., 326t Wynn, R.B., 28t, 48

X Xu, J.P., 356 357

Y Yadav, P.R., 261 Yang, C.-S., 359 Yang, J., 236 237

Author Index

Yang, T., 225 Yao, Z.G., 236 237 Young J.R., 63 Yu X., 314t

Z Zavala, C., 73 74, 122, 140, 142f, 200, 215 217, 225, 228, 236, 239, 245 248, 252, 259, 268 270 Zenk, W., 92, 310, 313, 338 340 Zenk, W.O., 326t Zhao, R., 378t, 402 403, 405 Zimbrick, G., 30t, 314t, 378t Zimmerman, H.P., 275t, 302 304 Zou, C., 90 92, 176 177, 244 Zuffa, G.G., 210 211, 366 370

Subject Index Note: Page numbers followed by “f” and “t” refer to figures and tables, respectively.

A AABW. See Antarctic bottom water (AABW) AAIW. See Antarctic intermediate water (AAIW) (Brazilian margin) Abyssal plain, 449, 450f contourites, 347 348 ACC. See Antarctic circumpolar current (ACC) ADFEX. See Arctic Delta Failure Experiment (ADFEX) Aeolianite, 429t Agat region, 28t Agulhas Current, 449 Agulhas Slump, 23 34 Airgun seismic reflection profile, 465f Alaknanda Valley (Garhwal Himalaya), 378t Alba Field, 378t Aleutian Trench, 407 Alika 1 and 2 Debris, 28t Alluvial fan, 449, 451f Alongslope processes, 2 3, 5 6, 146 148 Amazon Fan, Equatorial Atlantic high-amplitude reflection packet units, 153 lower-fan lobes, 153 155 sinuous channels, 150 153 Amazon River, 183t, 187, 219t Amirante Passage, 326t Anambra Basin (Southern Nigeria), 378t Anastomosing, 219t, 270 Anastomosite, 429t Ancient, 449 Ancient debrites identification, 99 Ancient hyperpycnites flawed principles, 248 249 grain size, 249 hyperpycnite facies model, 242 243 hyperpycnite fan models, 246 248 internal erosional surface, 244 inverse to normal grading, 243 244 lofting rhythmites, 245 246 massive sandstones, 245 modern analogs, 249 250 plant remains, 246 recognition of, 240f, 241 252, 241f, 242f submarine canyons, 250 252

traction structures, 244 245 Andrew Field, 30t Annot Sandstone (Eocene—Oligocene), Peira-Cava Area, Maritime Alps, SE France, 161 168 Antarctic bottom water (AABW), 100f, 145 146, 322, 449 Antarctic circumpolar current (ACC), 328 330 Antarctic intermediate water (AAIW) (Brazilian margin), 313 Antillean-Caribbean Basin, 326t Apulian Foreland (Southern Italy), 378t Aqua and Terra NASA’s Aqua (EOS PM-1), 463 464 Aqua satellite, 462f, 463 464 Arabian Sea, 289, 291t Arctic Delta Failure Experiment (ADFEX), 34 Argentine Basin, 347 348, 347f Atlantic Ocean, 192 193, 194f, 199, 199f Atlantic water (AW), 313 Atypical turbidite, 40t, 429t Avalanche, 449 Avalanching flow, 40t AW. See Atlantic water (AW) Ayabacas Formation, 378t

B Back-analysis, 449 Backwash, 281 282 Badwater Drive, valley off, 451f Baikal Rift Zone, 452 Baja California, 11 12, 100f, 133f Bajo Segura Basin, 378t Baroclinic currents, 363 364, 368 depositional framework, 365 366, 367f, 368f empirical data, 364 365, 364f, 365f, 366f Baroclinite, 429t, 449 Basal shear zone, 449 Basin-floor fans, 153 conventional, 172 model, tertiary, North Sea, 171 176 Basque-Cantabrian Basin (North Spain), 378t Bassein Slide, 28t Bathyal Ocean, 449 Bathymetry, 449

557

558

Subject Index

Bathymetry (Continued) Yellow River, China, 232f, 233 Bay of Bengal, 372 BCRS. See Bottom-current reworked sands (BCRS) Bedform-velocity matrix, 346 347, 346f Bentonite clay content, 477 Bering canyon, 473 Bering Sea, 465f, 468t, 473 Betsiboka River, 183t, 219t BICC. See Brazil intermediate counter current (BICC) Biological erosion, 302 304 Bioturbation, 342 343 Blake Bahama Outer Ridge, 333 334 Block slide, 55, 84, 116 117, 160 161 Bolide, 449 Boso Peninsula (Japan), 378t Bottom-current reworked sands (BCRS), 149 150, 352 355 Bottom-current reworking, 449 Bottom currents, 4f, 310 312, 449 baroclinic currents, 363 366 contourite problem, 327 349 deep-marine, 314t domain of, 310 maximum current velocities of, 326t reservoir quality, 370 372 sediment provenance, 366 370 thermohaline-induced geostrophic, 321 327 tidal, 356 363 types of, 320 321 wind-driven, 349 356 Bottom-simulating reflector (BSR), 304 Bottom-turbid layers, Yellow River (China), 233 234 Bouma sequence, 136 138, 138f, 149 150, 164, 241 242, 244, 324 325, 450 452 Braid delta, 203 204, 239 Braidite, 429t Brazil current (BC), 313 Brazil intermediate counter current (BICC), 313 Breccias, 417 435, 431f, 432f, 433f, 434f, 435f geological implications, 405, 406f meteorite, 417 435 sedimentary-collapse, 417 435 Brecciated clasts, 450 Brecciated zone, 450 Brisbane River, Moreton Bay, Australia, 203, 204f Broken Ridge (Indian Ocean), 397t Brunei Slide, 28t Brushy Canyon Formation, 178 Buoyancy-dominated flow, 225

C Calabar Estuary, 357 359

Calabrian Arc (Southern Italy), 378t California Baja, 11 12, 100f, 133f Death Valley, 451f Eel River, 191, 192f Miocene Monterey Formation, 414 416 Cangoa´ Field, 30t, 314t Canyon-fill facies, 119 120 Canyons, 116 117 Bering Canyon, 473 cyclone-induced hyperpycnal turbidity currents in, 253 254, 253f pliocene, 117 on slope of Gulf of Lion, 470f submarine, 117, 466f, 467f, 468t, 471f, 472f, 473 475, 474f Swatch-No-Ground Canyon, 470f Zhemchug Canyon, 473 Cape Fear MTD, 28t Carbonate clast, tabular, 103f Carnegie Ridge, 326t Cascading densewater event, 40t Cascadite, 429t Castile Evaporite, 378t CDS. See Contourite depositional system (CDS) CDW/CPDW. See Circumpolar deep-water (CDW/CPDW) Centre National d’Etudes Spatiales (CNES), 464 Chagos Ridge (Indian Ocean), 397t Chaibasa Formation, 378t, 436 Challenger Deep, 14 22 Chamais Slump, 28t Channelized lobes, 155 156 Channel-levee geometry, 150, 322 Chicxulub crater, 378t Chicxulub impact, 378t Chignik River, 183t, 219t Circumpolar deep-water (CDW/CPDW), 313 Clastic injections, 414 416, 437 439 Clastic sediment, 450 Clasts floating mudstone, 453 projected, 467 Clay content, bentonite, 477 Cloudy, 219t CNES. See Centre National d’Etudes Spatiales (CNES) Coalescing irregular, 219t Coalescing lobate, 183t plume, 260 Coarse-grained deltas vs.,, 228 229, 229f, 230f Coarse-grained subaerial debris flows, 37 38 Cohesionless debris avalanche, 40t Cohesionless debris flows, 102 105, 106f, 114 116

Subject Index

Cohesionless liquefied sandflow, 40t Cohesive debris flow, 99, 102 105, 106f, 325 327 Colorado Plateau, 378t Concud Fault (Eastern Spain), 378t Congo (Zaire) River, Atlantic Ocean, West Africa, 199, 199f Congruity, cosmic, 8 14, 8f Connecticut River, New England region, United States, Long Island Sound, US, 189, 192f Continental margin, 450 Continental margins, 34 35, 224 Continental rise, 450 Continental shelf, 450 Continental slope, 450 erosional features on, 473 Continental Slope Stability (COSTA), 34 Continuous creep, 40t Continuum, between turbidity currents and contour currents, 331 333 Contorted bedding, 450 Contour currents, 321 322, 333 336, 367 and turbidity currents, continuum between, 331 333 Contourite, 450 Contourite depositional system (CDS), 337 Contourite drift, 330 331 Contourite facies model, 341 344 bioturbation, 342 343 current velocities, 342 grain-size data, 344 internal divisions, 341 342 internal hiatuses, 342 multiple interactive processes, 343 344 Contourite problem, 327 349 abyssal plain contourites, 347 348 bedform-velocity matrix, 346 347 continuum between turbidity currents and contour currents, 331 333 contour currents, 333 336 contourite facies model, 341 344 erosional features, 337 338 global ocean circulation, 328 331 Gulf of Cadiz, 338 341 hiatuses in, 336 sandy intervals of contourite facies models, 348 349 shale clasts, 344 346 traction structures, 344 346 Contourites, 310, 311t, 444 hiatuses in, 336 seismic geometries of, 322 Contractional tectonics, 395 Conventional basin-floor fans, 172 Copper River, 28t, 183t, 219t, 443 444 Copper River Slide, 28t

Coral and Mamba gas fields, 372, 373f, 374f Coral reef, 261 Core A, 450 452 Cored reservoirs, 172 175 Core photography natural sunlight in, 461, 461f timing of, 461 Cores, reservoir facies in, 461 Cosmic congruity, 8 14, 8f, 10f COSTA. See Continental Slope Stability (COSTA) Costa Rica margin, 397t Creep, 48 continuous, 40t Current ripples, 169 170, 349 352 Current velocity, 323 324, 342, 349 352 submarine canyons, 357 359, 358f, 359t Currituck Slide, 28t Cyclone extratropical, 190 tropical, 473 475 Cyclone-induced hyperpycnal turbidity currents, 253 254, 253f Cyclonic waves, 288 292, 290t, 291t, 293f

D Dart River, Lake Wakatipu, South Island, New Zealand, 203 204, 205f Dead Sea Basin, 70 Dead Sea Graben, 378t Debris avalanche, 40t cohesionless, 40t Debris flows, 73 74, 146 coarse-grained subaerial, 37 38 cohesionless, 114 116 definition, 98 99 experimental, 110f facies models, 120 glacial, 63 68 identification, 99 101, 115f layer, deposition, 111f methods for recognizing, 456 457 muddy, 106f origin, 99 problem, 120 126 sandy, 99, 101 116, 114f, 115f flume experiments on, 477 478 strong, 109f subaqueous, 108 Debris slide, 40t, 48 Debris tongue, 452 453 Debrites, 452 gravelly, 99 identification of ancient, 99

559

560

Subject Index

Debrites (Continued) muddy, 99 101 sandy, 99 101, 161 De´collement, 466 Decomposition, gas-hydrate, 304 Deep creep, 40t, 45 Deep-lacustrine environments, 452 Deep-lacustrine sands, 453 Deep-marine bottom currents, 314t types of, 444 Deep-marine environments, 14 22 Deep-marine sediments, 453 Deep-marine tidal currents, 335, 368 Deep Sea Drilling Project (DSDP), 359, 469 471 Deep-sea sand deposition, 306 307 Deep-water, 453, 454f circumpolar deep-water (CDW/CPDW), 313 Deep-water genetic facies models, 442 443 Deep-water masses, 310 312 Deep-water processes, 340 341 Deep-water sands, water-escape structures in, 108 112 Deep-water sandstones, floating clasts in, 122 Deep-water sedimentation, 1 2 Deep-water tidal bottom currents, 320 321, 449 Deep-water turbidity currents, vital properties of, 93 95 Deep western boundary current (DWBC), 314t Deep western boundary undercurrent (DWBUC), 323 324 Deflecting, 191 192, 206 207 sediment plumes, 189, 191 192, 198, 206 207 Deformation, 395 prelithification, 377 396, 437 439 sediment, 471 soft-sediment, 128, 472 tectonic, 63 68, 88 Delaware Basin, 178 Delta formation, rational theory for, 443 444 Delta-front mass movements, 301 302 Delta vs. estuary, 231 232 Dense flow, 40t Densite, 429t Density currents, 227 Density-driven flows, 90 Density flows, 147f Density-modified grain flow, 49 Density plumes, 143f, 182 configurations of, 254 255 NASA for imaging, 462f types, 182, 188 Density-stratified flows, 108 Depositional loading, 301 302, 303f Depositional lobes, 153, 156

Depositional process sedimentology, 466 467 Deposits of gravity flows, 90 92 Devonian Lower Genesee Group, Central New York, 244 DGSRF, 313 318 Diamictite, 429t Dish structures, 453 Dispersive pressure, 243 Dissipating, 192 193 plume with irregular front, 256 259 DMLs. See Double mud layers (DMLs) Doon Valley (Garhwal Himalayas), 378t Double mud layers (DMLs), 164, 359, 444, 453 Downslope, 1 3, 5 6 Downslope gravity-driven process, 14, 92 97 Drained condition, 453 Drained slump, 40t DSDP. See Deep Sea Drilling Project (DSDP) Duplex-like structures, 40t, 63 70, 113, 114f DWBC. See Deep western boundary current (DWBC) DWBUC. See Deep western boundary undercurrent (DWBUC)

E Earth, 8 13 mass transport deposits (MTDs) on, 52t Earth flows, 40t, 179 Earth Observing System (EOS), 463 464 Earthquake-induced gravity tectonics, 70 Earthquakes, 274 277, 397t, 405 406, 409, 413 435 East Breaks, 28t East Breaks Slide, 28t Ebb-tidal currents, 292 296 Ebro Delta, Mediterranean Sea, Iberian Peninsula, 195, 195f Ecca Group (Karoo Basin, South Africa), 378t Edop Field, 30t, 362 363, 378t Eel River, California, Pacific Ocean, United States, 191, 192f Elastic behavior, 45 47, 92 93 Elastic mode of transport, 47 48, 93 Elwha River, 189, 210 Elwha sediment plume, Strait of Juan de Fuca, United States, 189, 190f, 191f Elwha sediment plume, Strait of Juan de Fuca, US, 189, 190f, 191f ENAM II (European North Atlantic Margin), 34 Eolian blanketing dust, 219t Eolian dust, 217, 263 EOS. See Earth Observing System (EOS) Erosional features on continental slope, 473 origin of, 337 338

Subject Index

Estuary Calabar, 357 359 defined, 232 delta vs.,, 231 232 ETOPO1, 453, 455f ETOPO2, 453 Ewing Bank Block 826 Field, 30t, 352 356, 353f, 354f, 355f, 356f, 370 372 Extrabasinal turbidites, 247 248 Extraterrestrial environment, 8, 11, 13 Extratropical cyclone, 190

F Facies model, 149 150 turbidite, 165 vertical, 149 150 Faeroe area, 30t, 172 175, 378t Faeroe Bank Channel, 326t Fazenda Alegre Field, 30t, 314t Fethiye Burdur Fault Zone (Southwest Turkey), 378t Fine-grained deltas vs. coarse-grained deltas, 228 229, 229f, 230f Fish activitity, 261 Flaser bedding, 352 355, 359 Floating armored mudstone balls, 163 Floating clasts, in deep-water sandstones, 122 Floating clay clasts, 156 158 Floating mudstone clast, 100f, 163, 453 Floating quartzose granules, 163 164 Flow, 453 buoyancy-dominated, 225 debris, 73 74 fluid-gravity, 453 fluidized, 131 laminar, 456 plume vs.,, 224 225 sediment, 471 sediment-gravity, 471 turbulent, 475 unsteady, 476 waning, 476 waxing, 476 Flowing-grain layer, 40t Flow slide, 48 Fluid dynamics, 453 Fluid-gravity flows, 453 Fluidized cohesionless particle flow, 40t Fluidized flows, 131 Fluid mechanics, 453 Fluid parcels, 363 364 Fluxoturbidite, 40t, 429t Flux, sediment, 471 Foinaven Field, 30t

561

Fonissa River, 183t Frigg Field, 30t, 370, 378t North Sea, 370 Froude-supercritical low, 178 Froude supercritical submarine fans, 177 178 Fundy rift basin, 378t Fyne Field, 30t, 378t

G Gannet Field, 30t, 378t Garden Banks, 30t, 314t Gas hydrate, 304 Gas-hydrate decomposition, 304 Genetic facies model, 342 Geohazards, 454 Glacial debris flows, 63 68 Glacial loading, 300 Glacial maxima and loading, 300 Glide, 123 Glide plane, 454 secondary, 471 Global ocean circulation, dual forcing of, 328 331 GLORIA (Geological Long-Range Inclined Asdic) system, 34, 52t, 349 352, 456 457 Goleta Slide, 300 Grain flows, 102 105, 106f, 131 133, 132f, 147 definition, 131 132 facies models, 133 identification, 132 origin, 132 problem, 133 Grainite, 429t Grain-size data, 344, 345f Grain-size scale graphic lithologic columns, 459, 460f on sedimentological logs, 457 459, 458f, 459f Granular flow, 79 Granular mass flow, 40t Graphic lithologic columns, grain-size scale, 459, 460f Gravelly debrites, 99 Gravite, 40t, 429t Gravitite, 40t, 429t Gravity-driven downslope process, 92 97 mass transport vs. turbidity currents, 92 93 Newtonian vs. plastic fluid rheology, 95 96, 96f sediment concentration, 97 sediment-gravity flows, 93 95, 94f turbulent vs. laminar flow state, 96 97, 97f Gravity-driven downslope processes, 336 Gravity-driven processes, 34 35 Gravity-driven sediment flows, 98f Gravity flows, 4f, 90, 147 148, 442 deposits of, 90 92

562

Subject Index

Gravity flows (Continued) sediment, 93 95, 94f types, 90 Gravity (density) flows, 92 Gravity tectonics, Earthquake-induced, 70 Great Bahama Bank, 370, 371f, 373 375 whitings plume, 261 Great Bahama Canyon, 469f, 473 Greater Antilles Outer Ridge, 326t Green Canyon, 30t, 206 207, 326t Greenland-Iceland-Faeroes Ridge, 326t Green River Formation, 378t Groundwater seepage, 298 Gryphon Field, 30t, 378t Guadalquivir River, Gulf of Ca´diz, Southern Spain, 196, 196f Guilin, 378t Gulf of Cadiz, 338 341, 338f, 372 375, 444 channel-current stage, 340 contour-current stage, 340 341 mixing and spreading stage, 340 Gulf of Lion, canyons on slope of, 470f Gulf of Mexico, 30t, 155 156 Gulf Stream, 349 Gyeongsang Basin, 378t

H Harding Field, 30t, 378t HARP units. See High-amplitude reflection packet (HARP) units HDTCs. See High-density turbidity currents (HDTCs) Hebrides Slope, 326t Hemipelagite, 119 120, 429t, 454 Hemiturbidite, 429t Heterolithic facies, 71, 454 Hiatuses, in contourites, 336 High-amplitude reflection packet (HARP) units, 153 High-concentration sandy turbidite, 429t High-density hyperpycnal plume, 270 High-density turbidity currents (HDTCs), 95, 98f, 99, 108 112, 110f, 122 123, 126f, 128f, 140, 149 150, 327 328, 454, 475 High Island, 30t, 314t High-resolution multibeam mapping system, 452 453 High-turbid mass flow, 270 Homogenite, 40t, 429t Homopycnal plume, 182, 443 444 Horizon Guyot, 314t Horse’s tail, 199 200, 219t Hueneme Canyon, 359t Hugli River, 183t, 219t Human activity, 298 299, 299f Hurricane Allen, 290t

Hurricane Camille, 289, 290t Hurricane Diana, 290t Hurricane Hugo, 289 Hurricane Isabel, 290t Hurricane Ivan, 289, 290t Hurricane Iwa, 290t Hurricane Joy, 290t Hurricane Katrina, 289 Hurricane Lili, 290t Hybrid flow, 40t, 325 327, 373 375, 444 Hybridite, 455 Hydrodynamics, 455 Hydrographer Canyon, 357 359 Hydroplaning, 233 234 Hydrostatic loading, 302 Hyperconcentrated flow, 95 Hyperpycnal flows, 140 144, 146, 187 189, 191 192, 200, 205 206, 209, 211, 455 456 academic discussions, 263 270, 264t challenges, 262 coalescing lobate plume, 260 continental margin, 224 cyclone-induced hyperpycnal turbidity currents in canyons, 253 254 definition, 214 density plumes configurations, 254 255 dissipating plume with irregular front, 256 259 external controls, 239 241 fine-grained deltas vs. coarse-grained deltas, 228 229 hyperpycnites ancient, 241 252 configurations of density plumes, 219t history, 216 217 incentive, 216 objective, 218, 218f problem, 217 and unresolved issues, 268 270 identification, 214 215 multilayer, 234 origin, 214, 215f plume vs. flow, 224 225 plunge point, 224 research on, 262 263 ring plume, 261 river currents transformation into turbidity currents, 228 vs. turbidity currents, 226 228 river-mouth flows, 225 226 scientific development, stages, 263 268 swirly cyclonic plume, 261 swirly plume, 262 tendril plume, 261 262

Subject Index

563

tidal lobate plume, 260 261 tide-modulated, 234 whitings plume, 261 Yangtze River, China, 237 239 Yellow River, China, 229 237 Hyperpycnal plume, 182, 443 444 Yangtze River, China, 237 Hyperpycnal turbidity currents, 224 cyclone-induced, 253 254 Hyperpycnal underflow, 233 Hyperpycnites, 224, 456 ancient, 241 252 configurations of density plumes, 219t facies model, 242 243 fan model, 246 248 history, 216 217 incentive, 216 objective, 218, 218f problem, 217 and unresolved issues, 268 270 Hypopycnal plumes, 182, 443 444 Yangtze River, China, 237

Inverse to normal grading, 243 244 IODP. See Integrated Ocean Drilling Program (IODP) IODP Expedition 339, 344 Isla Grande de Tierra del Fuego (Argentina), 378t Issyk-Kul lake (Kyrgyzstan, Central Asia), 378t

I

L

IGCP-511 (IUGS-UNESCO’s International Geoscience Programme 511), 34 IGCP-585 (E-MARSHAL), 34, 45 Impact cratering, 402 403, 405 Impactite, 429t Indian Ocean, 260, 397t Indian Ocean Tsunami (2004), 280 282 Indonesian Seas, 304 307 Inertia flow, 40t Inferred flow velocity, and sediment concentration, 128f Injectite, 161, 429t, 456 Insular Slope Slide, 28t Integrated Ocean Drilling Program (IODP), 469 471 Integrated Ocean Drilling Program (IODP) 339 cores, 444 Intercanyon facies, 125f Intermediate-term triggering events, 274, 444 Internal erosional surface, 244 Internal hiatuses, 342 Internalite, 429t Internal tides, 363 366, 456 Internal waves, 292, 294f, 295f, 363 366, 456 Yellow River, China, 234 International projects and symposiums, 34 35 Interpretite, 429t Intrabasinal sandstone clast, in mudstone matrix, 104f Intrabasinal turbidites, 247 248 Inverse grading, 456

Laacher See (Germany), 378t Lagoa Parda Field, 30t, 314t Lahar, 40t La Jolla Canyon, 362 363 Lake Baikal region (Russian Siberia), 378t Lake Erie, swirly plume, 262 Lake Michigan, tendril plume, 261 262 Lake Powell Delta, 378t Lakes MTDs in, 453 457 World’s deepest, 452t Laminar flow, 456 Laminar flow state vs. turbulent flow, 96 97, 97f Laminar inertia-flow, 40t, 97f, 128f, 177 178 Laminar mass flow, 40t Laminar sheared layer, 40t Laminite, 378t Landslide, 39, 88 transport vs. mass transport, 38 45 Lava delta, 279 280 Lenticular bedding, 359 Linear, 95 96, 228 229 Liquefaction, 128, 456 Liquefaction slide, 277 Liquefied cohesionless coarse-particle flow, 40t Liquefied flow, 40t Liquefied/fluidized flows, 146 case studies, 131 definition, 126 127

J Jackfork Group, Pennsylvanian, Ouachita Mountains, 169 171 Jupiter, 8 13

K Kaulakahi Canyon, 359t Kinetic sieving, 243 Kolankaya formation, 378t K-Pg boundary, 277 Krishna-Godavari (KG) Basin, Bay of Bengal, India, 116 120, 116f, 118t, 124t, 160 161 Krishna-Godavari Rivers, Bay of Bengal, India, 202, 203f, 204f Kuroshio Current, 200, 237, 326t

564

Subject Index

Liquefied/fluidized flows (Continued) facies models, 131 identification, 130 131 origin, 127 129 problem, 131 Liquidization, 130f, 377 396, 437 439 Lithofacies, 456 LIW, 319 Lixian (Eastern Tibetan Plateau), 378t Lobe, 456 Lobe fan, sieve, 451f Loess flow, 24t, 40t Lofting rhythmites, 245 246 Long-runout mechanisms, 74 82 concept, 74 79 extraterrestrial environments, 80 H/L ratio problems, 81 82 subaerial environments, 79 submarine environments, 79 80 Long-term triggering events, 444 Loop Current, 349 Los Frailes Canyon, 11 12, 100f Low-density hyperpycnal plume, 225, 233 Lowstand of sea level, 153, 304, 307, 444 Lucina Formation, 453

M Mackenzie River, 183t, 219t Makassar Strait, 30t, 304 307, 314t, 370 Mariana Trough, 397t Marimba Field, 30t Marlim Field, 30t Marnoso Arenacea Formation (Northern Italy), 378t Mars, 8 13 Mass creep, 40t Mass flow, 40t, 70, 94 Mass-flow lobes, Ulleung Basin, East Sea, Korea, 176 Massive, 183t, 219t Massive sand intervals, X-radiography for, 462 Massive sands, 113 reinterpretation of, 175 176 Massive sandstones, 245 Mass movement delta-front, 301 302 submarine, 22 23 Mass transport, 456 vs. landslide transport, 38 45 vs. turbidity currents, 92 93 Mass-transport complex (MTC), 28t, 49 Mass transport deposits (MTDs), 8 14, 22 23, 407 412, 409f, 441 442 deep-water case studies, 30t on Earth, 52t

in lakes, 453 457 sedimentology, 50 soft-sediment deformation structures (SSDS), 405 in subaerial environments, 28t submarine, 52t types, 49 50 worldwide large subaerial and submarine, 24t Mass-transport process, 47 48 fast-moving and slow-moving, 48 49 subaerial, 37 38 types, 40t Mass wasting, 49, 250, 300 M2 baroclinic tides, 304 307 Meanderite, 429t Mechanical behavior, subaqueous process on, 45 48 Mediterranean outflow water (MOW), 336, 339 channel-current stage, 340 contour-current stage, 340 341 mixing and spreading stage, 339f, 340 Mediterranean Sea Ebro Delta, 195, 195f Rhone River, 194 195, 195f Mediterranean undercurrent (MUC), 319 Megaturbidite, 40t, 49, 429t Melania formation, 30t Meltwater, 24t, 188 Messinian evaporates, 397t Meteorite breccias, 417 435 Meteorite impact, 277, 278f Microfolds, 378t, 397t Mid-Cape Basin (off South Africa), 397t, 406 Mid-Norway region, 30t Midway Sunset Field, 30t, 314t Miocene Monterey Formation (California), 414 416 Mississippi Canyon, 28t, 289, 290t Mississippi Fan, Gulf of Mexico channelized lobes, 155 156 Mississippi River, Gulf of Mexico, US, 191 192, 193f, 442f Miura Peninsula (Japan), 378t Mohr Coulomb failure criterion, 36 Monsoonal rainfall, 275t, 296 Monsoon flooding, 296 297, 296f, 297f Monterey Canyon, 156 160 Monterey Fan, North Pacific depositional lobe, 159 160 Monterey Canyon, 156 160 Monterey Formation, 378t, 414 416 Mornos River, 183t, 187 Mornos and Fonissa Rivers, Gulf of Corinth, Greece, 198, 198f Mount St. Helens, 24t, 28t, 52t, 79, 277 279 MTDs. See Mass transport deposits (MTDs)

Subject Index

M2 tidal dynamics, Yellow River (China), 236 237, 236f Mud-draped ripples, 118t Muddy debris flows, 106f Muddy debrites, 99 101 Muddy slumps, 119 120 Mudflow, 289, 300 302 Mud offshoot, 349 352, 463 Mudstone clast, 102f floating, 100f Mudstone clasts, floating, 453 Mudstone matrix, intrabasinal sandstone clast in, 104f Multibeam mapping system, 456 457 high-resolution, 452 453 Multilayer hyperpycnal flows, Yellow River (China), 234

N Namibe Basin (South West Angola), 378t Nankai Trough (Sea of Japan), 397t, 405 Narragansett Basin, 378t NASA (National Aeronautics and Space Administration), 182 for imaging density plumes, 462f satellites, 462f, 463, 463f National Geophysical Data Center (NGDC), US, 456 457 Natural debris flows vs. theoretical debris flows, 106f Natural sunlight, in core photography, 461, 461f Navajo Sandstone, 378t Newtonian fluids vs. plastic fluid rheology, 95 96, 96f Newtonian rheology, 133 135, 465 NGDC. See US National Geophysical Data Center (NGDC) Nigeria, 30t, 90 92, 359, 409 Niger River, 183t, 219t Nile River, 187 Nonuniform flow, 465 Normal grading, 465 466 North Adriatic dense water (NAdDW), 319 North Atlantic deep-water (NADW), 319, 328 330, 336 North Island (New Zealand), 378t North Pacific deep-water (NPDW), 319 Norwegian North Sea, 30t, 378t Norwegian sea deep-water (NSDW), 319 Nue´e ardente, 40t Nuuanu Debris Avalanche, 28t

O Ocean-bottom currents, 302 Ocean currents, Yangtze River (China), 237 Ocean Drilling Program (ODP), 469 471

565

Ocean Drilling Program (ODP) Site 1122 on Leg 181, 369 370 Oceanographic phenomena, 475 Ocean TOPography EXperiment, 464 ODP. See Ocean Drilling Program (ODP) Offshoot, 349 352, 463 Olistostrome, 40t, 49 Onibe River, 183t, 187, 219t Oockmarks, 411 412, 437 439 Opalo Field, 30t Orotava-Icod-Tino Debris, 28t Ouachita Mountains, 169 171 Outcrop, 466 Owyhee Mountains (Southwest Idaho), 378t

P Pacific Ocean Eel River, California, 191, 192f South Pacific Ocean, ring plume, 261 tidal lobate plume, 260 261 Paleocurrents, 208, 211 Pearl River, South China Sea, China, 200 201, 202f Pelagic sediment, 466 Pelagite, 466 Penecontemporaneous deformation structures. See Soft-sediment deformation structures (SSDS) Peroa´ Field, 30t, 314t Petroleum industry, 23 34 Petroleum reservoirs, 74, 176 177, 370, 417 435 Philippine Sea Plate, 397t, 409 Planar clast fabric, 466 Plant remains, 246 Plastic fluid, 45 47 Plastic fluid vs. Newtonian fluids rheology, 95 96, 96f Plastic rheology, 466 Pliocene age, erosional feature of, 124f Pliocene canyons, 117, 161 Plumes defined, 187 188 density, 182 types, 182, 188 homopycnal, 182 hyperpycnal, 182 hypopycnal, 182 river, 187 188 sediment, 187 188 wind forcing on, 205 207, 206f, 207f, 208f, 209f vs. flow, 224 225 Plumite, 429t Plunge point, 224 Pockmarks, 240, 378t Pore-water pressure, 37 38 excess, 38

566

Subject Index

Pore-water pressure (Continued) laboratory measurements of, 38 Prebetic Zone (Southeast Spain), 378t Prelithification deformation, 377 396, 437 439 Primary basal glide plane, 466 Process continuum, 331 333 Process sedimentology, 50, 342 343, 466 467 Progressive creep, 40t Ptygmatic folding, 71, 176 177, 378t Puga cap carbonate, 378t

Q Qua Iboe Canyon, 357 359, 362 363

R Razor-sharp planar margins, 56 Redondo Canyon, 357 359 Remote sensing technology, 456 457 Reservoir characterization, 83 86 facies in cores, 461 quality, 370 372, 371t sands, 119 120 Retrogressive flow slide, 40t Reynolds Number, 467 Rheology, 467 Newtonian, 465 of Newtonian fluids, 95 96, 96f Rhone River, Gulf of Lions, Mediterranean Sea, France, 194 195, 195f Ring plume, 261 Rio Balsas Canyon, 359t Rı´o de la Plata Estuary dissipating plume with irregular front, 256 259 South Atlantic Ocean, Argentina, and Uruguay, 192 193, 194f River currents transformation into turbidity currents, 228 vs. turbidity currents, 226 228, 227t River-mouth flow types, 225 226, 226f River-mouth process, Yellow River (China), 233 River-mouth study, 183t, 186 188 Brisbane River, Moreton Bay, Australia, 203, 204f Congo (Zaire) River, Atlantic Ocean, West Africa, 199, 199f Connecticut River, New England region, United States, Long Island Sound, US, 189, 192f Dart River, Lake Wakatipu, South Island, New Zealand, 203 204, 205f Ebro Delta, Mediterranean Sea, Iberian Peninsula, 195, 195f Eel River, California, Pacific Ocean, United States, 191, 192f

Elwha sediment plume, Strait of Juan de Fuca, United States, 189, 190f, 191f Guadalquivir River, Gulf of Ca´diz, Southern Spain, 196, 196f Krishna-Godavari Rivers, Bay of Bengal, India, 202, 203f, 204f Mississippi River, Gulf of Mexico, United States, 191 192, 193f Mornos and Fonissa Rivers, Gulf of Corinth, Greece, 198, 198f Pearl River, South China Sea, China, 200 201, 202f Rhone River, Gulf of Lions, Mediterranean Sea, France, 194 195, 195f Rio de la Plata Estuary, South Atlantic Ocean, Argentina, and Uruguay, 192 193, 194f Tiber River, Tyrrhenian Sea, Italy, 197 198, 197f Yangtze River, East China Sea, China, 200, 201f Yellow River, Bohai Bay, China, 199 200 River plume, 187 188 RMS amplitudes. See Root mean square (RMS) amplitudes Rock, 457 462, 469 avalanche, 40t, 49 fall, 40t, 45 47, 92 93 flow, 40t, 45 slide, 40t, 45 slump, 45 spread, 45 topple, 45 Rock-glacier creep, 40t Rock interpretation principles, 462 464 Rogue waves, 288 Roncador Field, 30t Root mean square (RMS) amplitudes, 467 Rotational landslide, 49 Rotational slip, 49, 83 84 Rukwa Rift Basin (Southwestern Tanzania), 378t Rupert River, 187 Ryukyu Trench, 326t

S Saffir Simpson Hurricane Scale, 475 476 Saharan Debris Flow, 28t, 52t Saidmarreh Slide, 28t, 52t Sakhalin Islands, 378t Salt movements, 300 301, 301f Salt River Canyon, 289 Samoan Passage, 326t Sand avalanche, 40t, 49 deep-lacustrine, 453 fall, 40t, 132 flow, 40t

Subject Index

injection, 55, 122f, 377 395, 397t Sand-clay-water mix, 107f Sand intervals massive, X-radiography for, 462 Sandy debris flow (SDF), 99, 101 116, 106f, 110f, 114f, 115f dominance of, 170 171 experiment, dimensions of flume, 106f, 107f flume experiments on, 477 478 composition of slurries, 477 deposits, 478 fronts of, 38 water-escape structures in, 111f Sandy debrites, 99 101, 114 117, 119 120, 161 internal layers in, 113 Sandy intervals of contourite facies models, 348 349 Sandy mass-transport deposits (SMTDs), 149 150, 301 302 deep-water case studies, 30t reservoirs, 23 34 types, 50 Sandy slide, 84 85, 279 280 Sandy slump, 50, 118t, 153 155, 161 San Francisco Bay, tidal lobate plume, 260 261 San Lucas Canyon, 133f Santa Barbara Channel, tidal currents in, 250 Santa Monica Canyon, 359t Santos Channel, 322 Santos Drift, 336 Sa˜o Luı´s Basin (Northern Brazil), 378t Satellite images, 182, 186, 190 192, 197 198, 200 201, 211 Saturn, 52t Scarp, 469 Scientific development stages, 263 268 Scientific drilling (SD), 395, 397t, 413, 437 439 at sea, 469 471 Scripps Canyon, 289, 302 304 SD. See Scientific drilling (SD) SDF. See Sandy debris flow (SDF) Sea, scientific drilling (SD) at, 469 471 Sea level, 306 307, 336, 444 Sea-level lowstand, 304 307, 305f, 306f SeaMARC 1 (Seafloor Mapping and Remote Characterization) system, 456 457 Sea overflow water (SOW), 319 Seasonal creep, 40t, 48 Sea surface temperature (SST), 256 260 Secondary glide plane, 471 Sediment, deep-marine, 453 Sedimentary-collapse breccias, 417 435 Sedimentary features, 461 Sedimentation, deep-water, 1 2

567

Sediment concentration gravity-driven downslope process, 97 inferred flow velocity and, 128f Sediment deformation, 471 tectonics environments on, 402 Sediment deformation structures (SSDS) breccias problem, 417 435, 431f, 432f, 433f, 434f, 435f Sediment failure and sliding, 35 Sediment flows, 471 gravity-driven, 98f Sediment flux, 471 Sediment-gravity flows, 93 95, 94f, 135 136, 336, 471 classification, 142f transformations for, 123 126 Sedimentological log, grain-size scale on, 457 459, 458f, 459f Sedimentological log sheet, 458f Sedimentology, 464f, 466 467, 471 depositional process, 466 467 mass transport deposits (MTDs), 50 Sediment, pelagic, 466 Sediment plumes, 187 188 satellite images of, 1 2 wind forcing on, 205 207, 206f, 207f, 208f, 209f Sediment provenance current directions, 366 369 detrital composition, 369 370 implications for, 210 211 Sediment slurries, 105 108, 477 Sediment-support mechanisms, 93 Sediment transport, implications for, 207 210, 210f Seiche, 239 240, 262 Seismic facies, 475 Seismic geometries, of contourites, 322 Seismic geometry, 322 Seismicity, 131 Seismite, 378t, 472 Seismogenic slump folds, 70 Seismoturbidite, 40t, 429t Sequence boundaries, 336 Sequence-stratigraphic model, 153 Sequence stratigraphy, 216 Sevier Basin, 30t, 314t Shale clasts, 344 346 Shallow-water processes, 340 341 Shear strength, 472 Sheet geometry, 363 Short-term triggering events, 444 Sieve lobe fan, 451f Sigmoidal cross-bedding, 164 165 Sigmoidal deformation structures, 63 68 Sigmoidal slices, tectonic origin for, 63 68 Silica sand, 477

568

Subject Index

Simple lobate, 219t, 254 Sinuous channels, 150 153 Slides case studies, 63 definition, 50 53 facies models, 63 identification, 55 62 methods for recognizing, 456 457 origin, 54 55 problems, 63 submarine, 50 52 Sliding, sediment failure and, 35 Slope continental, erosional features on, 473 movement, 38, 49 stability, 35 38 stability analyses, 472 Slump-creep, 40t Slump folds, 378t, 402, 409, 437 439 Slumps case studies, 72 definition, 63 deposit, 170 171 earthquake-induced gravity tectonics, 70 facies models, 72 identification, 71 methods for recognizing, 456 457 origin, 63 70 problems, 72 submarine channel deposition, 63 70 Slurry flow, 40t, 49, 475 SMTDs. See Sandy mass-transport deposits (SMTDs) Soft-sediment deformation, 128 Soft-sediment deformation structures (SSDS), 22, 128 131, 377 395, 437 439, 444 445, 472 advances, 402 405, 404f breccias problem, 417 435, 431f, 432f, 433f, 434f, 435f case study, 378t, 395, 396f, 397t classification, 401 402, 403f deep-water deposits, 437 439 definition, 395 396 future research, 413, 426f, 427f, 428f geological implications, 405 436 breccias, 405, 406f lateral extent, 405, 407f mass-transport deposits, 407 412, 409f, 410f, 411f, 412f, 413f, 414f, 415f, 416f, 417f, 418f, 419f, 420f, 421f, 422f, 423f, 424f ocean bottom currents, 406, 408f high-density turbidity currents, 439 origin, 396 400 as “seatmates”, 417 435 seismite problem, 414 416, 429t

tombolith problem, 436, 436f, 437f, 438f Soil strength, 35 37 wet sandy, 36 37 Soil creep, 40t Soil mechanics, 473 Solifluction, 40t South Atlantic central water (SACW), 319 South China Sea, 200 201, 202f Southern Ocean current (SOC), 322 South Marsh Island, 30t, 314t South Pacific Ocean, ring plume, 261 South Timbalier, 30t, 314t Spread, 40t Squamish Delta, 243, 274 SSDS. See Soft-sediment deformation structures (SSDS) Stable slope, 35 STEAM (Sediment Transport on European Atlantic Margins), 34 Storegga Slide, 302, 304 Straits of Florida, 314t, 373 375 Straits of Gibraltar, 326t STRATAFORM (STRATA FORmation on the Margins), 34 Stratified flows, 95 Stromatolite mats, 378t Sturzstrom, 40t, 48 Subaerial debris flows, coarse-grained, 37 38 Subaerial environments, 9 11, 13 mass transport deposits in, 28t Subaerial mass transport classification, 45 deposits, 24t process, 37 38 Subaerial process, on movement and material, 45 Subaqueous debris flows, 108 Subaqueous process, on mechanical behavior, 45 48 Subcritical fans, 177 179 Sublacustrine environment, 9 Submarine canyons, 117, 466f, 467f, 468t, 471f, 472f, 473 475, 474f classification, 251 252, 251f, 252f current velocity, 357 359, 358f, 359t identification, 359 363, 360f, 361f, 362f, 363f origin, 250 251 shelf-indenting, 116 117 types of, 356 357 Submarine channels, 150 Submarine environment, 8, 11 13 Submarine fan, 177 179, 216, 473 Submarine landslide, 40t, 49 Submarine mass failure (SMF), 49 Submarine mass movements, 22 23

Subject Index

Submarine mass transport deposits, 24t, 52t, 54 55 Submarine slides, 50 52 Supercritical fans, 177 179 Supercritical hyperpycnal flow, 225 Surface currents, 310 312 Surface transformation, 48 49 Suspended sediment concentration (SSC), 224 Suspensite, 429t Swatch-No-Ground Canyon, 470f Swirly, 183t Swirly cyclonic plume, 261 Swirly plume, 262 Synsedimentary deformation structures. See Softsediment deformation structures (SSDS)

T Tabular carbonate clast, 103f Talme Yafe Formation, 314t Talus creep, 40t Tectonics contractional, 395 deformation, 63 68, 88 environments on sediment deformations, 402 oversteepening, 300 Tectonite, 429t Tempestite, 429t Tendril plume, 261 262 Terra satellite, 462f, 469f Teruel half-graben basin (Eastern Spain), 378t The 1929 Grand Banks MTD, 24t, 28t, 456 Thermohaline circulation, 313 320 Thermohaline contour currents definition, 145 downslope initiation, 145 146 Thermohaline-driven contour currents, 334 335 Thermohaline-induced geostrophic bottom currents Antarctic bottom water, 322 current velocity, 323 324 seismic geometries of contourites, 322 traction structures, 324 327 Thermohaline water masses, 310 Tiber River, Tyrrhenian Sea, Italy, 197 198, 197f Tidal bottom currents, 356 363 Tidal currents deep-marine, 335, 368 ebb-tidal, 292 296 Santa Barbara Channel, 250 Tidalite, 118t, 252, 372, 429t Tidal lobate, 219t plume, 260 261 Tidal rhythmites, 245 246 Tidal river dynamics, Yangtze River (China), 237 239 Tidal shear front, Yellow River (China), 235 236, 235f

569

Tide-modulated hyperpycnal flows, Yellow River (China), 234 Tides, 292, 294f, 295f internal, 456 Tillite, 429t Tonga Trench, 326t, 397t Topex/Poseidon satellite, 464 Topple, 40t Toreva-block, 40t, 49 Total stress, 473 Traction carpet, 40t Tractionite, 429t Traction structures, 244 245, 324 327, 328f, 332f, 344 346, 444 Translational landslide, 49 Translational slide, 38, 83 84 Translational slump, 40t Transport velocity, 48 49 Tratropical cyclone, 190, 205 206 Triassic Yanchang formation, normal grading in, 244 Triggering processes, 273 274 biological erosion, 302 304 cyclonic waves, 288 292, 290t, 291t depositional loading, 301 302, 303f earthquakes, 274 277 ebb-tidal currents, 292 296 gas-hydrate decomposition, 304 glacial maxima and loading, 300 groundwater seepage, 298 human activity, 298 299 hydrostatic loading, 302 internal waves, 292, 294f, 295f meteorite impact, 277, 278f monsoon flooding, 296 297, 296f, 297f ocean-bottom currents, 302 rogue waves, 288 salt movements, 300 301, 301f sea-level lowstand, 304 307, 305f of sediment failures, 275t tectonic oversteepening, 300 tides, 292, 294f, 295f tsunami wave, 280 287, 280f, 281f, 283f volcanic activity, 277 280, 279f wildfire, 298 Triggers of sediment failures, 444 Tropical cyclone, 473 475 Tropical Storm Floyd, 290t Tropical Storm Ida, swirly cyclonic plume, 261 Tsunami, 475 Tsunamite, 40t, 429t Tsunami wave, 280 287, 280f, 281f, 283f Turbidite, 475 myths, 475

570 Turbidite channels, 337 Turbidite-dominated fan model, 170 171 Turbidite exuberance, 149 150 Turbidite facies model, 139f, 165, 175 176, 244 Turbidite-fan link, 138f Turbidite lobe, 363, 463 Turbidity currents case studies, 136 and contour currents, continuum between, 331 333 definition, 133 135, 142f facies models, 136 138 identification, 136 methods for recognizing, 457 origin, 135 136 problem, 138 140 river currents transformation into, 228 river currents vs.,, 226 228, 227t vs. mass transport, 92 93 Turbidity front, 225 Turbulent flow, 475 Turbulent state, 133 135

U Undaturbidite, 429t Underflow, 225, 227 Undrained condition, 475 Undrained slump, 40t Unifite, 429t Unsteady flow, 476 Upper Triassic Yanchang Formation, Ordos Basin, central China, 176 177 Upwelling, 200 201, 219t, 240 U.S. Geological Survey, 22 23 US National Geophysical Data Center (NGDC), 456 457 Usoy, 24t, 28t U-Turn, 183t, 196, 219t

V Var submarine canyon, 95 Velocity measurements, Yellow River (China), 234 transport, 48 49 Venus, 52t, 81 82 Victoria Land Basin (Antarctica), 378t Viscous fluid, 45 47, 92 93 Volcanic activity, 277 280, 279f Volcanic ash, 254, 466 Volcanic eruptions, 240, 417 435

Subject Index

W Waitemata Group, 378t Walther’s Law, 244 Waning flow, 476 Warm deep-water (WDW), 319 Water-escape structures in deep-water sands, 108 112 in sandy debris flows, 111f Wavy bedding, 118t Waxing flow, 476 WBU/WBUC. See Western boundary undercurrent (WBU/WBUC) WDW. See Warm deep-water (WDW) Weddell sea bottom water (WSBW), 319 Weddell sea deep-water (WSDW), 319 West Bermuda Rise, 326t Western boundary undercurrent (WBU/WBUC), 319 Wet sandy soil, 36 37 Whitings plume, 261 Wildfire, 298 Wilmington Canyon, 359t Wind-driven bottom currents, 367 368 current velocity, 349 352 Ewing Bank Block 826 Field, 352 356 Gulf Stream, 349, 350f, 351f Loop Current, 349 Wind-driven Loop Current, 335 Wind forcing, on sediment plumes, 205 207, 206f, 207f, 208f, 209f Winnowite, 429t Wireline logs, 84, 124f, 172 175 World’s deepest lakes, 452t WSBW. See Weddell sea bottom water (WSBW) WSDW. See Weddell sea deep-water (WSDW)

X X-radiography, for massive sand intervals, 462

Y Yangtze River, China, 200, 201f hyperpycnal and hypopycnal plumes, 237 ocean currents, 237 tidal river dynamics, 237 239 Yellow River, China, 199 200 bathymetry, 232f, 233 bottom-turbid layers, 233 234 delta vs. estuary, 231 232 drains, 230 internal waves, 234 M2 tidal dynamics, 236 237, 236f

Subject Index

multilayer hyperpycnal flows, 234 river-mouth processes, 233 tidal shear front, 235 236, 235f tide-modulated hyperpycnal flows, 234 velocity measurements, 234 Yucatan, 277, 349

Z Zafiro Field, 30t, 137f Zaire Canyon, 357 359 Zambezi River, coalescing lobate plume, 260 Zhemchug canyon, 473 Zion National Park (Utah, USA), 378t

571