You are on page 1of 38

4 Sediment-Gravity Flows and Their Processes

4
Definition of Sediment Gravity Flow
How are sediments transported to, and deposited in, deepwater environments? This
3 question has been the subject of discussion and debate since the 1870s and researched since
the 1950s (see Shanmugam, 2000, for a comprehensive review). Collectively, the primary pro-
cesses that transport sediment into deep water environments are called sediment gravity
TOC flows (Middleton and Hampton, 1973). Sediment gravity flows range from mass movements
(rock falls) and cohesive debris flows at one end of the spectrum to fully turbulent flows at the
other end of the spectrum (Fig. 4-1). Although the term turbidity current is often used to
describe deep water transport processes, it is only one of several types of flows that transport
Start
sediment into deep water, and rework them once deposited.
Unfortunately, transport and depositional processes are only rarely observed or measured
in deep marine and lacustrine environments. Thus, to understand sediment gravity flows, we
Ref. List must combine data sets from outcrops, subsurface well logs and cores, seismic reflection pro-
files, modern sea floor images and samples, laboratory flume experiments, and mathematical
models. Fortunately, rapidly evolving imaging, measurement, and computing technologies
Search mainly driven by petroleum exploration and developmentare providing new data and insights
that are leading to improved understanding of the complexities of sediment gravity flows and
their deposits.
This chapter summarizes our current state of knowledge by combining key historical
Help
concepts with more recent observations and analyses. The chapter is organized to discuss the
initiation of sediment gravity flows, followed by the spectrum of processes and resulting
deposits, flow combinations and transformations, and temporal and spatial variations in pro-
cesses. Processes of post-depositional reworking of sediment gravity flows on the sea floor are
also discussed in a separate section. Finally, mention is made of allocyclic and autocycle pro-
cesses, since vertical stacking patterns are important to interpreting larger scale evolution of
sedimentary sequences and basin fill.

The Sediment Gravity Flow Continuum


Sediment gravity flow Sediment Support Mechanism
Increasing sediment concentration

turbidity currents fluid turbulence


fluidized flow hindered settling
liquified flow hindered settling
grain flow dispersive pressure
debris flow matrix strength
slump-slide matrix strength
creep matrix strength
rock fall

Figure 4-1. Chart listing the different types of sediment gravity flows by their corresponding sed-
iment support mechanism. The classification is mainly after Lowe (1982). The figure was pro-
vided by D. Pyles (2002).
k

4-111
Sediment-Gravity Flows and Their Processes

We want to stress that understanding of sediment gravity flow processes is continually


4 evolving as results of new research are presented at technical meetings, research conferences
and in publications. Because of this, the sections in this chapter summarize key points of deep-
water processes and deposits and are not meant as exhaustive compilations of the existing
3 literature.

Generation and Frequency of Occurrence of Sediment Gravity Flows


TOC
Upslope sediment gravity flows can be triggered by: (a) seismically-generated slides;
(b) instability and slope failure resulting from rapid sedimentation, oversteepening and/or
change in pore pressure; (c) hyperpycnal flow (underflow) produced when dense, high sedi-
Start ment concentration river effluent discharges into the sea, and (d) fine-grained underflows
which trigger downslope sandy flows (Kneller and Buckee, 2000; Mulder et al., 2001a). Pro-
cesses (a) and (b) are collectively termed ignitive flows, and (c) and (d) are termed non-
Ref. List ignitive flows. Sudden discharge of gas hydrates (clathrates) upward through slope sediment
to the sea floor are also thought to be capable of initiating ignitive flows on submarine slopes.
The 1929 Grand Banks of Newfoundland, Canada earthquake generated an ignitive flow
Search by slope failure (Piper et al., 1988; Cochonat and Piper, 1995; Mulder et al, 1997; Piper et al,
1999). Here, a single flow transported sediment for several hundreds of km into the deep Atlan-
tic ocean floor. A second example of turbidity current generation by ignitive processes is the
1979 failure of part of the Nice, France airport (Piper and Savoye, 1993). The frequencies of
Help occurrence of ignitive sediment gravity flows are not recorded, but in seismically-active areas,
we may speculate that they could occur on the order of one flow every 10s to 100s of years.
At a smaller, and more recordable time scale are hyperpycnal flows, which are sustained
flows that form at a river mouth during periods of high river discharge and move along the sea
floor due to excess density relative to the ambient sea water (Mulder et al., 2003). Hyperpycnal
flows can be generated at a frequency of years from rivers with extremely high suspended load
(Mulder and Syvitski, 1995; Mulder et al., 2003). Eighty-four percent (84%) of the worlds riv-
ers can produce hyperpycnal flows and can account for 53% of the worlds oceanic sediment
load. During the time period between 1887 and 1937, 30 submarine cable breaks were recorded
in the Congo Submarine Canyon (West Africa), apparently due to turbidity currents generated
during times of high bed-load discharge from the Congo River (Heezen et al., 1964). On the
Amazon Fan, age-dating of a 240 m thick, 20 ka year interval of thin-bedded and laminated,
muddy sediment indicate an average occurrence of 1 flow event every 2 years (Pirmez et al.,
2000). The Amazon Fan flows also are thought to have been triggered by flood discharges from
the Amazon River system. In a core collected from the Var submarine canyon, located in the
western Mediterranean Sea, 1314 hyperpycnal flow deposits were recorded during the past
100 years and 910 were recorded during the past 50 years; this represents a frequency of
occurrence of one hyperpycnal turbidity current every 57.5 years (Mulder et al., 2001a). An
intense sediment gravity flow was recorded in the Zaire submarine valley in March 2001. This
flow demonstrated that even during periods of relatively high sea level, sediment can be trans-
ported to deep water, in this case because the Zaire River is connected to the canyon and fan
valley for a length of 760km (Khripounoff et al., 2003).

Sediment Support Mechanisms and Types of Sediment Gravity Flows


Once a sediment gravity flow is generated, a variety of processes may transport the sedi-
ment into the deep ocean, or into the deeper parts of lakes. Numerous studies suggest that the
nature of a particular flow and its resultant deposit are a function of: (a) gravity, (b) velocity
and fluid pressure within the flow in time and space (c) sediment support mechanism within the

4-112
Sediment Support Mechanisms and Types of Sediment Gravity Flows

flow, (d) size frequency distribution, composition, and concentration of particles available for
4 transport, (e) topography of the seabed over which the flow is transported and deposited, and (f)
frictional forces between the seabed and the flow.
Lowe (1982) classified sedment gravity flow types according to flow behavior and sedi-
3 ment support mechanism. Sediment support mechanisms are illustrated in Figure 4-2. Creep
and rockfall flow types, listed in Figure 4-1, were not included in Lowes (1982) original
classification.
TOC According to Lowe (1982), the support mechanism that prevails at any given instant dur-
ing flow is principally a function of the flow velocity and concentration of sediment within the
flow. Fluid turbulence (Newtonian flow) is the process of random motion of fluids within the
Start flow, and dominates under conditions of relatively low sediment concentration. Turbidity cur-
rents are characteristic of fully-turbulent flow. Sediment concentration is low enough that
upward-directed turbulence supports particles in suspension during transport.
Ref. List With increasing sediment concentration, particles begin to settle from the flow, but
upward flow of displaced water results in hindered settling of the particles (Fig. 4-2). With fur-
ther increases in sediment concentration, dispersive pressure, caused by grain-to-grain
Search collisions, becomes the predominant support mechanism (Fig. 4-2). At even higher sediment
concentrations, plastic (Bingham) flows with a high matrix strength predominate (Fig. 4-2).
The ratio of sediment to fluid is great enough to fully entrain the sediment as it flows in a plas-
tic or laminar fashion.
Help
Critical sediment concentrations, which distinguish plastic, fluidal or intermediate
flows, are not presently clearly defined. Based on a compilation of published sediment concen-
tration values of various flow types, Shanmugam (2000) suggests that the boundary between
laminar and turbulent flows occurs at about 2025% sediment volume within the flow. Recent
experiments by Baas and Best (2001) have suggested that 34% clay content in a flow is the
critical concentration distinguishing laminar from turbulent flows. Marr et al. (2001) place the
critical clay concentration in experimental flows at 0.75 wt. % when the clay is bentonite and
725 wt. % when the clay is kaolinite. Numerical modeling has suggested a critical sediment
concentration of 10% (Pratson et al., 2000).

Sediment Support Mechanisms

Fluid turbulence
-random motion of fluid in eddies
Increasing sediment concentration

Hindered settling
-sediment begins to settle out of the flow,
space required for a grain to fall makes
water move upward, providing a lift force
Dispersive pressure
-interaction of grains with one another,
rattling of grains against each other,
happens when shear occurs
Matrix strength
-cohesion, usually provided by fines

Figure 4-2. Diagrams illustrating the sediment support mechanisms (after Lowe, 1982). With
increasing concentration of sediment within a flow, the support mechanism changes from fluid
turbulence to hindered settling, to dispersive pressure to matrix strength. The figure was pro-
vided by D. Pyles (2002).
4-113
k
Sediment-Gravity Flows and Their Processes

Hyperpycnal flows differ from ignitive flows because the water within the flow is fresh
4 rather than denser sea water. To overcome density contrasts and buoyancy effects, suspended
sediment concentrations in excess of 36kg/m3 are required for the flow to sink to the sea floor
3 and move downslope. Sedimentation rates on the order of 12 m/100yrs. can be generated by
hyperpycnal flows (Mulder et al., 2003).

TOC Sediment Gravity Flow Processes and Deposits


Historically, the most commonly cited classifications of sediment gravity flow deposits
Start are those of Walker (1978) (Fig. 4-3), Lowe (1982) (Fig. 4-4), Mutti et al. (1999) (Fig. 4-5)
and Kneller (1995) (Fig. 4-6). These classifications are discussed in subsequent sections of this
chapter. Each author uses different terminology to describe the deposits of approximately
Ref. List equivalent flow types. An attempt at comparing the deposits of these flow types is shown in
Figure 4-7, although comparisons are not possible for all categories by all authors. In the fol-
lowing sections, the deposits associated with the three main flow typesplastic flows,
Search intermediate flows, and fluidal, turbulent flowsare discussed, along with more in-depth dis-
cussion of associated sediment gravity flow processes.

Help
Debris Pebbly ssts Massive Classical
flow Conglomerates sandstone turbidites
Time and/or space

Slump
Slide
Remolding
Liquefaction 4
(Traction)
2 Debris flow
Decreasing concentration

Traction
(Traction)
2
Upward flow
Grain of pore fluids 1
interaction
Fluid
1 3
turbulence 2
1 High concentration
turbidity current
Rivers
in flood 1 3 4
Low concentration
turbidity current Support mechanisms

Main long-distance
Flow initiation Late stage modifications
transport process

Figure 4-3. Chart illustrating the sediment transport processes of sediment gravity flows and resulting depos-
its. Sediment support mechanisms (inset) are those illustrated in Figure 4-2; i.e., 1= fluid turbulence, 2 = hin-
dered settling, 3 = dispersive pressure and 4 = matrix strength. The vertical axis depicts sediment
concentration, and the horizontal axis represents time and/or space. According to this diagram, flows can fol-
low a number of transport paths until sediments within them are deposited. (Modified fromWalker, 1978).
Reprinted with permission of American Association of Petroleum Geologists.
k

4-114
Sediment Gravity Flow Processes and Deposits
,

4 LOW-DENSITY
LIQUIFIED FLOWS [TURBIDITY CURRENTS] Te Te
Tt
Td S3
Td Tc
S2
Tc Tb
3

HIGH-DENSITY TURBIDITY CURRENTS


11 Tb 10 S3 =Ta 9 S1
Figure 4-4. Diagram illustrating the sediment grav-
ity flow continuum (vertical column on left). The
S1 continuum figures reflecting various deposits of the
continuum (113) move from bed 1 through bed 13.
TOC R3
12 13 8 Lines without arrows (e.g., 12 and 13) connect
members between which there probably exists a
continuous spectrum of flow and deposit types, but
Start R3 which are not part of an evolutionary trend of single
flows. Arrows connect members that may be parts
R3 of an evolutionary contnuum for individual flows.
7
The transition from disorganized cohesive flows (1
Ref. List R2
and 3), to thick, inversely graded, density-modified
6
grain flows and traction carpets (5) and to turbu-
GRAIN FLOWS

lent, gravelly, high-density turbidity currents (6) is


R2
Search speculative. (Modified from Lowe, 1982). Reprinted
4 5 with permission of Society of Sedimentary Geology
(SEPM).
COHESIVE FLOWS

2
Help
1 3

k
WF Facies F9b

Facies F4 Facies F2 Facies F4 Facies F5 Facies F7 Facies F8 Facies F9a


FT
FT
CDF FT HCF TC TC Bouma-a Bouma b-e
GHDTC SHDTC LDTC
Facies F3 Facies F6
Rip up mudstone clasts

Mud drape scours Tabular scours


Depression
Flow Direction
Gravelly high-density
Wavy-laminated facies composed TC Traction Carpets GHDTC
WF turbidity current
of poorly sorted gravel and sand
Sandy high-density
FT Flow transformation CDF Cohesive debris flows SHDTC
turbidity current
HCF Hyperconcentrate flow LDTC Low-density turbidity
Fluid escape structures
current

Figure 4-5. Diagram illustrating the deposits of the sediment gravity flow continuum. According to this classifi-
cation, flow processes change in the flow direction, so the type of sediment deposited also changes. Cohesive
debris flows (CDF) are precursors to other flow types, such as gravelly high dencity turbidity currents
(GHDTC), sandy high density turbidity currents (SHDTC) and low density turbidity currents (LDTC). Result-
ing deposits, such as debrites, massive sandstones (F7), and Bouma divisions are explained in the text. (Modi-
fied from Mutti et al., 1999). Reprinted with permission of American Association of Petroleum Geologists.
k

4-115
Sediment-Gravity Flows and Their Processes

4 WANING STEADY WAXING

3 ACCUMULATIVE

Yellow = Ta
Tan = Tb EROSION/
TOC Orange = Tc NON-DEPOSITION
Red = Td/e

UNIFORM
Start

Ref. List

DEPLETIVE
Search
direction of transport

Help Figure 4-6. Different types of turbidity current deposits formed under a variety of spatial (accumulative, uni-
form and depletive) and temporal (waning, steady and waxing) flow conditions. Bouma divisions are shown by
different colors. The horizontal thicknesses of individual deposits are indicators of thicknesses of the beds in
nature. Arrows point in the downcurrent direction. The nine flow types and their five deposits are explained
in the text (Modified from Kneller, 1995). Reprinted with permission of The Geological Society of London.

k
Bouma Walker Lowe Mutti Kneller
(Fig.14) (Fig.3) (Figs. 4/22) (Fig.5) (Fig.6)

Debris Flow Cohesive Flow F4(CDF)

[Slurry beds?]

Clast-supported R1-R3 F4/F7(?)


Conglomerate

Pebbly Sandstone S1-S2 F5/F7(?)

-Ta Massive Sandstone S3 F8 Steady


Waning
-Tb-e Classical Turbidites LDTC F9a Waxing

Figure 4-7. Chart comparing the common classifications of sediment gravity flow deposits shown in
Figures 4-34-6. Figures in which symbols and names of deposits are explained are provided beneath authors
names. The various deposits are described in the text.
k

4-116
Sediment Gravity Flow Processes and Deposits

Debris Flows and Debrites


4
Because of their high matrix strength, debris flows move in a plastic, laminar, cohesive
state. Their flow properties have been likened to that of wet concrete (Pratson et al., 2000). Debris
3 flows can be large or small, and can move for long distances down a slope and into a basin.
Debris flows are composed of a shear flow region and a plug flow region (Fig. 4-8). The
shear flow region is located in the lower part of the flow. Shear stresses in the bottom of the
TOC flow, generated by its movement, exceed the matrix shear strength and cause the flow to shear.
The shearing decreases upward through the flow. The plug flow region is located in the upper
part of the flow at the height at which the shear stress becomes less than the matrix shear
Start strength (Fig. 4-8). At this height, shearing stops and the flow moves as a plug with uniform
velocity. The uppermost portion of the flow may exhibit a decrease in velocity due to frictional
effects with overlying ambient sea water.
Ref. List Laboratory experiments by Mohrig et al. (1998) and Marr et al. (2001) have documented
the process of hydroplaning wherein the basal layer of a debris flow is lubricated by a
wedge-shaped layer of water forced beneath the flow during downslope movement (Fig. 4-8).
This water layer has the effect of deflecting upward the debris from the bed. The effect of basal
Search lubrication is that the head of a debris flow moves downslope at a higher velocity than the
body, attenuating and even detaching the head from the body. Other features associated with
laboratory-generated deposits of debris flows include structureless and ungraded grain size
Help distribution of the deposit, tension cracks, water-escape structures, compression ridges and
imbricate slices (Marr et al., 2001).
Whether or not debris flows have large erosive capabilities is debateable. Posamentier
(2003) have documented from shallow-subsurface 3D seismic data grooves with 40500m of
vertical relief that are associated with mass transport complexes 2530 km long, indicating
that erosion by mass movement can occur on the sea floor. On the other hand, Mohrig et al.s
(1998) laboratory experiments, as well as numerical modeling (Pratson et al., 2000), indicate
that debris flows are not capable of eroding the substrate over which they move. In these
experiments, the volume of sediment within the flow does not increase appreciably between
when the flow starts and when it ends. The dense, mud matrix of the flow also tends to inhibit
loss of sediment across its upper surface.

Figure 4-8. Schematic cross section illustrating


the internal flow structure of a debris flow.
Velocity of the shear flow region (Us) increases
upward from the base of the bed. Within the
hp
plug flow region, velocity (Up) remains con-
Up Plug F stant. The symbols hs and hp refer to thickness
z low
Region of each zone. Z is the vertical dimension or
thickness of the flow. Note that the front of the
hs flow is above the base of the bed. S is the angle
Us Shear of the bed relative to the horizontal plane. A
Flow
Region wedge of water lubricates this basal zone, giving
rise to the process of hydroplaning, explained
in the text. (Modified from Pratson et al., 2000).
S Reprinted with permission of Society of Sedi-
mentary Geology (SEPM) and American Asso-
x
ciation of Petroleum Geologists.
k

4-117
Sediment-Gravity Flows and Their Processes

If the constancy of sediment/water ratio throughout the length of a flow does occur in
4 the marine environment, as it does during laboratory experiments, then it provides one of the
most significant differences between debris flows and turbidity currents. However, the height
of the flow is inversely proportional to its velocity. If the flow velocity decreases, the flow will
3 thicken, and internal sediment concentration will decrease. If this velocity increases, the flow
will thin, and the internal sediment concentration will increase. The flow is driven forward
(downslope) by the weight of the flow and is retarded by friction acting on the seabed. Internal
TOC fluid pressure causes the flow to spread radially as it travels. The flow will continue downs-
lope to a lower gradient on the sea floor where the flow spreads radially and fluid pressure is
reduced below the frictional threshold for movement, so forward movement ceases.
Start Deposits of debris flows, often termed debrites, may be composed of mud, mixtures of
mud and sand, or mixtures of mud, sand, and gravel arranged in a disorganized or random
manner (Fig. 4-9). The high matrix strength is sufficient to hold gravel-size clasts within the
Ref. List flow and resulting deposit. Internally, debris flows may exhibit random orientation of clasts or
there may be some imbricate orientation of larger clasts. Because of the high matrix strength,
sedimentary structures resulting from fluidal movement are lacking. Slide and deformation
structures may be present.
Search

Help

Figure 4-9. Outcrop photograph of a poorly sorted debrite bed. Note the large boulders are sup-
ported within a dense, fine-grained matrix. California, U.S.A.
k

4-118
Sediment Gravity Flow Processes and Deposits

Intermediate Flows and Their Deposits


4
Lowe (1982) defined three types of intermediate flows on the basis of sediment support
mechanism: grain flows, liquified flows and fluidized flows (Fig. 4-1). Grain flows (Fig. 4-4)
3 are dispersions of particles maintained within a current solely by dispersive pressure arising
from grain-to-grain collisions (Fig.4-2). This process implies a relatively high concentration of
sediment within the flow. Lowe (1982) claims that grain flows can exist only on slopes
TOC approaching the angle of repose of subaqueous sand (1828 degrees). In deep water, grain
flows would form as thin beds of avalanche foresets on dune slipfaces (Fig. 4-4).
In liquified and fluidized flows (Figs. 4-1 and 4-4), pore fluids are forced upward during
Start sediment transport, as particles settle toward the base of the flow. Upward-rising fluid causes
the particles to remain in suspension. Deposits from liquified and fluidized flows may exhibit
fluid escape structures (e.g., dishes and vertical pipes) (Fig.4-10). Such structures form when
vertically-escaping water creates a cavity within the flow, causing internal, localized collapse.
Ref. List Sandstones that do not exhibit fluid escape structures are more likely to reflect grain-to-grain
collisions (dispersive pressure) during initial deposition, without forceful escape of fluids.
Pratson et al. (2000) and Marr et al. (2001) claim that with > 10% sediment concentra-
Search tion in a flow, grains will be constantly in mutual contact. Grain flows, liquified flows, and
fluidized flows probably fall within this category (>10% concentration), but are not considered
true debris flows in the sense of Lowe (1982).
Help

Figure 4-10. Photograph of dish structures in a sandstone bed. Stratigraphic top is toward the top of the
photo. Formation of dish structures is discussed in the text. The formation which contains the dish structures
is unknown. California, U.S.A.
k

4-119
Sediment-Gravity Flows and Their Processes

Subsequent to Lowes (1982) classification of flow types, Lowe and Guy (2000) recog-
4 nized an unusual deposit that they term a slurry bed, interpreted to form from a watery flow
transitional between a turbidity current and a debris flow. Such beds contain 1035% detrital
mud matrix, are enriched in water-escape structures, and are grain-supported. Lowe and Guy
3 (2000) interpret the slurry beds to have originated as low density turbidity currents containing
an abundance of flocculated or ripped-up, sand-size, cohesive mud particles that behave in a
hydrodynamically similar manner to more rigid silt- and sand-size quartz and feldspar grains.
TOC As the flow wanes, the mud particles settle toward the base of the flow. There, the grains
abrade against more rigid quartz and/or feldspar grains and disaggregate into their component
silt- and clay-size particles, giving rise to a mud-rich basal flow. The increased mud content
Start near the base of the flow increases viscosity and cohesion and suppresses near-bed turbulence,
thus creating a quasi-laminar basal flow, and the resulting slurry bed. Fluid escape structures
are common in slurry beds, owing to the upward flow of water as the density and viscosity in
Ref. List the basal layer increase during particle breakup. Slurry beds have been recognized by Lowe
and Guy (2000) in the Lower Cretaceous Britannia Formation of the North Sea, the Pennsylva-
nian Jackfork Group of Arkansas, and the Cretaceous Great Valley Group of California.
Search
Turbidity Currents and Turbidites

Help Processes

A turbidity current is a sediment gravity flow with fluidal (i.e., Newtonian) rheology in
which sediment is held in suspension by fluid turbulence (Figs. 4-11, 4-12). A turbidity current
contains a head, body, and tail. The head may erode the sea floor and entrain sediment back
into the body (Kneller and Buckee, 2000). Owing to viscosity differences between the turbid-
ity current and overlying ambient sea water, a series of billows, called Kelvin-Helmholtz
instabilities (Allen, 1985), form at the frictional interface of the fluids (Figs. 4-11, 4-12).
Laboratory experiments have indicated that once a turbidity current is initiated, its inter-
nal sediment distribution and velocity structure are quite complex (Kneller and Buckee, 2000).
A vertical velocity profile through an experimental turbidity current is shown in Figure 4-13,
along with a variety of experimental sediment concentration profiles. The velocity of the flow
is low near the base owing to frictional forces with the sea bed. Flow velocity reaches a maxi-
mum at some distance above the bed where the flow is fully turbulent, then decreases upward.
Sediment concentration profiles vary according to the absolute sediment concentration and the
near-bed processes of erosion and/or transport. Experiments measuring vertical grain size dis-
tributions show that fine-grained particles are distributed uniformly vertically through the
flow, whereas coarser-sized grains diminish in abundance upward from the base of the flow
(Garcia, 1994) (Fig. 4-13). Thus, within a flow containing a range of coarse- to fine-grained
particles, the base of the flow may be more poorly sorted than the overlying parts. Conversely,
well-developed slurry flows should exhibit greater mud concentrations near their base (Lowe
and Guy, 2000).
Numerical modeling by Pratson et al. (2000) suggests that turbidity currents move
downslope by the combined influence of the weight of the current (controlled by sediment
concentration) and internal fluid pressures that are counterbalanced by frictional forces
between the seabed and the current. The body may have velocities that are greater than those
of the head, so that with distance, the head may expand in thickness (Fig. 4-12) as the body
attempts to outrun the head (Talling et al., 2001).

4-120
Sediment Gravity Flow Processes and Deposits

4
AMBIENT WATER LOWER-DENSITY HIGHER-DENSITY
TURBULENT SUSPENSION TURBULENT SUSPENSION
Turbulent suspension clouds
Velocity of "thrusted" backwards
3 the head is
108 cm/s

TOC
LAMINAR
Pebble "tracers" INERTIA-FLOW
10cm

Suction due to Velocity profile limited in this interval Velocity profile


Start o strong pressure at 1-second distance (data in FIG. 3) at 2-seconds distance
Slope = 25 gradient in the from the flow head from the flow head
head of the flow

Ref. List Figure 4-11. Schematic cross section of a high-density turbidity current. Arrows point to the flow directions
within the current. Vertical flow velocity profiles are shown in shaded black. The flow is size-graded, with
coarsest grains at the base of the bed. The head of this flow is thicker than the body. The billows at the top of
Search the flow are termed Kelvin-Helmolz Billows. (Modified from Postma et al., 1988). Reprinted with permis-
sion of Elsevier Publishing Co.

k
Help

Figure 4-12. Photograph of a cross section of a turbidity current generated in a flume tank. The head and body
of the flow are clearly shown. The sediment box from which the sediment was dropped through a removable
bottom is in the upper right of the picture. Note that the head is thicker then the body, even after only a short
distance of transport. Photograph is from the flume tank of the St. Anthony Falls Laboratory, University of
Minnesota.
k

4-121
Sediment-Gravity Flows and Their Processes

4 Normalized height a b c d
Density/concentration Fine sediment
3 Coarse sediment
Velocity

TOC
Normalized velocity and density/concentration

Start Figure 4-13. Four graphs illustrating velocity (solid lines) and associated sediment concentration (dashed
lines) profiles for a variety of experimental turbidity currents: (a) two layer concentration model with a con-
stant concentration lower interval and an upper region of sediment detrained from the head of the flow; (b) a
smooth concentration profile, characteristic of low concentration, weakly depositional flows; (c) a stepped
Ref. List concentration profile observed in erosional flows; (d) a distribution observed in turbidity currents in which
coarse material is concentrated towards the lower part of the flow and the fine-grained material is evenly dis-
tributed thoughout the flow (Modified from Kneller and Buckee, 2000). Reprinted with permission of Interna-
Search tional Association of Sedimentologists.

k
Help Turbidity currents are capable of eroding the substrate. Thus, sediment can be continu-
ally entrained into the head of the flow, even as sediment is deposited from within the body
and tail. Also, Kelvin-Helmoltz billows may transport sediment from the head of the flow back
into the body. Ambient water can also be entrained within the flow, acting to dilute the sedi-
ment concentration while, at the same time, increasing the overall thickness of the flow.
Thus, the actual sediment concentration can decrease and increase in a non-systematic
fashion along the length of a flow, as well as vertically at any one position within the flow. If the
amount of new sediment entrained by erosion is less than the amount lost through deposition,
then the turbidity current eventually ceases. But, if the amount of new sediment entrained is
greater than the amount lost through deposition, the flow gains momentum, and further erodes
as it moves downslope. However, if sediment concentration becomes too high, turbulent flow
can be suppressed and a different sediment support mechanism becomes operative (Fig. 4-2).
Even with this complex flow behavior, or perhaps because of it, a turbidity current can
travel for long distances at high velocities. Single-event turbidity currents have been docu-
mented, which have transported sediment several hundred kilometers from their source. The best
documented of these, the 1929 Grand Banks of Newfoundland turbidity current, is calculated to
have traveled at the following velocities: approximately 20 m/sec at a distance of 300 km from
the earthquake epicenter, 14 m/sec at a distance of 500 km from the epicenter, and 11 m/sec at a
distance of 600 km from the epicenter. This single flow traveled over 600km in 13 hours (Uchupi
and Austin, 1979) at a velocity sufficient enough to transport particles up to 3 cm in diameter in
suspension. In a different flow, inferred flow velocities of the Nice airport turbidity current
reached 30 m/sec., forming deep submarine scours and transport of cobbles and boulders as bed
load, and coarse sand as suspended load (Piper and Savoye, 1993). On the Amazon Fan, individ-
ual flows are estimated over periods of several hours to days, at velocities of 13 m/sec (Pirmez
et al., 2000). The Zaire submarine valley flow was documented at an average velocity (integrated
over 1 hour) of 1.2m/sec (with higher instantaneous velocities) 150m above the sea floor, and
coarse sand and plant debris were collected 40m above the floor (Khripounoff et al., 2003).
The areal extent and volume of single-event turbidity current deposits can also be quite
large. For example, the Holocene Black Shell turbidite on the Hatteras abyssal plain covers an
area of approximately 25,000 sq. km. and comprises a minimum volume of 100 cu. km
(Elmore et al., 1979). 4-122
Sediment Gravity Flow Processes and Deposits

The Bouma Sequence


4
The Bouma Sequence (Bouma, 1962) has long been considered to be the fundamental
sand/mud deposit from a turbidity current (Fig. 4-14). Bouma (1962) defined this sequence as
3 the product of continuous deposition from a turbidity current. A complete Bouma sequence
consists of a grain-size fining-upward succession of (a) massive or normally size-graded, sandy
Bouma Ta division; (b) parallel laminated, sandy Bouma Tb division; (c) ripple/climbing-ripple
laminated/convoluted, sandy Bouma Tc division; (d) parallel laminated to massive, silty Bouma
TOC
Td division; and (e) silt-clay, often microfaunal-rich Bouma Te division (Fig. 4-14). Some out-
crop examples of Bouma divisions are provided in Figures 4-15, 4-16 and 4-17.

Start

Te SUSPENSION
Ref. List
Td MIXED
Figure 4-14. Sediment grain size, struc-
Tc tures, divisions of a complete Bouma
Search sequence, and sediment transport
TRACTION mechanisms. The different divisions
from Bouma Ta to Bouma Teare
Tb explained in the text. (Modified from
Help
Jordan et al., 1993). Reprinted with
permission of American Association of
Petroleum Geologists.
Ta SUSPENSION

k
The upward decrease in grain size and change in sedimentary structures are a result of
gradually waning flow velocities, ultimately leading to deposition of progressively finer-
grained sediment under progressively lower flow regime conditions. The Bouma Ta division is
thought to be deposited rapidly, directly from suspension. The Bouma Tb and Tc divisions are
the product of traction of grains along the sea bed under upper (Tb) and lower (Tc) flow
regime conditions. The Bouma Td division is deposited by suspension from the tail of a turbid-
ity current. The Bouma Te division probably is a mixture of fine-grained sediment from both
the tail of the current and slow settling of pelagic grains. Mixtures of shallow- and deep-
marine microfauna in the Te division are indicative of a turbidity current process, as is size-
grading of silt- and clay-sized particles. Turbidite mudstones and shales also exhibit a charac-
teristic suite of waning-flow sedimentary structures and textures (Fig. 4-18).
Laboratory experiments in which relatively dilute concentrations of sand were mixed
with varying amounts of clay and water simulated development of some of the common fea-
tures of Bouma divisions and provided insight into when these features form (Marr et al.,
2001). For example, grain-size grading was found to occur only during flow deceleration,
rather than during downslope movement of the flow (Marr et al., 2001). Water escape struc-
tures also formed soon after the flow came to rest, and continued for several minutes
afterward. Hydroplaning was not observed to occur in the dilute flows.

4-123
Sediment-Gravity Flows and Their Processes

TOC

Start
Figure 4-15. Photograph of beds illustrating
Bouma Ta, Tb, and Tc divisons in outcrop. The
Ref. List upward gradational decrease in grain size is
noted by a smoothing of the outcrop surface.
Location of the outcrop is unknown.
Search

Help

beds illustrating Bouma k


Figure 4-16. Photograph of

Ta-Te divisions in outcrop.


Note the normal size grad-
ing as evidenced by the
decrease in size and abun-
dance of coarse (white)
particles upward through
the Ta division. Name of
the formation which con-
tains this rock is unknown.
Newfoundland, Canada.
k

4-124
Sediment Gravity Flow Processes and Deposits

TOC

Start

Ref. List

Search

Help
Figure 4-17. Photograph of Bouma Tb-Tc division in outcrop. The upward change from parallel laminated to
ripple (including climbing-ripple) bedding is a result of decrease in traction current flow from upper to lower
flow regime velocities. Upper Miocene Mt. Messenger Formation, New Zealand.

k
70 T8 MICRO-BIOTURBATED MUD
Figure 4-18. Succession of sedimen-
UNGRADED MUD tary textures and structures in fine-
T7 +silt pseudonodules
60 grained turbidites Grain size and
sedimentary structures change sys-
T6 GRADED MUD
50 +silt lenses tematically upward much as they do
E in sandy turbidity current deposits.
T5 WISPY.CONVOLUTE LAMINAE In this diagram, the Bouma Tc and
GRADED LAMINATED MUD

40
T4 INDISTINCT LAMINAE Td divisions are the same as
described in Figure 4-14. However,
30 T3 THIN REGULAR LAMINAE the Bouma Td division is divided
into T1 and T2 subdivisions, and the
THIN IRREGULAR LAMINAE
20 T2 low-amplitude climbing ripples Bouma Te division is divided into
six subdivisions (T3-T8) based upon
D
T1 CONVOLUTE LAMINAE variations in grain size and small-
10
BASAL (LENTICULAR) LAMINA
scale sedimentary structures. (Mod-
BOUMA (1962)

T0 Fading ripples, micro-cross and parallel C ified from Stow and Shanmugam,
lamination, sharp scoured load-cast base
1980). Reprinted with permission of
SCALE
mm B Elsevier Publishing Co.
A
k

4-125
Sediment-Gravity Flows and Their Processes

Turbidity Current Classification Based Upon States of Flow


4
The implication of the Bouma Sequence is that turbidity currents will diminish in veloc-
ity over time and in the downcurrent direction, giving rise to deposition of progressively finer-
3 grained Bouma divisions, both vertically and laterally. Although this must be true where Bouma
divisions are present in their normal strataigraphic position (Fig. 4-14), Kneller and Branney
(1995) have challenged the general assumption of progressively waning flow on the grounds
that there is no real reason to believe that all turbidity currents in the deep ocean behave in this
TOC
manner. Kneller (1995) proposed that flows may wax, wane, or remain constant both over time
and distance (Fig. 4-19). Over time, at any one place on the sea floor, a flow can wax (increase
in velocity), wane (decrease in velocity), or remain steady (constant velocity). Over distance
Start along the sea floor, a flow can become accumulative (increase in velocity), depletive (decrease
in velocity), or remain uniform. Based upon these temporal and spatial variations in flow veloc-
ity, Kneller (1995) classified flow types into nine possible combinations (Fig. 4-6). Sediment
Ref. List concentration does not directly enter into this classification scheme. Thus, different Bouma
divisions can occur both vertically within a stratigraphic sequence or spatially along a deposi-
tional profile depending upon the temporal and spatial variations in flow velocity.
Search The topographic relief and gradient of the seafloor are thought to play major roles in
modifying the velocity of a flow as it travels down the depositional profile. Accumulative
flows (Figs. 4-6, 4-19, 4-20) might form with a downcurrent increase in slope gradient or
downcurrent convergence a of flow through a restriction on the sea floor. Uniform flows
Help
(Figs. 4- 6, 4-19, 4-20) might form over a floor with a progressively slight decrease in down-
current gradient. Depletive flows (Figs. 4-6, 4-19, 4-20) might form with a downcurrent
decrease in slope gradient or downcurrent divergence of flow as the flow moves beyond a con-
striction on the sea floor and becomes unconfined.
Also, when a sediment gravity flow meets a seafloor obstacle, the flow can either par-
tially or completely override the obstacle, be deflected around the side of the obstacle, be
confined or ponded by the obstacle, or reverse flow direction and flow back down the obstacle
(Fig. 4-21) (Kneller and Buckee, 2000). Which of these processes dominates is a function of
the velocity and density of the current, the flow stratification within the current, and the
dimensions of the obstacle.

waxing
waning

u
u=velocity Figure 4-19. Graphs of time (t) vs. velocity (u) and dis-
tance (x) vs. velocity (u), showing different types of
UNSTEADY STEADY flow under a variety of conditions. In the upper dia-
t = time
gram, flow velocity first waxes, then wanes, then
t
x= distance becomes steady over time. In the lower diagram, flow
accumulative velocity first increases (accumulative), then decreases
(depletive), then becomes uniform with downcurrent
depletive
distance. (Modified from Kneller, 1995). Reprinted
with permission of The Geological Society of London.
u

NON-UNIFORM UNIFORM

x
k

4-126
Sediment Gravity Flow Processes and Deposits

4 divergent flow decrease in slope

TOC

DEPLETIVE FLOW
Start
convergent flow increase in slope
Ref. List

Search

Help ACCUMLATIVE FLOW


Figure 4-20. Diagram illustrating some causes of spatially depletive (downcurrent decrease in flow velocity) and
spatially accumulative (downcurrent increase in flow velocity) sediment gravity flows. The upper diagram, (a) is
a plan view of depletive flow resulting from divergent flow on the sea floor as the flow becomes unconfined, and
(b) is a cross section view showing downcurrent decrease in flow velocity due to reduction in slope gradient. In
the lower diagram, (c) is a plan view illustrating accumulative flow resulting from convergence of flow on the sea
floor , and (d) illustrates accumulative flow resulting from a downcurrent increase in the slope gradient of the
sea floor. (Modified from Kneller, 1995). Reprinted with permission of The Geological Society of London.

k
flow path

Figure 4-21. Schematic diagram of possible flow paths (arrows) for sediment gravity flows when they encoun-
ter a sea floor obstacle. Flows can partially or completely override the obstacle, can divert around the obstacle
or flow partway up the obstacle then reverse flow. (Modified from Kneller and McCaffrey, 1999). Reprinted
with permission of Society of Sedimentary Geology (SEPM).
k

4-127
Sediment-Gravity Flows and Their Processes

Low- and High-Density Turbidity Current Deposits


4
Lowes (1982) classification of low- and high- density turbidity currents, differenti-
ated on the basis of concentration of sediments in the flow, is also widely cited (Fig. 4-4). He
3 applies the designation S to sediment deposited from high-density turbidity currents (Fig. 4-22).
Lowe (1982) further subdivided S sediments (Fig. 4-22). His S1 division is deposited
from traction currents, thus exhibiting traction structures, mainly planar laminations and cross
TOC stratification. The S2 division contains stacked, thin, inversely-graded beds deposited from bed
load. Grain collisions predominate during rapid sedimentation, thus suppressing turbulence.
The S3 division is deposited during high sediment fallout rates. Deposits are massive to nor-
Start mally graded, and may exhibit fluid escape structures.
Low-density turbidity currents contain dilute concentrations of clay, silt, and fine- to
medium-grained, sand-size particles (Lowe, 1982) (Fig. 4-4). Low-density turbidity currents
Ref. List
have been defined as containing 123% sediment by volume and high-density turbidity cur-
rents have been defined as containing 644% sediment by volume (Shanmugam, 1996). The
overlap in concentrations according to these definitions points to the present lack of clearly-
Search defined criteria to define these types of flows on the basis of sediment concentration.

Help Tt

S3

S3
Sand and Fine Gravel

S3 Suspension S3

S3

S2
S2 Traction Carpet

S1 Traction S3

R3 Suspension
S2
Gravel

R2 Traction Carpet
S1

Figure 4-22. Vertical profile of sediment grain size and sedimentary structures illustrating high-
to low-density turbidity current deposits using the terminology of Lowe (1982). S and R designa-
tions, and processes, are explained in the text and in Figure 4-4. Reprinted with permission of
Society of Sedimentary Geology (SEPM).
k

4-128
Sediment Gravity Flow Processes and Deposits

Massive (Structureless) Sandstones


4
Because the origin of massive (structureless) sandstones has generated considerable debate
in the literature, they are discussed separately in this chapter. Massive sandstones are the most
3 common type of sediment gravity flow sandstone observed in outcrops and cores (Fig. 4-23)
(Kneller,1995). Both Walker (1978) and Lowe (1982) differentiated massive sandstones from
graded Bouma Ta sandstones. Walker (1978) differentiated massive sandstone from the Bouma
Ta division (Fig. 4-3) by (a) the common presence of fluid escape structures, (b) fewer associated
TOC shale interbeds, (c) an increase in erosionally-based and irregularly-bedded sandstone, (d) coarser
grain size relative to associated sandstones, and (e) sandstone beds that are thicker than associ-
ated beds. Lowe (1982) referred to massive (S3) sandstones as fluidized or liquefied flow
Start deposits, depending upon the presence or absence of fluid escape structures (Fig. 4-4).
Shanmugam (1996,1997, 2000) has argued that massive sandstones are not turbidity
current deposits, but are the product of deposition from plastic or laminar flows. He uses the
Ref. List term sandy debrite for a massive sandstone, and claims that a mud matrix as low as 1% is
sufficient to provide cohesive strength to a flow. As mentioned previously, experimental work
by Marr et al. (2001) support this interpretation if the clay matrix is composed of bentonite and
the water content is 2540%. More clay is required if the mineral is not bentonite. Marr et al
Search
(2001) stress that their results apply only to laboratory-scale, and not field-scale flows.

Help

Figure 4-23. Core of unconsolidated sand (light


color) and lithified shale (darker color) from Long
Beach Unit, Wilmington oil field, California (Slatt et
al., 1993). Individual sand beds are massive, struc-
tureless, and of uniform grain-size from base to top.
Scale is in 0.1 ft. increments. Reprinted with permis-
sion of the Society of Sedimentary Geology (SEPM)
k

4-129
Sediment-Gravity Flows and Their Processes

Shanmugams argument is at least partially based on semantics. He claims that a Bouma


4 Ta bed must be size-graded to qualify as a turbidity current deposit, even though Boumas (1962)
definition alludes to the fact that some Ta beds are size-graded and others are not. But, if there is
not a range of particle sizes in the original flowsuch as might occur from a pre-sorted sand
3 then size-graded beds cannot be deposited from the turbulent flow. Shanmugam (2000) further
claims that the presence of shale clasts within a sandstone argues for cohesive, rather than turbu-
lent flow, yet he claims that debris flows do not have the capability to erode the sea floor.
TOC Because many shale clasts found within deep water sequences appear similar to underlying shale
beds, erosion of the muddy substrate by the flow must have occurred to generate the clasts.
Knellers (1995) classification shows that massive sand of uniform grain size will be depos-
Start ited under steady state flow (over time) in combination with depletive flow (over distance)
(Fig. 4-6). Even though grain size and bed thickness decrease in the downcurrent direction, the
deposit remains massive and of uniform grain size at any one depositional site on the sea floor.
Ref. List
It is reasonable to expect relatively steady flow of a turbidity current over time on the sea
floor. For example, steady flow has been suggested for a period of at least 2 hours for the
Grand Banks turbidity current (Piper et al. 1988). This should be sufficient time to deposit a
Search massive sand on the sea floor. The common occurrence of ungraded beds in the rock record
suggests steady state flow is common.

Help
Hyperpycnal Flows and Hyperpycnites

Hyperpycnal flows have been a topic of considerable discussion and interest at confer-
ences during the past few years. According to Mulder et al., (2003), the importance of
hyperpycnal floods as a sediment transport process in deep water has been underestimated for
many years. The following discussion summarizes current knowledge about hyperpycnal
flows and their deposits.
Mulder et al. (2003) differentiate turbidity currents that are generated by ignitive trans-
formation of a submarine slide into a turbulent flow and those that are generated by non-
ignitive conditions, such as from continuity of hyperpycnal discharge of a river during flood
stage (Fig. 4-24) Because the fluid in such flows is fresh water, the density contrast between
fresh water and sea water is such that a very high suspended sediment concentration is
required for the flow to sink or plunge to the sea floor (Fig. 4-24). Mulder et al. (2003) place
the critical sediment concentration at 3643kg sediment/m3 fluid. Variations within this range
are due to the variations in temperature and salinity of sea water at the river mouth.
Sediment concentrations measured from two sediment traps on the Zaire submarine fan
valley (one in the channel and one on the levee 18km away), were 3.28kg sediment/m2 of trap
area/day above the channel and a peak of 11kg sediment/m2/day at the levee site (Khripounoff et
al., 2003).. Sediment in the channel trap consisted of silt and large plant remains with finer
grained siliciclastic particles and 464 mg organic carbon/m2/day found in the levee sediment trap.
Hyperpycnal flows can be relatively long lived. During a major flood that lasted for
three days in 1994, the Var River in France generated a hyperpycnal turbidity current that
lasted for 18 hours (Mulder et al., 2003). The Var River transported 1114 times the rivers
normal particle load, which ultimately was deposited in the deep ocean. The total duration of
the sediment gravity flow in the Zaire submare valley is estimated to have been 10 days, with a
3 day delay between the channel and the levee station 18km away. Flow thickness exceeded
the 150m depth of the channel floor beneath its levee crest, allowing sediment to overflow
(Khripounoff et al., 2003).
4-130
Sediment Gravity Flow Processes and Deposits

4 Figure 4-24. Schematic cross section of main


types of flows debouching from a river mouth.
Hyperpycnal flows are denser than ambient
Interflow Hypopycnal Lofting water, and flow along the sea bed. Hypopycnal

3 flows are less dense than ambient water, so ride


on the sea water surface. Interflows are at some
intermediate density. and flow within the ambi-
ent water. Lofting occurs when the density of a
hyperpycnal flow decreases due to loss of sedi-
TOC ment within the flow by deposition, and the flow
Low-density hyperpycnal flow
rises into the ambient sea water. (Modified from
Mutti et al., 2003). Reprinted with permission of
Elsevier Publishing Co.
Start

k
Ref. List After a hyperpycnal flow debouches from the river mouth into marine water, it plunges to
the sea floor and moves downslope. If sediment is lost from the flow by deposition, the density
of the remaining flow can decrease to the extent that the flow detaches from the sea floor and
Search rides within the ambient water column at an appropriate depth (Fig. 4-24). This fact, due to the
initial fresh water nature of the internal fluid, provides a fundamental difference between normal
marine turbidity currents and hyperpycnal turbidity currents. Mulder et al. (2003) state that
hyperpycnal flows with relatively low sediment concentrations and density are the low-density
Help turbidity currents of Lowe (1982). In contrast, turbidity currents generated by ignitive processes,
in which the internal fluid is sea water, are the high-density turbidity currents of Lowe (1982).
The typical hyperpycnite deposit consists, from the base upward, of a coarsening-upward
successsion overlain by a fining-upward succession of beds (Fig. 4-25). The lower succession is
deposited during rising flood stage when the flow is waxing and river discharge is high at the
river mouth (Figs. 4-6, 4-19, 4-25). After the period of peak discharge, the flow wanes
(Figs. 4-6, 4-19, 4-25) and the upper succession is deposited. The stratigraphic continuity of a
single hyperpycnite bed can be disrupted by an erosional or bypass surface if the waxing-stage
flow reaches a threshold velocity capable of eroding the underlying rising-stage deposit
(Fig. 4-25). Grain size of a typical hyperpycnite increases vertically from silt to fine sand, and
then back to silt. Climbing ripples are a common sedimentary structure. Land-derived organic
material also is present in hyperpycnites. A hyperpycnite bed is shown in Figure 4-26.

progressive Figure 4-25. Graph illustrating flows with differ-


deposition of
Discharge type 4 bed ent discharge through time, and the representa-
tive lithofacies. 1. Low magnitude flood
generates a normally-graded bed. 2. Low magni-
3 tude flood which generates a complete inverse-
2 4 to-normally graded hyperpycnite due to waxing,
1 Critical discharge for the formation then waning flood stages. 3. Mid-magnitude
of a hyperpycnal turbidity current
flood generates a complete sequence which is
coarser grained than the example in 2. 4. High-
magnitude flood generates an inversely-graded
Time bed from waxing flow, followed by an erosion or
bypass surface due to peak flood flow velocity
LEGEND
Erosional contact sufficient to erode the substrate, followed by the
Sharp contact
Horizontal lamination
waning flow, normally-graded bed. (Modified
Climbing ripples from Mulder et al., 2001b). Reprinted with per-
1 2 3 4 Ripple cross lamination mission of Springer-Verlag Publishing Co.
Cl fs ms cs fsa msa
k

4-131
Sediment-Gravity Flows and Their Processes

3 Figure 4-26. Ouctrop photo of


a hyperpycnite of the type
shown in 4 of Figure 4-25. An
TOC erosional or bypass surface
caps an inversely-graded bed,
and is overlain by a normally-
graded bed. Marnoso-arena-
Start cea Formation, Apennine
Mountains, Italy.

Ref. List

k
Search
Whether the graded beds deposited by these flows should be termed hyperpycnites, hyper-
Help pycnal turbidites, or turbidites is subject to debate (Mutti et al, 2002). Mutti et al (2003) suggest
that the stratigraphic succession of beds deposited from hyperpycnal flows should be termed
mixed depositional systems to emphasize their transitional character between truly basinal,
ignitive turbidites and delta-fed, turbidite-like deposits derived from rivers in flood stage.
Modern hyperpycnites have been documented for a distance of 700km downcurrent
from linked river- canyon-fan systems in the central Japan Sea in water depths to 3350m
(Nakajima, 2006). Two 4m long cores contain alternating turbiditic silt and hemipelagic mud
beds. Some of the silt beds exhibit a typical fining-upward grain size trend. However, other
beds on the order of 5-10cm thick exhibit a distinct coarsening-upward trend with mean grain
size in the range 25-40 microns, capped by an erosional surface, then overlain by 5-10cm of
beds which decrease in grain size upward in the range 5-25microns. It is postulated that the
duration of these flows was on the order of several days to 3-4 weeks in order to travel 750km,
and that they could have maintained the density required for them to hug the sea floor by
entraining sea water and eroding sediment into the flow. Estimated velocities were on the
order of 0.3m/sec, and they have an estimated frequency of occurrence of 70 years.
Siltstones with similar characteristics have been documented for the Cretaceous Dad
Sandstone member of the Lewis Shale leveed-channel deposits (Chapter 6 and Chapter 7)
(Soyinka and Slatt, 2004; Soyinka, 2005). Laser grain size analyses of individual laminae and
thin beds reveal systematic changes in mean grain size within the silt size range (Fig. 4-27A).
In some intervals the change from coarsening- to fining-upward trend is separated by a subtle
erosional surface. This same surface has been documented in the modern hyperpycnites from
the Japan Sea, mentioned above (Fig. 4-27B). A major delta system is known to have fed the
Dad Sandstone slope and basinal facies (Pyles and Slatt, 2007), so the likelihood of hyperpyc-
nal processes and deposits is high.

4-132
Sediment Gravity Flow Processes and Deposits

4 B
0 10
Grain size (um)
20 30 40 0
hemipelagite
A 185
3
Grain size (um)

090
050
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70
TOC 0
Mode
Hb Fall

10
SC Peak
190 EC

Depth (cm)
HL Ha Rise
20 SBC
Start
30

Hemipelagites
40

Ref. List HL 50 GC
Hb Fall Peak
195
Ha Rise

60 SBC
Depth (mm) Coarsening upward
Fining upward
hemipelagite
Search

Figure 4-27. (A) Polished slab of siltstone and mudstone outcrop interval analyzed for grain size
Help distribution by laser grain size analysis (Soyinka, 2005); coin for scale. The graph to the right of
the slab is a plot of the modal grain size of individual samples analyzed over the length of the slab.
Ha = basal unit; Hb = top unit; HL = sandy horizontal laminae; GC = gradational contact; SC =
sharp contact; SBC = sharp basal contact. Coarsening-upward and fining-upward trends in
modal size are noted. An erosional surface is shown in the thin-section photomicrograph.
(B) Average grain size of samples over a 10+cm interval of siltstones and mudstones from the
floor of the Japan Sea; these beds and laminae are interpreted as hyperpycnites (Nakajima,
2006). EC = erosional surface denoting the vertical change in average grain size from coarsening-
upward to fining-upward. For comparative purposes, the vertical scales are the same in A and B.

k
Gravel Deposits from Turbidity Currents

Gravelly deposits of turbidity currents are not as common as the sandier and muddier
types described above. One reason is the general lack of pebbles and coarser grains within
shallower water areas from which sediment gravity flows originate. Suitable source areas usu-
ally are confined to tectonically active basins with narrow shelves, and relatively high
sedimentation rates (Reading and Richards, 1994). Although it might be surmised that such
large grains would only be found in proximal deep water settings, gravel has been cored in the
youngest channel of the Mississippi Fan a distance of approximately 220 km from the present
shelf edge (Stelting et al., 1985).
Walker (1978) defined a series of gravelly, sediment gravity flow deposits on the basis
of the abundance of pebbles and coarser grains and their sedimentary structures and stratifica-
tion style (Figs. 4-3, 4-28). One type, pebbly sandstone (Fig. 4-3), contains dispersed or
concentrated pebbles within a sandstone matrix. In outcrop, alternating pebble-rich and peb-
ble-poor beds (Fig. 4-29) and trough or planar-tabular cross beds are the most characteristic
internal sedimentary structures. Normal size grading (Fig. 4-30), large sole marks, lenticular-
ity, and scoured bases of beds (Fig. 4-31) are also common features.

4-133
Sediment-Gravity Flows and Their Processes

4 GRADED-
INVERSE-TO-
NORMALLY DISORGANIZED-
STRATIFIED GRADED-BED GRADED BED

TOC
NO INVERSE NO INVERSE NO STRAT., NO GRADING,
GRADING, GRADING, IMBRICATED NO INVERSE
Start STRAT., NO STRAT., GRADING,
CROSS-STRAT., IMBRICATED NO STRAT.,
IMBRICATED IMBRIC. RARE
Ref. List
THESE THREE MODELS SHOWN IN SUGGESTED
RELATIVE POSITIONS DOWNCURRENT

Search Figure 4-28. Diagram illustrating the sedimentary features of the spectrum of clast-supported conglomerates
with suggested downcurrent positions. (Modified from Walker, 1978). Reprinted with permission of American
Association of Petroleum Geologists.
Help

Figure 4-29. Photograph of a stratified, pebbly sandstone to sandy conglomerate. Lower Pennsylvanian Jack-
fork Group, Arkansas, U.S.A.
k

4-134
Sediment Gravity Flow Processes and Deposits

TOC

Start

Ref. List

Search

Help
Figure 4-30. Photograph of a graded, pebbly sandstone in outcrop. Top is toward the upper left corner. Lower
Pennsylvanian Jackfork Group, Arkansas, U.S.A.

Figure 4-31. Photograph of a series of amalgamated sandstone beds. The middle bed has been eroded, and the
resulting scour (arrow) has been filled by a normally-graded, pebbly sandstone. Top is toward the upper right
corner. Lower Pennsylvanian Jackfork Group, Arkansas, U.S.A.
k

4-135
Sediment-Gravity Flows and Their Processes

Walker (1978) classified other types of clast-supported conglomerates on the basis of


4 type of grading (normal or inverse), presence or absence of stratification, and presence or
absence of imbrication (Fig. 4-32). Based upon theoretical considerations, Walker defined
gravelly deposits ranging from (upcurrent): (a) inverse to normally graded, imbricated con-
3 glomerates; to (b) normally graded, non-imbricated conglomerates; to (c) graded-stratified,
imbricated conglomerates (downcurrent) (Fig. 4-28). Walker (1978) also recognized a fourth
class, composed of disorganized, non-stratified, and non-imbricated conglomerates. All of
TOC these conglomerates tend to be lenticular, with scoured bases.
According to Lowe (1982), most coarse gravel is probably transported near the bed
within a highly concentrated traction carpet, and in suspension in the lower part of a turbulent
flow.Intergranular dispersive pressure maintains coarser grains in suspension, whereas finer
Start grains filter to the sea bed, thus forming traction carpets near the bed. The coarser grains even-
tually fall from suspension to the bed, resulting in an inversely graded bed (Fig. 4-33).
However, Legros (2002) has suggested that size segregation by the kinetic sieving process and/
Ref. List or by progressive aggradation of increasingly coarse-grained particles on the sea bed are more
likely causes of inverse grading than is the maintainence of coarse grains within the flow by
dispersive pressure.Lowe (1982) applies the designations R1 for coarse gravel with traction
Search structures, R2 for inversely- graded gravel layers, and R3 for normally graded gravel layers
(Fig. 4-22).

Help

Figure 4-32. Photograph showing imbrication in a


shale-clast conglomerate stacked between two tan
sandstone beds. Rock hammer is oriented approxi-
mately parallel to the orientation of the brown shale
clasts. Upper Cretaceous Dad Sandstone Member,
Lewis Shale, Wyoming, U.S.A.

k
Figure 4-33. Diagram illustrating the sedimentary processes
of formation of inversely graded, traction carpet deposit. (A)
Basal part of a high density flow shows development of a
lower, inversely graded zone due to intergranular dispersive
pressure. (B) Fallout of grains from suspension increases the
clast concentration in the basal layer and results in formation
A B
of a traction carpet in which grains are supported by disper-
sive pressure. (C) Continued fallout from suspension
increases the density of grains in the traction carpet and
causes freezing in the upper part of the carpet. D) Final freez-
ing of traction carpet results in formation of a new inversely
graded basal layer above the deposit (Modified from Lowe,
1982). Reprinted with permission of Society of Sedimentary
Geology (SEPM).
C D
k

4-136
Sediment Gravity Flow Processes and Deposits

Flow Combinations and Transformations


4
The various classification schemes discussed above all imply that flows can vary tempo-
rally and spatially in a predictable, downcurrent continuum, giving rise to different sediment
3 gravity flow deposits along a single bed (Figs. 4-34-6). However, direct evidence of flow
transformations within a single bed are relatively rare because most outcrops are of insufficient
length and orientation with respect to bedding to trace a single bed laterally for a long distance.
TOC Some single bed flow transformations have been reported by Baruffini et al. (1994) and Drink-
water and Pickering (2001). Al-Siyabi (1998) documented a transformation between a slurry
bed and a massive sandstone in a single bed. Over a lateral distance of 5m (15ft.), this bed
changes from one with a mud content of 920% and an abundance of fluid escape pipes, to a
Start
massive sandstone with a mud content of 47% and without fluid escape structures. Whether
the slurry bed is upcurrent or downcurrent of the massive sandstone is not known, as no pale-
ocurrent indicators are present on the bed.
Ref. List
Remacha and Fernandez (2003) state that individual beds within the Eocene Hecho
Group (south-central Pyrenees, Spain) can be traced and correlated for tens of km in the
downcurrent direction, where they grade from coarse-grained channel fill, to finer-grained
Search channel-lobe transition deposits, to sheetlike lobes, and, finally, into basin plain deposits. Cor-
relations of individual beds is possible because of the presence of numerous key marker beds
that can be traced for these distances. Numbers of beds between markers, as well as their
Help stacking patterns and facies characteristics, provide the means for correlation. In this manner,
>50% of lobe beds (individual beds >9cm thick) have been traced downcurrent to form thin-
bedded basin-plain facies (individual beds <9cm thick). There is only a decrease of 2.5 thin
beds/km over these distances. The basin plain facies comprise classical Bouma turbidites
(Fig. 4-5) (F8 and F9 beds of Mutti et al., 1999).
Laboratory experiments support the observations of downcurrent flow transformations.
A flume experiment by Hampton (1972) provided a conceptual model for the evolution of a
turbidity current from a debris flow. Marr et al. (2001) and Mohrig and Marr (2003) have gen-
erated turbidity currents from parent debris flows in laboratory flume experiments. Two
processes that generate the turbidity currents from debris flows are: (1) grain-to-grain erosion
of sediment from the surface of a debris flow and its subsequent ejection into the overlying
water column, and (2) turbulent mixing of ambient water in front of a debris flow into its head,
and subsequent dilution to form a turbidity current. In the first case, erosion of sediment from
the head of the debris flow generates an overlying, less dense turbidity current that outruns the
debris flow and continues advancing down slope. The resultant deposit is a thin turbidite bed
on top of, and in front of the debrite. Whether the first or second process dominates in an indi-
vidual flow depends upon whether dynamic stresses at the head of the flow are of sufficient
magnitude to overcome the effective yield strength of the parent debris flow. If these stresses
are sufficiently large, then ambient water in front of the flow can be entrained into the head of
the flow, promoting transformation to a less dense turbidity current. If dynamic stresses are
insufficient to overcome yield strength, then grain-tograin erosion from the leading edge of
the flow occurs, and eroded grains inject into the overlying water column to generate a turbid-
ity current. The velocity at which the head is moving downslope appears to be a main factor in
which of the two processes dominates.
Contrary to the examples presented above, Pratson et al. (2000) claim that it is not possi-
ble for debris flows to transform into turbidity currents. Their experimental work indicates that
at a sediment concentration of > 10% , the flow is sufficiently cohesive to inhibit the exchange
of water and sediment across the surface. Thus, the sediment concentration remains constant.

4-137
Sediment-Gravity Flows and Their Processes

The presence of more than one gravity flow type within a single flow, documented
4 experimentally by Marr et al. (2001) and Mohrig and Marr (2003), had been suggested ear- lier
by others. Sanders (1965) interpreted the Bouma sequence as the deposit of two different kinds
of flow: a basal flowing grain layer and an overlying turbulent flow to which he restricted
3 the term turbidity current (Mutti et al., 1999). Allen (1985) and Postma et al. (1988) consid-
ered debris flows and low- and high-density turbidity currents as parts of a single sediment
gravity flow (Fig. 4-11). The most recent model (Wagerle, 2001) places debris flows at the
TOC base of the flow where sediment concentrations are >50% (Fig. 4-34). The body and head of
the flow contain 2050% sediment concentration (high-density flow) in turbulent suspen-
sion,whereas the low-density wake of the flow contains <20% sediment, also in turbulent sus-
pension, and consists of the finest-grained sediment fraction. Sedimentary particles within
Start each part of the flow are deposited at a different time and in a different position down the dep-
ositionalaxis.
The process by which sediment gravity flows split apart from an original single flow is
Ref. List termed flow stripping. Flow stripping is the process of separation of the components of a
sediment gravity flow as it travels within a sinuous channel (Piper and Normark, 1983), and is
thought to be unique to sinuous submarine channels (Peakall et al., 2000; Posamentier, 2001).
Search This process occurs along outside bends of sinuous channels, where turbidity current flows
accelerate, similar to flow in fluvial channels. Sediment gravity flows of high velocity are
unable to negotiate the bend, resulting in breaching of levees and deposition of sediment as
splays on the outside of the levee, immediately downcurrent from the bend (Fig. 4-35). Flow
Help
velocity would normally be insufficient for all sediment to overtop the levee bend, so that the
coarser-grained fraction of the flow remains within the channel while the finer-grained sedi-
ment is stripped out and transported to an extra-channel or levee location.
Mutti et al., (1999) refer to such flows as bipartite gravity flows, which include a
basal, fast-moving, relatively coarse granular layer and an overlying turbulent suspension of
finer-sized grains. Mutti et al. (1999) claims that the basal layer corresponds to the high-den-
sity turbidity current of Lowe (1982), in contrast to a low-density turbidity current which
probably originated directly as a turbulent flow. One direcct line of evidence for bipartite grav-
ity flows is the bidirectionality of paleocurrent indicators within an apparently single flow
(Fig. 4-36). According to Mutti et al. (2002), these current indicators result from the lower,
denser part of a bipartite flow moving in one direction within a channel, while the upper, less
dense part of the flow splits along the channel margin, and then recombines with the denser
part of the flow further downcurrent.

The Parts of a Gravity Current sediment concentration


(Vol. %)
0% W = wake
Decreasing Concentration

WS H = head
W
Decreasing Grain Size

LO
BIL
B = body
Decreasing Energy

Low-Density Flow
Fine Grained Fraction

20% upper concentration threshold


H High-Density Flow B Medium Grained Fraction

Debris Flow 50% lower concentration threshold


Coarse Grained Fraction

Figure 4-34. Conceptual model (cross section) for a single sediment gravity flow composed of debris flow, high-
density, and low-density turbidity current components. (Modified from Wagerle, 2001). Reprinted with per-
mission of R. Wagerle.
k

4-138
Sediment Gravity Flow Processes and Deposits

TOC

Start

Ref. List

Search

Help
Figure 4-35. Schematic illustration of a turbidity current in a channel and the process of flow stripping, where
the upper part of the turbidity current flow overtops the outer levee at a bend in the channel. The lower,
coarser-grained part of the flow remains in the channel. (Modified from Peakall et al., 2000). Reprinted with
permission of Society of Sedimentary Geology (SEPM) and American Association of Petroleum Geologists.

Figure 4-36. Photograph showing two different orientations of paleocurrent directions in a single bed. The
base of the sandstone exhibits flute casts which are oriented obliquely to the viewer. The upper part of the bed
exhibits climbing ripples oriented at approximate right angles to the viewer. Marnoso-arenacea Formation,
Apennine Mountains, Italy.
k

4-139
Sediment-Gravity Flows and Their Processes

According to Knellers (1995) classification, temporal variations in flow can result in a


4 variety of flow transformations and resulting deposits in one position on the sea floor
(Fig. 4-6). One example of a set of alternating Bouma Tb-Tc beds, interpreted to represent the
deposit of a single fluctuating flow, is shown in Figure 4-37. An unusual deposit has been
3 observed in some outcrops of deep water strata. The deposit consists of a lower, massive or
dewatered sandstone capped by a sandstone bed with shale and sandstone clasts, and with
complex soft sediment-deformed and/or undulatory structures (Fig. 4-38). Haughton et al.
TOC (2001) refer to such beds as a linked sediment couplet; possible origins for the upper part of
this couplet include: (a) vestige of a debris flow that only partially transformed to a turbidite,
and (b) bulking of a lagged tail of a sandy turbidity current and overriding of the flow. Another
possibility is flow stripping, whereby the lower and upper parts of the flow split in different
Start directions, then recombined further in the downcurrent direction (Mutti et al., 2002).

Ref. List

Figure 4-37. Photograph of


alternating planar laminated
Search
(P) Bouma Tb and ripple
bedded (R) Bouma Tc beds in
outcrop, suggesting surge
Help flow (alternating higher and
lower velocity flow) during
deposition. Upper Cretaceous
Dad Sandstone Member,
Lewis Shale, Wyoming,
U.S.A.

Figure 4-38. Outcrop photo-


k
graph of a linked couplet
(Haughton et al., 2001). The
lower part of this bed is evenly
bedded, but the upper part
has been contorted and the
beds have been detached. The
scale is in 0.1 ft. increments.
Lower Pennsylvanian Jack-
fork Group sandstone.
k

4-140
Post-Depositional Reworking of Sediment Gravity Flows

Post-Depositional Reworking of Sediment Gravity Flows


4
Reworking of deep marine sands on the ocean floor by a variety of currents is well docu-
mented (Heezen et al., 1966; Bouma and Hollister, 1973; Stow and Lovell, 1979; Stow and
3 Holbrook, 1984, Shanmugam 1997). In this section we discuss the two major types of currents,
large-scale contour currents and smaller-scale, local currents. Contour currents, which are major,
unidirectional oceanic currents that parallel slope contours (geostrophic currents), have been
TOC
invoked for large-scale reworking of turbidite deposits into contourite drifts (Mutti, 1992).
The internal stratigraphy of contourites has been discussed extensively by Stow and Lovell
Start (1979) and Stow and Holbrook (1984), who point to the difficulty in differentiating contourites
from fine-grained turbidites in the rock record owing to similar characteristics. Reworking of
sands by more localized, unidirectional bottom currents can winnow mud and improve sorting
Ref. List (Fig. 4-39), thus improving potential reservoir quality when buried to depth. As shown in Figure
4-39 if reworking is too prolonged or intense, the resulting sand beds may become very thin and
isolated, thus reducing their reservoir potential. An example of reworked Bouma Tb-Tc beds is
Search shown in Figure 4-40; in this example, reworking is noted by a sharp top to an asymmetric ripple
that is steep-sided in the opposite direction to primary cross-stratification.
In addition to unidirectional currents, internal waves and tidal currents within confined
Help
areas such as submarine canyons (Shepard et al., 1969; Bouma and Hollister, 1973) can
rework sediments and provide bi-directional ripples in sandstones. Complex ripples have been
noted on the modern deep sea floor (Fig. 4-41) as well as in deepwater sandstones (Fig. 4-42).

D D
C C C
B B B B
A A A A A

1 2 3 4 5

A A
6 7 8 9 10

Figure 4-39. Schematic illustrations showing post-depositional processes by which a complete


Bouma sequence (Fig. 4-14) might be progressively winnowed and reworked by ocean bottom
currents (after Stanley, 1993). Reprinted with permission of Elsevier Publishing Co.
k

4-141
Sediment-Gravity Flows and Their Processes

TOC

Start Figure 4-40. Outcrop photograph of a bottom-current


reworked Bouma Tb-Tc bed. Note the sharp top and
reverse orientation of the steep side of the ripple top
relative to the direction of internal cross-stratification.
Ref. List Precambrian turbidite, Newfoundland, Canada.

Search

Help

Figure 4-41. Photograph of


k
lunate- and lingoid-shaped rip-
ples from the Scotian Sea floor
at 4,010 m water depth, south-
east of Terra del Fuego, south-
ern Atlantic Ocean. (Reineck
and Singh, 1980, Fig. 664).
Reprinted with permission of
Elsevier Publishing Co.
k

4-142
Allocyclic vs. Autocyclic Processes and Vertical Stacking Patterns

TOC

Start

Ref. List

Search

Help
Figure 4-42. Outcrop photograph of ripples on a sandstone bedding plane surface. The ripples
are similar in appearance to those shown in Figure 4-41. Upper Cretaceous Dad Sandstone Mem-
ber, Lewis Shale, Wyoming, U.S.A.

k
Allocyclic vs. Autocyclic Processes and Vertical Stacking Patterns
The issue of transport and deposition of sediment gravity flows is significant to the
larger considerations of interpreting allocyclic (extrabasinal) and autocyclic (intrabasinal) pro-
cesses from vertical stacking patterns observed in outcrop or core. Vertical successions
represent the product of localized deposition at one position on the sea floor.
At the single bed scale, a progressive upward decrease in sediment grain size (waning
flow deposits of Fig. 4-6) implies decrease in depositional energy over time at one location.
Conversely, a progressive upward increase in grain size (waxing-flow deposits of Fig. 4-6)
implies an increase in depositional energy over time.
At the larger scale, a succession of sharp-based, amalgamated, massive or pebbly sand-
stones in sharp contact with underlying finer-grained Bouma Tb-d beds might be interepreted
as recording a rapid drop in relative sea level and resulting seaward shift in the axis of deposi-
tion (allocyclic process) (Mutti et al., 1994).
However, without other criteria, an equally plausible explanation for this vertical suc-
cession might be a lateral shift in the depositional axis, with higher energy flows being
deposited atop lower energy flows without any change in relative sea level (autocyclic pro-
cess). In this case, successive sedimentation events in one area would provide bathymetric
relief over time. When the relief becomes too high for sediment gravity flows to override, the
flows are diverted to the bathymetrically lower, adjacent sea bed (Fig. 4-21). This style of dep-
osition, which does not require a change in relative sea level, is termed compensation
bedding (Mutti and Sonnino,1981). Compensation bedding occurs at all scales, from major
sediment buildups (Walker, 1978, his Fig. 18), to thinner successions of beds (Jordan et al.,

4-143
Sediment-Gravity Flows and Their Processes

1993, their Fig. 69). This principle that the same vertical association of strata can be achieved
4 either by allocyclic or autocyclic processes is shown in Figure 4-43. This principle dictates that
caution should be used when making three-dimensional interpretation of processes from one-
or two dimensional data (logs, cores, outcrops, and 2D seismic).
3
(A)
TOC (Landward) (Basinward)

Start
Lateral shift
Ref. List (B)

Search

Coarser
Help grained

Finer grained (Basinward)

Figure 4-43. Schematic cross sections of depositional intervals with differing grain size and how
their vertical stacking arrangement (area in box) would be identical through either (A) landward
shift (allocyclic) or (B) lateral shift (autocyclic) in depositional axis between deposition of the lower
and upper intervals. (Modified from Al-Siyabi, 1998).

k
Summary: Lessons learned
Our understanding of sediment gravity flow processes continues to evolve at a rapid rate
owing to a resurgence of research into their hydrocarbon-bearing deposits. Key points pre-
sented in this chapter are summarized below.
1. Sediment gravity flows can be initiated in a number of ways, from seismically-induced
slides to flood stage discharge of rivers into the ocean. Such flows can occur with a fre-
quency of years to 100s of years between event, depending upon the type of flow.
2. Critical factors that contol the type of flow and deposit include gravity, sediment con-
centration and flow velocity over time and space, These factors are related to sediment
source area and type, climate, tectonics, and topography of the sea floor.
3. End member gravity flows, which are most common in the rock record, are cohesive
debris flows and fluidal turbidity current flows. Intermediate flow types between true
debris flows and turbidity currents also exist. A combination of observations of rocks and
rock sequences, measurements in modern sedimentary environments, laboratory flume
experiments, and computer modeling have illustrated the complexity of sediment gravity

4-144
References

flows. Contrary to historical perspective, flows can wax, wane, or remain steady in both
4 time and space, giving rise to complex sets of deposits, both laterally and vertically.
4. For descriptive, as well as some genetic classification, the Bouma sequence remains the
fundamental deposit from a turbidity current. Different terms sometimes have been
3 applied to divisions of the Bouma sequence by different authors. Other types of sediment
gravity flows that do not conform to this classification include muddy sandstones, slurry
beds, and massive sandstones (though the latter is sometimes classed as a Bouma Ta
TOC bed). Hyperpycnites (deposits from hyperpycnal flows) can be described by Bouma divi-
sions even though their vertical stratigraphy may differ from that of a classic Bouma
Sequence.
Start 5. Turbidity currents generated from hyperpycnal flows are probably of greater signifi-
cance than previously recognized, particularly off of major deltas. A distinctive coarsen-
ing-upward, followed by fining upward vertical sequence distinguishes hyperpycnites
Ref. List from other types of sediment gravity flows.
6. Downcurrent flow transformations of one type of sediment gravity flow to another type
of flow undoubtedly occur in nature. Deposits that document flow transformations are
Search difficult to find in outcrop, though rare outcrops have been identified. Flow transforma-
tions have been generated in flume tanks and by experimental modeling.
7. Once deposited on the sea floor, sediment gravity flows can be subjected to reworking by
long-lived bottom currents as well as by internal waves and tidal currents (in the shal-
Help
lower portions of submarine canyons). Reworking can winnow the finest-grained fraction
from sands, and sort them, to give rise to excellent potential reservoir rock upon burial.
8. Vertical stacking patterns of sediment gravity flows are important for interpreting the
larger-scale filling history of a basin. A similar vertical sequence of sediment gravity
flows can be produced by processes related to fluctuating base level (allocyclic pro-
cesses) or fluctuations in depositional processes and sites without a change in base level
(autocyclic processes). These differences should be considered when interpreting the
depositional history of a sequence of sediment gravity flow deposits.

References
Al-Siyabi, H.A., 1998, Sedimentology and stratigraphy of the Early Pennsylvanian Upper Jackfork interval in the
Caddo Valley Quadrangle, Clark and Hot Springs Counties, Arkansas, unpublished. Ph.D. dissertation, Col-
orado School Mines, 272 p.
Allen, J.R.L.,1985, Principles of physical sedimentology, London, Allen and Unwin, 272 p.
Baas, J.H., and J.L. Best. 2001, Can sedimentological flows with small amounts of clay particles form massive
sandstone beds?: AAPG. Convention Abstracts with Program, p. A9.
Baruffini, L., C. Cavalli, and L. Papani, 1994, Detailed stratal correlation and stacking patterns of the Gremiasco
and lower Castagnola turbidite systems, Tertiary Piedmont Basin, northwestern Italy, in P. Weimer, A.H.
Bouma, and B. F. Perkins, eds., Submarine fans and turbidite systems, Gulf Coast Section SEPM 15th
Annual Research Conference, Houston, p. 921.
Bouma, A. H, 1962, Sedimentology of some Flysch deposits: A graphic approach to facies interpretation: Amster-
dam, Elsevier, 168 p.
Bouma, A.H. and C.D. Hollister, 1973, Deep ocean sedimentation, in G.V. Middleton. and A.H. Bouma, eds, Tur-
bidites and deepwater sedimentation, SEPM Pacific Section Short Course, p. 79118.
Cochonat, P. and D.J.W. Piper, 1995, Source area of sediments contributing to the Grand Banks 1929 turbidity
current, in K.T. Pickering, R.N. Hiscott, N.H. Kenyon, F. Ricci-Lucchi and R.D.A. Smith, eds., London,
Chapman and Hall, p. 1213.
Das, H., J. Imran, C. Pirmez, and D. Mohrig, 2001, Numerical modeling of turbidity current and sedimentation in
meandering submarine channels, AAPG Convention Abstracts with Program, p. A47.

4-145
Sediment-Gravity Flows and Their Processes

Drinkwater, N.J. and K.T. Pickering, 2001, Architectural elements in a high-continuity sand-prone turbidite system,
4 late Precambrian Kongsfjord Formation, northern Norway: application to hydrocarbon reservoir character-
ization, AAPG Bulletin v. 85, p. 17311757.
Elmore, R.D., O.H. Pilkey, W.J. Cleary;, and H.A. Curran, 1979, Black Shell turbidite, Hatteras Abyssal Plain,
western Atlantic Ocean: Geological Society America Bulletin, v. 90, p. 11651176.
3 Garcia, M.H., 1994, Depositional turbidity currents laden with poorly sorted sediment: Journal Hydraulic Engineer-
ing, v. 120, p. 12401263.
Hampton, M.A., 1972, The role of subaqueous debris flows in generating turbidity currents: Journal Sedimentary
Petrology, v. 42, p. 775793.
TOC Haughton, P.D.W., W.D. McCaffrey, and M. Felix, 2001, Origin and significance of Linked Debrites: A key res-
ervoir heterogeneity in sandy turbidite systems: AAPG Convention Abstracts with Program, p. A83.
Heezen, B.C., R.J. Menzies, E.D. Schneider, M.W. Ewing, and N.C.L Granelli, 1964, Congo submarine canyon:
AAPG Bulletin v. 48, p. 11261149.
Start
Heezen, B.C., C.D. Hollister, and W.F. Ruddiman, 1966, Shaping of the continental rise by geostrophic contour
currents: Science, v. 152, p. 502508.
Jordan, D.W., D.R. Lowe, R.M. Slatt, C.G. Stone, A. D'Agostino, M.H. Scheihing, and R.H. Gillespie, 1993, Scales
Ref. List of geological heterogeneity of Pennsylvanian Jackfork Group, Ouachita Mountains, Arkansas: Applications
to Field Development and Exploration for Deep-Water Sandstones, Arkansas Geological Commission Field
Guidebook 93-1, 141 p.
Khripounoff, A., A. Vangriesheim, N. Babonneau, P. Crassous, B. Dennielou, and B. Savoye, 2003, Direct obser-
Search vation of intense turbidity current activity in the Zaire submarine valley at 400m water depth: Marine Geol-
ogy, v. 194, p. 151-158.
Kneller, B., 1995, Beyond the turbidite paradigm: physical models for deposition of turbidites and their implica-
tions for reservoir prediction: in, A.J. Hartley, and D.J. Prosser eds., Characterization of deep-marine clastic
Help systems, Geological Society Special Publication 94, p. 3149.
Kneller, B. and M.J. Branney, 1995, Sustained high density turbidity currents and the deposition of thick massive
sands: Sedimentology, v. 42, p. 2946.
Kneller, B. and C. Buckee, 2000, The structure and fluid mechanics of turbidity currents: a review of some recent
studies and their geological applications: Sedimentology, v. 47 (Supplement 1), p. 6294.
Kneller, B. and W.D. McCaffrey, 1999, Depositional effect of flow non-uniformity and stratification within turbid-
ity currents approaching a bounding slope: deflection, reflecton and facies variation: Journal Sedimentary
Research, v. 69, p. 980991.
Kolla, V., Ph. Bourges, J.-M Urruty, and P. Safa, 2001, Evolution of deep-water Tertiary sinuous channels off shore
Angola (west Africa) and implications for reservoir architecture: AAPG Bulletin v. 85, p. 13731405.
Legros, F., 2002, Can dispersive pressure cause inverse grading in grain flows?: Journal of Sedimentary Research,
v. 72, p. 166170.
Lowe, D.R. ,1982, Sediment gravity flows II: Depositional models with special reference to the deposits of high-
density turbidity currents: Journal Sedimentary Petrology, v. 52, p. 279297.
Lowe, D.R. and M. Guy, 2000, Slurry-flow deposits in the Britannia Formation (Lower Cretaceous ), North Sea: a
new perspective on the turbidity current and debris flow problem: Sedimentology, v. 47, p. 3170.
Marr, J., P. Harff, G. Shanmugam, and G. Parker, 1997, Experiments on subaqueous sandy debris flows: the role of
clay and water content in flow dynamics and depositional structures: Geological Society America Bulletin,
v. 113, p. 13771386.
Middleton, G.V. and M.A. Hampton, 1973, Sediment gravity flows: mechanics of flow and deposition, in G.V.
Middleton, and A.H. Bouma, eds., Turbidites and deep-water sedimentation, Pacific Section SEPM publica-
tion, p. 138.
Mohrig, D., K. X. Whipple, M. Hondzo, C. Ellis, and G. Parker, 1998, Hydroplaning of subaqueous debris flows:
Geological Society America Bulletin, v. 110, p. 387394.
Mohrig, D. and J.G. Marr, 2003, Constraining the efficiency of turbidity-current generation from submarine slides,
slumps and debris flows using laboratory experiments: Marine and Petroleum Geology.
Mulder, T. and J.P.M. Syvitski, 1995, Turbidity currents generated at river mouths during exceptional discharges to
the world oceans: Journal of Geology, v. 103, p. 285299.
Mulder, T., B. Savoye, and J.P.M Syvitski, 1997, Numerical modeling of a mid-sized gravity flow: the 1979 Nice tur-
bidity current (dynamics, processes, sediment budget and seafloor impact): Sedimentology, v. 44, p. 305326.
Mulder, T., S. Migeon, B. Savoye, and J.M. Jouanneau, 2001a, Twentieth century floods recorded in the deep Med-
iterranean sediments: Geological Society America, v. 113, p. 10111014.
Mulder, T., S. Migeon, B. Savoye, and J.-C. Faugeres, 2001b, Inversely graded turbidite sequences in the deep Medi-
terranean: a record of deposits from flood-generated turbidity currents?: Geo-Marine Letters, v. 21, p. 8693.

4-146
References

Mulder, T., J.P.M. Syvitski, S. Migeon, J.-C. Faugeres, and B. Savoye, 2003, Marine hyperphycnal flows: initia-
4 tion, behavior and related deposits: a review: Marine and Petroleum Geology.
Mutti, E. and F. Ricci Lucchi, 1975, Turbidite facies and facies associations, in Mutti, E. G.C Parea, F. Riccci Luc-
chi, M. Sagri, G. Zanzucchi, G. Ghibaudo and S. Iaccarino, eds., Examples of turbidite facies associations

3 from selected formations of northern Apennines, IX Int Cong. I.A.S., Nice France, Field Trip, p. 2136.
Mutti, E. and M. Sonnino, 1981, Compensation cycles: a diagnostic feature of turbidite sandstone lobes, Interna-
tional Association Sedimentologists, 2nd European Regional Meeting, Bologna, Italy, p. 120123.
Mutti, E., G. Davoli, S. Mora, and L. Papani, 1994, Internal stacking patterns of ancient turbidite systems from col-
TOC lisonal basins, in P. Weimer, A.H. Bouma, and B. F. Perkins, eds., Submarine fans and turbidite systems,
Gulf Coast Section SEPM, 15th Annual Research Conference, p. 257268.
Mutti, E., R. Tinterri, E. Remacha, N. Mavilla, S. Angella, and L. Fava, 1999, An introduction to the analysis of ancient
turbidite basins from an outcrop perspective: AAPG Continuing Education Course Note Series #39, 61p.
Start Mutti, E., F. Ricci Lucchi, and M. Roveri, 2002, Revisiting turbidites of the Marnoso-Arenacea Formation and their
basin-margin equivalents: problems with classic models, Excursion guidebook, Turbidite workshop, 2730
May, Parma, Italy.
Mutti, E., R. Tinterri, Giovanni Benevelli, S. Angella, D. diBiase, L. Fava, N. Mavilla, G. Caanna, and A. Cotti,
Ref. List 2003, Deltaic, mixed and turbidite sedimentation of ancient foreland basins: Marine and Petroleum Geology.
Nakajima, T., 2006, Hyperpycnites deposited 700 km away from river mouths in the central Japan Sea, Journal of
Sedimentary Research, v. 76, pp. 59-72.
Search Peakall, J., W.D. McCaffrey, B.C. Kneller, C.E. Stelting, T.R. McHargue, and W.J. Schweller, 2000, A process
model for the evolution of submarine fan channels: Implications for sedimentary architecture, in A.H.
Bouma and C.G. Stone, eds., Finegrained turbidite systems: AAPG Memoir 72; SEPM Special Publication
68, p. 7388.
Help Piper, D.J.W. and W.R. Normark, 1983, Turbidite depositional patterns and flow characteristics, Navy submarine
fan, California Borderland: Sedimentology, v. 30, p. 681694.
Piper, D.J.W., A.N. Shor, and J.E.H. Clarke, 1988, The 1929 Grand Banks earthquake, slump, and turbidity cur-
rent, in H.E. Clifton, ed., Sedimentologic consequences of convulsive events: Geologic Society of America
Special Paper, v. 229, p. 7792.
Piper, D.J.W. and B. Savoye, 1993, Processes of late Quaternary turbidity current flow and deposition on the Var
deep-sea fan, north-west Mediterranean Sea, Sedimentology, v. 40, p. 557582.
Piper, D.J.W., P. Cochonat, and M.L. Morrison, 1999, The sequence of events around the epicenter of the 1929
Grand Banks earthquake: initiation of debris flow and turbididty current inferred from sidescan sonar, Sedi-
mentology, v. 46, p. 7997.
Pirmez, C., R.T. Beaubouef, S.J. Friedmann, and D.C. Mohrig, 2000, Equilibrium profile and baselevel in subma-
rine channels: Examples from Late Pleistocene systems and implications for the archtecture of deepwater
reservoirs, in P. Weimer, R.M. Slatt, J. Coleman Jr , N.C. Rosen, H. Nelson, A. H. Bouma, M. Styzen,
and D.T. Lawrence, eds., Deepwater Reservoirs of the World, 20th Annual GCSSEPM Foundation Bob F.
Perkins Research Conference, p. 782805.
Posamentier, H.W., 2001. On the role of flow stripping and the deposition of channel levees and frontal splays in
deep-water systems; evidence from 3-D seismic data: AAPG Convention Abstracts with Program, p. A160.
Posamentier, H.W., 2003, Seismic geomorphology and stratigraphy of deep-water channelized turbidites, Marine
and Petroleum Geology, v. 20, p. 677-690.
Postma, G., W, Nemec, and K.L. Kleinspehn, 1988, Large floating clasts in turbidites: a mechanism for their
emplacement: Sedimentary Geology, v. 58, p. 4761.
Pratson, L.F., J. Imran, G. Parker, J.P.M. Syvitski, and E. Hutton, 2000, Debris flows vs. turbidity currents: a mod-
eling comparison of their dynamics and deposits, in A.H. Bouma, and C.G. Stone, eds., Fine grained tur-
bidite systems: AAPG Memoir 72; SEPM Special Publication 68, p. 5771.
Pyles, D.R. and R.M. Slatt, 2007, Stratigraphic evolution of the upper Cretaceous Lewis Shale, southern Wyoming:
Applications to understanding shelf to base-of-slope changes in stratigraphic architecture of mud-domi-
nated, progradational depositional systems: in J. Studlick, G. Steffens, R. Shew, and T. Nelson, eds., AAPG
Studies in Geology 56, Atlas of deep-water outcrops.
Remacha, E. and L. P. Fernandez, 2003, High-resolution correlation patterns in the turbidite systems of the Hecho
Group (South-Central Pyrenees, Spain): Marine and Petroleum Geology.
Reading, H.G. and M. Richards, 1994, Turbidite systems in deep-water basin margins classified by grain size and
feeder system: AAPG Bulletin, v. 78, p. 792822.
Reineck, H.E. and I.B. Singh, 1980, Depositional sedimentary environments, New York, Springer-Verlag, 549 p.

4-147
Sediment-Gravity Flows and Their Processes

Sanders, J.E., 1965, Primary sedimentary structures formed by turbidity currents and related resedimentation mech-
4 anisms, in Middleton, G.V. ed., Primary sedimentary structures and their hydrodynamic interpretation,
SEPM Special Publication 12, p. 192219.
Shanmugam, G., 1996, High-density turbidity currents: are they sandy debris flows?: Journal Sedimentary
Research, v. 66, p. 210.
3 Shanmugam, G., 1997, The Bouma Sequence and the turbidite mind set: Earth-Science Reviews, v. 42, p. 201229.
Shanmugam, G., 2000, 50 years of the turbidite paradigm (1950s1990s): deep-water processes and facies mod-
elsa critical perspective: Marine and Petroleum Geology, v. 17, p. 285342.
Slatt, R.M., S. Phillips, J.M. Boak, and M.B. Lagoe, 1993, Scales of geologic heterogeneity of a deep-water sand
TOC giant oil field, Long Beach Unit, Wilmington Field, California, in E.G. Rhodes and T.F. Moslow, eds.,
Marine clastic reservoirs, examples and analogous, New York, Springer-Verlag, p. 263292.
Shepard, F.P., R.F. Dill, and U. Von Rad, 1969, Physiography and sedimentary processes of La Jolla Submarine
Fan and Fan-Valley, California: AAPG Bulletin, v. 53, p,. 39420.
Start Soyinka, O.A., 2005, Thin-bedded turbidite and hyperpycnite mudstones in the Cretaceous Lewis Shale, Carbon
County, Wyoming: unpubl. M.S.thesis, Univ. Oklahoma, 90p.
Soyinka, O.A. and R.M. Slatt, 2004, Thin-bedded turbidite and hyperpycnite(?) mudstones in the Cretaceous Lewis
Shale, Carbon County, Wyoming: Preliminary results, in Scott, E.D. and A.H. Bouma, Depositional pro-
Ref. List cesses and characteristics of siltstones, mudstones and shales, Spec. Symposium AAPG-SEPM Annual
Meeting, Dallas.
Stanley, D. J. 1993, Model for turbidite-to-contourite continuum and multiple process transport in deep marine set-
tings: examples in the rock record: Sedimentary Geology, v. 82, p. 241255.
Search Stelting, C. E. , K.T. Pickering, A.H. Bouma, J. M. Coleman, M. Cremer, L. Droz, A.A. Meyer-Wright, W.R. Nor-
mark, S. OConnell, D.A.V. Stow, and DSDP Leg 96 Shipboard Scientists, 1985, Drilling results on the
middle Mississippi Fan,in Bouma, A.H, W. R. Normark, and N.E. Barnes, eds., Submarine fans and related
turbidite systems, New York, Springer-Verlag, p. 275282.
Help
Stow, D.A.V. and J.P.B. Lovell, 1979, Contourites: Their recognition in modern and ancient sediments: Earth-Sci-
ence Reviews, v. 14, p. 251291.
Stow, D.A.V. and G. Shanmugam, 1980. Sequence of structures in fine-grained turbidites: comparison of recent
deep-sea and ancient flysch sediments: Sedimentary Geology, v. 25, p. 2342.
Stow, D.A.V. and J.A. Holbrook, 1984, North Atlantic contourites: an overview, in D.A.V. Stow, and D.J.W. Piper,
eds, Fine grained sediments: Processes and facies; The Geological Society Special Publication 15, p. 245256.
Talling, P.J., J. Peakall, W. McCaffrey, L. Amy, and S. Sparks, 2001, How viscosity and density stratification
changes submarine mass flow behaviour: AAPG Convention Abstracts with Program, p. A197.
Uchupi, E. and J.A. Austin, Jr., 1979, The stratigraphy and structure of the aurentian Cone region: Canadian Journal
of Earth Sciences, v. 16, p. 17261752.
Wagerle, R., 2001, Deciphering proximal-distal from along-strike variations in the deep water architecture of the
middle Brushy Canyon Formation, Delaware Mountains, West Texas, unpublished M.S. thesis, Colorado
School of Mines, 253 p.
Walker, R.G., 1978, Deep-water sandstone facies and ancient submarine fans: Models for exploration for strati-
graphic traps: AAPG Bulletin, v. 62, p. 932966.

4-148

You might also like