Tag Archives: geostrophic flow

Storm surges and tempestites

Facebooktwitterlinkedininstagram
This EUMETSAT / Japanese Meteorological Agency (January 29, 2015) composite image of two tropical cyclones in the Indian Ocean between Madagascar and Australia is a good illustration of the relative sizes of these storms. These two (Diamondra and Eunice) are about 1,500 kilometres apart.

This EUMETSAT / Japanese Meteorological Agency (January 29, 2015) composite image of two tropical cyclones in the Indian Ocean between Madagascar and Australia is a good illustration of the relative sizes of these storms. These two (Diamondra and Eunice) are about 1,500 kilometres apart.

This is the second of three posts on tempestites:

1 Storms and storm surges: Forces at play

3 Evolving tempestite lithofacies models

 

If there is one suite of sedimentary structures that focuses our attention on shelf or platform dynamics, it is hummocky (HCS) and swaley cross-stratification (SWS). You might be forgiven for thinking that no shelf succession is complete without either or both of these bedforms.

There is consensus that this bedform duo is the product of storms, based on a substantial outcrop database, theoretical considerations (e.g., Myrow and Southard, 1996), and flow experiments (e.g., Dumas and Arnott, 2006), but only a sparse database of bottom-current storm-flow characteristics. They are part of a suite of sedimentary deposits and bedforms called tempestites. They represent depositional events that are distinct from fairweather conditions. There is general agreement that HCS deposits are the product of oscillatory storm-wave orbital bottom currents or bottom-hugging density currents (e.g., turbidity currents) that have been modified by storm waves (i.e., combined flow). Despite these overarching interpretations, the depositional processes responsible for these structures remain a matter of debate (e.g., Duke et al., 1991; Myrow, 2020).

The debate is centred around two important questions:

  1. How is sediment delivered and distributed across shelves and deltas during relatively short-lived storm events. – a question that requires us to evaluate the relative roles of downslope (shore-normal), unidirectional, bottom-hugging currents versus geostrophic currents that flow parallel or obliquely to shorelines?
  2. At the sediment-water interface, what are the relative contributions of storm-wave orbitals and storm-induced unidirectional currents to the development of bedforms like HCS (excluding ambient tidal currents and fairweather along-shore currents)?

 

Boundary layers in stormy waters

Myrow and Southard’s 1996 elegant description of storm wave and current dynamics is as relevant now as it was nearly three decades ago. An important component of their model deals with the delivery of sediment across a shelf. Ancient tempestite deposits can be 10s of centimetres and even metres thick, which means that substantial volumes of sediment were moved across a shelf during relatively short-lived events. How this is facilitated depends on how energy and shear stresses are partitioned through the water column and at the sediment-water interface, particularly in the bottom boundary layer of shelf and delta water masses.

The diagram below is modified (slightly) from Myrow and Southard (their Figure 1). It shows three distinct layers in waters at average shelf depths within the core geostrophic flow. The location of wave base across a shelf or delta platform is also an important boundary because it limits the depth of wave interaction at the sediment interface.

Myrow and Southard's boundary layer diagram, their Figure 1

  1. A surface boundary layer that is maintained by excessive turbulence and mixing. Fine sediment in suspension, for example that derived from wave reworking of the shoreface, or from hypopycnal flows, may reside in this layer for some time and may be deposited long after a storm has ended (which begs the question – should it be included in the definition of tempestites?).
  2. An inviscid middle layer in which viscous forces are small (i.e., Reynolds numbers are relatively high because inertial flow dominates).
  3. A bottom layer where the efficiency of sediment transport depends on two overlapping processes; this layer is where all the sedimentological action takes place:
    • A relatively thin layer (centimetres) at the boundary between wave orbitals and the sea floor, where shear stresses are directed alternately onshore and offshore and, because of wave refraction, the motion is normal to the shoreline.
    • A thicker layer of unidirectional currents (a few metres) that are directed offshore during intense storms. These bottom currents are the result of coastal setup. They overlap the wave-orbital motion to produce combined flow in the lower layer.

 

The critical forces at play

The forces that influence sediment transport and deposition in the bottom layer include:

  1. Offshore directed hydraulic pressures, the gradient of which depends on the magnitude of the coastal setup. Coastal setup refers to the elevation of sea level at the coast, where water masses pile up because of wind shear and Ekman Veering of currents that flow at right angles to the wind direction (deflecting to the right in the northern hemisphere). The magnitude of the setup depends on storm duration, wind direction and strength, wave fetch, and the amplifying effects of coastal geomorphology. A seaward pressure gradient develops because the elevated water mass is gravitationally unstable and will tend to dissipate as the storm wanes. Storm cells are also associated with low atmospheric pressures that cause sea levels to rise, but this component only accounts for about 5% of the coastal setup.
  2. Coriolis forces act at 90o to the wind, deflecting currents to the right in the northern hemisphere. As a storm develops, and depending on the wind direction, Ekman veering will push water masses shoreward and contribute to the coastal setup. Coriolis deflection of the seaward return flow from coastal setup produces isobath-parallel (shore-parallel) geostrophic flow.
  3. Friction forces exist where wave orbitals interact with the substrate. These forces increase from storm wave-base towards the shoreline. Wave orbitals tend to be more symmetrical in deep water, and increasingly elliptical towards the shore (flattened approximately horizontally). Orbital velocities also change with the passage of a storm, as wave heights build over time, but also because winds change direction as cyclonic wind flow passes landward. The magnitude of wave orbital velocity (and therefore shear stress) depends primarily on wave amplitude and period, that also depends on storm duration and fetch.
  4. Sediment suspended in the water column produces what Myrow and Southard call excess weight forces. The concentration of suspended sediment tends to be greatest near the shoreline, decreasing seaward with distance and water depth.  Thus, these forces tend to act downslope (seaward). Some of this sediment may be reworked from the sea floor, particularly in the surf zone. However, much sediment is also introduced by rivers and deltas as hypopycnal flows and hyperpycnal flows.

The proportion that each type of force or process influences bedforms and their lateral extent; the thickness of tempestites will vary from place to place and from one storm to the next depending on storm severity, the direction of storm approach along a coast, and coastal geomorphology (e.g., Myrow, 2020, op cit.). Thus, the range of possible lithofacies will also vary from one event to another. Geomorphic systems other than open-ocean shelves, such as large deltas and high-volume rivers, will impact the volume of suspension load and bedload sediment released to adjacent shelves. The response of storm surge encroachment over a delta will also be quite different to that over shelves where river input is minimal; in this case, marine processes will compete with fluvial flood-related processes – recent examples are Bhattacharya et al., 2020, looking at North America delta systems; Vaucher et al., 2023, examining a Late Pleistocene flood delta in Taiwan).

 

Shore-normal flows or geostrophic flows?

The term “shore-normal” here means bottom-hugging flow normal to shoreline: geostrophic flows parallel isobaths or may be oblique where deflected by seafloor topography. The problems associated with identifying shore-normal or shore-parallel bottom current flows as depositional modes for tempestite deposition is nicely encapsulated by two early publications: Leckie and Krystinik (1989), and Snedden and Nummedal (1991).

 

Shore-normal flows

Leckie and Krystinik proposed that combined shore-normal density currents (turbidity currents) and wave orbital flows were responsible for the majority of HCS-bearing beds in Cretaceous shelf deposits (Western Canada foreland basin), based on sole mark and parting lineation trends, together with offshore trends in stratigraphic thickness and grain size. Numerous studies have since shown the importance of shore-normal, wave-modified turbidity currents (i.e., combined flow), not only in ancient shelf settings but also across wave-dominated deltas and prodelta slopes where hyperpycnal flows are commonly generated, for example Myrow et al., (2002, Cambrian Shelf – PDF available), Jelby et al., (2019, Cretaceous delta ramp).

Density current flows may be triggered by excess weight forces (noted above), including those generated by plunging hyperpycnal flows, or by resuspension of sediment by storm waves; some excellent examples have been documented in Spitsbergen Cretaceous rocks (Jelby et al., 2019. Op cit.). Zavala (2020, open access) has incorporated HCS in some hyperpycnite lithofacies models.

 

Geostrophic flows

Storm-generated geostrophic currents develop when Coriolis forces deflect the return flow from coastal setup (the coastal setup pressure gradient is oriented seaward). The interaction of storm wave orbitals with these shore-parallel currents can produce combined flows that parallel or are oblique to the shoreline during the waning stages of storm surging. Snedden and Nummedal (1991, PDF available), while not dismissing the importance of shore-normal processes, made a concerted plug for shore-parallel or oblique geostrophic flows, based on mapping of a distinctive, graded, tempestite bed deposited by Hurricane Carla (1961). Their interpretation was based on measured shore-parallel isopachs and grain size distributions of a storm-deposited sand bed, observed wind forces, and modelled current directions. Likewise, Swift et al., (2006) posit shore parallel storm generated currents to explain bedform, paleocurrent, and sediment distribution in Cretaceous Book Cliff strata (Utah).

Geostrophic flows also interact with waves producing a range of combined-flow bedforms including ripples,  dunes, and asymmetrical HCS. However,  geostrophic currents are not restricted to the shoreface – they can extend across the shelf beyond storm wave-base where there is no interaction with wave orbitals in the lower boundary layer. In this case, all sediment transport is traction dominated, and bedforms like current ripples, dunes, and lower flow regime plane-bed lamination will form. This suite of bedforms will be coeval with the HCS-dominated tempestites although distinguishing them from other fairweather deposits might be difficult.

Geostrophic flow across continental shelves and deltas tends to be either parallel or oblique to shore, the latter depending on current deflection by shallow bathymetry, and possibly temperature and salinity barriers. Mapped geostrophic flows in modern seas commonly indicate significant departures from shore-parallelism, particularly in deeper waters beyond the shelf break. The example from East Sea (Japan Sea) shows flow vectors (measured, and calculated from sea level elevations) that define several eddies that coincide with seawater temperature anomalies. The eddies are located beyond the shelf-break, but they may complicate flow dynamics across the shelf during storms.

Surface geostrophic currents (black vectors) and sea surface heights (SSH) determined from VISO satellite altimeter data and coastal sea level data for the Sea of Japan between Korean Peninsula and Honshu Island. Sea surface heights are in metres relative to Global Mean Sea Level. The 500 m isobath has been added (black dashed line). Geostrophic currents are mostly coast-parallel across the Honshu shelf but develop 200-300 km diameter eddies farther offshore. UBIM is an ocean buoy. EKWC = East Korea Warm Current. SE = Sokcho Eddy. Image credit: Modified slightly from Son et al., 2014. Honshu I. License CC BY 3.0

Surface geostrophic currents (black vectors) and sea surface heights (SSH) determined from VISO satellite altimeter data and coastal sea level data for the Sea of Japan between Korean Peninsula and Honshu Island. Sea surface heights are in metres relative to Global Mean Sea Level. The 500 m isobath has been added (black dashed line). Geostrophic currents are mostly coast-parallel across the Honshu shelf but develop 200-300 km diameter eddies farther offshore. UBIM is an ocean buoy. EKWC = East Korea Warm Current. SE = Sokcho Eddy. Image credit: Modified slightly from Son et al., 2014. Honshu I. License CC BY 3.0

 

 

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Hummocky & swaley cross-stratification

Facebooktwitterlinkedininstagram
HCS mounds are exposed on bedding in the lower part of this shallowing-upward shelf parasequence (arrows). Mound amplitudes are 12-15 cm, and spacing is about 3 m. Lithofacies in the overlying beds include ripples, 2D dunes, and interbedded mudstone. The Jacob's staff is 1.5 m long. Eocene, Ellesmere Island.

HCS mounds are exposed on bedding in the lower part of this shallowing-upward shelf parasequence (arrows). Mound amplitudes are 12-15 cm, and spacing is about 3 m. Lithofacies in the overlying beds include ripples, 2D dunes, and interbedded mudstone. The Jacob’s staff is 1.5 m long. Eocene, Ellesmere Island.

HCS and SCS – useful paleobathymetric indicators

 

Use this link to read the introduction to the lithofacies series.

 

General occurrence

Hummocky cross-stratification and its close cousin swaley bedding (SCS) are two bedforms that offer sedimentologists a modicum of hope in their search for paleoenvironmental indicators. HCS (the commonly used acronym) forms on the lower shoreface – shelf during storms where wave orbitals combine with bottom-hugging flows to mould 3D bedforms. Thus, the presence of HCS in the rock record provides a reasonably reliable paleobathymetric indicator, forming somewhere between fairweather wave-base and storm wave-base.

 

External structure

HCS and SCS are conjoined bedforms; the hummocks are low amplitude mounds and the adjacent swales dish-shaped depressions. Both are approximately equidimensional to slightly asymmetric in map view. They are three-dimensional bedforms that require at least two cross-sectional profiles to describe their geometry. Hummock and swale amplitudes are measured in centimetres. The separation of hummocks, measured between mound apices, usually ranges from 10s of centimetres to 5-6 m. The hummocks differ from most other 3D bedforms in that there are no lee or stoss faces.

Parting lineation is common on exposed hummock surfaces.

 

Bedding exposure of hummocks and swales (yellow arrows). Hummock amplitude is 15-20 cm and spacing 3-4 m. An underlying swale is indicated by the red arrow. The HCS unit is underlain by a thin pebbly, normally graded sandstone (at the level of the hammer). The HCS-SCS are in gradational contact with the graded sandstone that is inferred to have been a turbulent, high suspension-load underflow. Mid-Jurassic Bowser Basin, northern British Columbia.

Bedding exposure of hummocks and swales (yellow arrows). Hummock amplitude is 15-20 cm and spacing 3-4 m. An underlying swale is indicated by the red arrow. The HCS unit is underlain by a thin pebbly, normally graded sandstone (at the level of the hammer). The HCS-SCS are in gradational contact with the graded sandstone that is inferred to have been a turbulent, high suspension-load underflow. Mid-Jurassic Bowser Basin, northern British Columbia.

Internal structure

Cross-section of two HCS-bearing sandstone beds (arrows). Laminae in the lower bed show subtle down-dip thickening (towards the adjacent trough). The amplitude of both mounds is 5-8 cm. The lower bed is overlain by mudstone and sandstone containing asymmetric current ripples (R). Campanian Monster Fm., Yukon.

Left: Cross-section of two HCS-bearing sandstone beds (arrows). Laminae in the lower bed show subtle down-dip thickening (towards the adjacent trough). The amplitude of both mounds is 5-8 cm. The lower bed is overlain by mudstone and sandstone containing asymmetric current ripples (R). Campanian Monster Fm., Yukon. Right: Diagrammatic view of hummock and swale geometry, the typical arrangement of laminae, and HCS-SCS set truncations. Modified from Swift et al., 1983.

HCS and SCS are found in sharp-based, fine- to medium-grained sandstone beds. In cross-section, sandstone laminae a few millimetres thick conform to the mound or swaley surfaces – i.e., they are continuous from trough to apex, and again to trough. Swaley bedding is concave upward. Individual laminae may thicken towards the troughs. The locus of lamina accretion is approximately vertical, conforming to the 3D surface expression of the mounds and swales. However, the mounds are not always symmetrical; cross-sections through many ancient HCS examples show some degree of lamination asymmetry. Truncations between successive sets of hummocks and swales are low-angle, commonly 10o to 15o. Some stratigraphic sections contain mostly SCS without the accompanying HCS.

The thickness of HCS-SCS bearing strata ranges from a few centimetres to 10s of metres. HCS-SCS sandstone units are usually interbedded with mudrocks. They may also overlie thin graded sandstone beds, and less commonly thin pebbly mudstones.

A stratigraphic section containing predominantly trough-like swaley cross-stratification in fine-grained sandstone. One HCS mound is visible near the top centre (arrow). Lens cap left centre is 55 mm diameter. Campanian Monster Fm., Yukon.

A stratigraphic section containing predominantly trough-like swaley cross-stratification in fine-grained sandstone. One HCS mound is visible near the top centre (arrow). Lens cap left centre is 55 mm diameter. Campanian Monster Fm., Yukon.

Formation – hydrodynamic behaviour

It is generally accepted that HCS-SCS are the depositional response to the passage of storm waves across the deeper parts of shelves and platforms, although there is still no consensus on their formation. Storm waves are higher amplitude and longer period than fairweather waves, and therefore the oscillatory motion of their orbitals will interact with deeper water substrates. However, based on theoretical considerations of wave dynamics and HCS geometry, P.A. Allen (1985) opined that wave motion cannot be the sole depositional agent; the implication here is that bottom-hugging currents must also be involved. Not everyone agrees with this conclusion.

One of the earliest experiments on HCS was conducted in a flume-like tunnel, where it was possible to introduce both oscillatory and unidirectional flow over a sand bed (Arnott and Southard, 1990). Low relief, HCS-like mounds were generated when oscillatory flow alone was applied. However, when a unidirectional current was added, even at low velocities, the HCS mounds became asymmetrical with slightly steeper faces downstream. With increasing velocity, the HCS bedforms were replaced progressively by ripples and larger 2D dunes. Thus, from an experimental perspective, both oscillatory and combined flow hypotheses are feasible.

 

Bedform stability under combined oscillatory and unidirectional flow determined from the experiments conducted Arnott and Southard, 1990 by – modified from Duke et al., 1991, who modified it from Arnott and Southard. HCS, as symmetrical and slightly asymmetrical mounds, occupies a narrow field at the low end of the unidirectional current velocity. At higher velocities the HCS washes out and forms 3D subaqueous dunes. The wave oscillation period was 8.5 seconds.

Bedform stability under combined oscillatory and unidirectional flow determined from the experiments conducted Arnott and Southard, 1990 by – modified from Duke et al., 1991, who modified it from Arnott and Southard. HCS, as symmetrical and slightly asymmetrical mounds, occupies a narrow field at the low end of the unidirectional current velocity. At higher velocities the HCS washes out and forms 3D subaqueous dunes. The wave oscillation period was 8.5 seconds.

The involvement of unidirectional bottom currents has since been verified from observations of recent storms on Atlantic and North Sea shelves. The jury is still deliberating on the exact mechanism, or mechanisms of HCS-SCS formation.

If bottom currents are involved, what kind of flows can we invoke?

Tidal currents (i.e., diurnal or semidiurnal) are always present, constantly interact with the substrate above fairweather wave-base, but do not produce HCS-like structures. Furthermore, they are bidirectional over the average life of a storm.

During major storms, water masses are pushed landward such that there is a temporary and local rise in sea level. Several authors have postulated that the subsequent storm surge ebb can generate seaward-directed, shore-normal turbidity currents (e.g., Hamblin and Walker, 1979). Perhaps these turbulent, high suspension-load flows combine with the wave-generated oscillatory flow to build HCS.

Based on paleocurrent evidence from HCS-bearing strata and associated turbidites (e.g., flute and groove casts, parting lineations), Leckie and Kristinik (1989) concluded that the turbidity currents have a significant component of shore-normal flow and that HCS forms under conditions where both density current and wave orbital flows combine. However, storm generated shore-parallel, or isobath-parallel flows have also been reported from the modern North American Atlantic shelf by Swift et al., 1983. Such flows are called geostrophic flows – flows that are deflected by Coriolis forces. The lower flow boundary of these geostrophic currents is turbulent and contains a high suspended sediment load (Duke et al., 1991)(PDF available). Thus, under certain conditions HCS-SCS can also be generated by a combination of shore-directed oscillatory flow and along-shore geostrophic currents that form in response to downwelling storm surges (Duke et al., 1991). Deposition of laminae occurs as the flows lose momentum such that sediment falls from suspension and is subsequently moulded by the oscillatory flow.

 

Common environments

By far the most common record of hummocky and swaley cross stratification is in shelf-platform successions. They tend to occur in the lower parts of shallowing-upward (and usually coarsening upward) parasequence-scale successions. Associated shoreface lithofacies include bedforms typical of shoreface environments (e.g., 2D and 3D crossbeds), indicators of tidal reversals, invertebrate and trace fossil assemblages, and mineralogical indicators such as ooids, glauconite, and marine carbonate cements.

HCS is generally found below fairweather wave-base in part because mounds are positive structures on the sea floor that at shallower depths are prone to reworking; its preservation potential above fairweather wave-base is significantly lower than at storm wave-base. SCS on the other hand accumulates in troughs, such that their preservation may extend to shallower depths across the shoreface.

 

Other posts in this series

Sedimentary lithofacies – An introduction

Ripple lithofacies: Ubiquitous bedforms

Climbing ripple lithofacies

Ripple lithofacies influenced by tides

Tabular and trough crossbed lithofacies

Laminated sandstone lithofacies

Low-angle crossbedded sandstone

Antidune lithofacies

Lithofacies beyond supercritical antidunes

Subaqueous dunes influenced by tides

Introducing coarse-grained lithofacies

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin