You are on page 1of 19

Marine and Petroleum Geology 26 (2009) 1900–1918

Contents lists available at ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Hybrid sediment gravity flow deposits – Classification, origin and significance


Peter Haughton a, *, Christopher Davis a, William McCaffrey b, Simon Barker c
a
UCD School of Geological Sciences, University College Dublin, Belfield, Dublin 4, Ireland
b
Institute of Geological Sciences, School of Earth and Environment, University of Leeds, Leeds LS2 9JT, UK
c
BG Group, 100 Thames Valley Park Drive, Reading, Berkshire RG6 1PT, UK

a r t i c l e i n f o a b s t r a c t

Article history: The deposits of subaqueous sediment gravity flows can show evidence for abrupt and/or progressive
Received 4 March 2008 changes in flow behaviour making them hard to ascribe to a single flow type (e.g. turbidity currents,
Accepted 15 November 2008 debris flows). Those showing evidence for transformation from poorly cohesive and essentially turbulent
Available online 1 April 2009
flows to increasingly cohesive deposition with suppressed turbulence ‘at a point’ are particularly
common. They are here grouped as hybrid sediment gravity flow deposits and are recognised as key
Keywords:
components in the lateral and distal reaches of many deep-water fan and basin plain sheet systems.
Sediment gravity flow
Hybrid event beds contain up to five internal divisions: argillaceous and commonly mud clast-bearing
Turbidite
Linked debrite sandstones (linked debrite, H3) overlie either banded sandstones (transitional flow deposits, H2) and/or
Transitional flow structureless sandstones (high-density turbidity currents, H1), recording longitudinal and/or lateral
heterogeneity in flow structure and the development of turbulent, transitional and laminar flow
behaviour in different parts of the same flow. Many hybrid event beds are capped by a relatively thin,
well-structured and graded sand–mud couplet (trailing low-density turbulent cloud H4 and mud
suspension fallout H5). Progressive bed aggradation results in the deposits of the different flow
components stacked vertically in the final bed. Variable vertical bed character is related to the style of
up-dip flow transformations, the distance over which the flows can evolve and partition into rheological
distinct sections, the extent to which different flow components mutually interact, and the rate at which
the flows decelerate, reflecting position (lateral versus distal) and gradient changes. Hybrid beds may
inherit their structure from the original failure, with turbidity currents outpacing debris flows from
which they formed via partial flow transformation. Alternatively, they may form where sand-bearing
turbidity currents erode sufficient substrate to force transformation of a section of the current to form
a linked debris flow. The incorporation of mud clasts, their segregation in near-bed layers and their
disintegration to produce clays that can dampen turbulence are inferred to be key steps in the generation
of many hybrid flow deposits. The occurrence of such beds may therefore identify the presence of non-
equilibrium slopes up-dip that were steep enough to promote significant flow incision. Where hybrid
event beds dominate the entire distal fan stratigraphy, this implies either the system was continually out
of grade in order to freight the flows with mud clasts and clays, or the failure mechanism and transport
path repeatedly allowed transmission of components of the initial slumps distally. Where hybrid beds are
restricted to sections representing fan initiation, or occur more sporadically within the fan deposits, this
could indicate shorter episodes of disequilibrium, due to an initial phase of slope re-adjustment, or
intermittent tectonically or gravity-driven surface deformation or supply variations. Alternatively,
changes between conventional and hybrid event beds may record changes in the flow generation
mechanism through time. Thus the vertical distribution of hybrid event beds may be diagnostic of the
wider evolution of the fan systems that host them.
Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction

A variety of sediment gravity flows shape underwater slopes and


deliver clastic sediment to deep-water basins (e.g. Nemec, 1990;
Stow et al., 1996; Mulder and Alexander, 2001). The flows have
* Corresponding author. different rheologies (fluidal versus plastic) and may exhibit laminar,
E-mail address: peter.haughton@ucd.ie (P. Haughton). mixed or turbulent behaviour giving rise to deposits with

0264-8172/$ – see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.marpetgeo.2009.02.012
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1901

a characteristic geometry, texture and internal structure (e.g. significantly overprinted by syn-to post-depositional water escape
debrites, ‘slurry flow’ deposits, turbidites). Observations from the and autodeformation. The deposit may merely reflect the condi-
modern sea floor (e.g. McCave and Jones, 1988; Zeng et al., 1991; tions operating during collapse of the flow and early consolidation,
Talling et al., 2007), cores (Haughton et al., 2003) and laboratory rather than during the prior transport phase. A further complica-
experiments (e.g. Marr et al., 2001), together with facies correlations tion is that single remobilisation events may generate more
in deep-water successions at outcrop (e.g. Talling et al., 2004) have than one type of flow, or flows that show complex longitudinal
shown that the inferred turbulence state and rheology can vary differentiation.
spatially and/or temporally during the course of a single transport Most classifications focus on stand-alone flow types and associ-
event. The term flow transformation describes changes in flow state ated particle support mechanisms. Mulder and Alexander (2001)
between laminar and turbulent states with or without intervening distinguish hyperconcentrated and concentrated density flows in
‘transitional flow’ conditions, or visa versa (Fisher, 1983; Waltham, between cohesive debris flows and turbidity currents. In their classi-
2004). Transitional flow here refers to the turbulence structure fication, turbidity currents are defined as currents in which fluid
developed as the flow transits from high to low or low to high turbulence provides the ‘main’ particle support, whereas particle
Reynolds number; this is currently poorly characterised for sediment interactions dominate support in hyperconcentrated flows, and
gravity flows but evidence from the deposits of decelerating clay- a combination of grain-to-grain interactions, fluid escape and
rich currents suggests it can involve cycling between more and less turbulence in the case of concentrated flows. Kneller and Buckee
turbulent state reflecting the development of cohesive near-bed (2000) point out that the manner of deposition from friction-domi-
layers (Lowe and Guy, 2000). As many event beds aggrade nated (i.e. non-cohesive) flows remains uncertain and it can be
progressively, they can capture temporal changes in flow structure difficult to deduce flow behaviour from the sedimentary record. They
(Kneller and McCaffrey, 2003) and thus give insights into longitu- instead advocate the broad use of the term turbidity current for any
dinal flow transformations and associated changes in flow rheology: gravity-driven, ‘turbid’ and not necessarily fully turbulent mixture of
the deposit may include vertically juxtaposed intervals deposited (suspended) sediment and water travelling through a less dense
under different flow regimes. To date, most attention has focussed fluid. However, it is still useful to retain the distinction between
on proximal transformation from plastic laminar to fluidal turbulent essentially fully turbulent (low-density turbidity flows) that emplace
behaviour, but as the distal ends of deep-water systems become generally planar and ripple laminated sand and finer grained beds
better characterised, evidence for the reverse transformation (i.e. and mixed mode, higher-concentration but still non-cohesive flows
turbulence suppression and a change via transitional flow regime to (high-density turbidity flows) that tend to emplace sand and/or gravel
cohesion-dominated transport and laminar flow) has been increas- (Lowe, 1982; Mutti, 1992). The key distinction between debris flows
ingly encountered (Chang and Grimm, 1999; Lowe and Guy, 2000; and turbidity currents is thus the mode of deposition viz a viz
Haughton et al., 2003; Sylvester and Lowe, 2004; Talling et al., 2004; en-masse freezing versus progressive aggradation, either grain by
Barker et al., 2008). The outer lateral and distal parts of many grain, or layer by layer (e.g. Kneller and Branney, 1995; Lowe and Guy,
systems are now recognised to include hybrid event beds emplaced 2000; Haughton et al., 2003; Amy and Talling, 2006).
by a combination of fluidal and plastic flow recording switching Whilst flow classification schemes have focussed on discrete
between turbulent, transitional and laminar behaviour. Such beds flow states defined by the degree of turbulence and the importance
challenge existing deposit classification schemes, and demand of cohesion, event beds displaying evidence for abrupt, progressive
a better understanding of the processes leading to the partitioning of or cyclical changes in flow behaviour during deposition are
the flow into sections with different rheology and turbulence state. commonplace yet hard to place in existing schemes. An alternative
The aims of this paper are (1) to re-examine the classification of or complementary classification focuses on the deposit and the
sediment gravity flow deposits in the light of evidence for hybrid inferred rheology(ies) and turbulence state of the flow type that
rheology and flow state within a single flow; (2) to document the created it (Fig. 1). The scheme distinguishes the deposits of flows
variable character and distribution of hybrid event beds across that were either cohesive (debris flows, mudflows), non-cohesive
a range of deep-water systems and (3) to consider the wider signif- (low- and high-density turbidity currents, cohesionless debris
icance of hybrid event beds for their host system and to examine why flows) or show evidence for mixed behaviour at a point (composite
they are well expressed in some systems, but not in others. flows). ‘At a point’ is specified because debris flows commonly
evolve to become turbidity currents due to flow transformations as
2. Classification of sediment gravity flows they travel basinward (e.g. Piper et al., 1999). Consequently, upslope
debrites can be lateral to turbidites down-dip. However, the
Sediment gravity flows are characterised by a wide range of emphasis here is on changes in flow behaviour as the flow passes
sediment concentration, particle cohesion, particle support mech- a point, and the record of longitudinal flow structure captured
anisms, duration and rheology. Several classification schemes vertically in a bed (cf. Kneller and McCaffrey, 2003). Note that ‘bed’
(e.g. Pickering et al., 1989; Nemec, 1990; Ghibaudo, 1992; Stow and is used throughout this paper in the sense of an ‘event bed’ – it may
Johansson, 2000; Mulder and Alexander, 2001) have been proposed include prominent internal lithological divisions that relate to
to capture the spectrum extending from dense, laminar cohesive stacking of different or evolving flow types.
flows (debris flows) to dilute fully turbulent flows (turbidity Three end-member types of mixed behaviour are evident from
currents) and their deposits. Whilst there is general agreement on vertical bed sequences (Fig. 1): (1) beds that show evidence for an
the flows and deposits at either end of the spectrum, the definition upward change from cohesive to non-cohesive transport; (2) beds
of intervening flow types has proved less satisfactory. This is partly that show an upward change from non-cohesive to cohesive
because such flows are hard to observe and instrument in nature behaviour, sometimes capped by deposits of non-cohesive flow
and the flow characteristics must be inferred indirectly and with again; and (3) beds that record more complex cyclical alternations
difficulty from the sedimentary record (Kneller and Buckee, 2000). between the deposits of more and less cohesive flow.
In addition, the behaviour of the flows can be critically controlled
by small amounts of suspended clay, the flows can be strongly 2.1. Outsized ‘megabeds’
vertically heterogeneous with deposition occurring mainly from
poorly characterised near-bed layers, they can evolve and change The first bed type comprises debrite–turbidite couplets that
character laterally, and once deposited structures may be tend to occur as isolated, unusually large-volume event beds, often
1902 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

Fig. 1. (A) Classification scheme for event beds emplaced by subaqueous sediment gravity flows recognising that many are the deposits ‘at a point’ of non-cohesive, cohesive and in
some cases transitional flow types. (B) Whereas ‘stand-alone’ debrites and turbidites (either high or low density) dominate the record of many deep-water systems and are arranged
in facies tracts recording downslope transformation and increasing dilution of the flow (debrites passing to high-density turbidites and eventually low-density turbidites
down-dip), other systems show a down-dip progression for non-cohesive flows (depositing turbidites) to flows that are partitioned into sections with radically different rheology,
with the deposits of cohesive flow components increasingly dominating in distal parts.

on the floors of confined basins or fjords (Kleverlaan, 1987; a trailing turbulent wake as illustrated experimentally by
Seguret et al., 1984). Some have been related to seismic triggering Hampton (1972). Alternatively, an initial cohesive failure may
(and referred to as seismoturbidites although they are not wholly trigger a subsequent sand failure by breaching or setting up
produced by turbidity currents; Mutti et al., 1984; Bouma, 1987; seiches in confined basins (Kleverlaan, 1987). Debrite–turbidite
Syvitski and Schafer, 1996). The stacking of turbidite sand on ‘megabeds’ tend to record rare large magnitude events against
debrite can be related to surface transformation due to reverse a background of ‘normal’ gravity current deposition (Mutti et al.,
shear whereby mixing at the front of a debris flow produces 1984; Haughton, 2000).
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1903

2.2. Turbidites with linked debrites 1988, their Fig. 1). Slurry flow in terrestrial settings is used in yet
another context to refer to plastic flows with behaviour akin to
The second category of mixed behaviour is expressed in beds debris flows (Pierson and Costa, 1987). In addition, the term slurry
whose thickness is not outsized compared to that of related event flow is also commonly used in engineering to refer to a wide
beds. This bed type is common, particularly in distal and lateral range of (generally pipe) flows including relatively dilute particle
margins of deep-water systems where it may be the dominant bed suspensions. It thus seems prudent not to identify slurry flow as
motif. Although the facies expression varies (see below), beds in a distinct type of flow nor its deposit as a stand-alone bed type.
which a transition from non-cohesive to cohesive rheology is Although the origin of the banding has been disputed (c.f. the
inferred are grouped as ‘hybrid’ event beds. Haughton et al. (2003) chlorite gelation model of Blackbourn and Thomson, 2000), the
identified a recurring facies association in the fringes of several idea that it forms due to processes operating during the transition
sand-prone, syn-rift Jurassic submarine fans in the North Sea Basin. between turbulent and laminar behaviour is borne out by the
Sandstone beds displaying full or partial Bouma sequences are context of banding in hybrid event beds where it tends to span the
surprisingly rare in these systems. Instead, the most common distal transition from cleaner lower sand to clay-prone upper sand or
bed-scale facies unit is a couplet comprising a lower, poorly graded, sandy mud (the deposit of turbidite and debrite respectively). Even
dewatered or structureless sandstone, capped by an argillaceous in Britannia, banding sits between clean ‘S3’ sandstone (interpreted
sandstone or sandy mudrock characterised by common (often as non-cohesive turbidites sensu lato) and muddy sandstone with
exotic) mud clasts and carbonaceous matter, and complex internal clasts attributable to debris flow processes (Barker et al., 2008);
fabrics including sheared sand injections. These occur lateral to and single Britannia beds thus include elements deposited by a range of
down-dip from amalgamated to isolated structureless and/or flow types (and partly overprinted by intrabed remobilisation) and
dewatered sandy turbidite beds – where conventional models to label these as the product of a single flow type obscures the
(e.g. Lowe, 1982) would predict the occurrence of well-laminated, rheological heterogeneity of the original flows. Beds with internal
relatively thin sandy beds deposited by dilute turbidity currents. banded divisions are here considered as a hybrid event bed variant.
These turbidite–debrite couplets occur repeatedly, always with the To summarise, from a pragmatic viewpoint and recognising that
debrite on top of the structureless sandstone, and with evidence the link between deposit and original flow is not straightforward,
that the underlying sand was still wet and actively dewatering sediment gravity flow deposits may be broadly classified as either:
when the debrite overrode it. Thus, they are thought to reflect (1) debrites emplaced en-masse by freezing of laminar or weakly
deposition from a rapidly decelerating sandy turbidity current, turbulent cohesive flows; (2) low-density turbidites deposited by
immediately followed by deposition from a clast-rich debris flow dilute and fully turbulent non-cohesive flows that aggraded
emplaced as part of the same event. The turbidity current and progressively; (3) high-density turbidites, deposited by concentrated
debris flow became partitioned up-dip. Because the debrite units non-cohesive flows commonly with suppressed near-bed turbu-
are most obvious and best developed in the more distal parts of the lence and particle support from a combination of turbulence,
fan systems, the presence of an underlying actively dewatering, particle interactions and fluid escape (Lowe, 1982). These may have
‘‘quick’’ sand bed may be important as this can both nourish the aggraded progressively either grain by grain, layer by layer, or
trailing debrite with water and sand, and provide a low-friction partly en-masse (cf. Shanmugam, 1996), and (4) hybrid flow deposits
basal contact, thus accounting for debrite mobility (see Mohrig deposited by flows that transition or switch from non-cohesive to
et al., 1998; Gee et al., 1999). The term ‘linked debrite’ was intro- cohesive behaviour during deposition, followed by a variably
duced for the upper clast-rich divisions (not the whole hybrid event developed, but generally volumetrically minor dilute turbulent
bed) to emphasise that they are part of the sand emplacement wake. Typically the lower part of the bed is attributed to a high-
event, as opposed to ‘stand-alone’ debrites (Haughton et al., 2003). density turbidite, whereas the upper part is a debrite. They are thus
Similar hybrid beds are recognised widely from the distal and partly progressively aggraded, and partly emplaced en-masse or in
lateral margins of other systems, although the facies expression can pulses. Where transformation was progressive, hybrid flow
vary, an issue we return to below. These event beds may be capped deposits may include thick intervals deposited under transitional
by the deposits of low-density turbidity currents (e.g. Tcde, Tde), to flow conditions (sensu Britannia Formation). Megabeds involve
produce turbidite–debrite–turbidite triplets (cf. McCaffrey and transformation from cohesive, laminar or weakly turbulent flow to
Kneller, 2001). non-cohesive turbulent flow and are classified with hybrid beds as
part of a wider family of composite event beds.
2.3. Britannia-style slurry/banded beds
3. Character of hybrid event beds
The third category of mixed behaviour occurs across a range of
scales, but is best expressed in relatively thick, clay-rich but sand- Since the description of hybrid event beds and linked debrites
prone event beds. Lowe and Guy (2000) and Lowe et al. (2003) from the Upper Jurassic of the North Sea (Haughton et al., 2003),
drew attention to repetitively stacked lighter and darker ‘bands’ in similar and related beds have been encountered in a range of other
cores through thick graded Lower Cretaceous event beds in the turbidite systems (e.g. Sylvester and Lowe, 2004; Talling et al.,
Britannia Field of the UK North Sea. Similar fabrics can also be seen 2004, 2007; Amy and Talling, 2006; Fig. 2). Such beds have also
in some outcrops (albeit often subtle; Sylvester and Lowe, 2004). been documented before in earlier outcrop studies (Wood and
The banded structure was attributed to flows that were interme- Smith, 1958, 1959; Ricci Lucchi and Valmori, 1980). These studies
diate between fully turbulent and laminar flow behaviour, with concur in suggesting hybrid event beds are preserved primarily in
episodic near-bed turbulence dampening forced by clay inter- the lateral and or distal edges of systems otherwise dominated by
particle electrical bonding following clay floc disintegration. Lowe turbulent flow processes, and reveal a wider spectrum of vertical
and Guy (2000) adopted the term ‘slurry flow’ for the depositing facies transitions and bed types which are summarised below,
flow. However, the term had already been applied by Wood and drawing on new subsurface data from NW European Jurassic,
Smith (1958, 1959) to mud-rich, commonly ungraded and unban- Cretaceous and Cenozoic deep-water systems of the North Sea, the
ded sandstones associated with sandier turbidite deposits from the Cenozoic of the UK Atlantic margin, offshore Norway and the Gulf
Aberystwyth Grits Series in Wales, and resembling the hybrid of Mexico. In addition, Carboniferous outcrops in Ireland and
turbidite–debrite couplets introduced above (see also Postma et al., northern England, the Appennine foredeep, Italy (Macigno Fm.,
1904 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

Fig. 2. Selected logged hybrid event bed profiles (note variable scale) from cores in the public domain across a range of North Sea deep-water systems (Jurassic, Cretaceous and
Cenozoic examples). These illustrate how widespread beds of this type are, the range of hybrid event bed styles and thicknesses, and the difficulty of assigning such beds to existing
deep-water process-based classification schemes. Examples are from (A) Upper Jurassic Magnus Sandstone Formation, northern North Sea, well 211/12a-M16; (B) Lower Cretaceous
Bed 78 in well 16/26-B2, Britannia, Field, Central North Sea; (C) Upper Jurassic, Kingfisher Field, well 16/8a-4, south Viking Graben, North Sea; (D) Palaeocene Forties Sandstones,
well 22/9-2, Everest Field, central North Sea; (E, F) Palaeocene Forties Sandstones, well 23/22a-3, Pierce Field, central North Sea; (G, H) Upper Jurassic, Kingfisher Field, South Viking
Graben, North Sea, well 16/8a-4; (I, J) Palaeocene Forties Sandstones, well 23/22a-3 Pierce Field, central North Sea.

Cilento flysch), as well as other published examples (see above) 3.1. Facies motifs
inform the descriptions. A key question concerns the significance of
the different hybrid bed types. In describing this range, cored The basic motif that all hybrid event beds share is a cleaner,
examples (both modern and ancient) are important as they sandier, lower portion to the bed and an upper, generally darker
can provide very clean sections through the argillaceous bed appearing, poorly sorted and/or chaotically structured muddier
tops that commonly are weathered and poorly preserved in outcrop sand or sandy mud division (Figs. 2–4). Unlike conventional
(although there are exceptions e.g. Marnoso Arenacea). For turbidites where clays are fractionated during flow and settle from
example, the development of banding is often best seen on cut suspension to generate a discrete mud cap, in hybrid beds clays are
surfaces (Sylvester and Lowe, 2004). A drawback, however, is that dispersed and mixed with sand grains through much of the upper
event definition can sometimes be ambiguous due to large, bed part of the bed (Figs. 2 and 3). Many of the muddy divisions show
parallel, core-traversing mud clasts in the debritic component of significant vertical fractionation of components such as mud clasts,
the beds that can masquerade as background intervals. Bio- outsized sand grains and carbonaceous matter. The latter may or
turbation can overprint/mimic the bed top fabrics; accordingly may not be capped by a thin (generally a few cm but sometimes
most of the examples examined are chosen from anoxic basins thicker) fine grained, laminated sand–mud couplet. The ‘ideal’
where the structures preserved are physical and the result of flow hybrid bed based on a compilation of bed motifs from a wide range
emplacement and not later disturbance. of systems (Figs. 3 and 4) thus displays up to five internal divisions,
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1905

Fig. 3. Schematic log showing ‘ideal’ organisation of a typical five-part (H1–H5) hybrid event bed. See text for discussion. Note that not all divisions may be present or have a consistent
character; H3 can sit directly on H1, and H1 and/or H4/5 may be missing. H3 may be clast-prone, extensively sand-injected or comprised of muddy sand without clasts. The proportion
of divisions also varies from bed to bed; some are dominated by expanded sandy H1 division, whereas others are dominated by H3. For key to log, refer to Fig. 2.

from base to top (H1) a lower coarser grained, clay-poor or at least Hybrid event beds are generally dcm to m thick, extending
less clay-rich sandstone, (H2) a typically banded mixed or transi- down to a few cm and up to >10 m thick whilst still retaining much
tional interval, (H3) a sandy mudstone or muddy sandstone with or of the structure of the ideal bed (Fig. 2). Generally, the lower
without mud clasts and sand injections, (H4) a structured finer sandstone is structureless or shows dewatering structures,
grained sandstone, often with dark laminae containing mud commonly dishes, or vertical dewatering sheets (Fig. 5A). Bed bases
chip (i.e. mm-scale aggregates) and plant matter segregations, and are generally sharp and planar, with loading or in some cases
(H5) a generally thin graded silt to clay cap that gradationally erosional sole structures. The sand can be normally graded or basal
overlies H4, the lower part of which can contain prominent sandy inverse-graded to ungraded. Typically, parallel and ripple lamina-
pseudonodules. tion are absent from the top of the lower clean sandstone, although

Fig. 4. Example of a typical five-part, sand-dominated hybrid event bed from the outer Palaeocene Forties fan system in the North Sea, well 23/22a-3, 2735 m driller’s depth (d.d.)
relative to Kelly’s Bushing (RKB). The lower main part of the bed is dominated by a structureless sandstone (H1) with a crude consolidation lamination. H2 shows a crude banding
defined by slightly paler (cleaner) band tops with floating mud clasts and deformed dewatering structures suggesting shearing by the overriding flow. H3 in this case contains only
vestiges of mud clasts (m) and is a uniform dark muddy sandstone with a carbonaceous flake (c) towards the top. The event bed is capped by a parallel to ripple laminated H4 very
fine-grained sandstone in which clay aggregates are effectively segregated (creating darker laminae and foresets). Note the pseudo paired drapes. These are not tidally influenced
deposits however (see Dreyer et al., 2004). For key to log, refer to Fig. 2.
1906 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

Sylvester and Lowe (2004) and Talling et al. (2004, 2007) identify to: (a) the relative proportions and presence of divisions H1–H5; (b)
dune scale cross-lamination and ripples before the transition to the the types and importance of mud clasts in H3; (c) the nature of the
upper argillaceous division in some systems. However, care must internal transition from cleaner H1 sand to more argillaceous H3
be taken in distinguishing banding (particularly micro-banding) sand; (d) the development of banded fabrics and H2 divisions
associated with transitional flow behaviour from upper flow (Fig. 6); (e) the texture of the argillaceous matrix in H3 (f) the
regime parallel lamination. Banding characteristically involves extent of syn-depositional sand injection sourced from H1 and
sharp-topped paler laminae/bands loaded into underlying darker, intruding H2 and H3; and (g) the expression of the event bed top
soft-sediment-sheared bands (Lowe and Guy, 2000). (H4 and H5).
Superimposed on the basic division of the bed into sandier The proportion of clean sand (H1) versus argillaceous facies (H2,
lower and argillaceous upper parts are a host of variations relating H3) varies widely from beds dominated by clean, structureless sand

Fig. 5. Core and outcrop examples illustrating a range of hybrid event bed styles and typical associated features. (A) Split core showing upper part of an Upper Jurassic event bed
from distal Magnus Sandstone Member, UK northern North Sea, well 211/13a-10, core depth 11,2820 d.d. RKB. The junction between H1 (dewatered sandstone forming lower 5 cm;
only top shown here) and H3, the upper central darker mud clast-rich unit is modified by a large ‘domal’ sand injection (i) that interfingers laterally with H3 sand-speckled muddy
sand. Note the carbonaceous clast (c), deformed sand-laminated mudrock clast (m) within H3 and pale vertical dewatering streaks (d) in the top of H1. At the top, there is a sharp
transition back into the H4 laminated very fine-grained sandstone (labelled Sl) that characteristically shows strong colour variation due to darker laminae rich in segregated mud
aggregates/chips. (B) Hybrid bed from the Pollica Formation at Santa Maria Castellabate in the Cilento Basin, southern Apennines, Italy. The lower part of bed is clean dewatered
sandstone, overlain by slightly darker mud clast-rich argillaceous sandstone and capped (either side of compass clinometer) by a thin ripple laminated sandstone. (C) Hybrid bed
from the Oligocene ‘Macingo costiero’ turbidite system, coastal Tuscany. Note the lower pale sandstone (H1) and chaotic upper argillaceous sandstone with abundant mud clasts and
carbonaceous matter. Isolated mud clasts occur surrounded by clean sandstone beneath the H1/H3 transition. Pen for scale. (D) Stacked thin hybrid beds from the central Ross
Formation at Kilbaha Bay, western Ireland. Note the paler coloured basal cleaner sand and darker muddy sand with sparse mud clasts towards top. A thin H4 sand occurs on top of
the event bed with the coin (for scale) on it. This has partly collapsed into the muddy sandstone beneath.
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1907

with very thin argillaceous caps (Fig. 2G), in some cases comprising Where mud clasts are small (<0.5 cm), they can be segregated into
just scattered flakey mud or carbonaceous clasts, to beds with laminae, producing a distinctive stratification of cleaner sandstone
either a very thin (cm-thick) basal sandy division or with mud- and mud fleck-charged sandstone, commonly towards the top of
prone sandstones at the bed base (Fig. 2J). Where event bed the argillaceous unit and transitional with the overlying laminated
populations have been examined (e.g. outer Forties fan system, cleaner sandstone. Where coarser grained extrabasinal gravel was
North Sea; Fig. 7), a full range of bed proportions is seen, and beds available up-dip, as in the case of the North Sea Jurassic systems,
containing a high proportion of argillaceous facies have a similar this does not appear in the H3 divisions.
thickness distribution to beds dominated by clean sandstone The transition from clean (H1) to argillaceous (H3) sandstone can
(although the latter class contains a number of thicker sand beds, take a variety of forms. Commonly in the case of the North Sea Jurassic
interpreted to relate to cryptic amalgamation). examples, the transition is sharp and inclined and often cross-cut by
Mud clasts are a key component in the argillaceous H3 parts of sand injections sourced from H1 (Fig. 5A). In the Palaeocene Forties
many event beds, but are rare or absent in others. In the North Sea Fan, either gradational or ‘stepped’ contacts are seen, and more of the
Jurassic fans (Miller, Magnus; Fig. 5A), argillaceous divisions contacts are relatively flat (Figs. 4 and 6). Cross-cutting sand injections
contain abundant mud (and sand) clasts. Generally, the mud clasts are rarer. Gradational contacts can be associated with the develop-
are elongate and define a bedding parallel fabric; where small, they ment of meso- or micro-banding (terminology of Lowe and Guy, 2000;
may show more chaotic arrangements. In some cases, the mud termed H2 here), with banding straddling the contact between
clasts do not resemble the local background mud, suggesting they cleaner and more clay-rich sandstone. In other cases banding is
cannot represent local erosion products. For example, clasts in restricted to the upper part of the basal H1 clean sandstone division
Jurassic H3 divisions in the Kingfisher Field include thin-bedded (Fig. 7). ‘Stepped’ contacts comprise several sharp contacts, spaced cm
mud and sandstone lithologies unlike the local ‘background’ to dcm apart, across which the lithology becomes increasingly darker
deposits. In the Forties H3 divisions, clasts of green siltstone and more mud-prone. Some of the step surfaces are associated with
resembling the underlying Lista Formation occur. Mud clast- sheared dewatering sheets and patches (Fig. 4).
charged beds are less common and many argillaceous sand divi- Although banding was first described in anomalously thick
sions contain only scattered mud clasts with cuspate, apparently (>10 m) event beds from the Britannia Sandstone Formation (Lowe
corroded boundaries, along with many fine mud flecks suggesting and Guy, 2000), it is an integral if more subtle part of hybrid beds in
extensive comminution (Fig. 4). The first occurrence of mud clasts other systems, where typically it is transitional with clean sand-
vertically within the bed is often as clusters of clasts surrounded stone beneath and argillaceous sandstone above (Fig. 6). It appears
by clean sand immediately beneath the transition to muddier to be best developed where mud clasts are rarer in bed caps and
sandstone (Figs. 5C and 6), even in cases where mud clasts are rare/ where clasts are extensively comminuted, and also where the lower
absent in the upper division. Generally these mud clasts are poorly sandstone division is thin and/or relatively fine grained with some
aligned and unstratified and they may be subrounded. Some dispersed clay. It is more common in the Forties event beds (Fig. 7)
argillaceous caps contain only dispersed clay and no mud clasts or than in the Jurassic examples. Similarly to Britannia banding (Lowe
evidence that they were once present (e.g. Talling et al., 2007). and Guy, 2000), lighter layers can have characteristic loaded bases

Fig. 6. Detail of a pair of stacked hybrid event beds from the Forties Sandstone, Everest Field, central North Sea, core depth 85690 –85640 d.d. RKB. Lower bed contains all five
divisions (H1–H5), whereas upper bed lacks evidence for a return to a turbulent tail (H4 and H5 missing). Both beds show well developed banded H2 divisions between H1 (clean
sandstone) and H3 muddy sandstone. Banding ranges from crude segregation of cm-scale lighter and darker intervals (cleaner and more clay-prone respectively) to more typical
sharp-topped lighter divisions with more diffuse loaded bases with underlying darker divisions (see upper part of D). Note how banding spans transition from paler, cleaner sand to
darker muddy sand. For key to log, refer to Fig. 2.
1908 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

Fig. 7. Compilation of event bed statistics for complete event beds from the Everest and Pierce fields on the medial lateral and distal lateral flank of the Palaeocene Forties Fan, North
Sea. (A) Pi charts illustrating the significance of hybrid beds. Bed types 1, 6 and 7 correspond to what would be classified as conventional high-density turbidite, stand-alone debrite,
and low-density turbidite, respectively. The remaining bed types are split into a range of types of hybrid bed on the basis of clean sand proportion (see Davis et al., 2009). Type 3 beds
are those that contain banded H2 divisions. Note the dominance of hybrid style beds in this system. Conventional turbidites comprises only a minor part of the fan fringe sampled in
these two locations, particularly when the relative thicknesses are considered. (B) Cross-plot showing the constituent facies make-up (as a proportion of clay-prone facies primarily
comprising H2 and H3 divisions of the beds) and the range of event bed thicknesses. Note that there is a continuum from sand-rich to clay-rich bed make-up, and that clay-rich Type 3
and Type 4 beds have broadly similar thicknesses to sand-dominated beds (some very thick sand-dominated ‘beds’ may reflect cryptic amalgamation).

and sharp tops; darker layers have a more chaotic appearance. latter case with clots or patches of paler sand balled or sheared
Unlike in Britannia, lighter bands tend not to have obvious internal through the darker matrix.
lamination, and strata-bound dewatering sheets are uncommon. Sand injections in the form of dykes, pipes and sills of pale sand
Significant variation in the range of grain sizes within the argil- can cross both matrix and clasts of the argillaceous division, but are
laceous H3 component of beds is seen. In some cases the H3 division not observed to cut the event bed top (Fig. 5A). Injection is
comprises dominantly fine or very fine-grained clay-rich sand, particularly common in event beds in syn-rift fans such as the
whereas elsewhere it can contain scattered relatively coarse sand Jurassic examples from the North Sea, but appears to be present is
grains (Fig. 5A), even coarser than in the underlying clean sand. The a smaller proportion of beds from the larger, post rift fans, such as
texture can be either homogenous or strongly heterogeneous, in the the larger Forties system.
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1909

Many hybrid event beds preserve thin parallel and/or ripple outcrop correlations (Talling et al., 2004; Amy and Talling, 2006)
laminated caps (Figs. 4 and 5B). Where mud chips are abundant, show that the absence of distinctive H3 divisions up-dip is not
these tend to be parallel stratified to laminated rather than rippled. simply a function of erosional removal. Where full bed profiles are
However, event beds do occur with no H4 sand cap; in some cases preserved, up-dip sandstones are capped by either very thin mud
the argillaceous unit contains foundered and contorted remnants of clast trains and or parallel and rippled laminated sandstone. For
well-laminated fine-grained sands within H3 (Fig. 5D). Rounded a given succession, event beds dominated by H3 divisions have
sandy pseudonodules (mm to cm across) are particularly common similar thicknesses to those with prominent H1 divisions. Were the
at the transition from laminated sand into the capping H5 debrites to come from source failures, they would therefore have to
mudstone (Haughton et al., 2003). surmount and bypass up-dip erosional relief left by the turbidity
current, or irregular relief where slopes are underlain by mobile salt
3.2. Interpretation or shale. As noted above, the transition between H1 and H3 can be
gradational and include an intervening H2 division, signifying
A number of mechanisms have been proposed to explain the a gradual rearward change in rheology along the length of the flow,
repeated association between a lower turbidite sandstone and rather than discrete debris flow and turbidity current components
overlying linked debrite in the same bed (e.g. Haughton et al., 2003; (Barker et al., 2008). In most cases, no tractional structures occur
Talling et al., 2004). These include (1) continued runout of a debris beneath the H3 division, implying the change to cohesive deposi-
flow following partial transformation to a forerunner turbidity tion occurred before the concentration and velocity of the fore-
current, (2) down-dip transformation from non-cohesive to more running turbidity current had dropped significantly, even at the
cohesive flow driven by incorporation of clay via erosion and (3) distal end of relatively large systems (e.g. the 300 km long North
simultaneous or retrogressive failure of heterogeneous (sandy and Sea Forties Fan), and in surprisingly thin event beds (<5 cm in some
muddy) slopes producing a co-genetic but otherwise unrelated cases; Fig. 5D). Hybrid event beds also occur in close association
sandy turbidity current and a muddy debris flow. The latter with turbidite channels and or bypass surfaces in some systems
mechanism is deemed less likely given the occurrence of hybrid (e.g. Carboniferous mid- to upper Ross Sandstone Fm, Ireland,
beds in a wide variety of settings, and the systematic stacking of Permian Brushy Canyon Fm, west Texas), suggesting a link to local
mud clast-rich debrite on top of underlying sand and never the erosion rather than to initial failure. Where flows are generated by
other way round. failure of gravelly syn-tectonic slopes, the linked debrites might be
Care must be taken to rule out bed amalgamation processes expected to carry extraformational gravel components, but this is
(Johansson and Stow, 1995), substrate deformation, injection ‘rip- not the case for Upper Jurassic footwall-sourced fans where coarser
down’ (Kawakami and Kawamura, 2002) or post-depositional grained gravels (Devonian sandstone cobbles and boulders) were
disturbance (Higgs et al., 1998) as alternative explanations for the available up-dip but do not occur in the down-dip H3 debrites.
juxtaposition of chaotic, mud clast-rich intervals and structureless The change in flow behaviour is often presaged by the arrival of
sandstone beds. Evidence for mud clasts that are lithologically scattered mud clasts that were deposited with clean sand in the
distinct from interbedded local mudstones argues against such upper part of H1 divisions, suggesting erosional bulking of the flow
processes, and the presence of undisturbed capping H4/5 divisions with mud clasts and/or clay may be important in driving changes in
points towards flow partitioning, as they imply temporally linked flow cohesion and in dampening turbulence. Bulking via erosion is
emplacement of the underlying divisions whose fixed sequence of known to be an important mechanism in several recent well-
occurrence is most plausibly explained as the vertical expression in characterised slope failures (e.g. 1929 Grand Banks, Piper and Asku,
the deposit of an evolved or inherited longitudinal flow structure 1987; 1979 Nice event, Mulder et al., 1997); the volumes of sedi-
(see below). The thin H4/5 caps appear never to be cross-cut by ment incorporated in the flow during bulking can be volumetrically
sand injections, implying the chaotic and often sand-injected similar to or even exceed that of the initial failure. It is this addi-
division beneath are not a product of post-depositional sand tional bulked component, particularly where muddy, that may be
injection. The injections that accompanied H3 emplacement important in controlling the subsequent evolution and longitudinal
predated H4 deposition. Importantly, the hybrid bed structure is structure of many gravity currents (mechanism 2 above; see
seen in event beds that are isolated in hemipelagic sediment and in Ricci Lucchi and Valmori, 1980; Haughton et al., 2003; Amy and
these cases substrate deformation beneath a succeeding overriding Talling, 2006). Observations of strongly erosional turbidites in
flow can be ruled out. outcrop (Fig. 8) show that mud clasts can be efficiently segregated
Continued runout of a debris flow that has partially transformed hydrodynamically on the bed and such examples may actually
to a forerunner turbidity current (mechanism 1 above) may explain capture the first stage in partitioning of the flow into sections with
deposition of hybrid event beds following large-volume slope different turbulence structure.
failures, and has been implicated in the emplacement of extensive The above observations indicate that the debrites occurring as
basin plain event beds in the Marnosa-Arenacea system, Italy (Amy H3 divisions can arise by flow transformations forced by bulking of
and Talling, 2006), and possibly on the recent Agadir basin floor originally mainly turbulent and hydraulically fractionated flows as
(Talling et al., 2007). Partitioning of the flow in this way can account opposed to runout of partly transformed debris flows from the
for the local occurrence of parallel and ripple laminated sandstone source failures. The upward change within event beds from less
beneath the H1/H3 contact, and a relatively sharp (over a few cm) clay to more clay-prone is most simply attributed to longitudinal
transition between the turbidite and debrite components of the heterogeneity in the structure of the passing flow. The flow front
bed. A forerunning turbidity current can move independently of the was either turbulent (accounting for the grading) or with sup-
debris flow behind it, and may decelerate and become sufficiently pressed turbulence but lacking cohesive effects (ungraded/
dilute to develop tractional structures before the arrival of the inversely graded H1 examples), but evidently the cohesion and
debris flow, particularly where the runout path is long. turbulence suppression became more important upwards, such
Some features of hybrid event beds are less easily explained by that the depositional regime passed through an intermediate or
the disintegration and distal propagation of the initial slope failure transitional flow state (H2) and then became laminar or very
as a trailing debris flow. H3 divisions thicken distally at the weakly turbulent (H3). The flows are inferred to have been
expensive of H1 and are absent or very thin over significant fundamentally turbulent originally as it is necessary to fractionate
distances (10s–100s km) up-dip. Core (Haughton et al., 2003) and components effectively along their length to explain the difference
1910 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

(2006) respectively, and turbidity currents are known to fractionate


clays from coarser grains effectively along their length, producing
for example the familiar Te division of the Bouma turbidite.
However, in the latter case the presence of clays has not forced the
sorts of transformations evident in hybrid event beds, so a signifi-
cant clay content in itself cannot be the key factor promoting the
change in flow behaviour. Although many hybrid beds contain
abundant carbonaceous matter, carbonaceous-rich but clay-poor
sandstones do not tend to indicate the same types of flow trans-
formation (Saller et al., 2006). Instead, the carbonaceous matter
becomes segregated into spaced laminae. Examples of hybrid event
beds without carbonaceous fragments also occur. Consequently,
although the presence of the carbonaceous matter suggests an
ultimate shallow water source, and where concentrated in partic-
ular layers, points to fractionation processes within the flow,
carbonaceous matter is probably not the primary cause of the
interpreted flow transformations.
The occurrence of mud clasts surrounded by clean sandstone in
the upper part of many H1 divisions may be significant. These clasts
are interpreted to have been deposited on an aggrading sand bed
(see Kneller and Branney, 1995) prior to the onset of flow trans-
formations and record the first arrival at the depositional location
of part of the flow charged with mud clasts, initially at low
concentrations (insufficient to influence the flow behaviour). The
isolated mud clasts foreran part of the flow where mud clasts were
more abundant and where the flow behaviour was modified. The
entrainment, transport and partial disaggregation of mud clasts
within near-bed layers may therefore be key to turbulence
suppression and formation of transitional and debrite deposits
linked to underlying sandstones (see Ricci Lucchi and Valmori,
1980; Lowe and Guy, 2000; Haughton et al., 2003; Amy and Talling,
2006; Amy et al., 2006). Postma et al. (1988) suggested that to
preserve mud clasts in the otherwise abrasive environment
provided by the base of a turbidity current required coupling
between erosion and turbulence dampening. Clays dispersed at low
concentrations high in the flow are unable to influence deposition
at the bed and settle out after passage of the flow (as H5 and in Te
Fig. 8. Examples of proximal mud clast entrainment and segregation from a Neogene divisions of Bouma-type beds). A link between clast disintegration
base-of-slope fan system in the Tabernas Basin, SE Spain. Such outcrops capture how and turbulence suppression implies that hybrid event beds are
turbulent gravity currents can acquire and segregate significant volumes of entrained likely to occur where flows are able to entrain, segregate and
mud blocks as a possible first step in linked debrite generation. (A) Scour-fill with
erosional base (b) overlain by fining up imbricated pebble conglomerate (p) and
abrade significant quantities of mud clasts. As such, the formation
sandstone (s) with numerous large mud rip-up clasts (m) spanning contact between of many hybrid type beds is probably intimately tied to mud
conglomerate and sandstone. Rambla Sierra 37 020 11.6000 N 2 220 01.0300 W (B) Close up erosion and to turbulent flows that are out-of-grade up-dip, an
of graded pebble conglomerate and very coarse sandstone with local concentration of issue we return to in a later section. The physical properties of the
mud rip-up clasts. Rambla Sierra 37 020 13.5000 N 2 210 26.3100 W. Note coin centre left for
entrained mud clasts must also partly determine the efficacy of clay
scale. In both examples the mud clast concentrations have been frozen proximally.
Linked debris flow formation may have been stifled by a high rates of deposition and/ release to the current.
or the coarse texture of the underlying sediment that would have inhibited basal The variable vertical structure of hybrid event beds outlined
lubrication. above may relate to a number of factors including: (1) the gener-
ation mechanism viz a viz source slumping versus erosional
bulking; (2) the length scale over which the flows are able to evolve
in behaviour captured in the bed. A ‘‘complete’’ hybrid event bed is and segregate components, a function of where along the transport
capped by deposition from a trailing waning low concentration and path the flow became partitioned and the overall scale of the
turbulent cloud (H4 and H5), implying a return to non-cohesive system; (3) the extent to which different flow components mutu-
behaviour before settling of a mud plume. Long distance correla- ally interact and mix with just-deposited sediment and (4) the rate
tions of hybrid event beds (Amy and Talling, 2006) show little at which the flows decelerate, reflecting position (lateral versus
evidence of compensation suggesting they can leave an essentially distal) and gradient changes.
flat basin floor with no evidence for mounding. This indicates little In smaller sand-rich systems (e.g. w10 km radius North Sea syn-
strength for the H3 linked debris flow deposits, despite the pres- rift Jurassic fans), H2 divisions are rare, H3 divisions are clast rich,
ence of large mud clasts. and contacts between debrite and underlying H1 sandstones are
The high mud clast content, poor sorting and lack of grading of sharp but irregular. In larger systems (e.g. the >300 km long Forties
the H3 divisions confirm an upward change to cohesive behaviour. fan, North Sea), transitional H2 divisions are better developed, mud
It is likely that the loss of turbulence is linked to increasing near- clasts in H3 divisions are less obvious (although they are often
bed clay concentrations. The ability of even small quantities of clay present as heavily ‘‘corroded’’ remnants), and internal contacts are
to suppress turbulence and modify the settling regime has been generally planar rather than foundered. The reason for such
shown experimentally by Baas and Best (2002) and Amy et al. differences could reflect the degree of internal longitudinal
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1911

Fig. 9. Explanation for the variable character of hybrid event beds based on differences in the scale of system. Inferred near-bed turbulence intensity with time illustrated
schematically by cross-plots. In small, sand-rich fan systems (e.g. North Sea Jurassic, offshore mid-Norway Cretaceous) (A), lateral partitioning of the flow into non-cohesive and
cohesive sections occurs over small length scales and hence the interval between linked debrite arrival and prior sand deposition is short, producing extensive sand injection and
inhibiting mud clast attrition. (B) In larger fan systems (e.g. 300 km long Forties fan system) flows partition and stretch over longer length scales, allowing a more transitional
change from non-cohesive to cohesive, with more extensive clast attrition and more modest autoinjection of H3 by sand. Note a similar down fan thickening of H3 at the expensive
of H1 is envisaged.

segregation within the depositing flows which will be sensitive to short distances, despite the large scale of the overall system. As the
flow length (Fig. 9). The short flows active on small fans may not lateral margins will be thinner and carry less turbulent energy than
have had sufficient time or space to develop efficient longitudinal the core of a flow, particularly where they passively impinge on
segregation of components, and entrained mud clasts would a confining slope, these may be more susceptible to turbulence
have undergone less mechanical breakage due to short transport dampening and flow transformation compared to the flow axis
distances. Partitioned debris flows would have developed as (Barker et al., 2008). Bed style in flow margins may differ from that
a consequence of up-dip erosion and overridden just deposited, along the flow axis – for example, banded sandstones and evidence
actively dewatering sand. They would thus have been prone to load for transitional flow may be more common in lateral as opposed
into, and to be autoinjected by, poorly consolidated sand from to distal margins (Barker et al., 2008; Davis et al., 2009).
beneath, producing the sharp foundered contacts, and H3 divisions
with common multiphase sand injections. At longer length scales, 4. Distribution of hybrid event beds at system scale
the flow head would have stretched away from the flow tail and
more efficient turbulent segregation of components could have Since the recognition of linked debrites and hybrid event beds, it
taken place. In addition, greater mud clast attrition would have has become apparent that they occur widely across a range of
occurred over the longer transport distances. Consequently, the systems, but vary in the nature and extent of their expression. They
turbulent flow head would have passed rearward into a zone of are common in some systems, but rare or absent in others. Here we
clay-induced turbulence suppression via an intervening part of the focus on their distribution at system scale, drawing attention to
flow showing transitional behaviour to account for banded H2 different fan scale arrangements that may reveal something of the
deposits between H1 turbidite and H3 debrite. As the bed aggraded wider fan character and supply mechanism. In this way, the limited
more slowly from a longer partitioned flow, there would have been information from facies distribution in a core or outcrop section
less scope for foundering and upward injection of H1 sand (the may be indicative of the wider fan type and system behaviour.
H1 sand was better consolidated and progressively – rather than Where the hybrid flows are triggered by bulking, the presence and
suddenly statically – loaded). However, during emplacement the style of linked debrites may have implications for the identification
linked debrite may have become coupled downwards with the of equilibrium versus non-equilibrium feeder and fan slopes, and
mechanically weak beds beneath and driven local shearing and their evolution.
modification of the transition zone (Fig. 4). Abundant hybrid beds are found in cores from the lateral and
A simple association between system scale and bed expression distal fringes of small Jurassic syn-rift, deep-water fans in the UK
is complicated by the potential for down-dip incision. When North Sea (Fig. 10). What is significant about their occurrence here
turbulent flows traverse irregular slopes, nick-point migration or is that they dominate the entire vertical record of the fans through
entrenchment through growing structures may produce erosion at a number of cycles of initiation, growth, aggradation and decay,
some point along the flow path, freighting a turbidity current with separated by regional condensed sections. Very few ‘conventional’
mud clasts. If the flow becomes overcharged and decelerates turbidite beds occur: nearly every sandstone bed >10 cm thick has
rapidly, linked debris flows may form and evolve over relatively an associated debrite cap. Important constraints on the origin of the
1912 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

Fig. 10. Detailed logged section of core from well 16/8a-4 from the outer part of the Upper Jurassic Miller–Kingfisher fan system in the south Viking Graben, North Sea. See Garland
et al. (1999) for context. The log illustrates the dominance of hybrid event beds with prominent H3 linked debrites throughout the stacked fan deposits (only a small section of
which is shown here). Conventional turbidite facies are rare and restricted to thin stacked low-density turbidites in background muds during periods of fan activity (e.g. the interval
at ca. 13,2980 d.d. RKB). For key to log, see Fig. 2.

‘linked debrite’ component of the couplets come from the rafted with sand grains coarser than those present in the immediately
clasts they transport. In the Jurassic examples, these include non- underlying sandstone bed, and such grains are also present in the
local slope lithologies so it is difficult to appeal to local seabed argillaceous matrix. Some of the mud clasts are either too large
erosion on the basin floor, or post-depositional foundering of clasts (>30 cm) to have been suspended in a turbulent current and/or too
due to sand injection. In addition, the clasts are locally armoured fragile and angular to have been turbulent current bedload, and so
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1913

must have travelled within debris flows for some distance. The mainly in the basal part of the succession, with an upward transi-
common association with abundant carbonaceous matter, some- tion to conventional turbidite facies. For example, the Carbonif-
times fractionated from the mud clasts, is also significant, as it erous Ross Sandstone in the Clare Basin, western Ireland, comprises
indicates probable linkage to the contemporaneous shelf (although up to 380 m of stacked deep-water sandstones (Elliot, 2000;
‘shelf floods’ can deliver carbonaceous matter to the upper slope; Martinsen et al., 2000; Fig. 11). For the most part, this sandy
Schwehr et al., 2007). Large footwall-derived Devonian sandstone succession was deposited by turbulent currents carrying
clasts in the Brae–Miller–Kingfisher system do not occur in the a restricted, primarily fine-grained, sand load, and poorly graded,
down-dip H3 debrites but characterise the fill to the proximal Brae unstructured sandstones are capped by relatively thin, ripple
channels. laminated divisions implying the flows had modest, low-concen-
A second facies pattern is seen in fans deposited in tectonically tration tails. However, the basal Ross Formation, seen north of
quiescent basins where the hybrid event beds are encountered Ballybunion, western Ireland (Fig. 11B and C) comprises a sandying-

Fig. 11. (A) Reconstruction of ‘acme’ Ross Sandstone fan system, western Ireland, and location of the Ballybunion section that preserves the system onset. (B) View from the sea of
the lower Ross Sandstone Formation illustrating the isolated sheet architecture at this level. (C) Summary logged section through the basal part of the Ross Sandstone north of
Ballybunion showing that hybrid flow deposits are common in the basal part of the fan stratigraphy, with conventional turbidites more important away from the base.
1914 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

upward, non-amalgamated sheet succession characterised by transfer from carbonate to clastic shelf-slope gradients; thereafter
hybrid event beds and sandstone beds showing ‘banding’ (sensu the slopes were broadly in equilibrium, and aggradation of the
Lowe and Guy, 2000). These are replaced higher in the succession basin floor discouraged renewed up-dip incision. Local occurrences
by conventional sheet turbidites with ripple laminated bed tops. of hybrid event beds in the upper parts of the formation may reflect
Repeated condensed sections demonstrate that the fan system was local disequilibrium due to lateral slope failures and/or periods of
episodically shutdown, but there was no recurrence of the debrite fan channel incision.
caps to beds on re-initiation of the fans in these instances. Local The distribution of hybrid event beds in the Forties fan system
hybrid beds occur in the upper Ross Formation (Fig. 5D), mainly indicates they were common during the initial establishment of
associated with channelisation and erosion surfaces. the fan. The lower Forties Formation is also characterised by
In a third style of turbidite succession, hybrid event beds occur mass transport complexes, and hence fan growth was initially
locally interleaved with conventional high or low-concentration across an uneven surface that would have promoted local inci-
turbidite beds during either fan aggradation or in the decay phase. sion and deposition. Halokinesis may also have contributed to
Cores in the lateral and distal reaches of the Cenozoic Forties fan departures from equilibrium slopes by deforming the basin floor.
system in the North Sea show common hybrid beds in the lower fan The upper Forties Formation contains bundles of hybrid beds
growth phase, but they also occur interleaved with intervals of towards the lateral and distal margins of the fan suggesting
stacked sandy turbidites in the upper part of the Forties Formation episodes when flows became choked with mud clasts, alternating
(Davis et al., 2009). Hybrid beds interleaved with conventional with times when flows remained largely turbulent. This may
basin floor turbidites are also seen in the Marnosa Arenacea correspond to the creation of fan valleys on the mid fan where
Formation of the Appennine Basin, Italy (Talling et al., 2004; Amy many of the producing reservoirs are in thick sandy turbidites
et al., 2006) and the Agadir Basin (Talling et al., 2004, 2007). filling incisional fan valleys. The hybrid beds down-dip may
Lastly, hybrid event beds can be important components of the correspond to valley-incising flows, and the sandy conventional
fill to slope channel complexes. Palaeocene (‘T34’) channels in the turbidites to valley-transmitting and valley-filling flows. In this
Schiehallion Field on the Atlantic margin west of the Shetland case, linked debrite deposition could be anticipated to have
Islands have been cored in well 204/25a-2. Towards the base of the switched on synchronously across the lower reaches of the fan
channel complex, abundant mud clasts occur mainly at sandy bed during episodes of fan-valley incision.
bases, whereas higher in the fill, mud clasts and carbonaceous The concentration of mud clasts and clays in hybrid beds in the
clasts are concentrated in bed tops in the form of bioturbation- upper parts of the Schiehallion channel fill (Fig. 12) is interpreted to
modified H3 divisions of hybrid beds. record the migration of the locus of erosion upslope during channel
filling. The mud clast-charged bed-bases low in the channel
4.1. Interpretation complex suggest local erosion and lag formation, whereas the
overlying hybrid beds require erosion farther up-dip to allow the
The character of the North Sea Jurassic hybrid beds indicates an flows to subsequently partition into sandy forward part and trailing
up-dip source for the H3 debrites – they cannot have been triggered turbulence dampened debrite. This style of fill may be character-
from distal or lateral slopes on account of their texture (oversized istic of backstepping of the system during filling and/or upslope
and/or exotic clasts, common carbonaceous flakes, coarse sand knickpoint migration.
grains). The dominance of hybrid beds through the cycles of
growth, aggradation and decay of the fans implies that these 4.2. Significance
systems were predisposed to producing beds of this type, either
because remnants of the source failures were easily transferred If hybrid bed development reflects periods when disequilib-
downslope to the basin floor, or the flows generated in shallow rium, over-steep, ‘out-of-grade’ slopes were present up-dip, their
water were prone to erosional bulking. The absence of footwall occurrence in fan deposits provides important clues to the wider
clasts in the Miller–Kingfisher linked debrites suggests the latter system behaviour and style. Disequilibrium slopes may occur for
was more likely. The implication is the delivery system slopes were a variety of reasons (Fig. 13A), including (1) flow magnitude
fundamentally out-of-grade through-out. Because these fans were increases; (2) tectonic deformation and modification of canyon or
all formed in the hanging wall of major extensional faults, the fan gradients; (3) gravity-driven surface deformation such as mass
slopes down which flows accessed the basin floor would have transport emplacement and rise of compressional toe-of-slope
remained steep through on-going fault displacements. Thin thrust ridges; (4) shale or salt diapirism; (5) a range of autocyclic
condensed horizons could also represent significant periods of time effects that operate on morphologically complex slopes and fan
during which the turbidite pathways were steepened by faulting. surfaces such as minibasin spilling and upslope knickpoint retreat,
Similar fan systems dominated by disequilibrium slopes and an and (6) occurrence of steep inherited slope profiles. The distribu-
abundance of hybrid flows include the Angel Formation of the NW tion of hybrid event beds within the fan deposits can identify
Australian shelf (Deller et al., 2006) and the syn-rift Cretaceous of whether the fan system was continuously out of grade, as for
offshore mid-Norway (Fugelli and Olsen, 2007). instance where tectonics prevents equilibrium slopes developing
Common hybrid event beds during the early phase of fan initi- at all, or intermittently ‘out of grade’, where fan establishment
ation in the basal Ross Sandstone Formation suggests that the first requires a discrete period of slope adjustment, or intermittent
flows to enter the basin either traversed ‘out-of-grade’ slopes and tectonic, surface deformation, or supply factors induce periods of
incorporated large volumes of mud from the Clare Shale formation disequilibrium (Fig. 13B). The characteristics of the debrite units
or were associated with initial failures that shed trailing linked themselves may give important additional clues as to where and
debrites downslope. Thereafter, flows deposited conventional why the flows were triggered; the clasts they carry potentially
turbidites, suggesting maintenance of equilibrium feeder and fan reveal something of the environment/degree of incision up-dip,
slopes through multiple cycles of fan initiation, growth, aggrada- and the sand grains incorporated in the debrites reflect where
tion and decay. The morphology of the basin in which the Ross along the transport path the uncoupling into turbidity current and
Sandstone was deposited was inherited from the earlier carbonate linked debris flow arose. Where the debris flow component
platform/basin and draped by the Clare Shale (up to 200 m thick). contains sand grains that are significantly coarser than the sand
Early disequilibrium may thus have been associated with the beneath it, this suggests debrite partitioning further up the system,
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1915

Fig. 12. Summary logged section through a fining-upward stack of Palaeocene (T34 age) sediment gravity flow deposits from a Schiehallion slope channel complex on the Atlantic
margin west of Shetland. For context, see Leach et al. (1999). Note the dominance of linked debrites in the upper part of the sandbody. Mud clasts in the lower part of the sandbody
are restricted to bed bases and record local erosion. Clasts in the linked debrites above must have come from erosion higher upslope and imply erosion migrated upslope as the
channel filled (backstepping and knickpoint migration?).
1916 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

Fig. 13. Summary of conditions leading to a dominance of hybrid style over conventional turbidite facies. Systems prone to mud clast erosion are deemed likely to show evidence for
flow partitioning and for the development of cohesive behaviour in the lateral and rearward part of the sediment gravity currents. Such a situation (A) is likely to arise where (1) slopes
are maintained out-of equilibrium by faulting or tectonic tilting (e.g. Upper Jurassic turbidites of the North Sea), (2) the sea floor is particularly irregular due to gravity sliding or
halokinesis (e.g. Palaeocene Forties fan, North Sea), (3) clastic systems drape prior carbonate-constructed basin margins (e.g. Ross Sandstone, western Ireland) and (4) where very
large-volume flows penetrate deep into muddy distal basins (Lower Cretaceous Britannia system, North Sea). (B) Schematic logs illustrating four characteristic distributions of hybrid
event beds at system scale. No vertical scale implied. Hybrid beds may either dominate the lateral and distal margins of the complete fan succession, only characterise fan start-up
sections, occur sporadically throughout the succession along with conventional turbidites, or characterise the upper part of the sand-prone section. See text for discussion.

and there is then an issue as to whether the flows arise from 5. Conclusions
disintegration of initial slumps or via proximal erosion and bulking.
Where there is less contrast between the sand grains in the debrites Although many turbidity currents become increasingly dilute as
and those in the underlying sand, the incision may have been they ingest water and drop sediment to the bed, others appear to
farther down the system, and bulking due to erosion may have evolve to more cohesive behaviour towards their lateral and distal
driven flow evolution. limit of runout. The latter emplace characteristic hybrid event beds
P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918 1917

that capture significant longitudinal variations in flow rheology as currents: Aptian Britannia Sandstone Member, U.K. continental shelf. Journal of
Sedimentary Research 78, 45–68.
the flow passes. A typical hybrid bed comprises up to five internal
Blackbourn, G.A., Thomson, M.E., 2000. Britannia Field, UR North Sea: petrographic
divisions from base to top: generally structureless sandstone constraints on Lower Cretaceous provenance, facies and the origin of slurry-
deposited from the frontal, high-density turbidity current part of flow deposits. Petroleum Geoscience 6, 329–343.
the flow, banded sandstone (succeeding transitional flow), clast- Bouma, A.H., 1987. Megaturbidite: an acceptable term? Geo-Marine Letters 7,
63–67.
rich muddy sandstone (debris flow), laminated sandstone (low- Chang, A.S., Grimm, K.A., 1999. Speckled beds: distinctive gravity-flow deposits in
concentration turbidity current) and mudstone cap (representing finely laminated diatomaceous sediments, Miocene Monterey Formation,
suspension fallout) – here termed H1–5, respectively. The common California. Journal of Sedimentary Research 69, 122–134.
Davis, C., Haughton, P., McCaffrey, W., Scott, E., Hogg, N., Kitching, D., 2009. Char-
distal expression, presence of gradational contacts between less acter and distribution of hybrid sediment gravity flow deposits from the outer
and more cohesive behaviour and evidence for hydraulic segrega- Forties Fan, Palaeocene Central North Sea. UKCS. Marine and Petroleum Geology
tion of components suggests that many hybrid event beds arise via 26, 1919–1939.
Deller, K.E., Payenberg, T.H.D., Lang, S.C., Daniel, R., 2006. Sedimentological insights
flow transformations during flow runout. The presence of mud into deepwater mass gravity sands from the North West Shelf, Australia. In:
clasts suspended in muddy sandstone augers the onset of laminar AAPG Annual Conference Abstract Volume, Houston, Texas.
or weakly turbulent flow conditions and implies that erosion and Dreyer, T., Bujak, J., Brunstad, H., Ramberg, J., Søyseth, L., 2004. Mixed deep- and
shallow-water depositional model for the Forties Sandstone Member in the
hydraulic concentration of the clasts within largely turbulent flows South Central Graben, North Sea. Norwegian Journal of Geology 84, 191–233.
is important to subsequently forcing flow transformations. Attrition Elliot, T.R., 2000. Depositional architecture of a sand-rich, channelized turbidite
and disintegration of mud clasts carried in near-bed layers can system: the upper carboniferous Ross Sandstone Formation, western Ireland.
In: Gulf Coast Section SEPM Annual Research Conference, Deep-water Reser-
release clay particles, whose interactions dampen turbulence
voirs of the World, pp. 342–373.
towards the base of the current, promoting cohesive effects, Fisher, R.V., 1983. Flow transformations in sediment gravity flows. Geology 11,
laminar or only weakly turbulent flow and en-masse rather than 273–274.
grain-by-grain deposition. The upper, low-concentration levels of Fugelli, E.G.M., Olsen, T.R., 2007. Delineating confined slope turbidite systems
offshore mid-Norway: the Cretaceous deep-marine Lysing Formation. American
the flow may remain turbulent and influence the final stages of Association of Petroleum Geologists Bulletin 91, 1577–1601.
deposition. The scale of the system and the site of erosion along the Garland, C.R., Haughton, P.D.W., King, R.F., Moulds, T.P., 1999. Capturing reservoir
flow pathway may influence the character of the deposit as it will heterogeneity in a small, sand-rich submarine fan, Miller field. In: Fleet, A.J.,
Boldy, S.A.R. (Eds.), Petroleum Geology of NW Europe: Proceedings of the 5th
determine the length scale and extent of longitudinal rheological Conference, pp. 1199–1208.
changes, and the time interval between the arrival of different flow Gee, M.J.R., Masson, D.G., Watts, A.B., Allen, P.A., 1999. The Saharan debris flow: an
components. Larger systems are characterised by less interaction insight into the mechanics of long runout submarine debris flows. Sedimen-
tology 46, 317–335.
between the different stages of the flow, greater mud clast attrition Ghibaudo, G., 1992. Subaqueous sediment gravity flow deposits – practical criteria
and in some cases by the development of banded divisions for their field description and classification. Sedimentology 39, 423–454.
resulting from a zone of transitional flow between the turbulent Hampton, M.A., 1972. The role of subaqueous debris flow in generating turbidity
currents. Journal of Sedimentary Petrology 42, 775–793.
front of the current and the following debris flow. Banded deposits Haughton, P.D.W., 2000. Evolving turbidite systems on a deforming basin floor,
previously described as a separate category of ‘slurry’ flow can be Tabernas, SE Spain. Sedimentology 47, 497–518.
seen as part of a wider spectrum of hybrid event bed types. In cases Haughton, P.D.W., Barker, S.P., McCaffrey, W., 2003. ‘Linked’ debrites in sand-rich
turbidite systems – origin and significance. Sedimentology 50, 459–482.
where the hybrid event beds arise by erosional bulking and
Higgs, R., Reading, H.G., Burne, R.V., 1998. Return of ‘‘The fan that never was’’;
turbulence suppression, their distribution at system scale can help Westphalian turbidite systems in the Variscan Culm Basin; Bude Formation
identify whether the system is in equilibrium or not. Hybrid beds (South-West England); discussion and reply. Sedimentology 45, 961–975.
are common in the outer reaches of systems that are out-of-grade Johansson, M., Stow, D.A.V., 1995. A Classification Scheme for Shale Clasts in Deep
Water Sandstones. In: Geological Society, Special Publications, vol. 94 221–241.
up-dip by virtue of changes in flow scale, active tectonics or Kawakami, G., Kawamura, M., 2002. Sediment flow and deformation (SFD) layers:
continuing substrate deformation. They are locally common in fan evidence for intrastratal flow in laminated muddy sediments of the Triassic Osawa
initiation stages during channel development and extension, and/ Formation, Northeast Japan. Journal of Sedimentary Petrology 72, 171–181.
Kleverlaan, K., 1987. Gordo megabed: a possible seismite in a Tortonian submarine
or prior to systems attaining grade when initiated over inherited fan, Tabernas Basin, province Almerı́a, southeast Spain. Sedimentary Geology
slope profiles. 51, 165–180.
Kneller, B.C., Branney, M.J., 1995. Sustained high-density turbidity currents and the
deposition of thick massive sands. Sedimentology 42, 607–616.
Acknowledgments Kneller, B., Buckee, C., 2000. The structure and fluid mechanics of turbidity
currents: a review of some recent studies and their geological implications.
Sedimentology 47, 62–94.
The research described here was partly funded under the Kneller, B.C., McCaffrey, W.D., 2003. The interpretation of vertical sequences in
auspices of Phase 5 of the Turbidites Research Group by Anadarko, turbidite beds: the influence of longitudinal flow structure. Journal of Sedi-
mentary Research 73, 706–713.
BG-Group, BHP Billiton, BP, Chevron, ConocoPhillips, Kerr McGee,
Leach, H.M., Herbert, N., Los, A., Smith, R.L., 1999. The Schiehallion development. In:
Devon, Maersk, Norsk Hydro, Shell, Statoil and Woodside. We Fleet, A.J., Boldy, S.A.R. (Eds.), Petroleum Geology of NW Europe: Proceedings of
would like to thank journal referees David Hodgson and Peter the 5th Conference, pp. 683–692.
Talling and editor Lawrence Amy for their perceptive comments on Lowe, D.R., 1982. Sediment gravity flows; II, depositional models with special
reference to the deposits of high-density turbidity currents. Journal of Sedi-
the original manuscript. mentary Petrology 52, 279–297.
Lowe, D.R., Guy, M., 2000. Slurry-flow deposits in the Britannia Formation (Lower
Cretaceous), North Sea: a new perspective on the turbidity current and debris
References flow problem. Sedimentology 47, 31–70.
Lowe, D.R., Guy, M., Palfrey, A., 2003. Facies of slurry-flow deposits, Britannia
Amy, L.A., Talling, P.J., 2006. Anatomy of turbidites and linked debrites based on Formation (lower Cretaceous), North Sea: implications for flow evolution and
long distance (120  30 km) bed correlation, Marnoso Arenacea Formation, deposit geometry. Sedimentology 50, 45–80.
Northern Apennines, Italy. Sedimentology 53, 161–212. Marr, J.G., Harff, P.A., Shanmugam, G., Parker, G., 2001. Experiments on subaqueous
Amy, L.A., Talling, P.J., Edmonds, V.O., Sumner, E.J., Lesueur, A., 2006. An experi- sandy gravity flows: the role of clay and water content in flow dynamics and
mental investigation of sand–mud suspension settling behaviour: implications depositional structures. Geological Society of America Bulletin 113, 1377–1386.
for bimodal mud contents of submarine flow deposits. Sedimentology 53, Martinsen, O.J., Lien, T., Walker, R.G., 2000. Upper Carboniferous deep water sedi-
1411–1434. ments, western Ireland: analogues for passive margin turbidite plays. In: Gulf
Baas, J.H., Best, J.L., 2002. Turbulence modulation in clay-rich sediment-laden flows Coast Section SEPM Annual Research Conference, Deep-water Reservoirs of the
and some implications for sediment deposition. Journal of Sedimentary World, pp. 533–555.
Research 72, 336–340. McCaffrey, W.D., Kneller, B., 2001. Process controls on the development of stratigraphic
Barker, S.P., Haughton, P.D.W., McCaffrey, W.D., Archer, S.G., Hakes, B., 2008. trap potential on the margins of confined turbidite systems and aids to reservoir
Development of rheological heterogeneity in clay-rich high-density turbidity evaluation. American Association of Petroleum Geologists Bulletin 85, 971–988.
1918 P. Haughton et al. / Marine and Petroleum Geology 26 (2009) 1900–1918

McCave, I.N., Jones, K.P.N., 1988. Deposition of ungraded muds from high-density Schwehr, K., Driscoll, N., Tauxe, L., 2007. Origin of continental margin morphology:
non-turbulent turbidity currents. Nature 333, 250–252. submarine-slide or downslope current-controlled bedforms, a rock magnetic
Mohrig, D., Whipple, K.X., Hondzo, M., Ellis, C., Parker, G., 1998. Hydroplaning of approach. Marine Geology 240 (1–4), 19–41.
subaqueous debris flows. Geological Society of America Bulletin 110, Seguret, M., Labaume, P., Madariaga, R., 1984. Eocene seismicity in the Pyrenees
387–394. from megaturbidites of the South Pyrenean Basin (Spain). Marine Geology 55,
Mulder, T., Alexander, J., 2001. The physical character of subaqueous sedimentary 117–131.
density flows and their deposits. Sedimentology 48, 269–299. Shanmugam, G., 1996. High-density turbidity currents: are they sandy debris flows?
Mulder, T., Savoye, B., Syvitski, J.P.M., 1997. Numerical modelling of a mid-sized Journal of Sedimentary Research 66, 2–10.
gravity flow: the 1979 Nice turbidity current (dynamics, processes, sediment Stow, D.V., Johansson, M., 2000. Deep-water massive sands: nature, origin and
budget and seafloor impact). Sedimentology 44, 305–326. hydrocarbon implications. Marine and Petroleum Geology 17, 145–174.
Mutti, E., 1992. Turbidite Sandstones. AGIP, 275 pp. Stow, D.A.V., Reading, H.G., Collinson, J.D., 1996. Deep seas. In: Reading, H.G. (Ed.),
Mutti, E., Ricci Lucchi, F., Seguret, M., Zanzucchi, G., 1984. Seismoturbidites: a new Sedimentary Environments: Processes, Facies and Stratigraphy. Blackwell
group of resedimented deposits. Marine Geology 55, 103–116. Science, Oxford, pp. 395–453.
Nemec, W., 1990. Aspects of sediment movement on steep delta slopes. In: Sylvester, Z., Lowe, D.R., 2004. Textural trends in turbidites and slurry beds from the
Colella, A., Prior, D.B. (Eds.), Coarse-grained Deltas. Special Publication of the Oligocene flysch of the East Carpathians, Romania. Sedimentology 51, 945–972.
International Association of Sedimentologists, vol. 10, pp. 29–73. Syvitski, J.P.M., Schafer, C.T., 1996. Evidence for an earthquake-triggered basin
Pickering, K.T., Hiscott, R.N., Hein, F.J., 1989. Deep Marine Environments: Clastic collapse in Saguenay Fjord, Canada. Sedimentary Geology 104, 127–153.
Sedimentation and Tectonics. Unwin Hyman Ltd, London, 416 pp. Talling, P.J., Amy, L.A., Wynn, R.B., Peakall, J., Robinson, M., 2004. Beds comprising
Pierson, T.C., Costa, J.E., 1987. A rheologic classification of subaerial sediment– debrite sandwiched within co-genetic turbidite: origin and widespread occur-
water flows. Geological Society of America Reviews of Engineering Geology 7, rence in distal depositional environments. Sedimentology 51, 163–194.
1–12. Talling, P.J., Wynn, R.B., Masson, D.G., Frenz, M., Schiebel, R., Akhmetzhanov, A.,
Piper, D.J.W., Asku, A.E., 1987. The source and origin of the 1929 Grand Banks Dallmeier-Tiessen, S., Benetti, S., Weaver, P.P.E., Georgiopoulou, A., Holz, C.,
turbidity current inferred from sediment budgets. Geo-Marine Letters 7, Cronin, B.T., Amy, L.A., 2007. Debris flow deposition from giant submarine flow
177–182. begins far away from original landslide. Nature 450, 541–544.
Piper, D.J.W., Cochonat, P., Morrison, M.L., 1999. The sequence of events around the Waltham, D., 2004. Flow transformations in particulate gravity currents. Journal of
epicentre of the 1929 Grand Banks earthquake: initiation of debris flows and Sedimentary Research, 129–134.
turbidity current inferred from sidescan sonar. Sedimentology 46, 79–97. Wood, A., Smith, A.J., 1958. Two undescribed structures in a greywacke series.
Postma, G., Nemec, W., Kleinspehn, K.L., 1988. Large floating clasts in turbidites: Journal of Sedimentary Petrology 28, 97–101.
a mechanism for their emplacement. Sedimentary Geology 58, 47–61. Wood, A., Smith, A.J., 1959. The sedimentation and sedimentary history of the
Ricci Lucchi, F., Valmori, E., 1980. Basin-wide turbidites in a Miocene, over-supplied Aberystwyth Grits (Upper Llandoverian). Quarterly Journal of the Geological
deep-sea plain: a geometrical analysis. Sedimentology 27, 241–270. Society, London 114, 163–195.
Saller, A., Lin, R., Dunham, J., 2006. Leaves in turbidite sands: the main source of oil Zeng, J.J., Lowe, D.R., Prior, D.B., Wiseman, W.J., Bornhold, B.D., 1991. Flow prop-
and gas in the deep-water Kutei Basin, Indonesia. American Association of erties of turbidity currents in Bute Inlet, British Columbia. Sedimentology 38,
Petroleum Geologists Bulletin 90, 1585–1608. 975–996.

You might also like