Saturday, June 1, 2013

14. Triggering mechanisms of sediment failures in deep water




Deep-water Processes by G. Shanmugam


Triggering mechanisms of sediment failures in deep water

 

The triggering of sediment failures is controlled by the pull of gravity, the weight of material, the slope, and the planes of weaknesses. Masson et al. (2006) and Feeley (2007) provide a general review of triggering mechanisms of MTDs. There are at least 21 triggering mechanisms of sediment failures that can initiate mass-transport deposits in deep water (Table 1). In many continental margins, several triggering mechanisms work concurrently, such as earthquakes and tsunamis.

 

Table 1. Triggering mechanisms of sediment failures. Modified after Shanmugam (2012a, b).

 

Triggering  Mechanisms
Environment of MTD Emplacement
Duration
  1. Earthquake
(Heezen and Ewing, 1952)
  1. Meteorite impact
(Claeys et al. 2002) (Figure 1)
  1. Volcanic activity
(Tilling et al., 1990)
  1. Tsunami waves
(Shanmugam, 2006a, b)
(Figure 2)
  1. Rogue waves
(Dysthe et al., 2008)
  1. Cyclonic waves (Bea et al., 1983; Prior et al., 1989; Shanmugam, 2008a) (Figure 3)
  2. Internal waves and tides (Shanmugam, 2013a, b)
  3. Ebb tidal current
(Boyd et al., 2008)
  1. Monsoonal rainfall
(Petley, 2012)
  1. Groundwater seepage
(Brönnimann, 2011)
  1. Wildfire
(Cannon et al., 2001)
  1. *Human activity (Dan et al., 2007).
Subaerial & submarine
 
Subaerial & submarine
 
Subaerial & submarine
 
Subaerial & submarine
 
 
Submarine
 
Subaerial & submarine
 
 
 
Submarine
 
Submarine
 
Subaerial
 
Subaerial & submarine
 
Subaerial
Subaerial & submarine
 
 
 
 
 
 
 
 
 
 
 
 
Short-term events:
a few minutes to several hours, days or months
  1. **Tectonic events: (a) tectonic oversteepening (Greene et al., 2006); (b) tensional stresses on the rift zones (Urgeles et al., 1997); (c) oblique seamount subduction (Collot et al., 2001), among others
  2. Glacial maxima and loading (Elverhoi et al., 1997, 2002; Lee, 2009)
  3. Salt movement (Prior and Hooper, 1999)
  4. Depositional loading (Coleman and Prior, 1982)
  5. Hydrostatic loading (Trincardi et al., 2003)
  6. Ocean-bottom currents (Locat and Lee, 2002)
  7. Biological erosion in submarine canyons (Dillon and Zimmerman, 1970)
  8. Gas hydrate decomposition (Popenoe et al., 1993; Sultan et al., 2004)
Subaerial & submarine
 
 
 
 
 
 
 
Submarine
 
 
Submarine
 
Submarine
 
Submarine
 
Submarine
 
Submarine
 
 
Submarine
 
 
 
 
 
 
 
 
 
 
 
 
Intermediate-term events: hundreds to thousands of years
  1. Sea-level lowstand (Shanmugam and  Moiola, 1988; Vail et al., 1991)
Submarine
 
 
Long-term events:
thousands to millions of years

* Although human activity is considered to be the second most common triggering mechanism (next to earthquakes) for known historic submarine mass (Mosher et al., 2010), it is Irrelevant for interpreting ancient rock record.

** Some tectonic events may extend over millions of years

 

              The triggering mechanisms are grouped into three major types based on their duration: (1) short-term triggering events that last for only a few minutes to several hours or days, (2) intermediate-term triggering events that last for hundreds to thousands of years, and (3) long-term triggering events that last for thousands to millions of years (Shanmugam, 2012a). Conceivably, some intermediate-term events may last for a longer duration. The importance here is that short-term events and long-term events are markedly different in their duration. Furthermore, these triggering mechanisms should not be confused with depositional processes, such as debris flows. Also, more than one triggering mechanism can cause debris flows at a given time in one site. However, there are no objective criteria to infer either the triggering mechanism or the transportational process from the depositional record.



Figure 1. Map showing the site of Chicxulub meteorite impact at the K-T boundary in Yucatan, Mexico. Stars represent approximate locations of mass-transport deposits and tsunami-related deposits associated with the Chicxulub impact at the K-T boundary boundary. Generalized outline of Lower Tertiary Wilcox Trend is from several sources. After Shanmugam (2012b).

 
 



Figure 2. Depositional model showing the link between tsunamis and deep-water deposition. (A) 1. Triggering stage in which earthquakes trigger tsunami waves. 2. Tsunami stage in which an incoming (up-run) tsunami wave increases in wave height as it approaches the coast. 3. Transformation stage in which an incoming tsunami wave erodes and incorporates sediment, and transforms into sediment flows. (B) 4. Depositional stage in which outgoing (backwash) sediment flows (i.e., debris flows and turbidity currents) deposit sediment in deep-water environments. Suspended mud created by tsunami-related events would be deposited via hemipelagic settling. After Shanmugam (2006a).
 
 
 
 
 
 

Figure 3. A. Highstand sedimentological model showing calm shelf waters and limited

extent of sediment transport in the shoreface zone (short green arrow) during fair-weather periods. Shoreface bottom-current velocities during fair weather are in the range of 10-20 cm s-1. The shelf edge at 200 m water depth separates shallow- water (shelf) from deep-water (slope) environments. (B) Highstand sedimentological model showing sediment transport on the open shelf, over the shelf edge, and in submarine canyons during periods of tropical cyclones (storm weather) into deep water (long red arrow). Mass-transport processes are commonly induced by intense cyclones (e.g., Hurricane Katrina, 2005). Modified after Shanmugam (2008a).

 

The obsolescence of sea-level lowstand model

 
In the petroleum industry, the sea-level lowstand model is the perceived norm for explaining the timing of deep-water sands (Figure 4A). 
 
 

Figure 4. A. Conventional sea-level model showing the popular belief that deep-water deposition of sand occurs only during periods of lowstand and deposition of mud occurs during periods of highstand. The present highstand is estimated to represent a period of 20,000 years. BP=before present. B. 200,000 cyclones are estimated to occur during the present highstand in the Indian Ocean (Bay of Bengal) and the Atlantic Ocean. C. 140,000 tsunamis are estimated to occur during the present highstand in the Pacific Ocean. The implication is that sands can be deposited during periods of sea-level highstands, thus making the conventional model obsolete.  After Shanmugam (2008a).

 

Saller et al. (2006), for example, attributed the timing of reservoir sands in the Kutei Basin in the Makassar Strait (Indonesian Seas) to a lowstand of sea-level. Nevertheless, the location of the Kutei Basin is frequently affected by earthquakes, volcanoes, tsunamis, tropical cyclones, monsoon floods, the Indonesian throughflow, and M2 baroclinic tides (Shanmugam, 2008b). These daily activities of the solar system (e.g., earthquakes, meteorite impacts, tsunamis, cyclonic waves, etc.) do not come to a halt during sea-level lowstands. In tectonically and oceanographically  tumultuous locations, such as the Indonesian Seas, the short-term events are the primary triggering mechanisms of deepwater sediment failures and they occur in a matter of hours or days during long periods of both highstands and lowstands (Shanmugam, 2008a).

              Deep-water Paleocene sand (100 m thick) of the Lower Tertiary Wilcox trend, which occurs above the K-T boundary in the BAHA #2 wildcat test well, has been interpreted as “lowstand” turbidite fan in the northern Gulf of Mexico (Meyer et al., 2007). However, because of the opportune location of the Lower Tertiary Wilcox trend and the stratigraphic position and age, the drilled Paleocene sand could alternatively be attributed to the Chicxulub impact and related seismic shocks and tsunamis (Figure 1). Tsunami-related deposition on continental margins has been discussed by Shanmugam (2006a) (Figure 2). Such alternative real-world possibilities are often overlooked because of the prevailing mindset of the sea-level lowstand model.

              Shanmugam (2008a) discussed the importance of tropical cyclones in understanding deep-water sand deposition during periods of sea-level highstands (Figure 4). The Hurricane Hugo (Hubbard, 1992), which passed over St. Croix in the U.S. Virgin Islands on 17 September 1989, had generated winds in excess of 110 knots (204 km hr-1, Category 3 in the Saffir-Simpson Scale) and waves 6–7 m in height. In the Salt River submarine canyon (>100 m deep), offshore St. Croix, a current meter measured net down-canyon currents reaching velocities of 2 m s-1 and oscillatory flows up to 4 m s-1. Hurricane Hugo had caused erosion of 2 m of sand in the Salt River Canyon at a depth of about 30 m. A minimum of 2 million kg of sediment were flushed down the Salt River Canyon into deep water (Hubbard, 1992). The transport rate associated with Hurricane Hugo was 11 orders of magnitude greater than that measured during fair-weather period. In the Salt River Canyon, much of the soft reef cover (e.g., sponges) had been eroded away by the power of the hurricane. Debris composed of palm fronds, trash, and pieces of boats found in the canyon were the evidence for storm-generated debris flows. Storm-induced sediment flows have also been reported in a submarine canyon off Bangladesh (Kudrass et al., 1998), in the Capbreton Canyon, Bay of Biscay in SW France (Mulder et al., 2001), in the Cap de Creus Canyon in the Gulf of Lions (Palanques et al., 2006), and in the Eel Canyon, California (Puig et al., 2003), among others. In short, sediment transport in modern submarine canyons is accelerated by tropical cyclones and tsunamis in the world’s oceans during the present sea-level highstand (Shanmugam, 2008a). In fact, empirical data clearly show that tropical cyclones and tsunamis are the two most important phenomena in transporting sediment into the deep sea during the present sea-level highstand (Figure 4). Therefore, the sea-level lowstand model is obsolete for explaining the triggering of deep-water SMTDs worldwide (Shanmugam, 2007, 2008a, 2013c).

 

References

 

Bea, R. G., S. G. Wright, P.Sicar, and A. W.  Niedoroda, 1983, Wave-induced slides in South Pass Block 70, Mississippi delta:  Journal of Geotechnical Engineering, v. 109, p. 619-644.

Boyd, R., K. Ruming, I. Goodwin, S M. Sandstrom, and C. Schröder-Adams, 2008, Highstand transport of coastal sand to the deep ocean: A case study from Fraser Island, southeast Australia: Geology, v. 36. p. 15–18.

Brönnimann, C.S., 2011, Effect of Groundwater on Landslide Triggering: Lausanne, Switzerland, École Polytechnique Fédérale de Lausanne, Ph.D. Dissertation, Unpublished, 239 p.

Cannon, S.H., R.M.  Kirkham, and M. Parise, 2001, Wildfire-related debris-flow initiation processes, Storm King Mountain, Colorado: Geomorphology, v. 39, p. 171-188.

Claeys, P., W. Kiessling, and W. Alvarez, 2002, Distribution of Chicxulub ejecta at the Cretaceous-Tertiary boundary, in C. Koeberl, and K.G. MacLeod, eds., Catastrophic Events and Mass Extinctions: Impacts and Beyond: Boulder, Colorado, Geological Society of America Special Paper 356, p. 55–68.

Coleman, J.M., and D.B. Prior, 1982, Deltaic environments, in P. A. Scholle, and  D. Spearing, eds., Sandstone Depositional Environments: AAPG Memoir 31, p. 139–178.

Collot, J.-Y., K. Lewis, G. Lamarche, and S. Lallemand, 2001, The giant Ruatoria debris avalanche on the northern Hikurangi margin, New Zealand: Result of oblique seamount subduction: Journal of Geophysical Research, v. 106 (B9), p. 19,271-19,297. doi:10.1029/2001JB900004.

Dan, G., N. Sultan, and B. Savoye, 2007, The 1979 Nice harbour catastrophe revisited: Trigger mechanism inferred from geotechnical measurements and numerical modeling: Marine Geology, v. 245, p. 40–64.

Dillon, W.P. and H.P.  Zimmerman, 1970, Erosion by biological activity in two New England submarine canyons: Journal of Sedimentary Petrology, v. 40, p. 542-547.

Dysthe, K., H.E. Krogstad, and P. Műller, 2008, Oceanic rogue waves: Annual Reviews of Fluid Mechanics, v. 40, p. 287-310.

Elverhoi, A., H. Norem, E. S. Anderson, J. A.  Dowdeswell, I. Fossen, H. Haflidason, N. H. Kenyon, J. S. Laberg, E. L. King, H. P. Sejrup, A. Solheim, and T. Vorren, 1997, On the origin and flow behavior of submarine slides on deep-sea fans along the Norwegian-Barents Sea continental margin: Geo-Marine Letters, v. 17, p. 119-125.

Elverhøi, A., F. De Blasio, F.A. Butt, D. Issler, C.B. Harbitz, L.Engvik, et al., 2002, Submarine mass-wasting on glacially-influenced continental slopes: processes and dynamics, in  J.A. Dowdeswell, and C.  O’Cofaigh, eds., Glacier-Influenced Sedimentation on High-Latitude Continental Margins: London, Geological Society, Special Publications 203, p. 73-87.

Feeley K., 2007, Triggering Mechanisms of Submarine Landslides. Research Report: Boston, MA, Department of Civil and Environmental Engineering, Northeastern University, 45 p.

Greene, H. G., L. Y. Murai, P. Watts, N. A. Maher, M. A. Fisher, C. E. Paull,  and P. Eichhubl, 2006, Submarine landslides in the Santa Barbara Channel as potential tsunami sources: Natural Hazards and Earth System Sciences, v. 6, p. 63–88.

Heezen, B. C., and M. Ewing, 1952, Turbidity currents and submarine slumps and the 1929 Grand Banks earthquake: American Journal of Science, v. 250, p. 849–873.

Hubbard, D.K. 1992, Hurricane-induced sediment transport in open shelf tropical systems–An example from St. Croix, U.S. Virgin Islands: Journal of Sedimentary Petrology, v. 62, p. 946–960.

Kudrass, H.R., Michels, K.H. and Wiedicke, M., 1998, Cyclones and tides as feeders of a submarine canyon off Bangladesh: Geology, v. 26, p. 715-718.

Lee, H.J., 2009, Timing of occurrence of large submarine landslides on the Atlantic Ocean margin: Marine Geology, v. 264, p. 53-64.

Locat, J. and H.J. Lee, 2002, Submarine landslides: advances and challenges: Canadian Geotechnical Journal, v. 39(1), p. 193-212.

Masson, D.G., C.B. Harbitz, R.B. Wynn, G. Pedersen, and F. Løvholt, 2006, Submarine landslides: processes, triggers, and hazard prevention: Royal Society of London Transactions, Series A 364 (1845), p. 2009-2039.

Meyer, D., Zarra, L. and Yun, J., 2007, From BAHA to jack, evolution of the lower tertiary Wilcox trend in the deepwater Gulf of Mexico: Sedimentary Record, v. 5, p. 4-9.

Mosher, D.C., L. Moscardelli, R.C. Shipp, J.D. Chaytor, C.D.P. Baxter, H.J. Lee, and R. Urgeles, 2010, Submarine Mass Movements and Their Consequences, in D.C. Mosher et al., eds., Submarine Mass Movements and Their Consequences: Advances in Natural and Technological Hazards Research, v. 28, p. 1-8.

Mulder, T., Migeon, S. Savoye, B. and Faugeres, J.-C., 2001, Inversely graded turbidite sequences in the deep Mediterranean. A record of deposits from flood-generated turbidity currents? Geo-Marine Letters, v. 21, p. 86-93.

Palanques, A., X. Durieu de Madron, P. Puig, J. Fabres, J. Guillén, A. M. Calafat, M. Canals, and J. Bonnin, 2006, Suspended sediment fluxes and transport processes in the Gulf of Lions submarine canyons. The role of storms and dense water cascading: Marine Geology, v. 234, p. 43– 61.

Petley, D., 2012, Global patterns of loss of life from landslides: Geology, v. 40(10), p. 927–930. doi:10.1130/G33217.1

Popenoe, P., E.A. Schmuck,  and W.P. Dillon,, 1993, The cape fear landslide; slope failure associatedwith salt diapirism and gas hydrate decomposition, in  W.C. Schwab, H.J. Lee, , and D.C. Twichell, eds., Submarine Landslides: Selected Studies in the U.S. Exclusive Economic Zone: U.S. Geological Survey Bulletin 2002, p. 40–53.

Prior, D. B., and J.R. Hooper, 1999, Sea floor engineering geomorphology: recent achievements and future directions: Geomorphology, v. 31, p. 411-439.

Prior, D. B., J. N. Suhayda, N.-Z. Lu, B. D. Bornhold, G. H. Keller, W. J. Wiseman, L. D. Wright, and Z.-S. Yang, 1989, Storm wave reactivation of a submarine landslide: Nature, v. 341, p. 47-50.

Puig, P., Ogston, A.S. Mullenbach, B.L. Nittrouer, C.A. and Sternberg, R.W., 2003, Shelf-to-canyon sediment-transport processes on the Eel continental margin (northern California):  Marine Geology, v. 193, p. 129-149

Saller, A. H., R. Lin, and J. Dunham, 2006, Leaves in turbidite sands: The main source of oil and gas in the deep-water Kutei Basin, Indonesia: AAPG Bulletin, v. 90, p. 15851608, doi:10.1306/04110605127.

.Shanmugam, G., 2006a, The tsunamite problem: Journal of Sedimentary Research, v. 76, p. 718730.

Shanmugam, G., 2006b, Deep-water processes and facies models: Implications for sandstone petroleum reservoirs: Amsterdam, Elsevier, Handbook of petroleum exploration and production, v. 5, 476 p.

Shanmugam, G., 2007, The obsolescence of deep-water sequence stratigraphy in petroleum geology: Indian Journal of Petroleum Geology, v. 16 (1), p. 1-45.

Shanmugam, G., 2008a, The constructive functions of tropical cyclones and tsunamis on deep-water sand deposition during sea level highstand: Implications for petroleum exploration: AAPG Bulletin, v. 92, p. 443471.

Shanmugam, G., 2008b, Leaves in turbidite sand: The main source of oil and gas in the deep-water Kutei Basin, Indonesia: Discussion: AAPG Bulletin, v. 92, p. 127137.

 

 

Shanmugam, G., 2012a, Process-sedimentological challenges in distinguishing paleo-tsunami deposits, in A. Kumar and I. Nister, eds., Paleo-tsunamis: Natural Hazards, v. 63, p. 530.

Shanmugam, G., 2012b, New perspectives on deep-water sandstones: Origin, recognition, initiation, and reservoir quality: Amsterdam, Elsevier, Handbook of petroleum exploration and production, v. 9, 524 p.

Shanmugam, G., 2013a, Comment on “Internal waves, an underexplored source of turbulence events in the sedimentary record” by L. Pomar, M. Morsilli, P. Hallock, and B. Bádenas [Earth-Science Reviews, 111 (2012), 56–81]: Earth-Science Reviews, v. 116, p. 195–205.

Shanmugam, G., 2013b, Modern internal waves and internal tides along oceanic pycnoclines: Challenges and implications for ancient deep-marine baroclinic sands: AAPG Bulletin, v. 97(5), p. 799-843.

Shanmugam, G. 2013c, Slides, Slumps, Debris Flows, and Turbidity Currents. In: Earth Systems and Environmental Sciences Reference Module. Elsevier (online), in press.

Shanmugam, G., and R. J.  Moiola, 1988, Submarine fans: Characteristics, models, classification, and reservoir potential: Earth-Science Reviews, v. 24, p. 383– 428.

Sultan, N., P. Cochonat, M. Canals, A. Cattaneo, B. Dennielou, H. Haflidason, J.S. Laberg, D. Long, J. J. Mienert, F. Trincardi, R. Urgeles, T.O. Vorren, and C. Wilson, 2004, Triggering mechanisms of slope instability processes and sediment failures on continental margins: a geotechnical approach: Marine Geology, v. 213(1-4), p. 291-321.

.Tilling, R.I., L. Topinka, and D.A. Swanson, 1990, Eruptions of Mount St. Helens: Past, Present, and Future: U.S. Geological Survey Special Interest Publication, 56 p.

Trincardi, F., A. Cattaneo, A. Correggiari, S. Mongardi,  A. Breda, and A. Asioli, 2003, Submarine slides during relative sea level rise: two examples from the eastern Tyrrhenian margin, in J. Locat, and J. Mienert, eds., Submarine mass movements and their consequences: Dordrecht, Kluwer Academic Publishers, p.  469-478.

Urgeles, R., M. Canals, J. Baraza, B. Alonso, and D. Masson, 1997, The most recent megalandslides of the Canary Islands: El Golfo debris avalanche and Canary debris flow, west El Hierro Island:  Journal of Geophysical Research, v. 102(B9), p. 20305-20323.

Vail, P. R., F. Audemard, S. A., Bowman, P. N. Eisner, and C. Perez-Cruz, 1991, The stratigraphic signatures of tectonics, eustacy and sedimentology - an overview, in G. Einsele, W. Ricken, and A. Seilacher, eds., Cycles and events in stratigraphy: Berlin, Springer-Verlag, p. 618-659.