You are on page 1of 14

Sedimentary Geology 271-272 (2012) 44–57

Contents lists available at SciVerse ScienceDirect

Sedimentary Geology
journal homepage: www.elsevier.com/locate/sedgeo

A facies model for internalites (internal wave deposits) on a gently sloping carbonate
ramp (Upper Jurassic, Ricla, NE Spain)
Beatriz Bádenas a,⁎, Luis Pomar b, Marc Aurell a, Michele Morsilli c
a
Dpto. Ciencias de la Tierra, Universidad de Zaragoza, 50009 Zaragoza, Spain
b
Departament de Ciències de la Terra, Universitat de les Illes Balears, 07122 Palma de Mallorca Illes Balears, Spain
c
Dipartimento di Scienze della Terra, Università di Ferrara, 44100 Ferrara, Italy

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

Article history: Internal waves are waves that propagate along the pycnocline, the interface between two density-stratified
Received 2 March 2012 fluids. Even though internal waves are ubiquitous in oceans and lakes, their impact in the sedimentary record
Received in revised form 27 May 2012 has remained largely unrecognized. Internal waves can remobilize the sediment from the depth at which the in-
Accepted 28 May 2012
ternal waves break onto the sea floor. In shelf, or ramp settings, internal wave deposits (internalites) have to be
Available online 6 June 2012
distinguished from tempestites while in slope and deeper settings internalites require distinction from turbi-
Editor: B. Jones dites. The Upper Kimmeridgian carbonate ramp succession cropping out near Ricla (NE Spain) provides some
key evidence to differentiate the depositional processes induced by breaking internal waves from those related
Keywords: to surface storm waves. Sandy-oolitic grainstone eventites, previously interpreted as tempestites, contain evi-
Internalites dence of reworking by turbulent events related to breaking internal waves. Underlying rationale are: 1) they
Breaking internal waves occur in distal mid-ramp position, detached from the coeval shallow-water successions; 2) they do not have
Carbonate ramp the characteristic coarsening- and thickening upward trend of storm deposits; 3) they gradually thin-out to dis-
Jurassic appear both up dip and down dip, interbedded with mid-ramp lime mudstones; and 4) they show little or no
Iberian Basin
erosion towards the shallower areas.
A facies model for internalites produced by two sediment populations, sand and mud, on a gently sloping car-
bonate ramp is proposed. The individual internalites occurring at Ricla include several architectural elements,
sequentially organized in dip direction, which can be related to the flows associated with breaking internal
waves: erosion in the breaker zone, swash run-up and tractive backwash flow. Individual internalites stack,
with down- and up-slope shingling configuration, in dm-thick packages thought to reflect the up-slope and
down-slope migration of the breaker zone, in turn related to depth variations of the palaeo-pycnocline. Pack-
ages occur in dm- to m-thick clusters suggested to reflect changes in sediment supply and/or variations in
water stratification affecting the energy of internal waves.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction (Small and Martin, 2002; Staquet and Sommeria, 2002; Nash and
Moum, 2005; Santek and Winguth, 2007). Solitons, or “solitary internal
Among eventites, sediments reflecting the effects of turbulent wave packets” sensu Apel (2002), are ubiquitous wherever water cur-
events (Seilacher, 1982, 1991), tempestites and turbidites are the rents and stratification occur in the neighbourhood of irregular topogra-
most frequently recognized. Tempestites and turbidites share some phy, particularly near shelf edges, seamounts, sills and submarine
common features, namely an erosional phase, reflected by the basal canyons (Munk, 1981; Ostrovsky and Stepanyants, 1989; Global-Ocean-
erosion surface, and a subsequent depositional phase during waning Associates, 2004; Wolanski et al., 2004; Santek and Winguth, 2007).
of the turbulence which produces a graded sediment succession due These conditions frequently happen in coastal regions, especially during
to different settling velocities of the sediment grains. summer months when a shallow thermocline develops. Solitons typically
Internal waves are gravity waves that propagate along a pycnocline, consist of rank-ordered wave packets, with number of cycles varying
the interface between two different density fluids (Munk, 1981; Apel, from a very few to a few tens, and with the largest amplitude and longest
2002). Any perturbation of the pycnocline, induced for example by sur- wave at the front and the smallest at the rear (Apel, 2002; Quaresma et
face waves, wind-stress fluctuations, tsunamis, tidal currents or river al., 2007). Their amplitudes vary from a few up to 140 m and their max-
plumes flowing into a coastal ocean, propagates as an internal wave imum wavelengths from less than 100 m to more than 5 km (Apel et al.,
2007; Quaresma et al., 2007; Santek and Winguth, 2007).
⁎ Corresponding author. Tel.: + 34 976762247.
Under thermohaline circulation, like in modern oceans, the
E-mail addresses: bbadenas@unizar.es (B. Bádenas), luis.pomar@uib.es (L. Pomar), pycnocline is induced primarily by temperature and secondarily by
maurell@unizar.es (M. Aurell), mrh@unife.it (M. Morsilli). salinity gradients in the water column. The depth of the pycnocline

0037-0738/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.sedgeo.2012.05.020
B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57 45

Fig. 1. A) Palaeogeography of western Europe during the Late Kimmerdigian (modified from Dercourt et al., 1993; paleolatitude from Osete et al., 2011). B) Main facies belts in the
northeastern Iberian Basin during the Late Kimmeridgian (adapted from Bádenas and Aurell, 2008). Numbers 1 to 6 indicate reference localities: 1 —Aldealpozo, 2 — Veruela, 3 —
Ricla outcrop, 4 — Aguilón, 5 — Ariño and 6 — Calanda. C) Synthetic stratigraphy of the Kimmeridgian in the northern Iberian Basin (see location of reference localities 1 to 6 in
B) including main facies belts and transgressive–regressive (T–R) sequences Kim1 and Kim2 (adapted from Aurell et al., 2010). Black box indicates the stratigraphic location of
the studied interval at Ricla outcrop within the transgressive deposits of the Kim2 Sequence enlarged in Fig. 2.

in open oceans is highly variable, changing with latitude and season pycnocline; Pomar et al., 2012). In small interior seas the seasonal
(from few to some 10's of metres in the case of the seasonal pycnocline, pycnocline is commonly shallow (few- to some tens of metres) and
from 100 to some 100's of metres in the case of the permanent strongly influenced by riverine discharge, wind regimes, and season,

Fig. 2. Facies architecture of the transgressive deposits (TST) of the third-order Kimmeridgian-2 Sequence (Kim2) in Ricla (updated from Bádenas and Aurell, 2001). Intervals rich in
sandy-oolitic eventites are concentrated in the distal mid-ramp setting and are detached from shallower facies. The two studied intervals rich in eventites located between logs 5
and 7 are indicated (see Fig. 1 for location in a broader context).
46 B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57

Fig. 3. A) Detail of the stratigraphic log 7 in Ricla (see Fig. 2 for location; adapted from Bádenas et al., 2005). The grey-shaded area denotes the two intervals with eventites analysed
in this paper. B) View of the two studied intervals rich in sandy-oolitic and oolitic eventites. C) Microfacies image of an oolitic grainstone eventite, with quartz grains in their cores.

being sharper and shallower during summer (Brown et al., 1989). In a world oceans (Wolanski et al., 2004; Quaresma et al., 2007; Lim et al.,
“greenhouse” world, the origin of the pycnocline was probably domi- 2010). Refraction on the shelf strongly orients the packet crests
nated by halothermal (salinity-driven) conditions (e.g., Kennett and along isobaths and retards their speed of advance (Emery and
Stott, 1991; Pak and Miller, 1992; Nunes and Norris, 2006). Gunnerson, 1973; Apel, 2002). Breaking of internal waves on sloping
Internal waves and solitons propagating along the shallow-water surfaces creates episodic and repetitive high-turbulent events and
pycnocline mostly dissipate over the continental shelf regions of the consequently erosion and transport of sediments (e.g., Pomar et al.,
B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57 47

Fig. 4. Architectural elements in the sandy-oolitic internalites at Ricla. An internalite commonly contains three basic elements sequentially organized in a down-slope direction: a–b–c–b–a.
Occasional elements (d and e) may also occur underlying element c.

2012). In contrast to tempestites (which occur in shallow shelf settings et al., 2005), are here reinterpreted as internalites (sensu Pomar et al.,
down to the storm-wave base sensu Seilacher, 1982), and turbidites 2012), the product of breaking internal waves.
(commonly accumulated on the basin floor), breaking of internal Evidence indicating sediment reworking by internal waves has
waves on sloping surfaces creates high-turbulent events in mid-shelf been obtained by characterization of the spatial distribution of the
settings, when depending on the seasonal thermocline, or deeper on sandy-oolitic eventites at outcrop scale and the analysis of architec-
the continental slope and in submarine canyons when it relates to the tural elements and stacking pattern of individual eventites in two
permanent thermocline (Cacchione and Wunsch, 1974). selected intervals rich in eventites. This work describes the detailed
Although internal waves in the interior of oceans and lakes are as distribution of the architectural elements of eventites generated by
common as surface waves (LaFond, 1966; Munk, 1981; Global-Ocean- breaking of internal waves on a gently sloping ramp sea floor. The
Associates, 2004; Thorpe, 2005), their impact in shaping sediments character and along-dip relationship of architectural elements
has remained largely unrecognized (see review in Pomar et al., 2012). within individual internalites are related to variable hydrodynamic
Key evidence to differentiate the depositional processes induced by processes, with a more precise description and development of the
breaking internal waves from those related to surface storm waves general model on sediment dynamics during breaking of internal
can be gathered in the Upper Kimmeridgian carbonate ramp succession waves previously proposed in Pomar et al. (2012). The significance
cropping out near Ricla (NE Spain). Sandy-oolitic deposits interbedded of the studied eventites in a broader context and the possible effects
with mid-ramp lime mudstones and marls, previously interpreted of depth variations of the pycnocline in carbonate ramp settings are
as tempestites and storm lobes (Bádenas and Aurell, 2001; Bádenas discussed.

Fig. 5. Example of internalites 1 to 14 (letters in white denote the type of architectural elements; see Fig. 4). Internalite 1 consists of a lower discontinuous interval of rip-up clast floatstone
(occasional d element) overlain by a bed with down-slope dipping cross-lamination (c element), with occasional rip-up clasts, that passes down slope into starved ripples (b element) and
thin laminae (a element). Internalites 2 to 7 correspond to attached and detached starved ripples (b elements) separated by muddy layers, deformed by compaction. Internalites 3, 4 and 5
infill a depositional depression in front of internalite 1. Internalites 8 and 9, with down-slope dipping cross-lamination (c element), are separated by an erosional surface marked by occasional
rip-up clasts; they pass down slope into starved ripples (b element). Note the erosion surface at the base of internalite 8. Internalites 10 to 13 are amalgamated levels of attached starved
ripples (b elements), with minor intercalated muddy intervals. Internalite 14 corresponds to starved ripples (b element) passing down slope into a down-slope dipping cross-laminated
bed (c element). An erosion surface underlying c element cuts into internalites 12 and 13, and an undulating top surface is blanketed by mud.
48 B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57

Fig. 6. Decimetre-thick package of individual internalites 1 to 6 (letters in white denote the type of architectural elements; see Fig. 4). Internalites 1 to 3 are lenticular beds with
down-slope dipping cross-lamination (c elements), separated by erosional surfaces. Internalite 4 corresponds to a rip-up clast floatstone bed (d element) with basal erosional
surface cutting into internalites 2 to 3. Note the discontinuous muddy interval (see M) preserved from erosion. Lime mudstone rip-up clasts are usually concentrated in erosive
furrows; they are variable in size and shape, from rounded (see rC) to angular (see aC). Clasts occasionally occur in down-slope dipping accumulations (see dC). Internalites 5
and 6 consist of down-slope dipping cross-laminated beds (c elements). Note basal erosional surfaces and occasional rip-up clasts (see oC).

2. Sedimentary context of the studied eventites The sandy-oolitic eventites of Ricla found in distal mid-ramp position
were previously related to storm events and, based on grain size, thick-
2.1. The Iberian Kimmeridgian carbonate ramp ness and sedimentary structures, were interpreted as storm lobes (dm-
thick cross-bedded deposits), proximal tempestites (cm-thick ripple
During the Kimmeridgian, a wide carbonate ramp developed in the beds) and distal tempestites (thin-sand laminae) (Bádenas and Aurell,
Iberian Basin (Fig. 1), an intracratonic basin located in the northeastern 2001; Bádenas et al., 2005). It is obvious that the main components of
part of the Iberian plate (e.g., Salas and Casas, 1993; Bádenas and these eventites (i.e., quartz grains, ooids) were derived from land or
Aurell, 2008; Aurell et al., 2010). This ramp was around 22° N palae- from the shallowest portion of the ramp, most probably via storm
olatitude (Osete et al., 2011) and opened to the Tethys Ocean to the resedimentation (e.g., Bádenas and Aurell, 2001). However, based on
east, facing the trade winds of hurricanes and winter winds (Marsaglia the distribution of eventites at outcrop scale and relationships with
and Klein, 1983; Price et al., 1995; Bádenas and Aurell, 2008). Shallow shallower facies, Pomar et al. (2012) suggested that the resedimented
(inner- to proximal mid-ramp) settings included a wide range of grains were reworked by high-turbulent events generated in the mid-
reefal- and grain-supported skeletal, oolitic, oncolitic, peloidal and ramp setting by breaking internal waves, so that the eventites actually
intraclastic facies, with intermittent input of siliciclastics supplied from correspond to internalites (Fig. 2). Their underpinning rationale includes:
the emerged Ebro and Iberian massifs. Offshore, the mid- to outer- (1) sandy-oolitic eventites are concentrated in discrete intervals
ramp settings were characterized by monotonous successions of lime detached from the shallower-water reef-apron successions; (2) the
mudstones and marls, including a distal mid-ramp “transitional belt” eventites do not show a coarsening-thickening upward trend typical of
characterized by sandy-oolitic-bioclastic eventites interbedded with storm-generated successions; (3) this distribution cannot be attributed
the mud-dominant successions. to erosion effects (“cannibalism”) of repeated storm events because the
In the Kimmeridgian succession cropping out near Ricla, the tran- eventites gradually thin and pinch out both up dip and down dip.
sition between shallow- and relatively deep deposits of the Iberian
carbonate ramp can be analyzed along a 6-km continuous exposure. 3. Architectural elements of internalites
Here, the “transitional belt” includes abundant eventites deposited
during a transgressive episode at the Eudoxus Zone (TST of the The architectural elements and stacking pattern of individual
third-order Kim2 Sequence: Fig. 1B and C). These deposits are over- internalites have been studied in two stratigraphic intervals (2-m in
lain by prograding (HST) shallow-water (high-energy and lagoonal) total thickness) located c. 11 m above the lower boundary of the stud-
oolitic and bioclastic facies (Bádenas and Aurell, 2001; Aurell et al., ied TST (Figs. 2 and 3). In this interval, the internalites are well devel-
2010). oped in a continuously exposed outcrop covering, down dip, a c. 2-km
wide mid-ramp segment (between reference logs 5 and 7 in Fig. 2).
The sandy-oolitic internalites commonly contain three basic archi-
2.2. Spatial distribution of eventites in the Ricla outcrop tectural elements (Fig. 4): (a) thin laminae, (b) starved ripples, and
(c) a lenticular bed with down-slope dipping cross-lamination.
The overall facies distribution in the TST deposits of the Kim2 Se- When these three elements are present in an individual internalite,
quence across the Ricla outcrop is shown in Fig. 2. In proximal areas, they are sequentially organized in the dip direction within the inter-
these deposits include aggradational metre-sized coral-microbial nalite, so that the thin laminae (a) pass down slope into starved rip-
buildups, with branching corals, chaetetid sponges, stromatoporoids ples (b) and then into a lenticular bed with down-slope dipping
and variable amounts of microbial crusts. These buildups are sur- cross-lamination (c) that thins out and passes again into starved rip-
rounded by skeletal- and oncolitic rudstone aprons composed of ples (b) and finally to thin laminae (a). Locally, other occasional ele-
poorly sorted reefal-derived debris and in-situ growing oncoids with- ments may be associated with element c, the cross-laminated bed,
in a sandy matrix with ooids and quartz grains (Bádenas et al., 2005). namely a lenticular bed of rip-up clast floatstone (d), and a lenticular
These facies gradually thin down dip to cm-thick skeletal- or oncolitic bed with up-slope dipping cross-lamination (e). A basal erosion sur-
rudstones interbedded with lime mudstones and marls. Down dip, face, with several centimetres incision in the underlying sediments,
and detached from the skeletal/oncolitic rudstones, sandy-oolitic usually occurs at the base of the elements c, d and e. These elements
eventites, located in a distal mid-ramp position, range from fine/medium also occur on depositional depressions left by previous internalites
sandstones to pure oolitic grainstones, with quartz grains and ooids usu- (see Section 5).
ally less than 1 mm in size, but locally with coarser grains such as echi- The down-slope dipping cross-laminated lenticular beds (c element),
noids, coral fragments, oncoids, quartzite pebbles and lime mudstone with S- to SE-directed palaeocurrents, are up to 20 cm-thick and extend
rip-up clasts. for less than 5 m in a dip direction (Figs. 5 and 6). Rip-up clasts (lime
B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57 49

undulating surface blanketed by muddy sediment. The c element thins


both up- and down slope and passes into starved ripples (b element;
Fig. 5).
An individual cm-thick train of attached or detached starved rip-
ples, with occasional rip-up clasts, forms the architectural b element
(Figs. 5 and 7). It may locally include or be replaced by a lower inter-
val with parallel- to low-angle lamination (Fig. 8). Extension in dip di-
rection is highly variable, from a few decimetres to several tens of
metres. Lower bounding surfaces are sharp and commonly flat, al-
though concave-up shapes may result from loading deformation. In-
dividual ripples are predominantly asymmetric, with down-slope
dipping internal cross-lamination where preserved, and straight to
sinuous crests with a NE–SW orientation (Fig. 7C). Occasional isolated
ripples with concave base and trough cross-lamination occur in tro-
ughs of starved ripples of previous internalites (Fig. 7A). Ripple tops
are sharp (non-gradational) and blanketed by muddy sediment.
Mud offshoots within sand laminae on the stoss side of ripples are
common (Figs. 7 and 8). Sand thready-fringes on the lee side of rip-
ples interpenetrating lateral muddy sediments also occur (Fig. 9A).
By analogy to “mud offshoots” definition (“mud drapes in ripple fore-
sets”: Shanmugam et al., 1993), we have termed these sand thready-
fringes as “sand offshoots”.
Thin, mm-thick, sand laminae (a element) are frequent within the
lime mudstones beds. Within the a–b–c–b–a down-slope succession
(Fig. 4), the thin laminae represent either the up-slope or the
down-slope element of a starved-ripple train (b element; Fig. 5).
The extent of thin laminae is difficult to assess because they are dis-
continuous, their exposure is not complete and they are frequently
bioturbated. Laminae are parallel to the bedding surfaces of limestone
beds in which they are included. However, the local presence of
down-slope dipping thin-sand laminae (Fig. 10), and thin‐sand off-
shoots in muddy-ripple foresets (i.e., mixed mud-sand ripples;
Fig. 9B) indicates a certain degree of depositional inclination. Thin-
sand laminae underlying internalites are often disrupted as a result
of interpenetration by fine sand- and muddy layers (see Fig. 8).
The rip-up clast floatstone (occasional d element; Fig. 4) consists
of a lenticular bed with thickness ranging from few- to 10 cm and var-
iable down-slope extension, usually less than 5 m (see Figs. 5 and 6).
Rip-up clasts are made from the lime mudstones in which internalites
are intercalated including, in some clasts, fine-sand laminae. Size and
shape of rip-up clasts are highly variable, from mm- to cm-sized sub-
rounded to cm-sized flat and angular clasts. The rip-up clasts are usu-
ally irregularly distributed within the bed and concentrated in
patches within erosive furrows, although they can occur as imbricat-
ed clasts dipping up slope, or forming down-slope dipping accumula-
tions. Amalgamation of rip-up floatstone beds can locally produce
complex internal structures.
A lenticular bed with up-slope dipping cross-lamination (e element;
Fig. 4) is occasionally recognized within the internalites (Fig. 10A).
Thickness is up to 10 cm and extension in dip direction is usually less
Fig. 7. Examples of starved-ripples and thin-sand laminae in internalites (architectural b than 0.5 m. Where preserved, this element is overlain by the down-
and a elements, in white; see Fig. 4). A) Internalites 1 to 5 form a package of amalgamated
slope dipping cross-laminated bed (c element) and separated by a flat
ripples (b elements), locally separated by muddy layers. Ripples have predominant down-
slope dipping cross-lamination. Internalite 3 consists of an isolated ripple with trough to undulated erosion surface (Fig. 10A). The internal structure may lo-
cross-lamination, located in the variably eroded depositional trough between detached cally be complex and formed by shingled stacking of individual beds
ripples of internalite 2. Muddy rip-up clasts occur in internalite 5. Internalites 6 to 8 are separated by thin muddy intervals or by some rip-up clasts (Fig. 10B).
discontinuous thin-sand laminae (a elements) within a muddy interval. Internalite 9 is
an isolated internalite formed by attached ripples with down-slope dipping cross-
lamination and sharp (non-gradational) lower and upper boundaries. B) An isolated inter-
4. Interpretation: generation of architectural elements by breaking
nalite consisting of attached asymmetrical starved ripples (b element), with mud off- internal waves
shoots on the stoss side and sand offshoots on the lee side. C) Top surface of an
internalite consisting of attached asymmetrical ripples (b element) with sinuous crests Breaking of internal waves on sloping surfaces creates episodic and
blanketed by muddy sediment. Note regularity of ripple crest spacing.
repetitive high-turbulent events and consequently erosion and trans-
port of sediments (Southard and Cacchione, 1972; Ribbe and
mudstone fragments) are occasionally abundant, resting parallel to the Holloway, 2001; Apel, 2002; Fringer and Street, 2003; Bogucki et al.,
inclined laminae. Grain-size trends of successive laminae are complex, 2005; Thorpe, 2005; Gilbert et al., 2007; Bourgault et al., 2008;
usually encompassing several fining-, but locally also coarsening- Boegman and Ivey, 2009; Lim et al., 2010). Thorpe and Lemmin
upward lamina sets. The upper boundary, when preserved, is a flat to (1999) have reported that the internal-wave surf zone has some, but
50 B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57

Fig. 8. Examples of internalites containing an interval with parallel- to low-angle lamination (architectural b element; see Fig. 4). A) Internalite (2) interbedded with lime mud-
stones, overlies an erosion surface and includes: a lower interval (2.1) of fine-grained parallel laminae, a coarser middle interval (2.2) with parallel laminae and an upper interval
(2.3) of asymmetrical starved ripples with internal down-slope dipping cross-lamination and mud offshoots on the stoss side. Interval (1) mostly consists of lime mudstones with
thin-sand laminae (architectural a element) that progressively increase upward; in the upper part (1.2), thin‐sand and mud-laminae are disrupted. Interval (3) consists of lime
mudstones with few thin-sand laminae. B) Internalite bed only consisting of a low-angle cross-laminated sandstone, isolated within muddy intervals.

possibly not all, the characteristics of the conventional “surface-wave” architectural elements of c, d or e. Differential pressures on the seafloor
surf zone. Bourgault et al. (2005) have shown direct field observations generated by the internal wave trains, in a similar way to that produced
of an internal solitary wave train impacting a shoaling bottom creating by surface storm waves (e.g., Foda, 2003), might also facilitate erosion.
boluses of turbulent water that move onshore from the breaking zone The pressure distribution on the bottom sediment underneath near-
and the creation of an intermediate layer that transports mixed water breaking or broken waves is transmitted into the porous bed, where it
away from the mixing site. Southard and Cacchione (1972), based on induces a flow and exerts a seepage force on the sediment grains within
laboratory experiments on breaking internal waves over a planar slop- the bed. The horizontal component of this seepage force is related to the
ing bottom, have shown the waves to break abruptly as they shoaled, horizontal pressure gradient of the pressure distribution on the bottom
producing a breaker in the form of a turbulent and rapidly dissipating and is therefore particularly pronounced under the steep front of a
vortex. In these experiments, sediment moved up slope by the breakers, forward-leaning breaking wave and may, if sufficiently large, cause dis-
partly in suspension, and down slope by the compensating return flow, ruption of the bed (Schwab and Lee, 1988; Puig et al., 2004; Chen and
as bedload (swash and backwash phases in Cacchione and Pratson, Hsu, 2005; Madsen and Durham, 2007; Chang, 2011). An excess of
2004). Predominant flow at the bed and, therefore, also net sediment pore-water pressure, and induced shear stress, exceeding the shear re-
transport, was down slope. Based on these experiments, Pomar et al. sistance of the overburden sediment can induce sediment disruption
(2012) proposed a preliminary model on sedimentary dynamics of and liquefaction. Traction currents can transport more sediment once
breaking internal waves for the Ricla eventites, which consider three liquefaction has occurred (Clukey et al., 1985). Strachan and Evans
sedimentary phases: breaking, swash run-up and backwash. (1991) suggested a sediment failure to have been triggered by excess of
Turbulence and vortices induced by breaking internal waves on a pore water pressures that might have been induced by internal wave
muddy/sandy sea floor (i.e., breaking: Fig. 11A) induced erosion, the ex- action; the sediment body is thought to have failed by liquefaction acting
pression of which is the irregular erosive surface underlying thicker retrogressively up slope. Laboratory experiments suggest fluidization of a

Fig. 9. Examples of sand offshoots in internalites and mixed mud-sand ripple lamination. A) Sand offshoots in the lee side of a ripple (architectural b element), passing down slope
into mixed sand-mud lamination. B) Muddy interval with down-slope dipping sand offshoots (mixed sand-mud ripple lamination). Note sand offshoots disappearing both up- and
down slope, and subsequent onlapping muds with parallel thin-sand laminae.
B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57 51

Fig. 10. Examples of internalites including up-slope dipping cross-laminated beds (architectural e element; see Fig. 4). A) Package of internalites 1 to 6. Internalite 1 is a ripple bed (b element)
which passes down slope into thin-sand laminae (a element). Ripples in internalite 1 are not well preserved due to erosion. Internalite 2 is a ripple train (b element) blanketed by a muddy
interval that becomes thicker down slope (right of the picture) and contains low-angle down-slope dipping thin-sand laminae. Internalite 3 is formed by a down-slope dipping cross-
laminated bed (c element) separated from internalite 2 by an erosion surface. Eventite 4 is formed by a lenticular bed with up-dipping cross-lamination (e element). Note lower erosion sur-
face cutting into internalites 1 to 4 and the muddy interval. Internalites 5 and 6 are amalgamated and consist of ripples (b element) passing down slope into down-slope cross-laminated beds
(c element). Internalites 4 and 5 are separated by an erosion surface; however, down slope, there is a muddy interval between them, preserved from erosion. B) Two packages of internalites
(p1 and p2) separated by a muddy interval, but laterally amalgamated (p1+ p2). Package 1 includes internalites 1 and 2 in down-slope shingling geometry; each internalite is formed by a
down-slope dipping cross-laminated bed (c element) passing down slope to a thin ripple and parallel-laminated bed (b element). Package 2 encompasses internalites 3 and 4. Internalite 3
consists of a lower part with a complex up-slope dipping cross-laminated bed (e element), deposited in the depositional depression left by package 1. Internalite 3 includes a thin upper part
with down-slope dipping cross-lamination (c element) passing up slope into ripples (b element).

bed may take place on a very short timescale (around a second) when in- following sedimentation of previous internalites), indicating the com-
duced by a sudden negative pressure loading (Foda et al., 1997). pensating return flow is the predominant agent in the net transport of
On the Ricla ramp, pressure changes as internal waves travelled up sediments (backwash phase: Fig. 11C). Imbricated clasts dipping up
ramp may have created high-shear stress on the seafloor, which in- slope, or forming down-slope dipping accumulations in the floatstone
duced also pressure changes in the pore waters and facilitated erosion bed, also reflect reworking by the backwash flow.
and generation of rip-up clasts. The disrupted intervals underlying The sediment-laden backwash flow led the deposition of the
internalites could then be attributed to convolution of bottom sedi- lenticular bed with down-slope dipping cross-lamination (c ele-
ment due to differential pressure. In the breaker zone, erosion of ment), the starved ripples with down-slope dipping internal cross-
muddy and sandy sediments generated a turbid sediment-laden lamination (b element), and the thin-sand laminae (a element).
water mass. Rip-up clasts may have moved up slope a short distance These sedimentary structures indicate transport of grains was pre-
but they rapidly settled near the generation area (i.e., “dumped dominantly by traction. The main body of the down-slope flow
heap”: Fig. 11B). The variable size and shape of rip-up clasts in the formed the cross-laminated bed (Fig. 11C), when it reached a depres-
floatstone bed (d element) indicate variable abrasion and transport, sion on the seafloor, such as the erosion trough produced in the
but even rounded clasts underwent little transportation from the breaker zone or a bathymetric low created by deposition in the fore-
site of origin (Smith, 1972). The lenticular bed with up-slope dipping front of previous internalites. Thin-sand laminae and starved ripples
cross-lamination (e element) is associated with the swash run-up occurring up slope of the cross-laminated bed accumulated from the
sediment transport. The bimodal character of the sediments facilitat- sand remnants in the queue of the backwash flow. Those down
ed segregation of grains from matrix, so that ooids and quartz grains slope of the cross-laminated bed formed from the sand remnants in
were transported as bedload while the mud component was the backwash flow, after deposition of the main cross-laminated
winnowed in suspension. Deposits generated during up-slope flow body. The existence of fining- and coarsening-upward lamina sets
are only preserved within depressions (mostly erosive depressions within the cross-laminated bed indicates a pulsating regime of the
produced in the breaker zone, but also depositional depressions left tractive backwash flow. Rip-up clasts within the inclined laminae
52 B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57

Fig. 11. Sketches depicting the sediment dynamics during breaking of internal waves on a shoaling surface (modified from Pomar et al., 2012), and the origin of the different ar-
chitectural elements in the Ricla internalites (see Fig. 4). A) Breaking of internal waves creates turbulent vortices and entrainment of sandy oolites and mudstone rip-up clasts.
B) Immediately after breaking, an intraclastic “dumped heap” (d element) and an up-slope dipping cross-laminated unit (e element) may form during the initial swash run-up.
C) The compensating backwash flow transports sediments down slope as bedload, producing thin-sand laminae (a element), starved ripples (b element) and a down-slope dipping
cross-laminated bed (c element). Thin-sand laminae and starved ripples occurring up slope accumulated from the sand remnants in the queue of the backwash flow; the main body
of the down-slope flow formed the cross-laminated bed, when it reached a depression on the seafloor; down-slope starved ripples (locally with mud- and sand offshoots) and thin
lamina formed from the sand remnants in the backwash flow, after deposition of the main c element.

also indicate some degree of reworking of previously generated can be transported in bedload at current velocities that would suffice
clasts, such as erosion by the backwash flow of the “dumped heap” to transport and deposit sand. Floccule ripples are cross-laminated
(e element) settled during the previous run-up flow. Up slope and with geometries very similar to those produced in sandy sediments,
down slope of the cross-laminated bed a “clearer” sediment-laden but have higher water content. Mixed mud-sand ripples and thin-
flow led to the deposition of starved ripples. Local intervals with sand inclined laminae occurring locally in Ricla evidence similar mud
parallel- to low-angle lamination underlying ripples probably were flocculation processes. An alternative to flocculation, in order to explain
deposited with higher flow velocity, with Froude number ≥ 1. the similar hydrodynamic behaviour and avalanching processes for
Ripples include both mud offshoots and sand offshoots interpen- mud and sand, is to consider erosion processes generating sand-size
etrating lateral muddy sediments, which cannot be explained by alter- mud fragments, similar to those obtained in flume experiments by ero-
nating low- and high-energy conditions as proposed for sigmoidal sion on soft-soupy muds (Schieber et al., 2010).
mud offshoots by Shanmugam et al. (1993). These offshoots point to
similar hydrodynamic behaviour and avalanching processes for mud
and sand, and this requires flocculation of mud. Flocculation is affected 5. Stacking of internalites
by particle concentration within the fluid and intensity of turbulence.
Floccule deposition is influenced by turbulence, bed shear stress, sedi- At the studied interval, internalites occur in two clusters encased in
ment concentration, and settling velocity. Swift moving (15–30 cm/s) the muddy mid-ramp succession (Fig. 12A), into which several inter-
muddy suspension is prone to flocculate over a wide range of salinities nalite packages gradually thin and, like fraying fringes, pinch out both
and sediment compositions, and the flocculated material travels as up dip and down dip. These clusters are dm- to m-thick and 1- to more
bedload and forms floccule ripples (Schieber et al., 2007; Schieber and than 2 km of extension in dip direction and, at outcrop scale, correspond
Yawar, 2009). These authors have experimentally shown that muds to the intervals rich in sandy-oolitic eventites (Fig. 2).
B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57 53

Fig. 12. Stacking of internalites. A) Internalites occur in packages, in which the architectural elements are commonly organized in a similar way. A lower interval, with thin-sand
laminae and ripples (a and b elements), is overlain by few cross-laminated beds (c elements) that, in turn, are often overlain by ripples. This vertical succession of architectural
elements is repeatedly related to a down-slope shingled stacking of individual internalites, but up-slope shingling of internalites also occurs. These internalite packages are grouped
in clusters encased in the muddy mid-ramp succession, detached from the shallower facies. Within these clusters, packages gradually thin and, interfingered with muds, pinch out
both up dip and down dip. B) Internalite packages can be, locally, stacked in downlapping configuration (“lobe-like” progradation), infilling the forefront depression of a bulge on
the seafloor created by a pile of previous internalites. C through E) Field sketch and photographs of two internalite packages (p1 and p2) with internal downlapping configuration of
c elements of internalites, with an intervening wedge of isolated ripples (b elements).

Individual internalites concentrate in packages (Fig. 12A). The thi- Fig. 5): intervals consisting of thin-sand laminae and ripples (a and b
ckest part of an internalite package contains several stacked c elements elements) are overlain by few cross-laminated beds (c elements, occa-
(dm-thick and some 10's m extension in dip direction). Mostly, c ele- sionally d and e elements), often in shingling configuration, which in
ments are stacked in down-slope shingling configuration, suggesting turn, can be overlain by ripples, and then by thin-sand laminae. This
each occurs in the forefront residual depression of the c element of a vertical stacking of internalites within a package results from the up-
previous internalite, partially modified by erosion (Fig. 12A, see also and down-slope shift of stacked internalites.
Figs. 6 and 10B). However, physical tracking, in dip direction, of both Unusually thick internalite stacks (up to 70 cm) occur in distal po-
single internalites and internalite packages, indicates that up-slope sition within the Ricla outcrops (Fig. 12B to D). These stacks are “lobe-
shingling also occurs. Within a package, the architectural elements are like”, with lensoidal geometry in the depositional strike direction and
commonly organized vertically in a similar way (Fig. 12A, see also marked sigmoidal shape having b20 m of extension in the dip
54 B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57

direction. Within the “lobe-like” stack, two packages can be differen- would generate perturbations in the pycnocline, as internal waves can be
tiated (p1 and p2 in Fig. 12C, D), based on the stacking patterns. Indi- excited by storms (e.g., Staquet and Sommeria, 2002; Santek and
vidual packages consist of downlapping c elements (down-slope Winguth, 2007).
dipping cross-laminated beds) of successive internalites bounded by
mm- to cm-thick mud laminae, preferentially at the toe (Fig. 12C). 7. Discussion: internalites vs. tempestites
At the toe of package p1 (Fig. 12D), a wedge of mud with oolitic-
sandy ripples (b elements) passes up dip to a thin mud layer and The coarse components of the mid-ramp eventites at Ricla were de-
down dip to muds with thin-sand laminae (a elements). On top of rived from land (quartz grains, occasional quartzite pebbles) and from
package p2, dm-wide hummocks occur and bioturbation has locally the shallowest portion of the ramp (ooids, occasional bioclasts) by differ-
destroyed the internal lamination. ent processes, including hyperpycnal flows or storm-induced currents
(Bádenas and Aurell, 2001). The analysis of the architectural elements
6. Interpretation: fluctuations of the pycnocline and the stacking pattern of these eventites indicate internal-wave
reworking and certain down-dip transport of sediments existing in mid
Up- and down-slope shift of internalite packages points to the up- ramp setting, in this case of previous resedimented deposits, and provide
and down-slope migration of the breaker zone, which can in turn be a facies model for deposits originated from breaking internal waves
related to the fluctuation of the pycnocline. Bedding configuration (Fig. 4). In mid-ramp settings, tempestites and/or internalites can occur
and stacking patterns in “lobe-like” bodies result from the infill, by interbedded within the mud-dominated successions. Some key criteria
successive internalites, of a larger depositional depression at the fore- can be used to differentiate tempestites and internalites (Table 1):
front of a package characterized by vertical stacking of internalites
(Fig. 12B). Vertical stacking of internalites would elevate the source (1) Stacking of eventites. Internalites concentrate in discrete inter-
point in the forefront of the package that would allow the successive vals in distal mid-ramp position detached from the shallow-
internalites to stack in downlapping configuration. This depositional water successions, gradually thinning-out to disappear both up
mechanism is coherent with the location of the “lobe-like” bodies in dip and down dip (Pomar et al., 2012; Fig. 2). Individual inter-
the distal position of the internalites accumulation. The two studied nalite beds stack in up- and down-dip shingling configuration,
clusters of grainy internalites at Ricla encased in the muddy mid- reflecting the up- and down-dip migration of the pycnocline
ramp facies are thought to reflect periods with enhanced sediment (and/or variable amplitude of the breaking wave). In Ricla, the
supply from shallow-water settings and/or variations of water strati- fact that eventite-rich intervals are located in the lower part of
fication affecting the energy of the internal waves. the TST, with a progressive landward migration, is coherent with
The origin of the pycnocline and ocean circulation in the studied the progressive landward shift of the pycnocline intersecting the
Kimmeridgian carbonate ramp was probably dominated by halothermal seafloor during transgression. In contrast, tempestites would in-
(primarily salinity-driven) conditions due to “greenhouse” climate, in crease to the shoreface and stack in thickening- and coarsening-
contrast to modern open oceans where the pycnocline is mainly induced upwards succession, reflecting the shallowing-upward character
by temperature, and secondarily by salinity gradients (e.g., Kennett and of the surface-storm beds accumulation (e.g., Burchette and
Stott, 1991; Pak and Miller, 1992; Nunes and Norris, 2006). In interior Wright, 1992).
seas, the seasonal pycnocline is commonly strongly influenced by river- (2) Sedimentary processes. Both tempestites and internalites en-
ine discharge. Low-salinity stratified water masses have been described compass an erosional phase and a depositional phase during
in the northern Indian Ocean (Sri Lanka east coast), caused by the high waning of turbulence.
seasonally varying precipitation and river runoff (Santek and Winguth, a) Internalites are associated to erosion in the mid-ramp settings
2007). Therefore, it is plausible to suggest for the Ricla succession that (breaker of the internal waves), whereas little or no erosion oc-
the clusters of internalites accumulated during periods of increased river- curs in shallower settings. Contrarily, erosion processes due to
ine discharge that resulted in both sharpening of the pycnocline and storm waves dominate in shallower areas, so that “cannibal-
hence internal waves as well as increased sediment delivery (quarz ism” and/or amalgamation increases inshore in tempestites
grains, quartzite pebbles) from the emerged areas of the Ebro and Iberian (Aigner, 1985). Rip-up clasts in internalites are derived from
massifs (Fig. 1). The possible effects of storm activity on sediment supply erosion in mid-ramp settings (breaker zone), but in tem-
and generation of internal waves in the Kimmeridgian carbonate ramp pestites can be derived from shallower settings.
can also be considered. Storms would increase the input of land- b) Internalite beds are deposited from tractive flows. Although
derived grains accumulated in shallower areas and shallow-water coarse predominantly down dip, internalites reflect both up- and
grains (ooids, skeletal grains) into mid-ramp setting. In addition, storms down-dip transport processes (run-up and backwash flows).

Table 1
Key differences between the Ricla internalites and tempestites.

Internalites generated by breaking internal waves in mid-ramp settings Tempestites

Stacking of eventites Concentrated in discrete intervals in distal mid-ramp position, detached from Tempestite beds increasing to the shoreface, attached to shallow-
the shallow-water successions water successions
Gradually thins-out to disappear both up dip and down dip Up-slope increase of tempestites; down-slope decrease of
tempestites
Bed stacking in shingling configuration, reflecting up- and down-slope mi- Thickening- and coarsening-upwards successions reflecting the
gration of the pycnocline and/or variable amplitude of the breaking wave shallowing-upward character of storm beds accumulation
Sedimentary Erosional Erosion localized in mid ramp position, associated to the breaker of the Predominant erosion in shallow areas
processes phase internal waves
Little or no erosion towards the shallower areas Increase of erosion (“cannibalism” and/or amalgamation) toward
the shallower areas
Rip-up clast from the internal-wave breaker zone (mid ramp) Rip-up clasts, if any, can be from shallower water
Depositional Derived from tractive flow Derived from density and oscillatory flows
phase Up-dip and down-dip transport, although preferentially down dip Down-dip transport
Down-dip organization of the architectural elements, as product of bedload Vertical grading of grain size and sedimentary structures
and the existence of depositional or erosional depressions indicating decreasing flow competence
B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57 55

Recurrent organization of the architectural elements in dip di- landward migration of the pycnocline during early transgression. The
rection (a–b–c–b–a; Fig. 4) responds to bedload differences in absence of the architectural element c could be related to small amount
the different parts of the tractive flow, and to the existence of of sand-grain sediments. Lower riverine discharge might have con-
depressions (depositional or erosional) on the seafloor. Cross- trolled decreased coarse sediment delivery to mid-ramp settings (via
bedded elements formed by the main tractive flow and accu- hyperpycnal flows or storm-induced currents).
mulated preferentially on seafloor depressions. Starved ripples
and thin‐sand laminae are related to the sand remnants in the
queue of the backwash flow. In contrast, tempestites derive 8. Conclusions
from storm-generated density flows travelling down slope,
combined with storm-wave action. The decreasing compe- Despite the ubiquity of internal waves in oceans and lakes, internal
tence of the density flow and the storm-wave relaxation is wave deposits (internalites) are seldom recognized in the sedimentary
reflected in the upward grading of grain size and associated record. For shelf and ramp deposits, they require differentiation from
sedimentary structures. tempestites, and from turbidites in deeper settings, with which they
share some common features of a basal erosion surface and a subse-
The defined facies model for the Ricla internalites includes all the ar- quent depositional phase during the waning of turbulent conditions.
chitectural elements (a–b–c–b–a) observed at the studied interval. Criteria for recognition of internalites, still to be fully developed,
However, other sandy-oolitic eventites in Ricla do not include the can be obtained by analyzing the spatial organization of architectural
down-slope dipping cross-laminated bed (element c), and only thin- elements and associated sedimentary structures within eventite beds.
sand laminae (element a) and cm-thick ripple beds (element b) are pre- A significant portion of eventites of Upper Kimmeridgian Ricla succes-
sent. This is the case of most of the eventites found at the lowermost sion can be reinterpreted according to the processes associated to the
part of the studied TST in Ricla (Fig. 13). Thin-sand laminae and cm- break of internal waves on a sloping sea floor with bimodal sediment
thick ripple beds are related in dip direction, indicating that they may type, sand and mud.
correspond to “a–b–a” internalite type. In addition, they concentrate Breaking of internal waves on sloping surfaces creates episodic high-
in distal mid-ramp position detached from shallower facies, with a turbulent events that remobilize sediment at the depth at which a
gradual up-dip and down-dip decrease in number. The shingling config- pycnocline intersects the sea floor. Internalites are concentrated in
uration of these intervals rich in “a–b–a” internalites is coherent with a discrete intervals in distal mid-ramp positions that are detached from

Fig. 13. Facies architecture based in the detailed logging and correlation across the lowermost part of the TST of the Kim2 Sequence at Ricla (see Fig. 2 for location; updated from
Aurell et al., 2011). The number of sandy-oolitic cm-thick eventites in correlatable intervals is higher in distal mid-ramp setting, detached from shallower facies, indicating that they
may correspond to internalites.
56 B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57

coeval shallow-water successions. They do not have the coarsening- Cacchione, D., Wunsch, C., 1974. Experimental study of internal waves over a slope.
Journal of Fluid Mechanics 66, 223–239.
thickening upward trend typical of storm-generated successions and Chang, F.Q., 2011. Shear failure of seabed induced by wave at the Chengdao Sea, the
they gradually thin-out to disappear both up- and down dip, inter- Yellow River Estuary. Advanced Materials Research 243–249, 4701–4704.
fingering with muddy mid-ramp lithofacies and little or no erosion to- Chen, C.Y., Hsu, J.R.C., 2005. Interaction between internal waves and a permeable sea-
bed. Ocean Engineering 32, 587–621.
wards shallower areas. Within internalites, diverse architectural Clukey, E.C., Kulhawy, F.H., Liu, P.L.-F., Tate, G.B., 1985. The impact of wave loads and
elements can be related to breaking internal waves and associated turbu- pore-water pressure generation on initiation of sediment transport. Geo-Marine
lent flows: internal wave breaker zone (basal erosion surface), “dumped Letters 5, 177–183.
Dercourt, J., Ricou, L.E., Vrielynck, B., (Eds.), 1993. Atlas Tethys Palaeoenvironmental
heap” in the surf zone (mud-chip floatstone bed: element d), up-dip Maps. BEICIP-FRANLAB, Gauthier-Vollars, Paris, 260 pp., 14 maps.
swash (up-slope dipping cross-laminated bed: element e) and backwash Emery, K., Gunnerson, C.G., 1973. Internal swash and surf. Proceedings of the Natural
flow (down-slope dipping cross-laminated bed: element c, isolated star- Academy of Sciences of the United States of America 70, 2379–2380.
Foda, M.A., 2003. Role of wave pressure in bedload sediment transport. Journal of Wa-
ved ripples: element b, and thin-sand laminae: element a).
terway, Port, Coastal, and Ocean Engineering 129, 243–249.
Individual internalites stack in packages, within which they are Foda, M.A., Hill, D.F., DeNeale, P.L., Huang, C.M., 1997. Fluidization response of sedi-
up- and down-slope shifted, often in shingling configuration. This ment bed to rapidly falling water surface. Journal of Waterway, Port, Coastal, and
staking pattern is suggested to reflect the landward and seaward mi- Ocean Engineering 123, 261–265.
Fringer, O.B., Street, R.L., 2003. The dynamics of breaking progressive interfacial waves.
gration of the internal waves breaker zone (fluctuation of the Journal of Fluid Mechanics 494, 319–353.
pycnocline). Clusters of internalite packages pinching out both up Gilbert, R.W., Zedler, E.A., Grilli, S.T., Street, R.L., 2007. Progress on nonlinear wave
dip and down dip and encased in the muddy mid-ramp facies reflect forced sediment transport simulation. IEEE Journal of Oceanic Engineering 32
(1), 236–248.
changes in sediment supply and/or variations in water stratification Global-Ocean-Associates, 2004. An atlas of internal solitary-like waves and their prop-
affecting the energy of the internal waves. erties. In: Jackson, C.R. (Ed.), Global Ocean Associates. Prepared for Office of Naval
Research – Code 322 PO, Alexandria, VA, p. 560. http://www.internalwaveatlas.
com/Atlas2_index.html.
Acknowledgements Kennett, J.P., Stott, L.D., 1991. Abrupt deep-sea warming, palaeoceanographic changes
and benthic extinctions at the end of the Palaeocene. Nature 353, 225–229.
LaFond, E.C., 1966. Internal waves. In: Fairbridge, R.W. (Ed.), The Encyclopedia of
This work is a contribution to the research projects Grupo Rec- Oceanography. Reinhold, New York, pp. 402–408.
onstrucciones Paleogeográficas (Aragón Government) and CGL2011- Lim, K., Ivey, G.N., Jones, N.L., 2010. Experiments on the generation of internal waves
over continental shelf topography. Journal of Fluid Mechanics 663, 385–400.
24546 (to B.B. and M.A.) and CGL2009-13254 (to L.P. and M.M.). We Madsen, O.S., Durham, W.M., 2007. Pressure-induced subsurface sediment transport in
are grateful to the editor Brian Jones, and to Paul (Mitch) Harris and an the surf zone. Proceedings Coastal Sediments ’07 Conference, 1, pp. 82–95.
anonymous reviewer for comments and suggestions on the original ver- Marsaglia, K.M., Klein, G.D., 1983. The paleogeography of Paleozoic and Mesozoic storm
depositional systems. Journal of Geology 91, 117–142.
sion of the manuscript. Munk, W., 1981. Internal waves and small-scale processes. In: Warren, B.A., Wunsch, C.
(Eds.), Evolution of Physical Oceanography. MIT Press, Cambridge, pp. 264–291.
Nash, J.D., Moum, J.N., 2005. River plumes as a source of large-amplitude internal
References waves in the coastal ocean. Nature 437, 400–403.
Nunes, F., Norris, R.D., 2006. Abrupt reversal in ocean overturning during the Pal-
Aigner, T., 1985. Storm depositional systems, dynamic stratigraphy in modern and an- aeocene/Eocene warm period. Nature 439, 60–63.
cient shallow-marine sequences. In: Friedman, G.C., Neugebauer, H.J., Seilacher, A. Osete, M.L., Gómez, J.J., Pavón-Carrasco, F.J., Villalaín, J.J., Palencia-Ortas, A., Ruiz-
(Eds.), Lecture Notes in Earth Sciences. Springer, Berlin, pp. 1–171. Martínez, V.C., Heller, F., 2011. The evolution of Iberia during the Jurassic from
Apel, J.R., 2002. Oceanic internal waves and solitons. In: Jackson, C.R. (Ed.), An atlas of palaeomagnetic data. Tectonophysics 502, 105–120.
internal solitary-like waves and their properties. Global Ocean Associates. Prepared Ostrovsky, L.A., Stepanyants, Y.A., 1989. Do internal solitons exist in the ocean? Re-
for Office of Naval Research – Code 322 PO, Alexandria, VA, pp. 1–40. http://www. views of Geophysics 27, 293–310.
internalwaveatlas.com/Atlas2_PDF/IWAtlas_Pg001_Background&Theory.pdf. Pak, D.K., Miller, K.G., 1992. Paleocene to Eocene benthic foraminiferal isotopes and as-
Apel, J.R., Ostrovsky, L.A., Stepanyants, Y., Lynch, J.F., 2007. Internal solitons in the semblages; implications for deepwater circulation. Paleoceanography 7, 405–422.
ocean and their effect on underwater sound. The Journal of the Acoustical Society Pomar, L., Morsilli, M., Hallock, P., Bádenas, B., 2012. Internal waves, an under-explored source
of America 121, 695–722. of turbulence events in the sedimentary record. Earth-Science Reviews 111, 56–81.
Aurell, M., Bádenas, B., Ipas, J., Ramajo, J., 2010. Sedmentary evolution o an Upper Juras- Price, G.D., Sellwood, B.W., Valdes, P.J., 1995. Sedimentological evaluation of general
sic carbonate ramp (Iberian Basin, NE Spain). In: van Buchem, F., Gerdes, K., circulation model simulations for the greenhouse Earth: Cretaceous and Jurassic
Esteban, M. (Eds.), Reference models of Mesozoic and Cenozoic carbonate systems case studies. Sedimentary Geology 100, 159–180.
in Europe and the Middle East — stratigraphy and diagenesis: Geological society of Puig, P., Ogston, A.S., Mullenbach, B.L., Nittrouer, C.A., Parsons, J.D., Sternberg, R.W.,
London, special publication, 329, pp. 87–109. 2004. Storm-induced sediment gravity flows at the head of the Eel submarine can-
Aurell, M., Bádenas, B., Pomar, L. Colombié C., Caline, B., Ipas, J., Martínez, V., San yon, northern California margin. Journal of Geophysical Research 109, C03019.
Miguel, G., Al-Nazghah, M.H., 2011. The Kimmeridgian-Lower Tithonian carbonate Quaresma, L.S., Vitorino, J., Oliveira, A., da Silva, J., 2007. Evidence of sediment
ramps (Upper Jurassic, NE Spain): architecture, facies distribution and cyclo- resuspension by nonlinear internal waves on the western Portuguese mid-shelf.
stratigraphy. Geo-Guías 8, 45–86. Marine Geology 246, 123–143.
Bádenas, B., Aurell, M., 2001. Proximal-distal facies relationships and sedimentary pro- Ribbe, J., Holloway, P.E., 2001. A model of suspended sediment transport by internal
cesses in a storm dominated carbonate ramp (Kimmeridgian, northern Iberian tides. Continental Shelf Research 21, 395–422.
basin). Sedimentary Geology 139, 319–340. Salas, R., Casas, A., 1993. Mesozoic extensional tectonics, stratigraphy and crustal evolu-
Bádenas, B., Aurell, M., 2008. Kimmeridgian epeiric sea deposits of northeast Spain: tion during the Alpine cycle of the eastern Iberian basin. Tectonophysics 228, 33–55.
sedimentary dynamics of a storm-dominated carbonate ramp. In: Holmden, C., Santek, D.A., Winguth, A., 2007. A satellite view of internal waves induced by the Indian
Pratt, B. (Eds.), Dynamics of Epeiric Seas: Geological Association of Canada, special Ocean tsunami. International Journal of Remote Sensing 28, 2927–2936.
paper, 48, pp. 55–71. Schieber, J., Yawar, Z., 2009. A new twist on mud deposition — mud ripples in experi-
Bádenas, B., Aurell, M., Gröcke, D.R., 2005. Facies analysis and correlation of high-order ment and rock record. The Sedimentary Record 7, 4–8.
sequences in middle–outer ramp successions: variations in exported carbonate on Schieber, J., Southard, J., Thaisen, K., 2007. Accretion of mudstone beds from migrating
basin-wide δ13Ccarb (Kimmeridgian, NE Spain). Sedimentology 52, 1253–1275. floccule ripples. Science 318 (5857), 1760–1763.
Boegman, L., Ivey, G.N., 2009. Flow separation and resuspension beneath shoaling Schieber, J., Southard, J., Schimmelmann, A., 2010. Lenticular shale fabrics resulting
nonlinear internal waves. Journal of Geophysical Research 114, C02018. from intermittent erosion of water-rich muds—interpreting the rock record in
Bogucki, D.J., Redekopp, L.G., Barth, J., 2005. Internal solitary waves in the Coastal the light of recent flume experiments. Journal of Sedimentary Research 80,
Mixing and Optics 1996 experiment: multimodal structure and resuspension. Jour- 119–128.
nal of Physical Oceanography 110, C02024. Schwab, W.C., Lee, H.J., 1988. Causes of two slope-failure types in continental-shelf sediment,
Bourgault, D., Kelley, D.E., Galbraith, P.S., 2005. Interfacial solitary wave run-up in the northeastern Gulf of Alaska. Journal of Sedimentary Research 58, 1–11.
St. Lawrence Estuary. Journal of Marine Research 62, 1001–1015. Seilacher, A., 1982. General remarks about event deposits. In: Einsele, G., Seilacher, A.
Bourgault, D., Kelley, D.E., Galbraith, P.S., 2008. Turbulence and boluses on an internal (Eds.), Cyclic and Event Stratification. Springer-Verlag, New York, pp. 161–174.
beach. Journal of Marine Research 66, 563–588. Seilacher, A., 1991. Events and their signatures — an overview. In: Einsele, G., Seilacher,
Brown, J., Colling, A., Park, D., Phillips, J., Rothery, D., Wright, J.D., 1989. Seawater: its A. (Eds.), Cycles and Events in Stratigraphy. Springer-Verlag, pp. 221–226.
composition, properties and behavior. Pergamon Press and The Open University, Shanmugam, G., Spalding, T.D., Rofheart, D.H., 1993. Traction structures in deep-
Oxford. 238 pp. marine, bottom-current-reworked sands in the Pliocene and Pleistocene, Gulf of
Burchette, T.P., Wright, V.P., 1992. Carbonate ramp depositional systems. Sedimentary Mexico. Geology 21 (10), 929–932.
Geology 79, 3–57. Small, J., Martin, J., 2002. The generation of non-linear internal waves in the Gulf of
Cacchione, D., Pratson, L., 2004. Internal tides. American Scientist 2, 130–137. Oman. Continental Shelf Research 22, 1153–1182.
B. Bádenas et al. / Sedimentary Geology 271-272 (2012) 44–57 57

Smith, N.D., 1972. Flume experiments on the durability of mud clasts. Journal of Sedi- Thorpe, S.A., 2005. The Turbulent Ocean. Cambridge University Press, Cambridge. 447
mentary Research 42, 378–383. pp.
Southard, J.B., Cacchione, D.A., 1972. Experiments on bottom sediment movement by break- Thorpe, S.A., Lemmin, U., 1999. Internal waves and temperature fronts on slopes.
ing internal waves. In: Swift, D.J., Duane, D.B., Pilkey, O.H. (Eds.), Shelf Sediment Trans- Annales Geophysicae 17, 1227–1234.
port: Process and Pattern. Hutchinson & Ross, Stroudsburg, Pa., pp. 83–97. Dowden. Wolanski, E., Colin, P., Naithani, J., Deleersnijder, E., Golbuu, Y., 2004. Large amplitude,
Staquet, C., Sommeria, J., 2002. Internal gravity waves: from instabilities to turbulence. leaky, island-generated internal waves around Palau, Micronesia. Estuarine, Coastal
Annual Review of Fluid Mechanics 34, 559–593. and Shelf Science 60, 705–716.
Strachan, P., Evans, D., 1991. A local deep-water sediment failure on the NW slope of
the UK. Scottish Journal of Geology 27, 107–111.

You might also like