Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

New Perspectives on Deep-water Sandstones: Origin, Recognition, Initiation, and Reservoir Quality
New Perspectives on Deep-water Sandstones: Origin, Recognition, Initiation, and Reservoir Quality
New Perspectives on Deep-water Sandstones: Origin, Recognition, Initiation, and Reservoir Quality
Ebook844 pages8 hours

New Perspectives on Deep-water Sandstones: Origin, Recognition, Initiation, and Reservoir Quality

Rating: 0 out of 5 stars

()

Read preview

About this ebook

This handbook is vital for understanding the origin of deep-water sandstones, emphasizing sandy-mass transport deposits (SMTDs) and bottom-current reworked sands (BCRSs) in petroleum reservoirs. This cutting-edge perspective, a pragmatic alternative to the conventional turbidite concepts, is crucial because the turbidite paradigm is built on a dubious foundation without empirical data on sandy turbidity currents in modern oceans. In the absence of evidence for sandy turbidity currents in natural environments, elegant theoretical models and experimental observations of turbidity currents are irrelevant substitutes for explaining the origin of sandy deposits as "turbidites." In documenting modern and ancient SMTDs (sandy slides, sandy slumps, and sandy debrites) and BCRSs (deposits of thermohaline [contour] currents, wind-driven currents, and tidal currents), the author describes and interprets core and outcrop (1:20 to 1:50 scale) from 35 case studies worldwide (which include 32 petroleum reservoirs), totaling more than 10,000 m in cumulative thickness, carried out during the past 36 years (1974-2010). The book dispels myths about the importance of sea level lowstand and provides much-needed clarity on the triggering of sediment failures by earthquakes, meteorite impacts, tsunamis, and cyclones with implications for the distribution of deep-water sandstone petroleum reservoirs.

  • Promotes pragmatic interpretation of deep-water sands using alternative possibilities
  • Validates the economic importance of  SMTDs and BCRS in deep-water exploration and production
  • Rich in empirical data and timely new perspectives 
LanguageEnglish
Release dateJan 25, 2012
ISBN9780444563552
New Perspectives on Deep-water Sandstones: Origin, Recognition, Initiation, and Reservoir Quality

Related to New Perspectives on Deep-water Sandstones

Titles in the series (3)

View More

Related ebooks

Earth Sciences For You

View More

Related articles

Reviews for New Perspectives on Deep-water Sandstones

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    New Perspectives on Deep-water Sandstones - G. Shanmugam

    New Perspectives on Deep-water Sandstones

    New Perspectives on Deep-water Sandstones

    Origin, Recognition, Initiation and Reservoir Quality

    G. Shanmugam, Ph.D.

    Department of Earth and Environmental Sciences, The University of Texas at Arlington, Arlington, Texas, U.S.A.

    ISSN  1567-8032

    Volume 9 • Number Suppl C • 2012

    New Perspectives on Deep-water Sandstones

    Handbook of Petroleum Exploration and Production, 9

    New Perspectives on Deep-water Sandstones

    Copyright

    Table of Contents

    Instructions for online access

    Cover

    Title Page

    New Perspectives on Deep-water Sandstones

    Handbook of Petroleum Exploration and Production, 9

    New Perspectives on Deep-water Sandstones

    Copyright

    Dedication

    Preface

    Acknowledgments

    Chapter 1: Introduction

    1.1 WHAT IS DEEP WATER?

    1.2 FLAWED TURBIDITE PARADIGM

    1.3 NEW PERSPECTIVES: SMTDS AND BCRS

    1.4 DATABASE

    1.5 SCOPE AND ORGANIZATION

    1.6 PROCESS SEDIMENTOLOGY

    1.7 SYNOPSIS

    Chapter 2: Origin and Classification of Sandy Mass-Transport Deposits

    2.1 INTRODUCTION

    2.2 LITERATURE

    2.3 CLASSIFICATION

    2.4 LANDSLIDE VERSUS MASS TRANSPORT

    2.5 SUBAERIAL PROCESSES BASED ON TYPES OF MOVEMENT AND MATERIAL

    2.6 SUBAQUEOUS PROCESSES BASED ON MECHANICAL BEHAVIOUR

    2.7 SUBAQUEOUS PROCESSES BASED ON SEDIMENT-SUPPORT MECHANISM

    2.8 SUBAQUEOUS PROCESSES BASED ON PROCESS CONTINUUM

    2.9 SUBAQUEOUS PROCESSES BASED ON TRANSPORT VELOCITY

    2.10 SYNOPSIS

    Chapter 3: Recognition of Sandy Mass-Transport Deposits

    3.1 INTRODUCTION

    3.2 SANDY SLIDE

    3.3 SANDY SLUMP

    3.4 SANDY DEBRITE

    3.5 ORIGIN OF MASSIVE SANDSTONE

    3.6 PROBLEMS WITH INTERPRETATION OF WIRELINE LOGS

    3.7 PROBLEMS WITH INTERPRETATION OF SEISMIC FACIES

    3.8 PROBLEMS ASSOCIATED WITH INTERPRETATION OF SEISMIC SINUOUS GEOMETRY

    3.9 SYNOPSIS

    Chapter 4: Bottom-Current Reworked Sands

    4.1 INTRODUCTION

    4.2 SURFACE CURRENTS, DEEP-WATER MASSES, AND BOTTOM CURRENTS

    4.3 BOTTOM CURRENTS VERSUS TURBIDITY CURRENTS

    4.4 GENETIC NOMENCLATURE

    4.5 THERMOHALINE-INDUCED GEOSTROPHIC BOTTOM CURRENTS

    4.6 WIND-DRIVEN BOTTOM CURRENTS

    4.7 DEEPWATER TIDAL BOTTOM CURRENTS

    4.8 BAROCLINIC CURRENTS (INTERNAL TIDES)

    4.9 PROBLEMATIC BEDFORM-VELOCITY MATRIX FOR DEEP-WATER BOTTOM CURRENTS

    4.10 PROBLEMS WITH INTERPRETATION OF SEISMIC FACIES AND GEOMETRIES

    4.11 SYNOPSIS

    Chapter 5: Initiation of Deep-Water Sediment Failures

    5.1 INTRODUCTION

    5.2 SHORT-TERM TRIGGERING EVENTS

    5.3 INTERMEDIATE-TERM TRIGGERING EVENTS

    5.4 LONG-TERM TRIGGERING EVENTS

    5.5 SYNOPSIS

    Chapter 6: Implications for Deep-water Sandstone Reservoirs

    6.1 GRAIN-SIZE DISTRIBUTION

    6.2 DIMENSIONS AND GEOMETRIES

    6.3 LONG-RUNOUT MTD

    6.4 TURBIDITES VERSUS DEBRITES

    6.5 TURBIDITES VERSUS TIDALITES

    6.6 SMTD AND BCRS, GULF OF MEXICO

    6.7 CHICXULUB METEORITE IMPACT, GULF OF MEXICO

    6.8 SAND INJECTION

    6.9 SEQUENCE STRATIGRAPHY

    6.10 SYNOPSIS

    Chapter 7: Reservoir Quality—Global Examples

    7.1 OFFSHORE CALIFORNIA

    7.2 OFFSHORE NIGERIA

    7.3 GULF OF MEXICO

    7.4 STRAITS OF FLORIDA

    7.5 U.K. NORTH SEA

    7.6 KRISHNA–GODAVARI BASIN, BAY OF BENGAL

    7.7 SYNOPSIS

    Chapter 8: Epilogue

    Appendix A: Concepts, Glossary, and Methodology

    DEEP-WATER SYSTEMS

    METHODOLOGY FOR RECOGNIZING MTD

    References

    Index

    About the Author

    Dedication

    Preface

    This volume is a follow-up to my earlier book "Deep-Water Processes and Facies Models: Implications for Sandstone Petroleum Reservoirs, published under Elsevier's Handbook of Petroleum Exploration and Production series (Volume 5). In that book and in my other publications, I have discussed the obsolescence of turbidite facies models and have advocated the importance of sandy mass-transport deposits (SMTDs) and bottom-current reworked sands (BCRSs) as deep-water reservoir facies. Although my views were considered dissident not so long ago, these views are becoming increasingly mainstream" today. In advancing this new perspective, the primary goal of this book is to consolidate the rock data and convince the reader of the importance of SMTD and BCRS in petroleum geology. This book is an antidote to the current complacency-ridden interpretation of deep-water sands based on turbidite and contourite facies models.

    In describing deep-water lithofacies using conventional core and outcrop, my experience (1978–2000) with Mobil Oil Corporation (now ExxonMobil, USA) and subsequent consulting experience with the petroleum industry worldwide (2002–2010), including Reliance Industries Limited (India), Oil and Natural Gas Corporation (India), Hardy Exploration and Production (India) Inc., and Research Institute of Petroleum Exploration and Development (RIPED) of PetroChina (China) have provided me a unique opportunity to examine the rock data and acquire an appreciation for the complexity of deep-water marine and lacustrine facies. I have included examples from published core studies conducted for Reliance. By design, this book is a cumulative reflection of my experience over the past 50 years. To maintain conceptual continuity, I have repeated critical portions from my previous publications.

    Although this volume is intended for a wide range of knowledge levels, including students, academics, researchers, and professional petroleum geoscientists, 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 new perspectives from a petroleum reservoir viewpoint, and (5) included an illustrated appendix on concepts, glossary, and methodology. In reflecting the proliferation of online publications, I have cited apt websites. The book contains 367 figures and 992 references. The print version of the book contains a color plate at the end of the book with 20 of original color figures. The remaining original color figures appear as greyscale images. The online version of the book contains all original color figures. Figure captions are written for color images.

    August 21, 2011

    G. Shanmugam, Ph.D.

    E-mail address: shanshanmugam@aol.com

    Acknowledgments

    Introduction

    Dr. G. Shanmugam

    Deep-water sandstone reservoirs have continued to be a major economic asset to the petroleum industry in the twenty-first century. In the deep-water Gulf of Mexico, for example, exploration efforts in 2008 alone have resulted in 15 new deep-water discoveries (Nixon et al., 2009). Exploratory drilling has found more than 6.6 billion barrels of oil equivalent (BBOE) since 2002, more than double the amount reported in 2005. On the production side, several large deep-water projects have transitioned from an exploratory and appraisal phase into production phase. At present, there are nearly 4,000 active oil and gas platforms (Figure 1.1). Since 1996, 17 deep-water discoveries have been made in the Lower Tertiary Trend in the northern Gulf of Mexico (Table 1.1). The significance of these 17 discoveries is that they not only occur in the present-day water depths of greater than 5,000 feet (1,524 m), but also comprise major reservoirs that were deposited in ancient deep-water settings, such as the BAHA discovery (Meyer et al., 2007).

    FIGURE 1.1 Map showing locations of nearly 4,000 active oil and gas platforms in the northern Gulf of Mexico.

    Image credit: National Oceanic and Atmospheric Administration, U.S. Department of Commerce. http://oceanexplorer.noaa.gov/explorations/06mexico/background/oil/media/platform_600.html. Accessed March 27, 2011.

    TABLE 1.1 List of 17 Discoveries in Present-Day Water Depths Greater than 5,000 ft (1,524 m) in the Northern Gulf of Mexico. Petroleum Reservoirs in these Discoveries are Associated with the Lower Tertiary Trend.

    Despite the enormous economic importance of deep-water reservoir sandstones worldwide, their origin is still mangled in controversies and confusion. I have made an attempt to untangle this sedimentological mess in an earlier volume (Shanmugam, 2006a). In continuing that effort, the primary purpose of this volume is to provide the much-needed clarity by explaining the inherent problems with the prevailing practice of interpreting deep-water sands as turbidites and by providing alternative options through systematic documentation of processes in modern oceans and their deposits in the geologic record using underwater photographs, side-scan sonar images, velocity measurements, conventional core, and outcrop data.

    1.1 WHAT IS DEEP WATER?

    Controversies still abound surrounding the interpretation of ancient sandstone as deep water in origin (Mulder et al., 2009a; Higgs, 2010; Mulder et al., 2010). The term deep water is used with different meanings by geoscientists and by drilling engineers in the petroleum industry. For example, geologists use the term to denote deepwater depositional origin of the subsurface reservoir, whereas drilling engineers use the term to denote the present-day drilling depths for the target reservoir, irrespective of its depositional origin (see Deep water in Appendix A).

    In general, the term deep water refers to areas in bathyal water depths (>200 m), which occur seaward of the continental shelf break, on the continental slope, continental rise, and abyssal plain environments. However, there is no consensus on the precise water depth that defines deep water. Pickering et al. (1989) use the term deep water to denote environments that occur exclusively below storm wave base. The depth of storm wave base is not a constant value, and it varies with the wind velocity of tropical cyclones. The maximum sustained wind velocity of cyclones changes from 61 km h−1 for a tropical depression to more than 249 km h−1 for a Category 5 hurricane in the Saffir–Simpson Scale (see Tropical Cyclone in Appendix A). Typically, storm wave base occurs in water depths ranging from 20 to 30 m on the continental shelf during low-intensity cyclones. However, the storm wave base may reach the shelf break and beyond (>200 m) during high-intensity cyclones, resulting in sediment transport beyond the shelf edge (Chapter 5). Therefore, the storm wave base, which varies between 20 m and >200 m, is not an objective criterion. Plus, the validity of hummocky cross-stratification, which is commonly used as a criterion for establishing storm wave base, is in dispute (Mulder et al., 2009a; Higgs, 2010).

    In the Gulf of Mexico, the threshold that separates shallow water from deep water ranges from 200 to 457 m (Richardson et al., 2004). The U.S. Department of the Interior uses the terms deep water and ultra-deep water for water depths greater than or equal to 1,000 feet (305 m) and 5,000 feet (1,524 m), respectively (Nixon et al., 2009). Gore (1992) considers continental shelf to occupy in water depths less than 180 m. On the continental margin off northwestern Africa, the shelf break is found invariably at 100–110 m (Seibold and Hinz, 1974). Hesse and Schacht (2011) prefer a water depth of 500 m for defining deep water in order to exclude deposition on the upper slope during periods of sea-level lowstand. In lacustrine basins that may not have well-developed shelf breaks, the defining criterion of deep water is problematic. In an attempt to resolve these basic issues, I suggest the following guidelines.

    1. Because shelf edge plays a critical role on continental margin sedimentation (Stanley and Moore, 1983), shelf edge is used here as the defining criterion of deep-water settings in modern oceans. Examples are the Gulf of Mexico (Figure 1.2) and the U.S. Atlantic margin (Figure 1.3, see color plate), among others. However, the seafloor topography (Figure 1.3, see color plate) and the corresponding variation in water depth are extremely complex and unpredictable (Figure 1.4). Shelf edge is a useful physiographic conceptual boundary in separating shallow water from deep-water settings. This is because shallow-water shelf setting is dominated by wave and tidal processes, whereas deep-water slope setting is dominated by gravity-driven processes. Although absolute water depths are difficult to interpret in older strata, shelf facies and slope facies may be distinguished with reasonable certainty.

    2. The shelf-edge criterion has its limitations. First, in areas such as the Indonesian seas, where a complex array of passages linking shallow and deep seas are complicated by the Indonesian throughflow (ITF) and dissipation of tidal energy (Gordon et al., 2010), distinguishing shelf facies from slope facies in the ancient counterpart could be a challenge (Section 4.8.5). Second, shelf-edge criterion is obsolete in submarine canyon settings because the shelf-slope break does not exist within submarine canyons. Canyons serve as a single environmental entity with increasing water depths from estuary to canyon (e.g., Zaire Canyon and West Africa, Section 4.7.1). In such cases, shelf facies and slope facies are absent. However, outside of the canyon, the shelf-slope break is an important physiographic boundary between the two major submarine provinces, namely shelf and slope (Vanney and Stanley, 1983). Most canyon-fill deposits are composed of mass-transport deposits (MTDs) and tidalites, mimicking deepwater and shallow-water facies, respectively (Shanmugam, 2003).

    3. In the modern Bering Sea (Figure 1.5, see color plate), the shelf-slope break straddles between 150 m and 175 m (Carlson and Karl, 1988). The importance of the Beringian margin in this discussion is that it contains the world's largest submarine canyon with a measured volume of 5,800 km³ of material removed from the shelf and slope (Carlson and Karl, 1988). The 1,400-km-long Beringian continental slope extends from the Aleutian Islands in the south to the Siberian margin in the north (Figure 1.5). The steep Beringian slope, with an average gradient of about 5°, separates the shallow (<150 m in water depth) Bering Shelf to the east and the deep (>3,600 m in water depth) Aleutian Basin to the west. The complicating factor here is the presence of earthquake-prone Aleutian Trench that to the south (Figure 1.6). In such a complex setting, both deep-water and shallow-water facies are likely to be severely deformed and be subjected to sand injection (Section 6.8).

    4. Conventionally, deep-water facies have been distinguished using physical, biological, and chemical parameters in the rock record (Rich, 1950; Krumbein and Sloss, 1963; Benedict and Walker, 1978; Shanmugam and Benedict, 1983). Stratigraphic position (Krumbein and Sloss, 1963; Shanmugam, 1978; Shanmugam and Walker, 1978) and facies association (Reading, 2001) are commonly used for inferring ancient depositional environments. Although various isotopes are useful for reconstructing paleoclimates and paleo-ocean circulations, they do not reveal information on water depths.

    5. Because modern deep-water slope environments in both the Gulf of Mexico (McAdoo et al., 2000) and in the U.S. Atlantic margin (Twichell et al., 2009), among others, are characterized by MTDs, the dominance of subaqueous MTDs in the rock record could be used as a criterion for interpreting ancient deep-water marine environments.

    6. Most lacustrine basins are shallow and low-relief entities; however, high-relief, deep-water lacustrine facies have been documented to be dominated by mass-transport deposits (Link and Osborne, 1978). Piston core obtained from 675 m water depth in the world's deepest Lake Baikal (maximum water depth: 1,637 m) in Siberia shows sandy slump (Charlet et al. 2005). In North America's second deepest (622 m) Crater Lake in Oregon, base-of-slope aprons are characterized by coarse-grained facies deposited by mass-transport processes that include rockfall, landslides, slumps, grain flows, and debris flows (Nelson et al., 1986). Therefore, the presence of MTDs is a reasonable criterion for interpreting ancient deep-water lacustrine settings as well.

    FIGURE 1.2 Three-dimensional perspective view of the Gulf of Mexico showing the complex bathymetry and submarine topography. The term deep water is used for areas that occur seaward of the shelf edge. For details on geology of the Gulf of Mexico, see Antoine (1972), Uchupi (1975), and Gore (1992). FL, Florida.

    Image credit: National Oceanic and Atmospheric Administration, U.S. Department of Commerce. http://oceanexplorer.noaa.gov/technology/tools/mapping/media/gis_gulf.html. Accessed March 27, 2011.

    FIGURE 1.3 Image showing shelf edge and deep-water part of the U.S. Atlantic margin. Note complex seafloor topography. The Latitude 31°30΄N Transect line that extends from the Georgia coast to a depth of 2,000 m (see Figure 1.4 for bathymetric profile along this transect). FL, Florida.

    Image courtesy of P. Weinbach, SCDNR. Image credit: National Oceanic and Atmospheric Administration, U.S. Department of Commerce. http://oceanexplorer.noaa.gov/explorations/04etta/background/plan/media/sig.html. Accessed March 27, 2011.

    FIGURE 1.4 Bathymetric profile along the Latitude 31°30΄N Transect (see Figure 1.3 for location of the Transect). This transect is part of NOAA's expedition Estuary to the Abyss: Exploring along the Latitude 31°30΄N Transect (August 20 to September 1, 2004). The purpose was to extend the transect out to 2,000 m depth, using a variety of sampling procedures (Chief Scientist: George Sedberry). Note shelf edge is at about 100 m along the transect. Seafloor is highly irregular due to deep-water banks, such as the Charleston Bump. This bump deflects the Gulf Stream offshore and causes eddy.

    Credit: National Oceanic and Atmospheric Administration, U.S. Department of Commerce. http://oceanexplorer.noaa.gov/explorations/04etta/background/plan/media/profiles.html. Accessed March 27, 2011.

    FIGURE 1.5 The Southern Alaska Coastal Relief Model showing bathymetry and topography of the Gulf of Alaska, Bering Sea, Aleutian Islands, and Anchorage city. Approximate positions of major submarine canyons (Navarinsky, Zhemchug, Pribilof, and Bering) are after Carlson and Karl (1988). This relief model was built from a variety of source datasets acquired from the National Geophysical Data Center, National Ocean Service, United States Geological Survey, National Aeronautics and Space Administration, and other U.S. and international agencies.

    Image credit: Lim, E., B.W. Eakins, and R. Wigley, Coastal Relief Model of Southern Alaska, National Geophysical Data Center, NESDIS, NOAA, 2009. http://www.ngdc.noaa.gov/mgg/coastal/s_alaska.html. Accessed March 31, 2011.

    FIGURE 1.6 Plot of earthquakes along the Aleutian trench and vicinity in the Southern Alaska Coastal Relief Model. This image was issued as part of a Tsunami Warning Bulletin Number 001 in response to an M=7.3 earthquake that occurred on June 24, 2011, in Fox Islands, Aleutian Islands. Note the location of an epicenter near the Zhemchug canyon.

    http://ptwc.weather.gov/ptwc/?region=1&id=pacific.2011.06.24.032024&msg. Accessed July 8, 2011.

    1.2 FLAWED TURBIDITE PARADIGM

    The turbidite paradigm has evolved into an influential conceptual entity during the past six decades (Figure 1.7). However, the paradigm is defective because it is built mostly on facies models (e.g., the Bouma Sequence, see Appendix A) derived from the ancient rock record, not on empirical data of turbidity currents from modern oceans. Various turbidite models have been so deeply entrenched in the psyche of geoscientists that virtually every deep-water sand can be and has been interpreted as some kind of turbidite (Figure 1.8). The paradigm has been so influential that even deposits of debris flows are called turbidites (Figure 1.9). The very basic tenet of field geology, which is to maintain a distinction between observation and interpretation, has been lost because of turbidite facies models and related mind-set (Shanmugam, 1997a). An extreme example of this turbidite mind-set can be found in a recent book chapter by Mulder (2011). In promoting the value of turbidite facies models, Mulder (2011) has reused one of my illustrations (Figure 1.10) that show the possibility of 16 divisions in an ideal turbidite bed. Ironically, I constructed this figure by correlating three turbidite facies models of (1) Lowe (1982), (2) Bouma (1962), and (3) Stow and Shanmugam (1980) to illustrate that an ideal turbidite bed should be composed of 16 divisions (Figure 1.10). My intention was to illustrate the absurdity of turbidite facies models with 16 divisions in a single bed that no one has ever documented either in the stratigraphic record or in flume experiments! But Mulder (2011) has totally neglected to reveal to the readers my true intention for constructing the figure. His misuse of my publication reminds me of a similar abuse in which the Creationists had cited my paper (Shanmugam, 1997a) entitled The Bouma Sequence and the turbidite mind set in support of their creation theory (Oard et al., 2006).

    FIGURE 1.7 Diagram showing history of turbidite paradigm. Note differing perspectives on stages of deep-water research. Walker (1973) and Stow (1985) believed that the normal science stage was achieved in 1950 and 1983, respectively. However, I contend that we are still in a crisis mode. Kuhn's (1996) stages of scientific development are used.

    After Shanmugam (2000), reproduced with permission from Elsevier.

    FIGURE 1.8 Compilation of published examples showing that any deep water deposit, no matter what its primary sedimentary features, can be interpreted as some kind of turbidite. Left-hand column shows published examples with various lithofacies and associated features. Middle column shows depositional processes suggested by the original authors. Right-hand column shows turbidite interpretation by the original authors. I have constructed diagrams of massive sand, rippled sand, and ungraded mud, and added information on fluid rheology, flow state, and nature of sediment load for this book (given in parentheses in the middle column). Note that sand with normal grading as wells as sand with inverse grading have been interpreted to be a turbidite.

    After Shanmugam (1997a), reproduced with permission from Elsevier.

    FIGURE 1.9 Two differing definitions of the term turbidites. According to Sanders (1965), turbidites are the exclusive deposits of turbidity currents. According to Mutti et al. (1999), deposits of all sediment-gravity flows, which include turbidity currents, fluidized flows, debris flows, and grain flows, are turbidites. In this book, Sanders’ (1965) definition is followed.

    After Shanmugam (2002a), reproduced with permission from Elsevier.

    FIGURE 1.10 Correlation of existing vertical facies models of (1) coarse-grained turbidites (Lowe, 1982), (2) classic turbidites (also known as the Bouma Sequence), and (3) fine-grained turbidites (Stow and Shanmugam, 1980) showing an ideal, but unrealistic, turbidite bed with 16 divisions (see text for details). Correlation of the S3 division of coarse-grained turbidites with the Ta division of the Bouma Sequence is after Lowe (1982). Correlation of various divisions between classic turbidites and fine-grained turbidites is after Pickering et al. (1989).

    After Shanmugam (2000), reprinted with permission from Elsevier.

    Under the influence of the turbidite paradigm, geoscientists routinely interpret thick (>1 m) layers of deep-water sand as turbidites (Bouma, 1962; Mutti, 1992). An example is the late Pleistocene turbidite sand bed that is 57 m in thickness (Normark and Reid, 2003). Another is the Miocene Marnoso-Arenacea turbidite sandstone beds in Italy that are 3–9 m in thickness (Ricci Lucchi, 1981). The implication in most cases is that large sandy turbidity currents in the geologic past, presumably triggered near the shelf-slope break, traveled hundreds of kilometers before emplacing thick turbidite sands across submarine fans (Mutti and Ricci Lucchi, 1972; Mutti, 1992) and over the basin plains (Ricci Lucchi and Valmori, 1980). However, the most troubling aspect of the turbidite paradigm is that no one has ever documented empirical data on active gravelly or sandy turbidity currents in modern oceans using vertical sediment concentration profiles and grain-size measurements. Nor has anyone replicated turbulent turbidity currents that could carry coarse sand and gravel in suspension in laboratory flume experiments.

    1.2.1 Lack of Empirical Data on Sandy Turbidity Currents

    There are articles that claim direct measurements of modern turbidity currents. But the data presented in those articles are not compelling. In addressing these lingering problems, let me begin with a clear definition of turbidity currents and their deposits (i.e., turbidites). The following narrow definition of turbidity currents and their deposits is necessary for preserving the original meaning and for preventing any digression into bizarre types of turbidity currents, such as fluxoturbidity currents (Dzulynski et al., 1959), seismoturbidity currents (Mutti et al., 1984a), megaturbidity currents (Labaume et al., 1987), atypical turbidity currents (Stanley et al., 1978), and high-density turbidity currents (Lowe, 1982). These fallacious turbidity currents have never been documented in modern oceans. There is only one type of turbidity current, and its attributes are as follows:

    1. Turbidity current is a sediment-gravity flow with Newtonian rheology and turbulent state in which sediment is supported by fluid turbulence and from which deposition occurs through suspension settling (Kuenen and Migliorini, 1950; Dott, 1963; Sanders, 1965; Middleton and Hampton, 1973; Shanmugam, 1996a, 2006a).

    2. Turbidity currents are capable of transporting mainly fine-grained sediment (mostly mud and fine sand) in suspension. The so-called coarse-grained turbidity currents are falsehoods, undocumented in modern oceans.

    3. Turbidity currents are catastrophic (Middleton, 1993), surge-type events that do not attain hydrodynamic equilibrium (Allen, 1973).

    4. Turbidity currents are unsteady and nonuniform flows (Allen, 1985).

    5. Deposition from turbidity currents occurs when the velocity of turbidity currents decreases with time (see Waning flow in Appendix A), resulting in normal grading (Figure 1.11). Normal grading should not include floating clasts (see Normal grading in Appendix A). Middleton and Hampton (1973) attributed development of normal grading in turbidites to decay of initial turbulence and to decreasing competency in the tail of the turbidity current.

    FIGURE 1.11 Core photograph showing turbidite units with sharp basal contact, normal grading, and gradational upper contact. Arrow marks a normally graded unit with fine-grained 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.

    After Shanmugam (2006a), with permission from Elsevier.

    Within the conceptual confinement as outlined above, let us examine the following claims of modern turbidity currents:

    1. Heezen and Ewing (1952) suggested that high-velocity turbidity currents may have caused the seaward succession of breaks in submarine cables following the 1929 Grand Banks earthquake of a magnitude 7.2 (Fine et al., 2005), offshore New Foundland, Canada (Figure 1.12). The key evidence for linking cable breaks to turbidity currents was a 1-m-thick graded Holocene sand layer on the Sohm Abyssal Plain (Heezen et al., 1954). Although Shepard (1954) was skeptical of this hypothesis, it is still popular (Piper et al., 1988; Fine et al., 2005). Nevertheless, this hypothesis is unsustainable for the following reasons:

    a. All cables were broken within a matter of about 13 hours of the main shock (Figure 1.13). But the timing of emplacement of the turbidite sand has not been correlated with the date of the earthquake on November 18, 1929. Such a precise correlation is not possible because available radiocarbon dating methods cannot resolve the emplacement of a sand layer to a particular day. It is quite conceivable that the turbidite sand layer may have been emplaced even prior to the November 18, 1929, seismic event.

    b. Piper et al. (1999, their Figures 16 and 17) proposed that following the earthquake, transformation of rotational slumps into debris flows, and further transformation of debris flows into turbidity currents had led to all cable breaks. In this model, turbidity currents were solely responsible for breaking all cables in both slump-dominated proximal and turbidite-dominated distal areas (Figure 1.13A). Flow transformation is a complex, progressive, and gradual phenomenon that may take place during long transport of sediment flows. Fisher (1983) proposed four types of transformations for sediment-gravity flows: (i) body transformation, (ii) gravity transformation, (iii) surface transformation, and (iv) elutriation transformation. But there are no objective sedimentological criteria for recognizing the type of flow transformation in the rock record (see Shanmugam, 2006a). Piper et al. (1999) did not provide any physical evidence for flow transformation into turbidity currents. Without the physical evidence, it is difficult to envision a double flow transformation (i.e., first slump into debris flow, and then debris flow into turbidity current) in quick succession near the epicenter. A pragmatic alternative explanation is that seismic shaking (see Section 5.2.1), fault movements, and resultant huge slumps of 200 km³ volume (Bornhold et al., 2003; Fine et al., 2005) could have easily broken the cables instantaneously near the epicenter, without turbidity currents (Figure 1.13B).

    c. The use of normally graded turbidite sand as the primary evidence for advocating cable breaks by turbidity currents in the entire area is tenuous. This is because normal grading from turbidity currents develops during waning flow conditions. Waning flows are not powerful enough to break cables.

    d. Alternatively, debris flows that passed through this area (Piper et al., 1999) could have easily broken the cables. The advantage of debris-flow hypothesis is the high sediment concentration (see Chapter 2). Debris flows with higher sediment concentration (i.e., flow strength) than turbidity currents have a greater power for breaking cables. Also, debris flows can be fast-moving agents. High velocities of up to 70 m s−1 have been estimated for subaerial debris avalanches that were triggered by the May 18, 1980, eruption of Mount St. Helens (Voight et al., 1983). Submarine debris flows also travel fast due to hydroplaning (Mohrig et al., 1998). High velocities of submarine debris flows include estimated values of 14 m s−1 (Elverhøi et al., 2000), 16 m s−1 (Imran et al., 2001), and 30 m s−1 (De Blasio et al., 2004). These velocity values are almost equal to or even higher than the estimated value of 19 m s−1 for the 1929 turbidity current (Piper et al., 1988).

    FIGURE 1.12 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).

    After Piper et al. (1988), courtesy of the Geological Society of America.

    FIGURE 1.13 Generalized cross-sectional view of the continental slope and rise showing the location of the epicenter of the 1929 Grand Banks earthquake and related cable breaks, off New Foundland, Canada. (A) Original model in which turbidity currents were considered to be the cause of all submarine cable breaks (Heezen and Ewing, 1952; and Piper et al., 1999). See Figure 1.12 for details on epicenter, slump zone, bathymetry, and timing of cable breaks. (B) Proposed alternative model in which seismic shaking and related slides and slumps are considered to be responsible for breaking cables instantaneously in the proximal area, and debris flows for breaking cables in the distal area. Diagrams are constructed using information

    from Piper et al. (1988 and 1999).

    For these reasons, I propose that cables near the epicenter were broken instantaneously by seismic shaking and related slides and slumps, and that cables in the remaining area were broken by debris flows (Figure 1.13B). The important point is that there are no empirical data for cable breaks either by turbidity currents or by debris flows.

    2. In the 1970s, the lack of empirical data on modern turbidity currents was attributed to the belief that turbidity currents were so powerful that they invariably destroyed instruments in submarine environments (Inman et al., 1976). This notion was partly based on a 1964 study in which a powerful down-canyon event that not only deformed a 2.5-cm-diameter steel rod of an instrument package by 90° (Figure 1.14), but also resulted in the loss of a current meter in Scripps Submarine Canyon, California (Inman et al., 1976). When the instrument-mounting bar was recovered by divers, the bar was bent parallel to the canyon wall with the free end pointing down-canyon direction. Inman et al. (1976) believed that the bent bar was the key evidence of the large forces associated with the down-canyon movement of sand and kelp by turbidity currents. On November 24, 1968, a sustained event with a velocity of 190 cm s−1 was measured for over 2.5 hours. Inman et al. suggested that the sustained high-velocity event was turbidity current. But sustained high-flow velocity is not the defining criterion of turbidity currents (discussed earlier). In these cases, there are no empirical data to prove that turbidity currents were indeed responsible for deforming the steel rod. Alternatively, the event that deformed the steel rod could have as well been a debris flow. Indeed, Zakeri et al. (2008) experimentally demonstrated that submarine debris flows of non-Newtonian rheology are fully capable of breaking pipelines. On the outer continental shelf (OCS) of the Gulf of Mexico, the 2005 Category 5 Hurricane Katrina destroyed 46 petroleum platforms and damaged 20 others (MMS, 2006). Katrina-induced mudflows damaged, at least, six pipelines (Alvarado, 2006).

    3. Hay et al. (1982) claimed that they recognized turbidity currents in modern environments using geophysical methods of remote acoustic detection. However, such techniques are incapable of resolving the turbulent state or the Newtonian rheology of turbidity currents. Importantly, Hay et al. (1982) did not provide any data on sediment concentration or grain size.

    4. Dengler and Wilde (1987) published a paper entitled "Turbidity currents on steep slopes: application of an avalanche-type numeric model for ocean thermal energy conversion design. On page 411, Dengler and Wilde state, On November 23, 1982, during the passage of hurricane Iwa, current sensor moorings in place along the proposed pipe-line route for the Ocean Thermal Energy Conversion (OTEC) Pilot Plant at Kahe Point, Oahu, Hawaii, moved downslope during a sequence of slump and/or turbidity current events…" The issue remains whether the event was a slump or turbidity current. Again, there is no direct evidence for turbidity currents.

    5. Normark (1989) published a paper entitled "Observed parameters for turbidity-current flow in channels, Reserve Fan, Lake Superior." But these flows were not natural events. They were man-made events triggered by the discharge of taconite railings (i.e., byproducts of extracting low-grade iron from the parent rock) by the Reserve Mining (Company) operation on the north shore of Lake Superior. Importantly, Normark (1989) did not document the turbulent state and the Newtonian rheology of the flow. Also, normal grading is absent in sediment core. In short, whether these man-made flows were turbidity currents or not remain unresolved. The discharge of tailings into Lake Superior was terminated due to environmental regulations. Therefore, it is no longer possible to verify the nature of these flows.

    6. In commenting on turbidity currents in modern fjords reported by Hay et al. (1982) and by Prior et al. (1987), Middleton (1993) concluded that these currents are weaker and more continuous than those of most interest to geologists. Furthermore, Middleton (1993, pp. 98 and 99) observed that currents in modern lakes have been relatively continuous, relatively slow, and of relatively low density, implying that these lake currents are different from deep-water turbidity currents. In evaluating high-velocity (3 m s−1) turbidity currents inferred by Dengler et al. (1984), Middleton (1993, p. 100) commented that "… there are essentially no direct measurements of flow thickness, or velocity and sediment concentration profiles from such strong flows."

    7. Khripounoff et al. (2003) published an article entitled "Direct observation of intense turbidity current activity in the Zaire submarine valley at 4000 m water depth." In this case, the criterion for classifying an event as turbidity current was high particle flux and maximum measured velocity of 121.4 cm s−1. But they did not provide the critical data on vertical sediment concentration profiles with grain-size variations. Again, velocity is not the defining criterion of turbidity currents.

    8. Parsons et al. (2003, p. 839) claimed, "In fact, one of us (JDP) has personally observed (via an ROV) a dilute turbidity current associated with internal wave resuspension in the Eel Canyon." But the authors did not provide any information on sediment concentration in establishing turbidity currents. Nor did they provide information on the Reynolds number of the resuspension.

    9. Xu et al. (2004) published an article entitled "In-situ measurements of velocity structure within turbidity currents." They documented four events in which one of them was a man-made event caused by dumping of dredged material from the nearby Moss Landing Harbour at the head of the Monterey Canyon, California. The other three natural events had sediment concentration value of less than 1 kg m−3. The required sediment concentration for defining hyperpycnal flows (akin to turbidity currents) in ocean waters is 40 kg m−3, (Mulder and Syvitski, 1995; Imran and Syvitski, 2000), an order of magnitude higher than the measured value in the Monterey Canyon. Although Xu et al. (2004, their Figure 3) provided vertical velocity profiles for the four events, they did not provide any vertical sediment concentration profile with grain-size data.

    10. Mulder

    Enjoying the preview?
    Page 1 of 1