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

Only $11.99/month after trial. Cancel anytime.

Deep Marine Systems: Processes, Deposits, Environments, Tectonics and Sedimentation
Deep Marine Systems: Processes, Deposits, Environments, Tectonics and Sedimentation
Deep Marine Systems: Processes, Deposits, Environments, Tectonics and Sedimentation
Ebook2,180 pages22 hours

Deep Marine Systems: Processes, Deposits, Environments, Tectonics and Sedimentation

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Deep-water (below wave base) processes, although generally hidden from view, shape the sedimentary record of more than 65% of the Earth’s surface, including large parts of ancient mountain belts. This book aims to inform advanced-level undergraduate and postgraduate students, and professional Earth scientists with interests in physical oceanography and hydrocarbon exploration and production, about many of the important physical aspects of deep-water (mainly deep-marine) systems. The authors consider transport and deposition in the deep sea, trace-fossil assemblages, and facies stacking patterns as an archive of the underlying controls on deposit architecture (e.g., seismicity, climate change, autocyclicity). Topics include modern and ancient deep-water sedimentary environments, tectonic settings, and how basinal and extra-basinal processes generate  the typical characteristics of basin slopes, submarine canyons, contourite mounds and drifts, submarine fans, basin floors and abyssal plains.

LanguageEnglish
PublisherWiley
Release dateOct 23, 2015
ISBN9781118865422
Deep Marine Systems: Processes, Deposits, Environments, Tectonics and Sedimentation

Related to Deep Marine Systems

Related ebooks

Earth Sciences For You

View More

Related articles

Reviews for Deep Marine Systems

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

    Deep Marine Systems - Kevin T. Pickering

    Table of Contents

    Cover

    Title Page

    Copyright

    Preface

    About the companion website

    Part 1: Process and product

    Chapter One: Physical and biological processes

    1.1 Introduction

    1.2 Shelf-edge processes

    1.3 Deep, thermohaline, clear-water currents

    1.4 Density currents and sediment gravity flows

    1.5 Turbidity currents and turbidites

    1.6 Concentrated density flows and their deposits

    1.7 Inflated sandflows and their deposits

    1.8 Cohesive flows and their deposits

    1.9 Accumulation of biogenic skeletons and organic matter

    1.10 Summary

    Chapter Two: Sediments (facies)

    2.1 Introduction

    2.2 Facies classifications

    2.3 Facies Class A: Gravels, muddy gravels, gravelly muds, pebbly sands, ≥5% gravel grade

    2.4 Facies Class B: Sands, >80% sand grade, <5% pebble grade

    2.5 Facies Class C: Sand–mud couplets and muddy sands, 20–80% sand grade, <80% mud grade (mostly silt)

    2.6 Facies Class D: Silts, silty muds, and silt–mud couplets, >80% mud, ≥40% silt, 0–20% sand

    2.7 Facies Class E: ≥95% mud grade, <40% silt grade, <5% sand and coarser grade, <25% biogenics

    2.8 Facies Class F: Chaotic deposits

    2.9 Facies Class G: Biogenic oozes (>75% biogenics), muddy oozes (50–75% biogenics), biogenic muds (25–50% biogenics) and chemogenic sediments, <5% terrigenous sand and gravel

    2.10 Injectites (clastic dykes and sills) (Figs 2.46–2.50)

    2.11 Facies associations

    Chapter Three: Deep-water ichnology

    3.1 Introduction

    3.2 General principles of ichnology

    3.3 Colonisation of SGF deposits: Opportunistic and equilibrium ecology

    3.4 Ichnofacies

    3.5 Ichnofabrics

    3.6 Trace fossils in core

    3.7 Case study I: Trace fossils as diagnostic indicators of deep-marine environments, Middle Eocene Ainsa–Jaca basins, Spanish Pyrenees

    3.8 Case study II: Subsurface ichnological characterisation of the Middle Eocene Ainsa deep-marine system, Spanish Pyrenees

    3.9 Summary of ichnology studies in deep-water systems

    3.10 Concluding remarks

    Chapter Four: Time–space integration

    4.1 Introduction

    4.2 Submarine fan growth phases and sequence stratigraphy

    4.3 Tectono-thermal/glacio-eustatic controls at evolving passive continental margins

    4.4 Eustatic sea-level changes at active plate margins

    4.5 Changing relative base level and sediment delivery processes

    4.6 Autocyclic processes

    4.7 Palaeo-seismicity and the stratigraphic record

    4.8 Deconvolving tectonic and climatic controls on depositional sequences in tectonically active basins: Case study from the Eocene, Spanish Pyrenees

    4.9 Problems in determining controls on sediment delivery

    4.10 Carbonate versus siliciclastic systems

    4.11 Computer simulations of deep-water stratigraphy

    4.12 Laboratory simulations of deep-water stratigraphy

    4.13 Supercritical versus subcritical fans

    4.14 Hierarchical classification of depositional units

    4.15 Concluding comments

    Chapter Five: Statistical properties of sediment gravity flow (SGF) deposits

    5.1 Introduction

    5.2 Cloridorme Formation, Middle Ordovician, Québec

    5.3 Vertical trends

    Part 2: Systems

    Chapter Six: Sediment drifts and abyssal sediment waves

    6.1 Introduction

    6.2 Distribution and character of contourites and sediment drifts, North Atlantic Ocean

    6.3 Facies of muddy and sandy contourites

    6.4 Seismic facies of contourites

    6.5 The debate concerning bottom-current reworking of sandy fan sediments

    6.6 Ancient contourites

    6.7 Facies model for sediment drifts

    Chapter Seven: Submarine fans and related depositional systems: modern

    7.1 Introduction

    7.2 Major controls on submarine fans

    7.3 Submarine canyons

    7.4 Architectural elements of submarine-fan systems

    7.5 The distribution of architectural elements in modern submarine fans

    7.6 Modern non-fan dispersal systems

    7.7 Concluding remarks

    Chapter Eight: Submarine fans and related depositional systems: ancient

    8.1 Introduction

    8.2 Ancient submarine canyons

    8.3 Ancient submarine channels

    8.4 Comparing modern and ancient channels

    8.5 Ancient lobe, lobe-fringe, fan-fringe and distal basin-floor deposits

    8.6 Seafloor topography and onlaps

    8.7 Scours

    8.8 Basin-floor sheet-like systems

    8.9 Prodeltaic clastic ramps

    8.10 Concluding remarks

    Chapter Nine: Evolving and mature extensional systems

    9.1 Introduction

    9.2 Models for lithospheric extension

    9.3 Subsidence and deep-water facies of rifts and young passive margins

    9.4 The post-breakup architecture of passive margins

    9.5 Failed rift systems

    9.6 Fragments of ancient passive margins

    9.7 Concluding remarks

    Part 3: Plate tectonics and sedimentation

    Chapter Ten: Subduction margins

    10.1 Introduction

    10.2 Modern subduction factories

    10.3 Arc–arc collision zones

    10.4 Forearc summary model

    10.5 Marginal/backarc basins

    10.6 Ancient convergent-margin systems

    10.7 Forearc/backarc cycles

    10.8 Concluding remarks

    Chapter Eleven: Foreland basins

    11.1 Introduction

    11.2 Modern foreland basins

    11.3 Ancient deep-marine foreland basins

    11.4 Concluding remarks

    Chapter Twelve: Strike-slip continental margin basins

    12.1 Introduction

    12.2 Kinematic models for strike-slip basins

    12.3 Suspect terranes

    12.4 Depositional models for strike-slip basins

    12.5 Modern strike-slip mobile zones

    12.6 Ancient deep-marine oblique-slip mobile zones

    12.7 Concluding remarks

    References

    Index

    End User License Agreement

    List of Illustrations

    Chapter One: Physical and biological processes

    Figure 1.1 Simplified conceptual overview of the evolution of sediment gravity flows and other deep-marine transport processes as a function of concentration. The horizontal axis is time and/or space, but no units are implied because the evolution of some flows is much longer than for others. For example, turbidity currents might flow for only hours to days, whereas contour-following geostrophic bottom-currents (i.e., thermohaline currents) have velocity fluctuations lasting thousands of years. Notice that non-turbulent flows tend to deposit en masse, so that the deposit is simply the original flow, arrested in place when driving forces are no longer adequate to keep the material moving. In contrast, turbulent flows lose their sediment load by settling, and therefore become increasingly less concentrated during the depositional phase. Modified from Middleton and Hampton (1973) and Walker (1978).

    Figure 1.2 Four snapshots during the deceleration and eventual deposition (Time 4) of a non-turbulent debris flow, showing how textures, fabrics and internal structures of the eventual deposit are locked into place by the progressive downward thickening of a ‘rigid plug’ (arrested debris). These are streamwise vertical cross-sections through the flow. In the ‘rigid plug’, there is little to no internal deformation because gravity-induced shear stress (τ) is less than the critical shear stress (τc) needed to overcome resisting forces (due to internal grain friction and electrostatic cohesive forces). Decreasing slope explains the decreasing shear stress. Profiles of shear stress (τ) and velocity (U) are shown in each case. In the ‘rigid plug’, the change in velocity with depth is zero, although the velocity of the plug itself is positive up until Time 4. In SGFs of this type, the material is effectively deposited ‘from the top downward’, and the base of the flow is the last part to be deformed by shearing (e.g., Time 3).

    Figure 1.3 Schematic representation of lutite flows cascading downslope. The increase in length of the arrows indicates an increase in concentration toward the base of the continental slope. The nepheloid layer is a part of the water column near continental margins where the suspended sediment concentration is particularly high because persistent currents prevent deposition of fine-grained suspended load. Redrawn from McCave (1972).

    Figure 1.4 Model for hemipelagites (Silurian Bailey Hill Formation) and turbidite sandstones (Brimmon Wood Member) in the Welsh Basin. Suspended silts, fine sands and organics (blue ‘tongues’) were advected off the shelf by waves and currents (large arrows and smaller downslope-oriented arrows), forming dilute bottom- and mid-water flows. The particulate materials then settled vertically from these flows (short vertical arrows). Annual seasonal layering in the hemipelagites was preserved under anaerobic/dysaerobic conditions. Redrawn from Dimberline and Woodcock (1987).

    Figure 1.5 Diagrams to show the periodicity of oscillating up- and down-canyon currents. Tide relationship obtained from the predicted tide at the nearest reference station. (a) Kaulakahi Canyon between Kauai and Niihau islands, Hawaii; (b) Hueneme Canyon, California; (c) Congo Canyon, west Africa. Redrawn from Shepard and Marshall (1978).

    Figure 1.6 Relationship of wind speed and swell height to the magnitude of up- and down-canyon currents during a storm period in Hydrographer Canyon, off Massachusetts. The slowest currents occurred during periods of reduced wind speeds and reduced swell. Redrawn from Shepard and Marshall (1978).

    Figure 1.7 Graphs to show the most likely direction and approximate speed of internal waves up and (less frequently) down the axes of various submarine canyons. The speed of wave advance is approximate because of errors in matching wave crests between current-meter stations, particularly in cases where the current appears to be up-canyon. Redrawn from Shepard and Marshall (1978).

    Figure 1.8 Schematic representation of a sediment slide. Circled numbers show (1) headwall scarp, (2) extensional ridges and blocks, (3) lateral margins, (4) basal shear surface ramps and flats, (5) basal shear surface grooves, (6) basal shear surface striations, (7) remnant blocks, (8) translated blocks, (9) outrunner blocks, (10) folds, (11) longitudinal shears (= first-order flow fabric), (12) second-order flow fabric, (13) pressure ridges, (14) fold and thrust systems. From Bull et al. (2009).

    Figure 1.9 Conceptual model of submarine slide evolution (Gee et al. 2006). Stage 1 shows seafloor rupture. Stage 2 shows tabular blocks, basal striations, debris flow and a turbidity current generated in the headwall area. Downslope of the headwall area, turbidity currents erode furrows in the seafloor. Stage 3 shows the development of secondary slide events within the headwall, triggering secondary debris flows and turbidity currents.

    Figure 1.10 Dip seismic-reflection profile through a failed part of the wall of Munson Canyon, US Atlantic coast. The failure surface is overlain by a chaotic MTC to the left, and is ∼150 m below the seabed to the right. The apparent downward step in the failure surface at the edge of the depression is an artifact (‘pullup’) created by the differing acoustic travel-time in water and sediment. Modified from O'Leary (1993).

    Figure 1.11 Frequency distribution of submarine slides (MTCs) on the US Atlantic margin as a function of seabed slope at the site of initiation. Redrawn from Booth et al. (1993).

    Figure 1.12 Graph of the horizontal-component peak acceleration versus distance from the epicentre of the 1994 Northridge earthquake, California. The green envelope encloses >150 data points plotted by Mueller (1994). The curves represent the average peak acceleration (solid line) and the ±1 standard deviation accelerations (dashed lines) expected for a magnitude 6.7 earthquake. Redrawn from Mueller (1994).

    Figure 1.13 Horizontal ground acceleration, ax, required to reduce the static safety factor to a value of 1.0 for given slope angles and sediment density in the range 1.5–2.0 g cm−1 (simplified from Booth et al. (1985). For example, if SF = 2.0 and α = 10°, ground accelerations of about 0.07 g or greater will reduce safety factor to 1.0 or less, and failure will be likely. On the same slope with SF = 4.0, accelerations of at least 0.2 g would be needed to cause failure. Note that SF is itself a function of bottom slope and excess pore pressure (Eq. 1.1).

    Figure 1.14 Approximate tracks of Western Boundary Undercurrent (WBU) and Deep Western Boundary Current (DWBC) along the eastern continental margin of North America. The HEBBLE area was the site of detailed long-term measurements of bottom currents, and is an acronym for High Energy Benthic Boundary Layer Experiment. Redrawn from Hollister and McCave (1984).

    Figure 1.15 Schematic model showing upward coarsening and fining of contourite facies from the Faro Drift, offshore southern Portugal. Redrawn from Gonthier et al. (1984).

    Figure 1.16 Diagnostic criteria for the recognition of fine-grained turbidites. Rapid deposition from a decelerating SGF produces both wet-sediment deformation structures like load casts, and climbing ripples. Accumulation rates under thermohaline currents are much lower, preventing the development of these types of sedimentary structures. The co-occurrence of several of these structures is sufficient to rule out deposition by clear-water bottom currents (contour currents). Redrawn from Piper and Stow (1991).

    Figure 1.17 Sequential development of a turbulent density current formed by release of an alkaline saline solution containing a pH indicator into an acidified freshwater ambient environment. The pH indicator stains the saline current purple, changing to red once mixing with the overlying ambient fluid results in a neutral pH. Mixing is strongly developed in the upper part of the flow and in the wake behind the head of the current. There, the visual contrast between red and colourless regions provides a detailed image of the shapes of the turbulent eddies. See Hallworth et al. (1996) for details.

    Figure 1.18 Sequential development of a three-layered saline, turbulent density current in which the lower layer has the greatest density. ρ* = density contrast between layers; ρc = average flow density; t = time after flow initiation. The three starting colours (red, yellow, blue) were created by artificial dyes, whereas transitional colours show the extent of fluid mixing during evolution of the flow (e.g., red+yellow = orange; blue+yellow = green). See Gladstone et al. (2004) for details.

    Figure 1.19 Relative importance of particle-support mechanisms for the four varieties of SGF recognised in this book, and defined in Section 1.4.1. Where orange and red symbols are superimposed, the support varies from significant to dominant. See also Figure 1.22.

    Figure 1.22 Summary of flow characteristics, typical deposits, and grain-support mechanisms for cohesive and frictional (non-cohesive) SGFs, modified from Mulder and Alexander (2001).

    Figure 1.20 (a) Simplified depiction of a turbidity current or concentrated density flow divided into head, body and tail regions. (b) Conceptual view of the vertical stratification of velocity and concentration expected in the body of such a flow. Rapid particle fallout during deposition can increase the near-bed concentration to the point that turbulence is damped, grain collisions become common, and deposits become poorly organised. The eventual deposit is much thinner than the flow that created it.

    Figure 1.21 Approximate solids concentrations typical of SGFs, modified from Mulder and Alexander (2001). Dashed lines show the possible extensions of sediment concentrations to lower and higher values than those deemed to be typical. Note the overlaps between different flow types, which result from the effects of flow stratification and different textures of the sediment load.

    Figure 1.23 The gradual transition in grain-support mechanisms between inflated sandflows, concentrated density currents and turbidity currents. Cross-overs in the relative importance of these mechanisms are used to define these SGFs. As is the case for cohesive flows, the thickness of the deposit of an inflated sandflow is similar to the thickness of the flow itself. Turbidites, in contrast, are much thinner than the associated turbidity current. Modified from Mulder and Alexander (2001).

    Figure 1.24 Multiphase hybrid SGF proposed by Pickering and Corregidor (2005) to explain certain disorganised beds in the Ainsa Basin, southern Pyrenees.

    Figure 1.25 Schematic drawing of the Little Manly Slump flow transformation model from Strachan (2008). T1 to T4 represent progressive development of transformation with time. Vertical velocity profiles for T1 to T4 are indicated and, where two are present, indicate the differences between slump and debris flow vertical velocity profiles. T5 shows deposition of units following slump cessation, together with a log showing a vertical profile. Surface and body transformations are inferred during T2 and T3. (FS = fine sand; MS = medium sand; CS = coarse sand.)

    Figure 1.26 Limited distribution of debrite units when compared with associated turbidites within tri-partite beds in the Marnoso arenacea, Italy. From Amy and Talling (2006).

    Figure 1.27 Examples of linked debrites and Facies Class B sandstones (see Section 2.4) from the fringe of a Jurassic sand-prone fan in the subsurface of the North Sea. From Haughton et al. (2003).

    Figure 1.28 (a) Haughton et al. (2009) classification scheme for event beds emplaced by subaqueous sediment gravity flows. (b1) Debrites, concentrated density-flow deposits and turbidites dominate the record of many deep-water systems and record increasing downdip dilution of the flow (debrites passing to concentrated density-flow deposits and eventually turbidites). (b2) In some systems there is instead a downdip progression from non-cohesive flows (depositing concentrated density-flow deposits and turbidites) to flows transformed into components with radically different rheology, with the deposits of the cohesive flow components increasingly dominant distally. From Haughton et al. (2009) who used ‘high- and low-density turbidity currents’ instead of ‘concentrated density-flows and turbidity currents’.

    Figure 1.29 Summary of depositional origin for hybrid event beds as a result of (a) loss of turbulence and deceleration of a clay-rich flow, (b) longitudinal segregation of clays and clay flakes to suppress turbulence in the rear and margins of an otherwise turbulent flow, (c) bulking and disintegration of clay clasts to release clay near-bed in an otherwise turbulent flow and (d) downdip runout of a flow that was either synchronously triggered with a concentrated density flow, or that partially transformed to generate a forerunning concentrated density flow. From Haughton et al. (2010).

    Figure 1.30 Contrasting near-bed grain support between a turbidity current and a concentrated density flow. Although both types of SGF are turbulent away from the sedimentation surface, the latter has strong grain interaction at the aggrading bed, and turbulence is sufficiently damped to prevent the development of tractional sedimentary structures.

    Figure 1.31 Idealised streamwise cross-section of a turbidity current, divided into head, body and tail regions. Settling from the wake behind the head produces a lateral size grading in the flow.

    Figure 1.32 Ratio of head velocity to body velocity plotted against bottom slope for experimental turbidity currents (5-m flume) of Middleton (1966a). Horizontal bars show extent of scatter in the data. Ranges of bottom slope for continental margins are superimposed.

    Figure 1.33 Measured flow velocities of turbidity currents in Bute Inlet compared with velocity estimates inferred from the texture of turbidites along the flow path. Redrawn from Zeng et al. (1991).

    Figure 1.34 Distribution of the sand turbidite generated by the 1929 Grand Banks earthquake. Numbers beside cable break positions are the time of each break, in minutes, after the earthquake. Bathymetric contours are in metres. Turbidite thickness is contoured in centimetres. Redrawn from Piper et al. (1988).

    Figure 1.35 Giant gravel waves in the Eastern Valley of the Laurentian Fan, likely formed under a hyperpycnal flow produced by an outburst flood during decay of continental ice sheets. (a) Fields of gravel-rich waves in lanes of larger and smaller waves, buried by sand ribbons. (b) Detail of rather sinuous gravel-rich waves of various sizes, locally covered by sand patches, using 1-km swath system. From Wynn et al. (2002b). Copies of original graphics courtesy of D.J.W. Piper, Geological Survey of Canada.

    Figure 1.36 Optimum conditions for ignition of turbidity currents, from Normark and Piper (1991). The plot shows critical velocity and critical concentration for various flow thicknesses (h, metres) and representative grain sizes (Ds, mm) on a slope of S = 0.05, assuming a drag coefficient of CD = 0.004 (based on four-equation model results of Parker et al. (1986)). Also shown is the approximate offset in the concentration scale resulting from an increase in S to 0.1. Circled numbers are measured suspended-sediment concentrations from a number of settings: 1 = Huanghe River, China, in flood (Wright et al. 1988); 2 = Sustina River, Alaska, in flood (Hoskin & Burrell 1972); 3 = maximum measured discharge into Glacier Bay, Alaska (Hoskin & Burrell 1972); 4 = ignitive condition for sand in the head of La Jolla Canyon (Fukushima et al. 1985); 5 = nearshore sediment concentration after flood of Santa Clara River (Drake et al. 1972). The ‘cold underflow’ is the approximate sediment concentration in cold river water required to produce a marine underflow (Gilbert 1983).

    Figure 1.37 Comparison of normally graded surge-generated turbidites and inverse-to-normally graded hyperpycnal-flow deposits produced by the rising and then falling flood stage of a river. The grain size scale below column 1 has divisions of fine (f), medium (m) and coarse (c) silt, and fine (f) and medium (m) sand. Redrawn from Mulder et al. (2001).

    Figure 1.38 Conceptual diagram to explain autosuspension. If gravitational energy input = energy losses, the flow will be self maintaining, and grains with settling velocity w will be kept in suspension by vertical velocity fluctuations of average strength c01-math-0008 , approximated by u*.

    Figure 1.39 Thickness of the sedimentary deposit in an experimental tank in the presence of an obstacle oriented oblique to the radiating flow. Contours are in millimetres of thickness. The spreading turbidity current (purple vectors) came from the bottom of this map. The obstacle (green) is a wedge with leading edge 2.4 cm high. Initial suspension density was 1.222 g cm−3 (10%), maximum head velocity was 24 cm s−1, and grain size was 80 µm. The obstacle induced a hydraulic jump. Downflow of this jump, flow expansion caused a velocity drop and enhanced deposition (beyond the red line marked ‘abrupt thickness increase’). Flow vectors (purple) changed sharply at the position of the jump. Redrawn from Morris and Alexander (2003).

    Figure 1.40 Definition diagram for spatially and temporally accelerating and decelerating turbidity currents. The former accumulate or deplete, whereas the latter wax and wane. Arrows point downflow. Predicted grading profiles are shown, with white = sand and black = mud. Redrawn from Kneller (1995).

    Figure 1.41 Subdivision of acceleration space into fields characterised by deposition, non-deposition (or erosion) and bypass. Redrawn from Kneller (1995).

    Figure 1.42 Graphical representation of the velocity history of a turbidity current (arrow) as it changes with time (t) and distance along the flow path (x). Redrawn from Kneller (1995).

    Figure 1.43 Ideal sequence of sedimentary structures in a turbidite bed (from Bouma 1962, with interpretation from Harms & Fahnestock 1965; Walker 1965; Middleton 1967; Walton 1967; Stow & Bowen 1980).

    Figure 1.47 Summary of schemes for subdivision of fine-grained turbidites, based on Hesse (1975), Piper (1978), van der Lingen (1969) and Stow and Shanmugam (1980).

    Figure 1.44 Explanation of Allen (1969 1982) for the absence of dune-scale cross-stratification in most turbidites. Once the rate of grain fallout declines sufficiently for tractional bedforms to develop, the stability field for dunes has either been bypassed, or there is not enough time for dunes to grow before current ripples become the stable bedform.

    Figure 1.45 Generalised sequence of internal structures produced beneath large reflected turbidity currents (Pickering & Hiscott 1985). The thick mud cap is deposited after the flow becomes ponded within a confined basin. Single beds of this type may be >10 m thick.

    Figure 1.46 3.5-kHz profile of sediment waves on western levée of Monterey Fan valley. The levée crest is 20 km to the right. Arrows show migration of sediment-wave crests and troughs. From Normark et al. (2002).

    Figure 1.48 (a–d)Schematic representation of the four stages of silt and mud deposition through the boundary layer of a turbidity current to form silt and mud laminae. From Stow and Bowen (1980).

    Figure 1.49 Maximum grain size plotted against bed thickness for Lower Carboniferous turbidites of (a) the Rhenaer Kalk (498 beds), and (b) the Posidonienkalk (235 beds), from Sadler (1982). The left-hand diagrams show contours of density of data points representing total bed thickness and corresponding maximum grain size. Arrows are selected vertical grading curves using the x-axis scale as a ‘distance to top of bed’. The right-hand diagrams show fields for Bouma divisions Ta through Td found at the base of the beds for which maximum size was determined. Green zones = overlap of fields for divisions Tb, Tc and Td. Dashed line = lower limit of field for division Ta, which has extensive overlap with the other fields. Arrow = modal horizontal grading curve, based on contour pattern in left-hand diagram.

    Figure 1.50 Spaced stratification within a graded bed of coarse to medium sandstone. The base and top of the bed are not shown. Each 5–10 cm-thick stratification band is inversely graded at its base and then essentially structureless and ungraded above. The sandstone has a strong grain fabric with high grain imbrication angle. Scale divisions are 10 cm.

    Figure 1.51 (a) Preferred transitions between nine ‘states’ in 214 thick sandstone beds (average thickness 380 cm) deposited from concentrated density flows, Tourelle Formation, Québec, Canada. Numbers indicate the number of transitions between states; T25 signifies that a rippled division is the highest division in 25 beds. Complete beds (tops preserved) only occur 62 times – the other 152 beds had their tops bevelled during amalgamation. Structural divisions were produced either by rapid mass deposition (‘freezing’) or by selective grain-by-grain deposition with traction transport. In some cases, post-depositional reworking of the top of the bed is indicated by medium-scale cross-stratification. Arrow weight indicates the statistical significance of transitions (based on method of Powers & Easterling 1982: p. 922) with the levels of significance being >98% (solid arrow) and >93% (dashed arrow). No arrows are shown if there is only one transition between states, regardless of significance level. (b) Generalised bed model based on the transition diagram, with interpretation. Where appropriate, structural divisions have Bouma (1962) Ta–e labels.

    Figure 1.52 Mechanism proposed by Postma et al. (1988) for the transport of large mud clasts within a strongly stratified SGF. At 25°, the slope for the experimental demonstration of this effect is unreasonably high for natural SGFs.

    Figure 1.53 Relationship between deposit characteristics, shear rate and flow concentration in sandy gravity-flow deposits. From Branney and Kokelaar (2002).

    Figure 1.54 Process cube to explain deposits from sandload SGFs. High concentrations (the two upper scenarios) form typical deposits of inflated sandflows and some concentrated density flows. From Branney and Kokelaar (2002).

    Figure 1.55 Relation of Bingham Number to critical Reynolds Number for turbulence in a Bingham plastic. Experimental data are for pipe flow but scales have been adjusted to the correct values for 2D SGFs using the thickness of the flow as the length scale. From Hiscott and Middleton (1979), based on Hampton (1972).

    Figure 1.56 Experimental subaerial debris flow with a series of advancing surface waves (arrows). The grid squares have sides of 1 m, and a human Figure in the top right also provides scale. From Major (1997: his Figure 7E).

    Figure 1.57 Amalgamated experimental debris flow deposits separated by a cryptic surface (arrow) that could easily be missed in a natural exposure. Leftmost scale divisions are in centimetres. Photograph courtesy of J. J. Major. A different view of the arrowed amalgamation surface is shown in Major (1997: his Figure 10A).

    Figure 1.58 100 kHz side-scan sonar image (and relief profile), and matching sub-bottom profile of hummocky debrite lobe in Bute Inlet, Canada. From Prior et al. (1984).

    Figure 1.59 TOBI 30 kHz sonograph of ‘woodgrain’ texture, consisting of flow-parallel banding, pressure ridges and longitudinal shears, on the surface of the Saharan Debris Flow deposits at ∼4350 m water depth. Light tones are high backscatter and the vehicle track is along the centre-line of the image. From Weaver et al. (1995).

    Figure 1.60 SeaMarc1A mosaic showing shrub-shaped SGF deposits at the terminations of dendritic channel network on the Mississippi Fan, and interpreted by Schwab et al. (1996) and Talling et al. (2010) as the deposits of mudflows. From Paskevich et al. (2001) and http://pubs.usgs.gov/of/2000/of00-352/htmldocs/images.htm.

    Figure 1.61 Schematic diagram of relationships among debrite types. See text for abbreviations. Redrawn from Shultz (1984).

    Figure 1.62 World distribution of dominant sediment types in the oceans. Redrawn from Jenkyns (1986).

    Figure 1.63 Schematic diagram, based on the Pacific Ocean, to show the main location of sediment types relative to the CCD and near-surface organic productivity. Redrawn from Ramsay (1977).

    Figure 1.64 Fluctuations in the depth of the CCD since 150 Ma in the Atlantic, Pacific and Indian oceans and the variation in global sea level. Redrawn from Kennett (1982).

    Figure 1.65 Rates of sediment accumulation of some modern and ancient pelagic and hemipelagic sediments. Note that estimated rates for ancient sediments are ∼60–70% of modern rates. Redrawn from Scholle and Ekdale (1983).

    Figure 1.66 Schematic reconstruction of the Palaeocene and Oligocene distribution of continents and ocean surface-circulation patterns. Numbers refer to following: (a) Palaeocene; 1, Proto Gulf Stream; 2, Tethys Current; (b) Oligocene; 1, Norwegian-Greenland Sea; 2, North Labrador Passage; 3, Greenland-lceland-Faroe Ridge; 4, South Tasman Rise; 5, Drake Passage. From Leggett (1985) after Haq (1981).

    Figure 1.67 Reconstruction of the central Atlantic Ocean in the Early Cretaceous, showing location of some DSDP sites, the inferred surface circulation and probable distribution of varve-type laminated sediments. Redrawn from Robertson (1984).

    Figure 1.68 Summary of the evolution of sediment gravity flows and other deep-marine transport processes as a function of particle concentration. Compare with Figure 1.1. Deposits are arranged across the top of the diagram at the end of each evolutionary pathway. The box labelled with the letter a relates to dense, cohesive SGFs that never became turbulent. Box b corresponds to concentrated density flows which, during their depositional phase, become strongly stratified into a non-turbulent lower part and a thicker turbulent upper part. The lower part of this flow eventually deposits en masse rather than by suspension settling. Boxes with letters c–e correspond to turbulent SGFs of variable concentration – these lose their sediment load by selective (grain-by-grain) deposition, forming graded beds with tractional structures. Eventually, all sediment is dropped and concentration approaches zero. Grey ‘dumbells’ link each evolving flow with its deposit. The box and flow labelled f is a thermohaline contour current with low concentration and very long-period velocity fluctuations (e.g., thousands of years). Clearly, the time scale for these currents and their fluctuations is much longer than the time scales for other flows in the diagram, which might be no longer than hours to days. Modified from Middleton and Hampton (1973) and Walker (1978).

    Chapter Two: Sediments (facies)

    Figure 2.1 Summary of relative grain sizes (diameters of filled circles), geometry of basal and upper contacts, primary depositional structures, and emplacement processes (superscripts to the right) for deep-marine facies A–E of Mutti and Ricci Lucchi (1972). Subscripts to the left indicate whether the deposits are organised (O) or disorganised (D); subscripts to the right are subfacies numbers.

    Figure 2.2 Gravity-flow facies F1–F9 and WF of Mutti (1992) with a summary of transport and depositional processes associated with cohesive debris flows to low-density turbidity currents. F1 = beds with muddy matrix, plastic deformation, dispersed large clasts (some projecting from bed top). F2 = beds with basal scours, dispersed large mudstone clasts, crude grading of muddy sand/gravel mixtures. F3 = clast-supported gravels, mostly unstratified, commonly inversely graded. F4 = thick coarse-grained sands with spaced stratification (Hiscott 1994b). F5 = thick coarse-grained sands with fluid-escape structures and poor sorting. F6 = coarse-grained planar and cross-stratified sands, ungraded (same as Mutti 1977 Facies E). F7 = thin, coarse-grained sands with horizontal, commonly very thin, laminae akin to spaced stratification (Hiscott 1994a). F8 = deposits of structureless medium- to fine-grained sand, possibly graded (equivalent to Bouma 1962 Ta division). F9 = deposits of thoroughly current-laminated very fine sand and silt capped by mudstone (equivalent to Bouma 1962 Tb through Te divisions). WF = <20 cm-thick deposits of very poorly sorted coarse sand to pebbles with faint wavy laminae generally overlying F2 and underlying F4. CgRF = conglomerate remnant facies with convex top. TC = traction carpets. FT = flow transformation. Note that some facies in this Figure constitute 100% of an event-deposit, whereas others form divisions within the succession of deposits left by a single passing gravity flow. In this book we use a different classification of flow types (see Table 1.1 for equivalencies).

    Figure 2.3 Facies classification of Ghibaudo (1992) employing strings of upper and lower case letters as codes for gravity-flow deposits. G = gravel; GS = gravel–sand couplet; S = sand, M = mud and T = silt, either alone or as part of a sand–mud (SM), silt–mud (TM), mud–sand (MS) or mud–silt (MT) couplet; Gy = gravelly; My = muddy; g = graded, s = plane-stratified, x = cross-stratified, l = laminated, r = ripple-laminated, m = massive; t = thin-bedded.

    Figure 2.4 Summary of facies scheme used in this book. Modified from Pickering et al. (1995a).

    Figure 2.5 Descriptive facies for resedimented conglomerates proposed by Walker (1975a, 1976, 1977, 1978).

    Figure 2.6 Facies A1.1: disorganised gravel. (a) Ordovician fill of slope channel at Grosses Roches, Gaspé Peninsula, Quebec. (b) Middle Eocene Gerbe I System, Ainsa Basin, Spanish Pyrenees. (c) Middle Eocene Charo Canyon fill, Ainsa Basin, Spanish Pyrenees. Hammer ∼35 cm long. Note, near-vertical erosional edge of these gravel deposits cut into thin-bedded facies. (d) Middle Eocene Morillo I fan System, Ainsa Basin, Spanish Pyrenees. Note that Facies A1.1 occurs as a decimetre-thick layer of cobbles and gravel ∼25 cm above the human scale.

    Figure 2.7 Facies A1.2: disorganised muddy gravel. (a) Plio–Pleistocene Hata Formation, Ashigara Group, central Japan. Scale divisions 10 cm. (b) Well A3 core photograph, Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. Core width ∼6.5 cm. Depth marker in well is metres below ground level.

    Figure 2.8 Facies A1.3: disorganised gravelly mud. (a) Miocene (Tortonian) Tabernas Basin, southeast Spain. Camera lens cap for scale. (b) Pliocene–Pleistocene Hata Formation, Ashigara Group, central Japan. Notebook 19 cm high. (c) Upper Precambrian Kongsfjord Formation, Finnmark, Arctic Norway. Camera lens cap for scale. (d) Eocene Richmond Formation, Jamaica. Camera lens cap for scale.

    Figure 2.9 Facies A1.4: disorganised pebbly sand. (a) Middle Eocene, Gerbe I System, Ainsa Basin, Spanish Pyrenees. (b) Upper Precambrian, Rybachi-Sredni Peninsula, northern Russia. Camera lens cap for scale. (c) Late Cretaceous–Palaeocene Carmelo (Pigeon Point) Formation, Point Lobos, California (cf. Anderson et al. 2006). Note the crude stratification defined by pebble stringers. Mobile telephone for scale. (d) Well A3 core photograph, Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. Core width ∼6.5 cm. Depth marker in well is metres below ground level.

    Figure 2.10 Facies A2.1: stratified gravel. (a) Cambrian–Ordovician Cap Enragé Formation, Quebec Appalachians. Top is to the left. (b) Late Cretaceous–Palaeocene Carmelo (Pigeon Point) Formation, Point Lobos, California (cf. Anderson et al. 2006). Note the low-angle stratification defined by pebble-rich layers.

    Figure 2.11 Facies A2.2: inversely graded gravel. (a) Cretaceous Nakaminato Formation, Ibaraki Prefecture, Japan. Scale divisions 10 cm. (b) Well A3, Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. Core width ∼6.5 cm. (c) Lower Silurian (Llandovery) Caban Coch conglomerates, Elan Valley, Welsh Basin, UK. Camera lens cap scale.

    Figure 2.12 Facies A2.3: normally graded gravel. (a) Middle Eocene Morillo System, Ainsa Basin, Spanish Pyrenees. Compass-clinometer for scale. (b) Cretaceous Nakaminato Formation, Ibaraki Prefecture, Japan. Scale divisions 10 cm. Note dish structures in underlying sandstone. (c) Middle Eocene Gerbe System, Ainsa Basin, Spanish Pyrenees. 15-cm scale.

    Figure 2.13 Facies A2.4: graded-stratified gravel. (a) Middle Eocene Morillo System, Ainsa Basin, Spanish Pyrenees. Compass-clinometer for scale. (b) Lower Ordovician succession at Grosse Roches, Gaspé Peninsula, Quebec. Gerard Middleton provides scale. (c) Upper Precambrian Kongsfjord Formation, Finnmark, Arctic Norway. Camera lens cap for scale.

    Figure 2.14 Facies A2.5: stratified pebbly sand. (a) Oligocene Annot Sandstone Formation, Contes sub-basin, southeast France. Mobile phone left of centre for scale ∼10 cm long. (b) Close-up of left-hand picture to show detail of stratification in Facies A2.4. Mobile phone at lower left for scale ∼10 cm long. (c) Eocene Richmond Formation, Wagwater Group, Jamaica. Camera lens cap for sale.

    Figure 2.15 Facies A2.6: inversely graded pebbly sand. Eocene, ODP Site 1276A, Newfoundland Basin (210-1276A-9R2, 55–74 cm). Above the large pebbles, the 50 cm-thick deposit becomes normally graded. This SGF deposit is at the base of ∼2 m of amalgamated sand beds. Scale bar 10 cm.

    Figure 2.16 Facies A2.7: normally graded pebbly sand. Pliocene Kiyomusi Formation, Boso Peninsula, Japan. Scale divisions 10 cm.

    Figure 2.17 Facies A2.8: graded-stratified pebbly sand. (a) Lower Ordovician Tourelle Formation, Gaspé Peninsula, Quebec. Arrow 10 cm long. (b) Silurian Milliners Arm Formation, New World Island, Newfoundland, Canada. Note load and flame structures created during deposition with residual shear by the overriding flow, suggesting palaeoflow towards right. Camera lens cap scale.

    Figure 2.18 Facies B1.1: thick/medium-bedded, disorganised sands. All scales in centimetres. (a), (b) Cambrian–Ordovician Cap Enragé Formation, Quebec Appalachians. Paler sediment between dish structures, and paler sediment in fluid-escape pillars, is deficient in matrix because of flushing by escaping pore-water. (c) Lower Silurian (Llandovery) Aberystwyth Grits, Welsh Basin. Amalgamated bed including at least four scour-based event-deposits of structureless sand above a basal unit of Facies B2.1 parallel-stratified sand.

    Figure 2.19 Facies B1.2: thin-bedded, coarse-grained sands. (a) Middle to Upper Ordovician (Caradoc–Ashgill) Point Leamington Formation, north-central Newfoundland, Canada. (b) Nummulite-rich bed, Well L2, fan lateral-margin and basin-slope deposits, Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. Core width ∼6.5 cm.

    Figure 2.20 Facies B2.1: parallel-stratified sands. (a) Fine-grained sandstone example, Permian Karoo System, South Africa. (b), (c) and (d) are from the Lower Ordovician Tourelle Formation, Quebec Appalachians. All scales in centimetres. (b) 165 cm-thick bed, graded overall but consisting of inversely graded spaced stratification ‘bands’. (c) Lower part of graded bed showing stratified division. (d) Detail of part of a division of spaced stratification with arrows at the base of each stratum. Each stratum is inversely graded in its basal 1–1.5 cm from fine sand to medium or coarse sand. (e) Middle Eocene, Ainsa System, Ainsa Basin, Spanish Pyrenees, Well A1. Core width ∼6.5 cm.

    Figure 2.21 Facies B2.2: large-scale cross-stratified sands. (a) Cambrian Blow-me-down Brook Formation, Woods Island, Bay of Islands, western Newfoundland, Canada. Notebook (centre right) is ∼20 cm high. (b) Upper Precambrian Kongsfjord Formation, Finnmark, Arctic Norway. (c) Bedform (∼25 cm thick) of very coarse-grained, cross-stratified, sandstone, Oligocene Grès d'Annot Formation, Peira Cava, southeast France. (d) Close-up of part of Facies B2.2 shown in part (c) to show multi-storey sets; 1 Euro coin for scale. (e) Cross-stratified sands in core. Middle Eocene, Ainsa System, Ainsa Basin, Spanish Pyrenees. Core width ∼6.5 cm.

    Figure 2.22 Facies C1.1: poorly sorted muddy sand. (a) Cretaceous (Albian), ODP Site 1276A, Newfoundland Basin (210-1276A-69R2, 43–65 cm). Note floating and flattened shale clasts and sharp bed top. Scale bar 10 cm. (b) Close-up of part of a 2–3 m-thick SGF deposit, Cretaceous Nakaminato Formation, Ibaraki Prefecture, Japan. Scale divisions 10 cm.

    Figure 2.23 Facies C1.2: mottled muddy sands in core. (a) Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. (b) Middle Miocene, ODP Site 792E, Izu-Bonin forearc basin (126-792E-21R2, 6–18 cm). Scale bar 10 cm.

    Figure 2.24 (a) Facies association of mainly Facies Class C beds (Class C2.1 with minor C2.2). Upper Precambrian Kongsfjord Formation, Finnmark, Arctic Norway. (b) Facies association of mainly Facies Class C beds (Facies C2.2). Lower Silurian (Llandovery), Lake District, England. (c) Facies association of mainly Facies Class C beds (Facies Class C2.2 and C2.3), Upper Carboniferous (Namurian) Ross Formation, County Clare, western Ireland. Compass-clinometer scale right of centre.

    Figure 2.25 Facies C2.1, 2.2 and 2.3. (a) Graded Facies C2.1 sand–mud couplet without tractional structures, Lower Miocene, ODP Site 793, Izu-Bonin forearc basin (126-793B-19R3, 70–95 cm). Scale 10 cm. (b) Bouma Tbcd bed of Facies C2.2, Lower Oligocene, ODP Site 792, Izu-Bonin forearc basin (126-792E-56R4, 13–34 cm). Scale 10 cm. (c) Rippled lenses of very fine sand, ODP Site 942, Amazon Fan (155-942A-6H5, 93–109 cm). (d, e) Variant of Facies C2.3, thin-bedded sand–mud couplets as very coarse-grained beds, Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. (f) Climbing ripple-lamination or ripple drift, Middle–Upper Ordovician Cloridorme Formation, Gaspé Peninsula, Quebec, Canada. Scale divisions 10 cm.

    Figure 2.26 Facies C2.4: thick- to very thick-bedded, mud-dominated sand–mud couplets. (a) Facies C2.4 with lighter-coloured, thick, structureless, silty mudstone caps interbedded with mainly Facies Class C turbidites, Upper Cretaceous, Flysch di Monte Cassio, belonging to the so-called ‘Helminthoid Flysch’. Note river and village for scale. (b) Basal part of Facies C2.4 bed, Cloridorme Formation, Middle–Upper Ordovician (Caradoc–Ashgill), Gaspé Peninsula, Quebec, Canada. Note flow reversals shown by bi-directional cross-stratification at base of bed, (see arrows); 5 cm-scale. (c) Middle part (below mud cap) of Facies C2.4 bed, Cloridorme Formation, Middle–Upper Ordovician (Caradoc–Ashgill), Gaspé Peninsula, Quebec, Canada. Gradually thinning of laminated to muddy/pseudonoduled couplets is attributed to the seiche created when a large-volume gravity flow was ponded in a basinal depression. (d) Inverted ∼12 m-thick Facies C2.4 bed near Cotefablo, Middle Eocene Jaca Basin, Spanish Pyrenees. Person (∼2 m high) has top of head coincident with the base of sandstone part of bed, with prominent grey-coloured structureless upper part to right. Note well-developed slaty cleavage in silty part of bed.

    Figure 2.27 Facies C2.5: medium to very thick-bedded, mud-dominated, slurried sand–mud couplets. (a) Sedimentary structure divisions M6 and M2b of slurry-flow beds. The M6 division shows a mixed and deformed unit of very fine-grained sand and silt that passes upwards into deformed mudstone. From Lowe et al. (2003). (b) Sedimentary structure divisions M1, M2, M3, M4 and M5 of slurry-flow beds. M1 shown here is a thin flat-laminated layer that marks the base of slurry bed 50, well 16/26-B1. It immediately overlies a mudstone unit and is overlain by an M2a mixed slurried division. M2a (base of bed 46, well 16/26-B8) shows a thin M2c mesobanded division overlain by a very thick mixed slurried to megabanded M2a unit. M2b and M2c (bed 46, well 16/26-24) consist of macrobanded and mesobanded sandstone respectively. The M3 division of wispy laminations (beds 34/36, well 16/26-B8) shows characteristic fine subvertical water-escape channels (light streaks). Dish-structured M4 division (bed 62, well 16/26-B5) shows moderately curved dishes cross-cut by large, vertical and subvertical water-escape channels (light). Core at far right (top of bed 86, well 16/26-B10) shows an M4 division of faint flat dish structures overlain by an M5 division of fine, flat sandy microbanding that grades upwards into laminated silt and mud without cross-laminations or other current structures. Scale units shown on left side of core segments are 0.1 foot (30 mm) long. From Lowe et al. (2003).

    Figure 2.28 Facies D1.1: structureless silts. Structureless, closely spaced, thin coarse silt beds from levée deposits of the Yellow Channel–Levée System, Amazon Fan (ODP Leg 155: Interval 155-937C-6H-3, 36–58 cm). After Normark et al. (1997).

    Figure 2.29 Facies D1.2: muddy silts interbedded with Facies B1.2 and C2.3 sandstones, Miocene Monterey Formation, California, USA (pencil 14 cm long). After Surpless et al. (2009).

    Figure 2.30 Facies D1.3: mottled silt and mud. Cretaceous (Cenomanian), ODP Site 1276A, Newfoundland Basin (210-1276A-32R1, 48–62 cm). Scale bar 10 cm.

    Figure 2.31 Facies D2.1: graded-stratified silt. (a) Lower Silurian (Llandovery) Aberystwyth Grits, Welsh Basin; 15 cm scale bar. (b) Cretaceous (Cenomanian–Albian), ODP Site 1276A, Newfoundland Basin (210-1276A-37R3, 95–114 cm). Scale bar 10 cm. The highest essentially structureless bed above the scale bar is Facies D1.1. (c) Permian Karoo System, South Africa. Coin scale ∼2 cm diameter. (d) Upper Precambrian Kongsfjord Formation, Finnmark, Arctic Norway. Camera lens cap for scale.

    Figure 2.32 Facies D2.2: thick irregular silt and mud laminae. (a) and (b) Haydrynian (upper Proterozoic) Conception Group, False Cape, Avalon Peninsula, Newfoundland. Thickness 18 cm at left, scale in centimetres at right. (c) Upper Jurassic (Kimmeridge) Boulder Beds, northeast Scotland. This has been referred to as ‘tiger-stripe’ facies.

    Figure 2.33 Facies D2.3: thin regular silt and mud laminae in core. (a) Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. Inclination of beds due to coring through dipping section. (b) ODP Site 931, Amazon Fan (155-931B-4H6, 116–121 cm). Scale bar 10 cm. (c) Photographic print from thin section of silt (dark) and mud (pale) laminae like those in the middle image, ODP Site 931, Amazon Fan (155-931A-4H2, 78–80 cm). Thickness shown 1.5 cm.

    Figure 2.34 Facies E1.1: structureless muds. Clayey nannofossil ooze, interpreted as turbidite mud (Te), Upper Pliocene–Pleistocene ODP Leg 135 (Lau Basin, southwest Pacific) Site 835 (Interval 135-835A-1H-4, 20–45 cm). (a) Core photograph and (b) X-radiograph of same interval. Core width ∼6 cm. ODP, Site 835 (cf. Rothwell et al. 1994).

    Figure 2.35 Facies E1.2: varicoloured muds. ODP Site 934, Amazon Fan (155-934A-1H3, 7–22 cm). The colour banding is accentuated by FeS staining. Scale bar 10 cm.

    Figure 2.36 Facies E1.3: mottled muds. (a) ODP Site 941, Amazon Fan (155-941B-1H4, 12–31 cm). Scale bar 10 cm. (b) Well A6, Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. Scale bar in centimetres.

    Figure 2.37 Facies E2.1: graded muds. (a) Pale (green shades) and dark grey banded muds, Cap des Rosiers Formation, Cap Ste-Anne, Quebec. The pale units have the sharpest bases so are inferred to be the SGF deposits, capped by dark grey, presumably hemipelagic muds. Scale ∼45 cm long. (b) Silty mud, poorly sorted and carbonaceous near base grading up into fine mud, DSDP Leg 96, Mississippi Fan, Gulf of Mexico (Sample 615-22-2, 15–45 cm). After Stow et al. (1986).

    Figure 2.38 Facies E2.2: laminated muds and clays from ODP Leg 169S, Saanich Inlet, off Vancouver Island, British Columbia (see Bornhold et al. 1998). (a) Alternations of paler laminae (diatom ooze) and darker laminae of diatomaceous mud with rare very light-coloured sub-laminae of silty muds (Core interval 169S-1034B-3H-6, 85–98 cm). (b) X-ray radiograph of alternating laminated sediments showing cm-scale pale (diatom-rich) and dark layers (containing terrigenous sediment) interpreted as annual ‘varves’, within which there are up to 10 intra-annual sub-laminae (Core interval 169S-1034B-4H-3, 14–34 cm). (c) Cretaceous (Albian) ODP Site 1276A, Newfoundland Basin (210-1276A-72R5, 103–111 cm). Pyrite nodules attest to the sulfidic nature of these organic-rich laminated muds. Scale bar 6.5 cm long.

    Figure 2.39 Facies F1.1: rubble. (a) Deep-towed camera photograph of rock talus with attached coral on the volcanic slope of Mata Fitu submarine volcano in the NE Lau Basin ∼150 km SW of Samoa. Water depth ∼2440 m. The two green dots just above coral are lasers with separation of 15 cm. NOAA- Earth Ocean-Interactions and WHOI-MISO TowCam (http://www.whoi.edu/page.do?pid=17619). (b) Oligocene, ODP Site 793, Izu-Bonin forearc basin (126-793B-87R2, 67–84 cm). Primary voids between volcanic clasts are filled with diagenetic cement. (c) Rubble as unsorted talus covering a larger block, Logatchev hydrothermal field (14°45′N, Mid-Atlantic Ridge). From Petersen et al. (2009).

    Figure 2.40 Facies F1.2: dropstones and isolated ejecta. (a) Dropstone in background mudstone from ODP Leg 119 Site 739, continental shelf of Prydz Bay, East Antarctica, Core 119-739C-34R-3, 68–87 cm (see Hambrey et al. 1991). (b) ODP Site 645, Baffin Bay, Canadian Arctic (105-645F-3H4, 31–61 cm). Isolated dropstones occur in mud above a pale-coloured ice-rafted unit rich in detrital carbonate. Scale bar 10 cm.

    Figure 2.41 Facies F2.1: coherent folded and contorted strata. (a) Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. (b) ‘Black Flysch’ (Deva Formation), Arminza, northern Spain. Scale in middle fold axis 1 m long. (c) Pliocene–Pleistocene, Miura Peninsula, southeast Japan. (d) Middle Ordovician Cloridorme Formation, Gaspé Peninsula, Quebec. Scale divisions 10 cm. (e) Ross Formation (‘Ross Slide’), Upper Carboniferous (Namurian), Bridges of Ross, County Clare, western Ireland. Human scale in centre of image.

    Figure 2.42 Facies F2.2: brecciated and balled strata. (a) Middle to Upper Ordovician (Caradoc–Ashgill) Point Leamington Formation, north-central Newfoundland, Canada. Hammer for scale. (b) Small-scale syn-sedimentary faults within Facies F2.1, Upper Jurassic (Kimmeridgian), northeast Scotland.

    Figure 2.43 Facies G1.1: biogenic ooze. Siliceous bioturbated radiolarian (biogenic) mudstones (chert), occurring as both red (a) and grey deposits (b), Middle to Upper Ordovician (Caradoc–Ashgill) Point Leamington Formation, north-central Newfoundland, Canada. Examples of biogenic oozes have been interpreted as associated with major oceanographic events, for example from ODP Leg 207 in a transect at Demerara Rise, western tropical Atlantic (cf. Mosher et al. 2007). (c) δ¹³Corg record from ODP Site 1258 showing the excursion considered to represent the Late Cretaceous OAE 2 (Oceanic Anoxic Event 2; cf. Erbacher et al. 2005). The adjacent photo is a highly vertically compressed photomosaic of recovered cores. Depths are in metres below the seafloor. The inset plate is a sample core image within OAE 2, displayed with the correct aspect ratio. (d) δ¹³C record at ODP Site 1258 through the Palaeocene/Eocene Thermal Maximum (PETM) (after Nuñes & Norris 2006). The adjacent core photo is highly vertically compressed. The inset core photograph is of the boundary and is shown with the correct aspect ratio.

    Figure 2.44 Facies G1.2: muddy ooze. Structureless mud-bearing diatom ooze (top part) with one lamina in interval 178-1099A-2H-3, 31–32 cm, of Chaetoceros spp. spores, and the top part of a turbidite in interval 178-1099A-2H-3, 7–42 cm. Palmer Deep, ODP Leg 178 Site 1099. After Barker et al. (1999).

    Figure 2.45 (a) Facies G2.1: biogenic mud. Radiolarian ooze (interval 191-1179C-20H-6, 30–50 cm). Northwestern Pacific Ocean, ODP Site 1179. After Kanazawa et al. (2001). (b) Holocene diatomaceous ooze, IODP Expedition 318 Site U1357. Note the distinct seasonal laminations.

    Figure 2.46 Meso- and small-scale sandstone dykes and sills. (a) Sandstone clastic dyke, Upper Precambrian Konsgfjord Formation, Finnmark, Arctic Norway. Lower thick sandstone bed shows local normal bedding orientation and was donor bed for the dyke that intruded up to just below human scale. (b) Pipe or pillar structure in fluidised bed, Upper Precambrian Konsgfjord Formation, Finnmark, Arctic Norway. The consistent sense of overturning of the pipes/pillars is approximately downcurrent as deduced from sole marks and current ripples in surrounding beds. (c) Top surface of sandstone bed that shows pervasive dish-and-pillar structure, Upper Precambrian Konsgfjord Formation, Finnmark, Arctic Norway. Penknife for scale. Note the preferred orientation of the sandstone injections that has the same orientation as the strike of palaecurrents in surrounding beds. (d; facing page) Clastic dyke of scoriaceous and pumiceous material cutting through volcaniclastic deposits and offset by layer-parallel faults (left of scale), Miocene Misaki Formation, Miura Peninsula, Japan; 15-cm bar scale. (e) Top view of sandstone injections (dykes) with a preferred orientation immediately below an MTD (fallen block), Ainsa I Fan, Middle Eocene Ainsa Basin, Spanish Pyrenees. The emplacement of the MTD may have been responsible for the injections; 10 cm scale bar.

    Figure 2.50 Sand volcano, Ross Formation, Namurian Clare Basin, western Ireland. Pen for scale ∼14 cm long.

    Figure 2.47 Large-scale sandstone dykes and sills from the Upper Cretaceous–Lower Palaeocene Marca Canyon, Panoche Hills, California. In these plates, the sandstones are lighter grey and the mudstones dark grey. See Hurst et al. (2011).

    Figure 2.48 Part of large clastic dyke–sill complex in the Miocene Monterey Formation correlative stratigraphic interval, Yellow Bank Beach, near Santa Cruz, California. Note near-vertical, dark-coloured, hydrocarbon-bearing dyke in centre. The yellow/orange-coloured lower half of the outcrop left of centre is a clastic sill that fed the dyke seen in the centre. From Boehm and Moore (2002).

    Figure 2.49 Synoptic diagram of a sand injectite complex with a tripartite architecture based on outcrop and subsurface observations in host rock (dark grey). Remobilised parent sandstone units (yellow arrows); sandstone dykes (red arrows) and sills (blue arrows); irregular sandstone intrusions (orange arrow); sandstone extrudites, for example, sand volcanoes (green arrow). From Hurst et al. (2011).

    Figure 2.51 Example of channel facies association, Ainsa II Fan, Middle Eocene Ainsa System, Ainsa Basin, Spanish Pyrenees. ‘Base’ refers to the inferred erosional base of one of the nested (laterally offset-stacked) submarine channels observed in the nearby outcrop (cf. Pickering & Corregidor 2005; Pickering & Bayliss 2009). Core depths in metres below ground level, younging from right to left.

    Chapter Three: Deep-water ichnology

    Figure 3.1 Examples of bio-deformational structures and trace fossils in core (Ainsa core L1, core width ∼6 cm). (a) Trace fossils associated with fine- to medium-grained sandstones and laminated siltstone–mudstone intervals (Th = Thalassinoides; Oph = Ophiomorpha; Sk = Skolithos; Pl = Planolites). (b) Bio-deformational structures overprinted by later trace fossils (Ch = Chondrites).

    Figure 3.2 Stratinomic classification of trace fossils in relation to the casting medium (e.g., sandstone). Redrawn from Seilacher (1964a), Martinsson (1965) and Bromley (1996).

    Figure 3.3 Block diagrams of common trace fossils in deep-water environments.

    Figure 3.6 Images of common trace fossils in Ainsa Basin cores. The core is ∼6 cm wide. (a) Ophiomorpha (O) in coarse-grained sandstone. (b) Phycosiphon (P), Planolites (Pl), Thalassinoides (T) and Chondrites (C) in mudstones. (c) Planolites (Pl) and Scolicia (S) in medium-grained sandstones and siltstones. (d) Skolithos (Sk) and Scolicia (S) in fine-grained sandstones, siltstones and mudstones. (e) Nereites (N) and Phycosiphon (P) in fine-grained sandstones and mudstones. (f) Planolites (Pl), Thalassinoides (T) and Palaeophycus (Pal) in fine-grained sandstones and siltstone–mudstone couplets. (g) Scolicia (S) in fine-grained sandstones and siltstone–mudstone couplets. (h) Thalassinoides (T), Halopoa (H), Planolites (Pl), Ophiomorpha (O) and Palaeophycus (Pal) in medium-grained sandstones and siltstone–mudstone couplets. (i) Teichichnus (Tei) in medium-grained sandstones. (j) Thalassinoides (T), Halopoa (H) and Scolicia (S) in medium-grained sandstones and siltstone–mudstone couplets.

    Figure 3.4 Outcrop images of trace fossils from the Middle Eocene Ainsa–Jaca basins, Spanish Pyrenees. (a) Halopoa storeana isp. n. (Uchman 2001). Hypichnial full-relief. Scale bar: 2 cm. See Uchman (2001) for discussion. (b) Thalassinoides suevicus (Rieth 1932). Hypichnial full-relief. Scale bar: 2 cm. See Howard and Frey (1984) for discussion. (c) Chondrites intricatus Brongniart 1823. Endichnial full-relief. Scale bar: 2 cm. See Uchman (1998) and Fu (1991) for discussion. (d) Teichichnus isp. Epichnial full-relief. Scale bar: 2 cm. See Seilacher (1955) for discussion. (e) Planolites ispp. Endichnial full-relief. Scale bar: 2 cm. See Pemberton and Frey (1982) for discussion. (f) Ophiomorpha rudis (Książkiewicz 1977). Epichnial full-relief. Scale bar: 2 cm. See Uchman (2001) for discussion.

    Figure 3.5 Outcrop images of trace fossils from the Ainsa–Jaca basins. (a) Zoophycos insignis (Squinabol 1890). Endichnial full-relief. Scale bar: 2 cm. See Uchman (1999) for discussion. (b) Helminthopsis ispp. Hypichnial full-relief. Scale bar: 2 cm. See Wetzel and Bromley (1996) for discussion of ichnogenus. (c) Scolicia plana (Książkiewicz 1970) Endichnial full-relief. Scale bar: 2 cm. See Uchman (1998) for discussion. (d) Phycosiphon incertum (Fischer-Ooster 1858). Endichnial full-relief. Scale bar: 2 cm. See Wetzel and Bromley (1994). (e) Nereites missouriensis (Weller 1899) Endichnial full-relief. Scale bar: 2 cm. See Uchman (1995) for discussion.

    Figure 3.7 Block diagrams of common graphoglyptids in deep- water environments.

    Figure 3.8 Outcrop images of trace fossils from the Ainsa–Jaca basins. (a) Lorenzinia nowaki Książkiewicz 1970. Hypichnial semi-relief. Scale bar: 2 cm. See Uchman (1998) for discussion. (b) Helminthorhaphe japonica (Tanaka 1970) Hypichnial semi-relief. Scale bar: 2 cm. See Uchman (1998) for discussion. (c) Spirohaphe involuta (De Stefani 1895). Hypichnial full-relief. Scale bar: 2 cm. See Seilacher (1977) for discussion. (d) Cosmorhaphe sinuosa (Azpeitia Moros 1933) Hypichnial semi-relief. Scale bar: 2 cm. See Seilacher (1977) for discussion. (e) Megagrapton submontanum (Azpeitia Moros 1933). Hypichnial semi-relief. Scale bar: 10 cm. See Uchman (1998). (f) Paleodictyon strozzii (Meneghini 1850). Hypichnial semi-relief. Scale bar: 1 cm. See Uchman (1995) for discussion.

    Figure 3.9 Examples of deep-water trace-fossil assemblages. (a) Post-depositional deposit assemblages within and at the top of SGF deposits. (b) Post-depositional and pre-depositional assemblages at the base of SGF deposits.

    Figure 3.10 Diverse trace-fossil assemblage on the sole of a SGF sandstone bed, Middle Eocene Ainsa Basin, Spanish Pyrenees. The pre-depositional trace-fossil assemblage (Paleodictyon, Helminthorhaphe, Gordia) is preserved by the depositing SGF and is later cross-cut by a post-depositional trace-fossil assemblage (Ophiomorpha).

    Figure 3.11 Examples of ichnofabric indices (sensu Droser & Bottjer 1986), Ainsa 6 core, Ainsa Basin, Spanish Pyrenees (see text for details). The core is ∼6 cm wide. From Heard et al. (2008).

    Figure 3.12 Main zones of bioturbation associated with SGF deposits (see text for details). From Uchman (1999).

    Figure 3.13 Summary diagram of average bioturbation, ichnodiversity, number of pre- and post-depositional trace fossils and number of graphoglyptids for the most characteristic environments of the deep-water complex of the Ainsa–Jaca basins. Redrawn after Heard and Pickering (2008).

    Figure 3.14 Summary of characteristic trace-fossil assemblages in submarine fan and related environments of the Ainsa–Jaca basins. (L = Lockeia; Sk = Skolithos; Ar = Arenicolites; Nu = Nummulites-lined burrow; H = Halopoa; Pl = Planolites; Ch = Chondrites; Oph = Ophiomorpha; Th = Thalassinoides; S = Saerichnites; Lor = Lorenzinia; Gl = Glockerichnus; Z = Zoophycos; Phy = Phycosiphon; Lo = Lophectenium; N = Nereites; Sc = Scolicia; Tae = Taenidium; Pr = Protovirgularia; Cos = Cosmorhaphe; He = Helicolithus; H = Helminthorhaphe; Hel = Helminthopsis; Sp = Spirohaphe; Pa = Palaeomeandron; De = Desmograpton; Ur = Urohelminthoida; M = Megagrapton; Pal = Paleodictyon). Redrawn from Heard & Pickering (2008).

    Figure 3.15 Block diagram illustrating common trace fossils in individual ichnofabrics from channel-axis to external levée-overbank and interfan. The average bioturbation intensity associated with each ichnofabric is illustrated in the graph, whilst the relative proportion of sandstone (yellow) versus marlstone (grey) associated with each ichnofabric is also provided.

    Chapter Four: Time–space integration

    Figure 4.1 Cenozoic part of the Haq et al. (1987) sea-level record.

    Figure 4.2 Comparison of back-stripped sea-level records from New Jersey (blue, Miller et al. 2005a; brown, Kominz et al. 2008), Russian platform (pink, Sahagian et al. 1996), and Scotian margin (grey, Steckler & Watts 1978) with Exxon Production Research Company (EPR; green = Vail et al. 1977a; black = Haq et al. 1987; maroon = Haq & Al-Qahtani 2005). Note the much higher amplitude of the EPR estimates. From Miller et al. (2011).

    Figure 4.3 (a) Components of the lowstand systems tract and lowstand fan. (b) The lowstand systems tract, lowstand wedge. From (Posamentier et al. (1988). (c) Comparison of the original depositional model of Vail (1987) and Posamentier and Vail (1988a), and more recent ‘in reality’ perceptions about the position of siliciclastic systems in deep-water basins from the SEPM Stratigraphy Web (http://www.sepmstrata.org/page.aspx?pageid=41). HST = highstand system tract; LST = lowstand system tract, and TST = transgressive system tract.

    Figure 4.4 (a) Candidate Type I sequence boundary and palaeo-valley fill, Namurian Tulig cyclothem, Clare Basin, western Ireland. Above the unconformity (arrow), there are non-marine fluvial (deltaic?) sediments, including coal horizons. Below the unconformity, there is an interval of offshore marine prodelta fine-grained deposits, below which several packets of sandstones are seen. Cumulative cut-down visible in this image is ∼4 m, but total observed cut down is ∼15 m. (b) Flute casts in fluvial sandstones at the base of the palaeo-valley fill shown in (a) (c) Regional marine flooding surface at top of palaeo-valley-fill sandstones shown in (a). Surface characterised by abundant Zoophycos and phosphate concretions.

    Figure 4.5 The Early Toarcian section from Ponta do Trovão and Praia do Abalo, Peniche, Portugal, currently the sole candidate for the Global Stratotype Section and Point (GSSP). The graphic log and sample heights are based on the section measured only in the cliff exposures at this locality. (Ammonite biostratigraphy from Mouterde (1955).) (a) Total Organic Carbon (TOC) data from hand specimens; Sr-isotope data from belemnites, normalised to a value of 0.710250 for NIST 987. (b) Stratigraphic distribution of SGF deposits, including debrites, shown together with insoluble residue from bulk rock samples. (c) High-resolution C-isotope data from bulk sediment, fossil wood and belemnites. Horizons labelled 1, 2, 3 and 4 represent key levels correlatable between Yorkshire (England) and Peniche (Portugal). The single black shale at 21.5 m in the section with a TOC value of 2.6% has a δ¹³Corg value of −29.5‰. From Hesselbo et al. (2007). See text for discussion.

    Figure 4.6 Changing character of sequence boundaries of Mutti (1985) from the basin margin into the deeper parts of the basin. In this hypothetical example, the sequences are complete and show an evolution from Type I to Type II to Type lII systems. See text for explanation.

    Figure 4.7 Sea-level changes and resulting facies cyclicity in a submarine fan system comprising channel-fill and lobe elements. Sub-stages result from minor oscillations in sea level on a major fall/rise excursion. Falls in sea level can result in proximal unconformities and distal sandstone packets (Facies Class B and C) 3–15 m thick. Minor rises can result in proximal backfilling of channels and distal deposition of thin-bedded lobe-fringe deposits (Facies Class D). From Mutti (1985). See text for explanation.

    Figure 4.8 Channel-fill models to show three principal phases in the history of a channel. Changes in the channel phases are related to changes in relative base level (accommodation space, eustasy and tectonic uplift/subsidence). (a) Channel–levée–overbank complexes. (b) Canyon and non-levéed channels. Modified from Pickering et al. (1995b).

    Figure 4.9 Schematic depiction of the relationship between relative sea level and type of dominant mass-flow process. The succession comprises cohesive-flow deposits at the base (corresponding to the initial period of relative sea-level fall), overlain by frontal-splay-dominated and then levéed-channel dominated sections (corresponding to the subsequent period of early and late relative sea-level lowstand, respectively). The succession is capped by deposition of cohesive-flow and condensed-section deposits (corresponding to periods of rapid sea-level rise and highstand, respectively). From Posamentier and Kolla (2003).

    Figure 4.10 (a) Seismic reflection profile from offshore Indonesia illustrating the stratigraphic succession of a deep-water sequence. Cohesive-flow deposits (1) are overlain by frontal-splay deposits (i.e., sandy lobes) (2), channel–levée deposits (3), and again cohesive-flow deposits (4). The entire succession is then inferred to be mantled by a thin veneer of condensed-section deposits (5). (b) Schematic depiction of an idealised deep-water depositional sequence, with two hypothetical log profiles. From Posamentier and Kolla (2003).

    Figure 4.11 Summary diagrams and palaeogeographic reconstructions illustrating the influence of shelf width between canyon heads and the littoral zone on canyon–channel system activity. From Covault et al. (2007).

    Figure 4.12 Age model for key reflectors showing timing of principal coarse-sediment supply to the Hueneme, Mugu and Dume fans. Sea-level curve from Lambeck and

    Enjoying the preview?
    Page 1 of 1