Turbidity current

Last updated
Turbidites are deposited in the deep ocean troughs below the continental shelf, or similar structures in deep lakes, by turbidity currents which slide down the slopes. Turbidite formation.jpg
Turbidites are deposited in the deep ocean troughs below the continental shelf, or similar structures in deep lakes, by turbidity currents which slide down the slopes.
Longitudinal section through an underwater turbidity current NOAA Turbidity Current Diagram.jpg
Longitudinal section through an underwater turbidity current

A turbidity current is most typically an underwater current of usually rapidly moving, sediment-laden water moving down a slope; although current research (2018) indicates that water-saturated sediment may be the primary actor in the process. [1] Turbidity currents can also occur in other fluids besides water.

Contents

Researchers from the Monterey Bay Aquarium Research Institute found that a layer of water-saturated sediment moved rapidly over the seafloor and mobilized the upper few meters of the preexisting seafloor. Plumes of sediment-laden water were observed during turbidity current events but they believe that these were secondary to the pulse of the seafloor sediment moving during the events. The belief of the researchers is that the water flow is the tail-end of the process that starts at the seafloor. [1]

In the most typical case of oceanic turbidity currents, sediment laden waters situated over sloping ground will flow down-hill because they have a higher density than the adjacent waters. The driving force behind a turbidity current is gravity acting on the high density of the sediments temporarily suspended within a fluid. These semi-suspended solids make the average density of the sediment bearing water greater than that of the surrounding, undisturbed water.

As such currents flow, they often have a "snow-balling-effect", as they stir up the ground over which they flow, and gather even more sedimentary particles in their current. Their passage leaves the ground over which they flow scoured and eroded. Once an oceanic turbidity current reaches the calmer waters of the flatter area of the abyssal plain (main oceanic floor), the particles borne by the current settle out of the water column. The sedimentary deposit of a turbidity current is called a turbidite.

Seafloor turbidity currents are often the result of sediment-laden river outflows, and can sometimes be initiated by earthquakes, slumping and other soil disturbances. They are characterized by a well-defined advance-front, also known as the current's head, and are followed by the current's main body. In terms of the more often observed and more familiar above sea-level phenomenon, they somewhat resemble flash floods.

Turbidity currents can sometimes result from submarine seismic instability, which is common with steep underwater slopes, and especially with submarine trench slopes of convergent plate margins, continental slopes and submarine canyons of passive margins. With an increasing continental shelf slope, current velocity increases, as the velocity of the flow increases, turbulence increases, and the current draws up more sediment. The increase in sediment also adds to the density of the current, and thus increases its velocity even further.

Definition

Turbidity currents are traditionally defined as those sediment gravity flows in which sediment is suspended by fluid turbulence. [2] [3] [4] However, the term "turbidity current" was adopted to describe a natural phenomenon whose exact nature is often unclear. The turbulence within a turbidity current is not always the support mechanism that keeps the sediment in suspension; however it is probable that turbulence is the primary or sole grain support mechanism in dilute currents (<3%). [5] Definitions are further complicated by an incomplete understanding of the turbulence structure within turbidity currents, and the confusion between the terms turbulent (i.e. disturbed by eddies) and turbid (i.e. opaque with sediment). [6] Kneller & Buckee, 2000 define a suspension current as 'flow induced by the action of gravity upon a turbid mixture of fluid and (suspended) sediment, by virtue of the density difference between the mixture and the ambient fluid'. A turbidity current is a suspension current in which the interstitial fluid is a liquid (generally water); a pyroclastic current is one in which the interstitial fluid is gas. [5]

Triggers

Hyperpycnal plume

When the concentration of suspended sediment at the mouth of a river is so large that the density of river water is greater than the density of sea water a particular kind of turbidity current can form called a hyperpycnal plume. [7] The average concentration of suspended sediment for most river water that enters the ocean is much lower than the sediment concentration needed for entry as a hyperpycnal plume. Although some rivers can often have continuously high sediment load that can create a continuous hyperpycnal plume, such as the Haile River (China), which has an average suspended concentration of 40.5 kg/m3. [7] The sediment concentration needed to produce a hyperpycnal plume in marine water is 35 to 45 kg/m3, depending on the water properties within the coastal zone. [7] Most rivers produce hyperpycnal flows only during exceptional events, such as storms, floods, glacier outbursts, dam breaks, and lahar flows. In fresh water environments, such as lakes, the suspended sediment concentration needed to produce a hyperpycnal plume is quite low (1 kg/m3). [7]

Sedimentation in reservoirs

The transport and deposition of the sediments in narrow alpine reservoirs is often caused by turbidity currents. They follow the thalweg of the lake to the deepest area near the dam, where the sediments can affect the operation of the bottom outlet and the intake structures. [8] Controlling this sedimentation within the reservoir can be achieved by using solid and permeable obstacles with the right design. [8]

Earthquake triggering

Turbidity currents are often triggered by tectonic disturbances of the sea floor. The displacement of continental crust in the form of fluidization and physical shaking both contribute to their formation. Earthquakes have been linked to turbidity current deposition in many settings, particularly where physiography favors preservation of the deposits and limits the other sources of turbidity current deposition. [9] [10] Since the famous case of breakage of submarine cables by a turbidity current following the 1929 Grand Banks earthquake, [11] earthquake triggered turbidites have been investigated and verified along the Cascadia subduction Zone, [12] the Northern San Andreas Fault, [13] a number of European, Chilean and North American lakes, [14] [15] [16] Japanese lacustrine and offshore regions [17] [18] and a variety of other settings. [19] [20]

Canyon-flushing

When large turbidity currents flow into canyons they may become self-sustaining, [21] and may entrain sediment that has previously been introduced into the canyon by littoral drift, storms or smaller turbidity currents. Canyon-flushing associated with surge-type currents initiated by slope failures may produce currents whose final volume may be several times that of the portion of the slope that has failed (e.g. Grand Banks). [22]

Slumping

Sediment that has piled up at the top of the continental slope, particularly at the heads of submarine canyons can create turbidity current due to overloading, thus consequent slumping and sliding.

Convective sedimentation beneath river plumes

Laboratory images of how convective sedimentation beneath a buoyant sediment-laden surface can initiate a secondary turbidity current. Convective sedimentation beneath a buoyant plumes can lead to a secondary turbidity current.tif
Laboratory images of how convective sedimentation beneath a buoyant sediment-laden surface can initiate a secondary turbidity current.

A buoyant sediment-laden river plume can induce a secondary turbidity current on the ocean floor by the process of convective sedimentation. [24] [4] Sediment in the initially buoyant hypopycnal flow accumulates at the base of the surface flow, [25] so that the dense lower boundary become unstable. The resulting convective sedimentation leads to a rapid vertical transfer of material to the sloping lake or ocean bed, potentially forming a secondary turbidity current. The vertical speed of the convective plumes can be much greater than the Stokes settling velocity of an individual particle of sediment. [26] Most examples of this process have been made in the laboratory, [24] [27] but possible observational evidence of a secondary turbidity current was made in Howe Sound, British Columbia, [28] where a turbidity current was periodically observed on the delta of the Squamish River. As the vast majority of sediment laden rivers are less dense than the ocean, [7] rivers cannot readily form plunging hyperpycnal flows. Hence convective sedimentation is an important possible initiation mechanism for turbidity currents. [4]

An example of steep submarine canyons carved out by turbidity currents, located along California's Central Coast MBNMS Davidson Map Update.pdf
An example of steep submarine canyons carved out by turbidity currents, located along California's Central Coast

Effect on ocean floor

Large and fast-moving turbidity currents can carve gulleys and ravines into the ocean floor of continental margins and cause damage to artificial structures such as telecommunication cables on the seafloor. Understanding where turbidity currents flow on the ocean floor can help to decrease the amount of damage to telecommunication cables by avoiding these areas or reinforcing the cables in vulnerable areas.

When turbidity currents interact with regular ocean currents, such as contour currents, they can change their direction. This ultimately shifts submarine canyons and sediment deposition locations. One example of this is located in the western part of the Gulf of Cadiz, where the ocean current leaving the Mediterranean Sea (also known as the Mediterranean outflow water) pushes turbidity currents westward. This has changed the shape of submarine valleys and canyons in the region to also curve in that direction. [29]

Deposits

Turbidite interbedded with finegrained dusky-yellow sandstone and gray clay shale that occur in graded beds, Point Loma Formation, California. Turbidite-California.jpeg
Turbidite interbedded with finegrained dusky-yellow sandstone and gray clay shale that occur in graded beds, Point Loma Formation, California.

When the energy of a turbidity current lowers, its ability to keep suspended sediment decreases, thus sediment deposition occurs. When the material comes to rest, it is the sand and other coarse material which settles first followed by mud and eventually the very fine particulate matter. It is this sequence of deposition that creates the so called Bouma sequences that characterize turbidite deposits.

Because turbidity currents occur underwater and happen suddenly, they are rarely seen as they happen in nature, thus turbidites can be used to determine turbidity current characteristics. Some examples: grain size can give indication of current velocity, grain lithology and the use of foraminifera for determining origins, grain distribution shows flow dynamics over time and sediment thickness indicates sediment load and longevity.

Turbidites are commonly used in the understanding of past turbidity currents, for example, the Peru-Chile Trench off Southern Central Chile (36°S–39°S) contains numerous turbidite layers that were cored and analysed. [30] From these turbidites the predicted history of turbidity currents in this area was determined, increasing the overall understanding of these currents. [30]

Antidune deposits

Some of the largest antidunes on Earth are formed by turbidity currents. One observed sediment-wave field is located on the lower continental slope off Guyana, South America. [31] This sediment-wave field covers an area of at least 29 000 km2 at a water depth of 4400–4825 meters. [31] These antidunes have wavelengths of 110–2600 m and wave heights of 1–15 m. [31] Turbidity currents responsible for wave generation are interpreted as originating from slope failures on the adjacent Venezuela, Guyana and Suriname continental margins. [31] Simple numerical modelling has been enabled to determine turbidity current flow characteristics across the sediment waves to be estimated: internal Froude number = 0.7–1.1, flow thickness = 24–645 m, and flow velocity = 31–82 cm·s−1. [31] Generally, on lower gradients beyond minor breaks of slope, flow thickness increases and flow velocity decreases, leading to an increase in wavelength and a decrease in height. [31]

Reversing buoyancy

The behaviour of turbidity currents with buoyant fluid (such as currents with warm, fresh or brackish interstitial water entering the sea) has been investigated to find that the front speed decreases more rapidly than that of currents with the same density as the ambient fluid. [32] These turbidity currents ultimately come to a halt as sedimentation results in a reversal of buoyancy, and the current lifts off, the point of lift-off remaining constant for a constant discharge. [32] The lofted fluid carries fine sediment with it, forming a plume that rises to a level of neutral buoyancy (if in a stratified environment) or to the water surface, and spreads out. [32] Sediment falling from the plume produces a widespread fall-out deposit, termed hemiturbidite. [33] Experimental turbidity currents [34] and field observations [35] suggest that the shape of the lobe deposit formed by a lofting plume is narrower than for a similar non-lofting plume

Prediction

Prediction of erosion by turbidity currents, and of the distribution of turbidite deposits, such as their extent, thickness and grain size distribution, requires an understanding of the mechanisms of sediment transport and deposition, which in turn depends on the fluid dynamics of the currents.

The extreme complexity of most turbidite systems and beds has promoted the development of quantitative models of turbidity current behaviour inferred solely from their deposits. Small-scale laboratory experiments therefore offer one of the best means of studying their dynamics. Mathematical models can also provide significant insights into current dynamics. In the long term, numerical techniques are most likely the best hope of understanding and predicting three-dimensional turbidity current processes and deposits. In most cases, there are more variables than governing equations, and the models rely upon simplifying assumptions in order to achieve a result. [5] The accuracy of the individual models thus depends upon the validity and choice of the assumptions made. Experimental results provide a means of constraining some of these variables as well as providing a test for such models. [5] Physical data from field observations, or more practical from experiments, are still required in order to test the simplifying assumptions necessary in mathematical models. Most of what is known about large natural turbidity currents (i.e. those significant in terms of sediment transfer to the deep sea) is inferred from indirect sources, such as submarine cable breaks and heights of deposits above submarine valley floors. Although during the 2003 Tokachi-oki earthquake a large turbidity current was observed by the cabled observatory which provided direct observations, which is rarely achieved. [36]

Oil exploration

Oil and gas companies are also interested in turbidity currents because the currents deposit organic matter that over geologic time gets buried, compressed and transformed into hydrocarbons. The use of numerical modelling and flumes are commonly used to help understand these questions. [37] Much of the modelling is used to reproduce the physical processes which govern turbidity current behaviour and deposits. [37]

Modeling approaches

Shallow-water models

The so-called depth-averaged or shallow-water models are initially introduced for compositional gravity currents [38] and then later extended to turbidity currents. [39] [40] The typical assumptions used along with the shallow-water models are: hydrostatic pressure field, clear fluid is not entrained (or detrained), and particle concentration does not depend on the vertical location. Considering the ease of implementation, these models can typically predict flow characteristic such as front location or front speed in simplified geometries, e.g. rectangular channels, fairly accurately.

Depth-resolved models

With the increase in computational power, depth-resolved models have become a powerful tool to study gravity and turbidity currents. These models, in general, are mainly focused on the solution of the Navier-Stokes equations for the fluid phase. With dilute suspension of particles, a Eulerian approach proved to be accurate to describe the evolution of particles in terms of a continuum particle concentration field. Under these models, no such assumptions as shallow-water models are needed and, therefore, accurate calculations and measurements are performed to study these currents. Measurements such as, pressure field, energy budgets, vertical particle concentration and accurate deposit heights are a few to mention. Both Direct numerical simulation (DNS) [41] and Turbulence modeling [42] are used to model these currents.

Notable examples of turbidity currents

See also

Related Research Articles

<span class="mw-page-title-main">Sediment</span> Particulate solid matter that is deposited on the surface of land

Sediment is a naturally occurring material that is broken down by processes of weathering and erosion, and is subsequently transported by the action of wind, water, or ice or by the force of gravity acting on the particles. For example, sand and silt can be carried in suspension in river water and on reaching the sea bed deposited by sedimentation; if buried, they may eventually become sandstone and siltstone through lithification.

<span class="mw-page-title-main">Turbidite</span> Geologic deposit of a turbidity current

A turbidite is the geologic deposit of a turbidity current, which is a type of amalgamation of fluidal and sediment gravity flow responsible for distributing vast amounts of clastic sediment into the deep ocean.

<span class="mw-page-title-main">Forearc</span> The region between an oceanic trench and the associated volcanic arc

Forearc is a plate tectonic term referring to a region in a subduction zone between an oceanic trench and the associated volcanic arc. Forearc regions are present along convergent margins and eponymously form 'in front of' the volcanic arcs that are characteristic of convergent plate margins. A back-arc region is the companion region behind the volcanic arc.

<span class="mw-page-title-main">Submarine canyon</span> Steep-sided valley cut into the seabed of the continental slope

A submarine canyon is a steep-sided valley cut into the seabed of the continental slope, sometimes extending well onto the continental shelf, having nearly vertical walls, and occasionally having canyon wall heights of up to 5 km (3 mi), from canyon floor to canyon rim, as with the Great Bahama Canyon. Just as above-sea-level canyons serve as channels for the flow of water across land, submarine canyons serve as channels for the flow of turbidity currents across the seafloor. Turbidity currents are flows of dense, sediment laden waters that are supplied by rivers, or generated on the seabed by storms, submarine landslides, earthquakes, and other soil disturbances. Turbidity currents travel down slope at great speed, eroding the continental slope and finally depositing sediment onto the abyssal plain, where the particles settle out.

<span class="mw-page-title-main">Bouma sequence</span> Set of structures in sediments or sedimentary rocks

The Bouma sequence describes a classic set of sedimentary structures in turbidite beds deposited by turbidity currents at the bottoms of lakes, oceans and rivers.

<span class="mw-page-title-main">Abyssal fan</span> Underwater geological structures associated with large-scale sediment deposition

Abyssal fans, also known as deep-sea fans, underwater deltas, and submarine fans, are underwater geological structures associated with large-scale sediment deposition and formed by turbidity currents. They can be thought of as an underwater version of alluvial fans and can vary dramatically in size, with widths from several kilometres to several thousands of kilometres. The largest is the Bengal Fan, followed by the Indus Fan, but major fans are also found at the outlet of the Amazon, Congo, Mississippi and elsewhere.

<span class="mw-page-title-main">Marine sediment</span>

Marine sediment, or ocean sediment, or seafloor sediment, are deposits of insoluble particles that have accumulated on the seafloor. These particles either have their origins in soil and rocks and have been transported from the land to the sea, mainly by rivers but also by dust carried by wind and by the flow of glaciers into the sea, or they are biogenic deposits from marine organisms or from chemical precipitation in seawater, as well as from underwater volcanoes and meteorite debris.

<span class="mw-page-title-main">Sedimentary structures</span> Geologic structures formed during sediment deposition

Sedimentary structures include all kinds of features in sediments and sedimentary rocks, formed at the time of deposition.

Abyssal channels are channels in Earth's sea floor. They are formed by fast-flowing floods of turbid water caused by avalanches near the channel's head, with the sediment carried by the water causing a build-up of the surrounding abyssal plains. Submarine channels and the turbidite systems which form them are responsible for the accumulation of most sandstone deposits found on continental slopes and have proven to be one of the most common types of hydrocarbon reservoirs found in these regions.

<span class="mw-page-title-main">Submarine landslide</span> Landslides that transport sediment across the continental shelf and into the deep ocean

Submarine landslides are marine landslides that transport sediment across the continental shelf and into the deep ocean. A submarine landslide is initiated when the downwards driving stress exceeds the resisting stress of the seafloor slope material, causing movements along one or more concave to planar rupture surfaces. Submarine landslides take place in a variety of different settings, including planes as low as 1°, and can cause significant damage to both life and property. Recent advances have been made in understanding the nature and processes of submarine landslides through the use of sidescan sonar and other seafloor mapping technology.

<span class="mw-page-title-main">Contourite</span> Type of sedimentary deposit

A contourite is a sedimentary deposit commonly formed on continental rises in lower slope settings, although it may occur anywhere that is below the storm wave base. Countourites are produced by thermohaline-induced deepwater bottom currents and may be influenced by wind or tidal forces. The geomorphology of contourite deposits is mainly influenced by the deepwater bottom-current velocity, sediment supply, and seafloor topography.

<span class="mw-page-title-main">Sediment gravity flow</span> Sediment transport mechanism

A sediment gravity flow is one of several types of sediment transport mechanisms, of which most geologists recognize four principal processes. These flows are differentiated by their dominant sediment support mechanisms, which can be difficult to distinguish as flows can be in transition from one type to the next as they evolve downslope.

<span class="mw-page-title-main">Powder snow avalanche</span>

A powder snow avalanche is a type of avalanche where snow grains are largely or completely suspended and moved by air in a state of fluid turbulence. They are particle-laden gravity currents and closely related to turbidity currents, pyroclastic flows from volcanoes and dust storms in the desert. The turbulence is typically generated by the forward motion of the current along the lower boundary of the domain, the motion being in turn driven by the action of gravity on the density difference between the particle-fluid mixture and the ambient fluid. The ambient fluid is generally of similar composition to the interstitial fluid, and is water for turbidity currents and air for avalanches. These flows are non-conservative in that they may exchange particles at the lower boundary by deposition or suspension, and may exchange fluid with the ambient by entrainment or detrainment. Such flows dissipate when the turbulence can no longer hold the particles in suspension and they are deposited on the lower boundary. When the turbulence is strong enough to suspend new material from the bed or the underlying dense flow then current is said to be auto-suspending. Particle concentrations in the suspension cloud are usually sufficiently low that particle-particle interactions play a small or negligible role in maintaining the suspension. In powder snow avalanches, even at these low concentrations, the extra density of the suspended particles is large relative to that of air, so the Boussinesq approximation, where density differences are considered negligible in inertia terms, is invalid, so that the snow grains carry most of the flows momentum. This is in contrast to turbidity currents and laboratory experiments in water where the extra inertia of the particles can usually be neglected. Nonetheless, due to the extreme difficulty in estimating particle concentrations in natural flows there remains considerable uncertainty—and debate—concerning the particle loading in large submarine turbidity currents and the validity of the Boussinesq approximation.

<span class="mw-page-title-main">Lowe sequence</span>

The Lowe sequence describes a set of sedimentary structures in turbidite sandstone beds that are deposited by high-density turbidity currents. It is intended to complement, not replace, the better known Bouma sequence, which applies primarily to turbidites deposited by low-density turbidity currents.

The term gravity current intrusion denotes the fluid mechanics phenomenon within which a fluid intrudes with a predominantly horizontal motion into a separate stratified fluid, typically along a plane of neutral buoyancy. This behaviour distinguishes the difference between gravity current intrusions and gravity currents, as intrusions are not restrained by a well-defined boundary surface. As with gravity currents, intrusion flow is driven within a gravity field by density differences typically small enough to allow for the Boussinesq approximation.

<span class="mw-page-title-main">Quinault Canyon</span> Submarine canyon in the Pacific Ocean near Washington, United States

The Quinault Canyon is a submarine canyon, off Washington state, in Olympic Coast National Marine Sanctuary.

Cascadia Channel is the most extensive deep-sea channel currently known of the Pacific Ocean. It extends across Cascadia Abyssal Plain, through the Blanco Fracture Zone, and into Tufts Abyssal Plain. Notably, Cascadia Channel has tributaries, akin to river tributaries.

<span class="mw-page-title-main">Southern Hydrate Ridge</span>

Southern Hydrate Ridge, located about 90 km offshore Oregon Coast, is an active methane seeps site located on the southern portion of Hydrate Ridge. It extends 25 km in length and 15 km across, trending north-northeast-south-southwest at the depth of approximately 800 m. Southern Hydrate Ridge has been the site of numerous submersible dives with the human occupied Alvin submarine, extensive visits by numerous robotic vehicles including the Canadian ROV ROPOS, Jason , and Tiburon (MBARI), and time-series geophysical studies that document changes in the subsurface distribution of methane. It is also a key site of the National Science Foundations Regional Cabled Array that is part of the Ocean Observatories Initiative (OOI), which includes eight types of cabled instruments streaming live data back to shore 24/7/365 at the speed of light, as well as uncabled instruments.

Madeira Abyssal Plain, also called Madeira Plain, is an abyssal plain situated at the center and deepest part of the Canary Basin. It is a north-northeast to south-southeast elongated basin that almost parallels the Mid-Atlantic Ridge. Its western boundary is marked by a chain of seamounts known as the either Seewarte Seamounts or Atlantis-Great Meteor Seamount Chain. Its eastern boundary is a distinct break of slope that marks the foot of the African Continental Rise. This abyssal plain occupies an area of about 68,000 km2 (26,000 sq mi). Across this basin, slope angles are generally less than 0.01°.

The Kaikōura Canyon is a geologically active submarine canyon located southwest of the Kaikōura Peninsula off the northeastern coast of the South Island of New Zealand. It is 60 kilometres (37 mi) long, and is generally U-shaped. The canyon descends into deep water and merges into an ocean channel system that can be traced for hundreds of kilometres across the deep ocean floor. At the head of the Kaikōura Canyon, the depth of water is around 30 metres (98 ft), but it drops rapidly to 600 metres (2,000 ft) and continues down to around 2,000 metres (6,600 ft) deep where it meets the Hikurangi Channel. Sperm whales can be seen close to the coast south of Goose Bay, because the deep water of the Kaikōura Canyon is only one kilometre (0.62 mi) off the shoreline in this area.

References

  1. 1 2 "'Turbidity currents' are not just currents, but involve movement of the seafloor itself". EurekAlert!. Monterey Bay Aquarium Research Institute. 5 October 2018. Retrieved 8 October 2018.
  2. Sanders, J.E. 1965 Primary sedimentary structures formed by turbidity currents and related resedimentation mechanisms. In: Primary Sedimentary Structures and Their Hydro-Dynamic Interpretation – a Symposium Middleton, G. V.), SEPM Spec. Publishers, 12, 192–219.
  3. Meiburg, E. & Kneller, B. 2010, "Turbidity currents and their deposits", Annual Review of Fluid Mechanics, vol. 42, pp. 135–156.
  4. 1 2 3 Wells, Mathew G.; Dorrell, Robert M. (2021-01-05). "Turbulence Processes Within Turbidity Currents". Annual Review of Fluid Mechanics. 53 (1): 59–83. Bibcode:2021AnRFM..53...59W. doi:10.1146/annurev-fluid-010719-060309. ISSN   0066-4189. S2CID   224957150.
  5. 1 2 3 4 Kneller, B. & Buckee, C. 2000, "The structure and fluid mechanics of turbidity currents: A review of some recent studies and their geological implications", Sedimentology, vol. 47, no. SUPPL. 1, pp. 62–94.
  6. McCave, I.N. & Jones, K.P.N. 1988 Deposition of ungraded muds from high-density non-turbulent turbidity currents. Nature, 333, 250–252.
  7. 1 2 3 4 5 Mulder, T. & Syvitski, J.P.M. 1995, "Turbidity currents generated at river mouths during exceptional discharges to the world oceans", Journal of Geology, vol. 103, no. 3, pp. 285–299.
  8. 1 2 Oehy, C.D. & Schleiss, A.J. 2007, "Control of turbidity currents in reservoirs by solid and permeable obstacles", Journal of Hydraulic Engineering, vol. 133, no. 6, pp. 637–648.
  9. 1 2 Adams, J., 1990, Paleoseismicity of the Cascadia subduction zone: Evidence from turbidites off the Oregon-Washington Margin: Tectonics, v. 9, p. 569–584.
  10. Goldfinger, C., 2011, Submarine Paleoseismology Based on Turbidite Records: Annual Review of Marine Science, v. 3, p. 35–66.
  11. Heezen, B.C., and Ewing, M., 1952, Turbidity currents and submarine slumps, and the 1929 Grand Banks earthquake: American Journal of Science, v. 250, p. 849–873.
  12. 1 2 Goldfinger, C., Nelson, C.H., and Johnson, J.E., 2003, Holocene Earthquake Records From the Cascadia Subduction Zone and Northern San Andreas Fault Based on Precise Dating of Offshore Turbidites: Annual Review of Earth and Planetary Sciences, v. 31, p. 555–577.
  13. Goldfinger, C., Grijalva, K., Burgmann, R., Morey, A.E., Johnson, J.E., Nelson, C.H., Gutierrez-Pastor, J., Ericsson, A., Karabanov, E., Chaytor, J.D., Patton, J., and Gracia, E., 2008, Late Holocene Rupture of the Northern San Andreas Fault and Possible Stress Linkage to the Cascadia Subduction Zone Bulletin of the Seismological Society of America, v. 98, p. 861–889.
  14. Schnellmann, M., Anselmetti, F.S., Giardini, D., and Ward, S.N., 2002, Prehistoric earthquake history revealed by lacustrine slump deposits: Geology, v. 30, p. 1131–1134.
  15. Moernaut, J., De Batist, M., Charlet, F., Heirman, K., Chapron, E., Pino, M., Brümmer, R., and Urrutia, R., 2007, Giant earthquakes in South-Central Chile revealed by Holocene mass-wasting events in Lake Puyehue: Sedimentary Geology, v. 195, p. 239–256.
  16. Brothers, D.S., Kent, G.M., Driscoll, N.W., Smith, S.B., Karlin, R., Dingler, J.A., Harding, A.J., Seitz, G.G., and Babcock, J.M., 2009, New Constraints on Deformation, Slip Rate, and Timing of the Most Recent Earthquake on the West Tahoe-Dollar Point Fault, Lake Tahoe Basin, California: Bulletin of the Seismological Society of America, v. 99, p. 499–519.
  17. Nakajima, T., 2000, Initiation processes of turbidity currents; implications for assessments of recurrence intervals of offshore earthquakes using turbidites: Bulletin of the Geological Survey of Japan, v. 51, p. 79–87.
  18. Noda, A., TuZino, T., Kanai, Y., Furukawa, R., and Uchida, J.-i., 2008, Paleoseismicity along the southern Kuril Trench deduced from submarine-fan turbidites: Marine Geology, v. 254, p. 73–90.
  19. Huh, C.A., Su, C.C., Liang, W.T., and Ling, C.Y., 2004, Linkages between turbidites in the southern Okinawa Trough and submarine earthquakes: Geophysical Research Letters, v. 31.
  20. Gràcia, E., Vizcaino, A., Escutia, C., Asiolic, A., Garcia-Orellanad, J., Pallàse, R., Lebreiro, S., and Goldfinger, C., 2010, Holocene earthquake record offshore Portugal (SW Iberia): Applying turbidite paleoseismology in a slow-convergence margin: Quaternary Science Reviews, v. 29, p. 1156–1172.
  21. Pantin, H.M. 1979 Interaction between velocity and effective density in turbidity flow: phase-plane analysis, with criteria for autosuspension. March Geol., 31, 59–99.
  22. Piper, D.J.W. & Aksu, A.E. 1987 The source and origin of the 1929 Grand Banks turbidity current inferred from sediment budgets. Geo-March Lett., 7, 177–182.
  23. Jazi, Shahrzad Davarpanah; Wells, Mathew (2020). "Dynamics of settling-driven convection beneath a sediment laden buoyant overflow: implications for the length-scale of deposition in lakes and the coastal ocean". Sedimentology. 67 (1): 699–720. doi: 10.1111/sed.12660 .
  24. 1 2 Parsons, Jeffrey D.; Bush, John W. M.; Syvitski, James P. M. (2001-04-06). "Hyperpycnal plume formation from riverine outflows with small sediment concentrations". Sedimentology. 48 (2): 465–478. Bibcode:2001Sedim..48..465P. doi:10.1046/j.1365-3091.2001.00384.x. ISSN   0037-0746. S2CID   128481974.
  25. Burns, P.; Meiburg, E. (2014-11-27). "Sediment-laden fresh water above salt water: nonlinear simulations". Journal of Fluid Mechanics. 762: 156–195. doi:10.1017/jfm.2014.645. ISSN   0022-1120. S2CID   53663402.
  26. Davarpanah Jazi, Shahrzad; Wells, Mathew G. (2016-10-28). "Enhanced sedimentation beneath particle-laden flows in lakes and the ocean due to double-diffusive convection". Geophysical Research Letters. 43 (20): 10, 883–10, 890. Bibcode:2016GeoRL..4310883D. doi:10.1002/2016gl069547. hdl: 1807/81129 . ISSN   0094-8276. S2CID   55359245.
  27. Davarpanah Jazi, Shahrzad; Wells, Mathew G. (2019-11-17). "Dynamics of settling-driven convection beneath a sediment-laden buoyant overflow: Implications for the length-scale of deposition in lakes and the coastal ocean". Sedimentology. 67 (1): 699–720. doi: 10.1111/sed.12660 . ISSN   0037-0746.
  28. Hage, Sophie; Cartigny, Matthieu J.B.; Sumner, Esther J.; Clare, Michael A.; Hughes Clarke, John E.; Talling, Peter J.; Lintern, D. Gwyn; Simmons, Stephen M.; Silva Jacinto, Ricardo; Vellinga, Age J.; Allin, Joshua R. (2019-10-28). "Direct Monitoring Reveals Initiation of Turbidity Currents From Extremely Dilute River Plumes". Geophysical Research Letters. 46 (20): 11310–11320. Bibcode:2019GeoRL..4611310H. doi: 10.1029/2019gl084526 . ISSN   0094-8276. PMC   6919390 . PMID   31894170.
  29. Mulder, T., Lecroart, P., Hanquiez, V., Marches, E., Gonthier, E., Guedes, J.-., Thiébot, E., Jaaidi, B., Kenyon, N., Voisset, M., Perez, C., Sayago, M., Fuchey, Y. & Bujan, S. 2006, "The western part of the Gulf of Cadiz: Contour currents and turbidity currents interactions", Geo-Marine Letters, vol. 26, no. 1, pp. 31–41.
  30. 1 2 Völker, D., Reichel, T., Wiedicke, M. & Heubeck, C. 2008, "Turbidites deposited on Southern Central Chilean seamounts: Evidence for energetic turbidity currents", Marine Geology, vol. 251, no. 1–2, pp. 15–31
  31. 1 2 3 4 5 6 Ercilla, G., Alonso, B., Wynn, R.B. & Baraza, J. 2002, "Turbidity current sediment waves on irregular slopes: Observations from the Orinoco sediment-wave field", Marine Geology, vol. 192, no. 1–3, pp. 171–187.
  32. 1 2 3 Hürzeler, B.E., Imberger, J. & Ivey, G.N. 1996 Dynamics of turbidity current with reversing buoyancy. J. Hydraul. Eng., 122, 230–236.
  33. Stow, D.A.V. & Wetzel, A. 1990 Hemiturbidite: a new type of deep-water sediment. Proc. Ocean Drilling Program, Scientific Results, 116, 25–34.
  34. Steel, Elisabeth; Buttles, James; Simms, Alexander R.; Mohrig, David; Meiburg, Eckart (2016-11-03). "The role of buoyancy reversal in turbidite deposition and submarine fan geometry". Geology. 45 (1): 35–38. doi:10.1130/g38446.1. ISSN   0091-7613. S2CID   132607431.
  35. Steel, Elisabeth; Simms, Alexander R.; Warrick, Jonathan; Yokoyama, Yusuke (2016-05-25). "Highstand shelf fans: The role of buoyancy reversal in the deposition of a new type of shelf sand body". Geological Society of America Bulletin. 128 (11–12): 1717–1724. Bibcode:2016GSAB..128.1717S. doi:10.1130/b31438.1. ISSN   0016-7606.
  36. Mikada, H., Mitsuzawa, K., Matsumoto, H., Watanabe, T., Morita, S., Otsuka, R., Sugioka, H., Baba, T., Araki, E. & Suyehiro, K. 2006, "New discoveries in dynamics of an M8 earthquake-phenomena and their implications from the 2003 Tokachi-oki earthquake using a long term monitoring cabled observatory", Tectonophysics, vol. 426, no. 1–2, pp. 95–105
  37. 1 2 Salles, T., Lopez, S., Eschard, R., Lerat, O., Mulder, T. & Cacas, M.C. 2008, "Turbidity current modelling on geological time scales", Marine Geology, vol. 248, no. 3–4, pp. 127–150.
  38. Rottman, J.W. & Simpson, J.E. 1983, "Gravity currents produced by instantaneous releases of a heavy fluid in a rectangular channel", Journal of Fluid Mechanics, vol. 135, pp. 95–110.
  39. Parker, G., Fukushima, Y. & Pantin, H.M. 1986, "Self-accelerating turbidity currents", Journal of Fluid Mechanics, vol. 171, pp. 145–181.
  40. Bonnecaze, R.T., Huppert, H.E. & Lister, J.R. 1993, "Particle-driven gravity currents", Journal of Fluid Mechanics, vol. 250, pp. 339–369.
  41. Necker, F., Hartel, C., Kleiser, L. & Meiburg, E. 2002, "High-resolution simulations of particle-driven gravity currents", International Journal of Multiphase Flow, vol. 28, pp. 279–300.
  42. Kassem, A. & Imran, J. 2004, "Three-dimensional modeling of density current. II. Flow in sinuous confined and uncontined channels", Journal of Hydraulic Research, vol. 42, number. 6, pp. 591–602.
  43. Bruce C. Heezen and Maurice Ewing, "Turbidity Currents and Submarine Slumps, and the 1929 Grand Banks Earthquake," American Journal of Science, Vol. 250, December 1952, pp. 849–873.
  44. 1 2 3 Piper, D.J.W., Cochonat, P. & Morrison, M.L. 1999, "The sequence of events around the epicentre of the 1929 Grand Banks earthquake: Initiation of debris flows and turbidity current inferred from sidescan sonar", Sedimentology, vol. 46, no. 1, pp. 79–97.
  45. 1 2 Goldfinger, C., Nelson, C.H., Morey, A., Johnson, J.E., Gutierrez-Pastor, J., Eriksson, A.T., Karabanov, E., Patton, J., Gracia, E., Enkin, R., Dallimore, A., Dunhill, G., and Vallier, T., 2011, Turbidite Event History: Methods and Implications for Holocene Paleoseismicity of the Cascadia Subduction Zone, USGS Professional Paper 1661-F, Reston, VA, U.S. Geological Survey, 332 p, 64 Figures.
  46. Atwater, B.F., 1987, Evidence for great Holocene earthquakes along the outer coast of Washington State: Science, v. 236, p. 942–944.
  47. Atwater, B.F., and Hemphill-Haley, E., 1997, Recurrence intervals for great earthquakes of the past 3500 years at northeastern Willapa Bay, Washington, Professional Paper, Volume 1576: Reston, VA., U.S. Geological Survey, p. 108 p.
  48. Kelsey, H.M., Witter, R.C., and Hemphill-Haley, E., 2002, Plate-boundary earthquakes and tsunamis of the past 5500 yr, Sixes River estuary, southern Oregon: Geological Society of America Bulletin, v. 114, p. 298–314.
  49. Kelsey, H.M., Nelson, A.R., Hemphill-Haley, E., and Witter, R.C., 2005, Tsunami history of an Oregon coastal lake reveals a 4600 yr record of great earthquakes on the Cascadia subduction zone: GSA Bulletin, v. 117, p. 1009–1032.
  50. Nelson, A.R., Sawai, Y., Jennings, A.E., Bradley, L., Gerson, L., Sherrod, B.L., Sabean, J., and Horton, B.P., 2008, Great-earthquake paleogeodesy and tsunamis of the past 2000 years at Alsea Bay, central Oregon coast, USA: Quaternary Science Reviews, v. 27, p. 747–768.
  51. 1 2 3 Hsu, S.-., Kuo, J., Lo, C.-., Tsai, C.-., Doo, W.-., Ku, C.-. & Sibuet, J.-. 2008, "Turbidity currents, submarine landslides and the 2006 Pingtung earthquake off SW Taiwan", Terrestrial, Atmospheric and Oceanic Sciences, vol. 19, no. 6, pp. 767–772.
  52. Vincent, Warwick F.; Bertola, Carinne (2014). "Lake Physics to Ecosystem Services: Forel and the Origins of Limnology". Limnology and Oceanography E-Lectures. 4 (3): 1–47. doi:10.4319/lol.2014.wvincent.cbertola.8. ISSN   2164-0254.
  53. Forel (1887). "Le ravin sous-lacustre du Rhône dans le lac Léman". Bulletin de la Société vaudoise des ingénieurs et des architectes. 11: 1–2.
  54. Amos, Jonathan (7 June 2021). "Underwater avalanche continued for two whole days". BBC News . Retrieved 7 June 2021.