207
Views
4
CrossRef citations to date
0
Altmetric
Short communication

Reconsidering the seawater-density parameter in hydrodynamic flow transport equations for coastal boulders

ORCID Icon & ORCID Icon
Pages 363-370 | Received 27 Jun 2019, Accepted 12 Jan 2020, Published online: 29 Jan 2020

ABSTRACT

Existing hydrodynamic flow transport equations for coastal boulder transport are useful for estimating post-event the characteristics of extreme storm waves and tsunamis. However, the effect of suspended sediment concentration (SSC) on seawater density is normally ignored. This is unrealistic given that turbulent runup flows easily entrain available fine sediment. Proper consideration of SSC can be encouraged by including a mixed-fluid density coefficient (Cρ) as a multiplier for clear-seawater density, where elevated sediment content can be assumed and estimated. Minimum flow velocities required for boulder transport are shown to reduce as sediment concentrations increase.

Introduction

Coastal boulder proxies for extreme wave events

Geological evidence is increasingly being utilised by coastal scientists to identify and characterise the nature of palaeo-high energy wave events on coasts such as tsunamis and storms. Such evidence enables comparisons in magnitude between events and, through cautious interpretation of age-datable material, provides insight into the temporal frequencies of the strongest wave events experienced on exposed coastlines over timescales much longer than written records permit (Yu et al. Citation2009; Terry et al. Citation2015, Citation2016; Lau et al. Citation2016). Coastal boulder deposits represent one kind of geological evidence that may be investigated (; Hearty Citation1997; Etienne Citation2012; Yu et al. Citation2012). First described as a potential proxy for big storms in the early twentieth century (Hedley and Taylor Citation1907), boulder analysis is now a well-established method in coastal geomorphological research, and has been applied across a range of tropical and extra-tropical regimes (Etienne and Paris Citation2010; Goto et al. Citation2010; Engel and May Citation2012; Salzmann and Green Citation2012; Lau et al. Citation2015; Kennedy et al. Citation2017). Measuring the size, position and orientation of transported coastal boulders reveals clues concerning the power and direction of the onshore flows that deposited them, as generated by extreme waves striking the coast (Scicchitano et al. Citation2007; Goto et al. Citation2009; Terry et al. Citation2013; May et al. Citation2015; Shah-Hosseini et al. Citation2016).

Figure 1. Sampling and measurement of coastal boulder fields: reef-derived carbonate clasts near Chaweng, Ko Samui island, southern Thailand (above), and large clasts of volcanic breccia and limestone on Ludao island, SE Taiwan (below).

Figure 1. Sampling and measurement of coastal boulder fields: reef-derived carbonate clasts near Chaweng, Ko Samui island, southern Thailand (above), and large clasts of volcanic breccia and limestone on Ludao island, SE Taiwan (below).

Hydrodynamic flow transport equations

Hydrodynamic flow transport (HFT) equations quantifying relationships between coastal boulder size and minimum wave height required for transport were originally developed by Nott (Citation1997, Citation2003). Nott’s HFT equations were rearranged by Nandasena et al. (Citation2011) to enable calculation of the minimum flow velocity (MFV) necessary to initiate boulder movement, according to various possible modes of transport: sliding, rolling or lifting. In the HFT equations below (Nandasena et al. Citation2011), u is the calculated MFV in m/s and other parameters are as listed in .

Table 1. Parameters in the widely-used hydrodynamic flow transport equations for transport initiation of coastal boulders.

(A) For boulder sliding:(1) u22(ρs/ρw1)gc(μscosθ+sinθ)Cd(c/b)+μsCl(1) (B) For boulder rolling:(2) u22(ρs/ρw1)gc(cosθ+(c/b)sinθ)Cd(c2/b2)+Cl(2) (C) For boulder lifting:(3) u22(ρs/ρw1)gccosθCl(3)

The seawater density problem

Modelling fluid mechanics at both small and large scales necessarily requires approximation of the processes involved. Models depicting the objects of study are therefore simplifications, in part to avoid superfluous detail and to reduce expensive computational effort in their application (Munson et al. Citation2013; Çengel and Cimbala Citation2015). Broadly speaking, the current HFT equations represent a one-dimensional depiction of boulder transport, which is appropriate for pragmatic approximations.

Regarding the application of HFT equations, however, a persistent issue can be identified with the seawater density parameter (ρw). Although ρw is a variable, most authors treat it as a de facto constant by simply substituting a density value of 1.025 g/cm3 or similar. This value corresponds to the density of clear seawater and ignores any contribution of suspended sediments in the turbulent flow. Yet, sediment and seawater mixtures are more viscous fluids than clear seawater, potentially altering the fluid dynamics between the flow and boulders interacting with the flow stream.

González et al. (Citation2007) emphasised that sediment content must be taken into account when computing tsunami hydrodynamic forces. For modelling tsunami impacts on coastal structures, Jiffry et al. (Citation2015) and Attary et al. (Citation2017) are among the few studies using fluid density values greater than for clear seawater (1200 kg/m3 and 1080–1320 kg/m3, respectively). Nonetheless, the effect of elevated sediment concentrations on hydrodynamics is recognised as a continuing source of epistemic uncertainty in tsunami sediment transport modelling, and sediment-enhanced fluid density is not yet addressed in models for coastal boulder transport (Jaffe et al. Citation2016). Kain et al. (Citation2012) believe that water alone is incapable of moving large clasts, and that movement is only possible by debris-rich flows. They also advocated including the effects of increased fluid density from fine sediment in the water.

We agree with these views. In response, our aim here is to draw closer attention to the contribution of suspended sediment on fluid density. One way to avoid suspended sediment being overlooked or ignored in the application of existing HFT equations for coastal boulder transport is to introduce a simple coefficient that adjusts the clearwater fluid density according to the suspended sediment concentration (SSC), where this is known, assumed or can be estimated.

Anticipating elevated SSC in tsunamis and breaking storm waves

Although direct empirical evidence is limited, it is intuitive that sediment fractions will be non-zero values in high-energy waves and fast-moving runup (and backwash) on affected coastlines. Beaches, reefs and deltas provide sources of unlithified sediment for erosion and entrainment near the shoreline. The most intense sediment transport in the coastal zone is often beneath breaking waves in the surf on a beach. Neglecting wave breaking processes might therefore lead to an underestimate of SSC in the turbulent surf zone (Soulsby Citation1997). Accounts of black or brown tsunami waves (Houtz Citation1962; González et al. Citation2007) and deposition of sand and mud layers inland (Etienne and Terry Citation2012; Yamada et al. Citation2014) likewise indicate elevated SSC. From investigation of the 2018 Indonesian (Sulawesi) tsunami, Sassa and Takagawa (Citation2019) suggest that significant earthquake-generated coastal sediment liquefaction can result in a liquefied gravity flow-induced tsunami. Yoshii et al. (Citation2018) similarly conceive that earthquake shock can liquefy sediment prior to tsunami arrival. For localised tsunamis caused by subaerial or submarine landslides (Tappin et al. Citation2008; Goff and Terry Citation2016; Lau et al. Citation2018), it is plausible that the landslide mass would contribute to elevated suspended sediment concentrations.

Methods

Incorporating suspended sediments

Consider the three situations illustrated in . a shows an evenly distributed flow stream where the sediments are equally mixed. Such a scenario is possible in highly turbulent conditions. b represents another situation where a depth-profile exists. Sediment-induced density stratification occurs where large amounts of sediment are suspended in the water column (Apotsos et al. Citation2011), although stratification is largely absent in published tsunami sand transport models (Jaffe et al. Citation2016). If the entrained sediments are a mix of fines and coarse material, transported in suspension and in traction, then the flow will potentially have this type of gradient. c represents a heterogeneous mixture with an instantaneous profile showing high levels of concentration in the upper layer. Such situations will temporarily create an uneven momentum profile. The purpose of this exercise is to demonstrate that there are various possible within-flow sediment distributions during flow contact with any coastal boulders near the shoreline.

Figure 2. Three simplified variations of an incoming flow stream with different distributions of sediments, as represented by black dots. Side views are shown. (a) A homogenously mixed flow stream where the sediment content is evenly distributed. (b) A depth-gradient where gravitational effects give higher sediment concentrations in the lower layers of the flow than the uppermost layers. (c) An unevenly mixed flow stream with a densely concentrated area at the top left ‘corner.’

Figure 2. Three simplified variations of an incoming flow stream with different distributions of sediments, as represented by black dots. Side views are shown. (a) A homogenously mixed flow stream where the sediment content is evenly distributed. (b) A depth-gradient where gravitational effects give higher sediment concentrations in the lower layers of the flow than the uppermost layers. (c) An unevenly mixed flow stream with a densely concentrated area at the top left ‘corner.’

b and 2c are two-dimensional problems that would introduce extra complexity if we were to model them. An alternative, however, is to continue using the current HFT equations in their one-dimensional form, but to multiply clear seawater density (ρw) by a ‘mixed-fluid density coefficient’ (Cρ), where appropriate. This avoids a de facto assumption of a zero-sediment concentration. A mixed-fluid density value (ρm) can be calculated as follows, based on volume ratios of sediment and seawater:(4) ρm=(fw×ρw)+(fs×ρs)g/cm3(4) where, ρm is the average density of the seawater and sediment mix, fw is the seawater fraction by volume between 0 and 1, fs is the sediment fraction by volume between 0 and 1, ρw is the density of clear seawater (1.025 g/cm3), ρs is the sediment density, and fw + fs = 1

The mixed-fluid density coefficient (Cρ) is then calculated:(5) Cρ=ρm/ρw(5)

The Cρ coefficient can now be incorporated into Equations (1)–(3) as a multiplier for clear seawater density (ρw). The revised HFT equations for boulder transport are written thus:

(A) For boulder sliding:(6) u22(ρs/(Cρ.ρw)1)gc(μscosθ+sinθ)Cd(c/b)+μsCl(6)

(B) For boulder rolling:(7) u22(ρs/(Cρ.ρw)1)gc(cosθ+(c/b)sinθ)Cd(c2/b2)+Cl(7)

(C) For boulder lifting:(8) u22(ρs/(Cρ.ρw)1)gccosθCl(8)

Results and Discussion

Effects of increasing sediment concentrations on fluid density

shows the increase in mixed-fluid density (ρm) with increasing seawater sediment concentrations. The density coefficient (Cρ) increases correspondingly (). Appropriate ranges of Cρ values can be selected from , according to the sediment type, density (ρs) and estimations of SSC (see below).

Figure 3. Values of the mixed-fluid density coefficient (Cρ) for a flow stream of seawater containing suspended sediment, according to sediment concentrations and sediment (grain) density, based on equation (5).

Figure 3. Values of the mixed-fluid density coefficient (Cρ) for a flow stream of seawater containing suspended sediment, according to sediment concentrations and sediment (grain) density, based on equation (5).

Table 2. Relationships between the concentration of example types of sediments in seawater and the resulting density of the seawater and sediment mix (ρm), based on equation (4).

It is seen that incorporating increasing Cρ values into the HFT equations has an inverse influence on the calculated minimum flow velocity (MFV) required to transport coastal boulders. This means that lower flow speeds are sufficient for boulder transport in seawater containing greater proportions of sediment (). For example, for a reef limestone boulder subjected to flow in seawater containing 5% or 20% sediment concentration of similar carbonate lithology (i.e. coral sands), the MFV required for boulder movement decreases to 0.96 and 0.85 of the equivalent MFV in clear seawater, respectively. At the same sediment concentrations, for a basalt boulder in a flow stream containing basaltic sands, the corresponding reductions in MFV required for boulder transport are 0.94 and 0.77, as compared to flow transport in clear seawater. has an important implication for MFV results reported in the literature from coastal boulder analysis: MFVs are likely to be overestimations because boulder transport in clear seawater is normally assumed.

Figure 4. Reduction in the minimum flow velocity (MFV) required for coastal boulder transport in a fluid mixture of seawater and sediment. The MFV is expressed as a fraction of the MFV calculated for clear seawater without suspended sediment. Two examples are given for coastal boulders of different lithologies: limestone and basalt. The sediments entrained in seawater are assumed to have the same lithology as the coastal boulders. The graphs apply to any mode of boulder transport – sliding, rolling or lifting.

Figure 4. Reduction in the minimum flow velocity (MFV) required for coastal boulder transport in a fluid mixture of seawater and sediment. The MFV is expressed as a fraction of the MFV calculated for clear seawater without suspended sediment. Two examples are given for coastal boulders of different lithologies: limestone and basalt. The sediments entrained in seawater are assumed to have the same lithology as the coastal boulders. The graphs apply to any mode of boulder transport – sliding, rolling or lifting.

Inferring realistic sediment concentrations

Inherent challenges exist in collecting empirical measurements of sediment concentrations during HEW events. Defining sediment concentrations that are representative of all natural situations is therefore difficult. Goto et al. (Citation2014) suggested 0–10% (average 2%) sediment concentrations within the 2011 Tohoku-oki (Japan) tsunami runup as it crossed the flat topography of the Sendai Plain. Greater concentrations might be anticipated at the shoreward edge at initial wave impact. Battisto et al. (Citation1998) reported SSC (sand and mud) values <1% in the surf zone at the Duck research station in North Carolina, measured in a sampling campaign during a storm in April 1997, but believed their concentrations were significant underestimates. For tsunami transport of sediment released by underwater landslides, a sediment concentration of 55 ± 15 kg/m3 (∼3% concentration, assuming ρs = 2.0 g/cm3) was given by Brizuela et al. (Citation2019) as their most conservative estimate. In the ‘Tsunami Sand Transport Laboratory Experiments’ of Yoshii et al. (Citation2018), sediment concentrations were mostly <5%, but considerable differences existed between experimental cases, with some situations producing very high concentrations up to 14% and 30%. A guideline value of 7% was assumed for vertically-averaged tsunami sediment concentration (Applied Technology Council Citation2019), but such averages mask the possibility of much higher concentrations at the base of flows produced by suspended sediment-induced density stratification and hindered particle settling (Li and Huang Citation2013).

In modelling experiments by Tehranirad et al. (Citation2016), tsunami SSCs reached maximum values of 100 kg/m3 (∼4% concentration, assuming ρs = 2.65 g/cm3). According to process-based modelling studies by Apotsos et al. (Citation2011), erosive capabilities of tsunami flows may produce SSCs >200 kg/m3 (>8% concentration, assuming ρs = 2.65 g/cm3) in runup and up to 600 kg/m3 (∼23% concentration, ρs = 2.65 g/cm3) in high-velocity backwash in the offshore zone. Li and Huang (Citation2013) also modelled SSC values up to >750 kg/m3 (>28% concentration) at the base of backwash flows, when allowing for density stratification and hindered settling. In their numerical simulations of sediment transport in the 2011 Tohoku tsunami, Yamashita et al. (Citation2016) set maximum permissible SSC to 38%, while Larsen et al. (Citation2017) imposed a maximum boundary value of 60% concentration.

Whether or not instantaneous maximum sediment concentrations above 20% in breaking storm waves or tsunamis are realistic remains to be confirmed. Until such time as further empirical measurements are available for real HEW events or from wave tank experiments, conservative maximum SSC values within the range 2–10% are suggested. However, recall that other types of hyperconcentrated flows do exist in nature, such as sea bottom fluid-mud layers (Sheremet et al. Citation2005a, Citation2005b), erosion gully flow (Jiongxin Citation2004), and lahars that are able to maintain sediment concentrations >50% (Pierson and Scott Citation1985; Cronin et al. Citation1997; Lavigne and Thouret Citation2002).

Caveats to findings

Several caveats need to be mentioned. (1) Any change in fluid properties will have implications for the contact dynamics with obstacles positioned within the flow. Chosen values for the coefficient of drag (Cd) and the coefficient of lift (Cl) to be substituted in HFT equations would also need to incorporate the increased density of the mixed fluid. (2) It could be argued that temperature and salinity also affect seawater density. However, reasonable estimations indicate that such effects are minimal. For seawater temperatures of 1–35°C, representing environments ranging from high-latitude winters to the hot tropics, seawater density (ρw) only changes by 0.77% (Haynes Citation2017). Similarly, 0–4% salt percentage (by mass) for freshwater or seawater causes ρw to vary by just 2.72%. Accordingly, salinity and temperature considerations can largely be discounted. (3) The contribution to boulder movement by the transfer of momentum from collisions of pebbles or cobbles entrained as traction load is unknown. (4) A fluidised bedload of mobile sands and gravels at the base of the flow stream might reduce friction, thereby increasing the efficiency of boulder transport.

Addressing the issues above is beyond the present scope, but the purpose of mentioning them is to recognise the implications and limitations of including a mixed-fluid density coefficient to represent elevated sediment concentrations in seawater. Such additional effects of sediment content on boulder transport processes could be investigated and incorporated into the HFT equations in future refinements. Field measurements during extreme storm waves and quasi-empirical approaches in wave-tank experiments offer good possibilities for further work.

Conclusions

We recognise the popularity, value and simplicity of existing hydrodynamic flow transport (HFT) equations for coastal boulder transport to estimate characteristics of tsunamis or storm waves as they strike coastlines. One-dimensional HFT equations are easily applied and are widely used. Yet, many workers overlook or ignore the influence of sediment content on fluid density. This is an omission, given the probability of significant sediment entrainment in the high-velocity turbulent flows generated by tsunami runup or breaking storm waves. Incorporating a simple ‘mixed-fluid density coefficient’ (Cρ) as a multiplier for clear seawater density (ρw) can easily adjust for elevated sediment concentrations, where these are known, assumed or can be estimated. The additional coefficient introduces no greater complexity when applying the existing HFT equations, but encourages proper consideration of sediment concentrations in the flow responsible for coastal boulder transport.

Acknowledgements

Miss Maitha Eisa Almeer is thanked for her efforts with literaturesearching. Two anonymous reviewers provided helpful advice for improving the original manuscript.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work was supported by Zayed University [Grant Number R19088].

References

  • Apotsos A, Gelfenbaum G, Jaffe B. 2011. Process-based modeling of tsunami inundation and sediment transport. Journal of Geophysical Research: Earth Surface. 116:F01006. doi: 10.1029/2010JF001797
  • Applied Technology Council. 2019. Guidelines for Design of Structures for Vertical Evacuation from Tsunamis (3rd ed.). FEMA P-646. Federal Emergency Management Agency. https://www.fema.gov/media-library-data/1570817928423-55b4d3ff4789e707be5dadef163f6078/FEMAP646_ThirdEdition_508.pdf.
  • Attary N, Unnikrishnan VU, Lindt JWVD, Cox DT, Barbosa AR. 2017. Performance-based tsunami engineering methodology for risk assessment of structures. Engineering Structures. 141:676–686. doi: 10.1016/j.engstruct.2017.03.071
  • Battisto GM, Friedrichs CT, de Kruif A, Rijks DC. 1998. Data Report: Pump Sampling and Sediment Analysis in Support of the Sensor Insertion System Duck, N.C. April and October, 1997. VIMS Data Report No. 56, Virginia Institute of Marine Science, VA 23062, 68pp.
  • Brizuela N, Filonov A, Alford MH. 2019. Internal tsunami waves transport sediment released by underwater landslides. Nature Scientific Reports. 9:10775. doi: 10.1038/s41598-019-47080-0
  • Çengel Y, Cimbala JM. 2015. Fluid mechanics. 5th ed. New Delhi: McGraw Hill.
  • Cronin S, Neall V, Lecointre JA, Palmer A. 1997. Changes in Whangaehu river lahar characteristics during the 1995 eruption sequence, Ruapehu volcano, New Zealand. Journal of Volcanology and Geothermal Research. 76:47–61. doi: 10.1016/S0377-0273(96)00064-9
  • Engel M, May SM. 2012. Bonaire's boulder fields revisited: evidence for Holocene tsunami impact on the Leeward Antilles. Quaternary Science Reviews. 54:126–141. doi: 10.1016/j.quascirev.2011.12.011
  • Etienne S. 2012. Marine inundation hazards in French Polynesia: geomorphic impacts of tropical cyclone Oli in February 2010. In: Terry JP, Goff J, editors. Natural Hazards in the Asia-Pacific Region. London: Geology Society of London. Special Publication, vol. 361; p. 21–39.
  • Etienne S, Paris R. 2010. Boulder accumulations related to storms on the south coast of the Reykjanes Peninsula (Iceland). Geomorphology. 114:55–70. doi: 10.1016/j.geomorph.2009.02.008
  • Etienne S, Terry JP. 2012. Coral boulders, gravel tongues and sand sheets: features of coastal accretion and sediment nourishment by Cyclone Tomas (March 2010) on Taveuni Island, Fiji. Geomorphology. 175–176:54–65. doi: 10.1016/j.geomorph.2012.06.018
  • Goff J, Terry JP. 2016. Tsunamigenic landslides – the Pacific Islands blind spot? Landslides. 13:1535–1543. doi: 10.1007/s10346-015-0649-3
  • González FI, Bernard E, Dunbar P, Geist E, Jaffe B, Kânoǧlu U, Locat J, Mofjeld H, Moore A, Synolakis C, et al. (Science Review Working Group). 2007. Scientific and technical issues in tsunami hazard assessment of nuclear power plant sites. NOAA Tech. Memo. OAR PMEL-136, Pacific Marine Environmental Laboratory, Seattle, WA, 125 pp. + appendices on CD.
  • Goto K, Hashimoto K, Sugawara D, Yanagisawa H, Abe T. 2014. Spatial thickness variability of the 2011 Tohoku-oki tsunami deposits along the coastline of Sendai Bay. Marine Geology. 358:38–48. doi: 10.1016/j.margeo.2013.12.015
  • Goto K, Okada K, Imamura F. 2009. Characteristics and hydrodynamics of boulders transported by storm waves at Kudaka island, Japan. Marine Geology. 262:14–24. doi: 10.1016/j.margeo.2009.03.001
  • Goto K, Shinozaki T, Minoura K, Okada K, Sugawara D, Imamura F. 2010. Distribution of boulders at Miyara Bay of Ishigaki island, Japan: a flow characteristic indicator of tsunami and storm waves. Island Arc. 19:412–426. doi: 10.1111/j.1440-1738.2010.00721.x
  • Haynes WM. 2017. CRC handbook of chemistry and physics: a ready-reference book of chemical and physical data, 97th ed. Boca Raton, FL: CRC Press.
  • Hearty P. 1997. Boulder deposits from large waves during the Last Interglaciation on North Eleuthera Island, Bahamas. Quaternary Research. 48:326–338. doi: 10.1006/qres.1997.1926
  • Hedley C, Taylor TG. 1907. Coral reefs of the Great Barrier, Queensland. a study of their structure, life-distribution, and relation to mainland physiography. Australia and New Zealand Association for the Advancement of Science. Rep. Meeting 1907. Proceedings of Section C, p. 397–413.
  • Houtz RE. 1962. 1953 Suva earthquake and tsunami. Bulletin of the Seismological Society of America. 52:1–12.
  • Jaffe B, Goto K, Sugawara D, Gelfenbaum G, La Selle S. 2016. Uncertainty in tsunami sediment transport modelling. Journal of Disaster Research. 11:647–661. doi: 10.20965/jdr.2016.p0647
  • Jiffry H, Pilakoutas K, Garcia R. 2015. Structural assessment of low-rise reinforced concrete frames under tsunami loads. International Journal of Civil and Environmental Engineering. 9:171–176.
  • Jiongxin X. 2004. Hyperconcentrated flows in the slope-channel systems in gullied hilly areas on the Loess Plateau, China. Geografiska Annaler: Series A. Physical Geography. 86(4):349–366.
  • Kain CL, Gomez C, Moghaddam AE. 2012. Comment on ‘Reassessment of hydrodynamic equations: minimum flow velocity to initiate boulder transport by high energy events (storms, tsunamis), by N.A.K. Nandasena, R. Paris and N. Tanaka [Marine Geology 281, 70–84]. Marine Geology. 319–322:75–76. doi: 10.1016/j.margeo.2011.08.008
  • Kennedy AB, Mori N, Yasuda T, Shimozono T, Tomiczek T, Donahue A, Shimura T, Imai Y. 2017. Extreme block and boulder transport along a cliffed coastline (Calicoan Island, Philippines) during Super Typhoon Haiyan. Marine Geology. 383:65–77. doi: 10.1016/j.margeo.2016.11.004
  • Larsen B, Fuhrman D, Baykal C, Sumer B. 2017. Tsunami-induced scour around monopile foundations. Coastal Engineering. 129:36–49. doi: 10.1016/j.coastaleng.2017.08.002
  • Lau AYA, Terry JP, Switzer A, Pyle J. 2015. Advantages of beachrock slabs for interpreting high-energy wave transport: evidence from Ludao Island in southeastern Taiwan. Geomorphology. 228:263–274. doi: 10.1016/j.geomorph.2014.09.010
  • Lau AYA, Terry JP, Ziegler A, Pratap A, Harris D. 2018. Boulder emplacement and remobilisation by cyclone and submarine landslide tsunami waves near Suva City, Fiji. Sedimentary Geology. 364:242–257. doi: 10.1016/j.sedgeo.2017.12.017
  • Lau AYA, Terry JP, Ziegler AD, Switzer AD, Lee Y, Etienne S. 2016. Understanding the history of extreme wave events in the Tuamotu Archipelago of French Polynesia from large carbonate boulders on Makemo atoll, with implications for future threats in the central South Pacific. Marine Geology. 380:174–190. doi: 10.1016/j.margeo.2016.04.018
  • Lavigne F, Thouret J-C. 2002. Sediment transportation and deposition by rain-triggered lahars at Merapi Volcano, Central Java, Indonesia. Geomorphology. 49:45–69. doi: 10.1016/S0169-555X(02)00160-5
  • Li L, Huang Z. 2013. Modeling the change of beach profile under tsunami waves: a comparison of selected sediment transport models. Journal of Earthquake and Tsunami. 7(1):1350001. doi: 10.1142/S1793431113500012
  • May SM, Engel M, Brill D, Cuadra C, Lagmay AMF, Santiago J, Suarez JK, Reyes M, Brückner H. 2015. Block and boulder transport in Eastern Samar (Philippines) during Supertyphoon Haiyan. Earth Surface Dynamics. 3:543–558. doi: 10.5194/esurf-3-543-2015
  • Munson BR, Okishi TH, Huebsch WW, Rothmayer AP. 2013. Fluid mechanics, 7th edition. Singapore: Wiley-Blackwell.
  • Nandasena NAK, Paris R, Tanaka N. 2011. Reassessment of hydrodynamic equations to initiate boulder transport by high energy events (storms, tsunamis). Marine Geology. 281:70–84. doi: 10.1016/j.margeo.2011.02.005
  • Nott J. 1997. Extremely high wave deposits inside the great barrier reef, Australia: determining the cause – tsunami or tropical cyclone. Marine Geology. 141:193–207. doi: 10.1016/S0025-3227(97)00063-7
  • Nott J. 2003. Waves, coastal boulders and the importance of the pre-transport setting. Earth and Planetary Science Letters. 210:269–276. doi: 10.1016/S0012-821X(03)00104-3
  • Pierson T, Scott K. 1985. Downstream dilution of a lahar: transition from debris flow to hyperconcentrated streamflow. Water Resources Research. 21:1511–1524. doi: 10.1029/WR021i010p01511
  • Salzmann L, Green A. 2012. Boulder emplacement on a tectonically stable, wave-dominated coastline, Mission Rocks, northern KwaZulu-Natal, South Africa. Marine Geology. 323–325:95–106. doi: 10.1016/j.margeo.2012.07.001
  • Sassa S, Takagawa T. 2019. Liquefied gravity flow-induced tsunami: first evidence and comparison from the 2018 Indonesia Sulawesi earthquake and tsunami disasters. Landslides. 16:195–200. doi: 10.1007/s10346-018-1114-x
  • Scicchitano G, Monaco C, Tortorici L. 2007. Large boulder deposits by tsunami waves along the Ionian coast of south-eastern Sicily (Italy). Marine Geology. 238:75–91. doi: 10.1016/j.margeo.2006.12.005
  • Shah-Hosseini M, Saleem A, Mahmoud AMA, Morhange C. 2016. Coastal boulder deposits attesting to large wave impacts on the Mediterranean coast of Egypt. Natural Hazards. 83:849–865. doi: 10.1007/s11069-016-2349-2
  • Sheremet A, Mehta AJ, Kaihatu JM. 2005a. Wave sediment interaction on a muddy shelf. Proceedings of the 5th International Symposium on Ocean Wave Measurement and Analysis, WAVES 2005, 3–7 July 2005, Madrid, Spain.
  • Sheremet A, Mehta AJ, Liu B, Stone GW. 2005b. Wave-sediment interaction on a muddy inner shelf during Hurricane Claudette. Estuarine, Coastal and Shelf Science. 63:225–233. doi: 10.1016/j.ecss.2004.11.017
  • Soulsby R. 1997. Dynamics of marine sands. A manual for practical applications. London: Thomas Telford. 249pp.
  • Tappin DR, Watts P, Grilli ST. 2008. The Papua New Guinea tsunami of 17 July 1998: anatomy of a catastrophic event. Natural Hazards and Earth System Sciences. 8:243–266. doi: 10.5194/nhess-8-243-2008
  • Tehranirad B, Kirby JT, Fengyan S. 2016. Does morphological adjustment during tsunami inundation increase levels of hazard? Center for Applied Coastal Research, University of Delaware, Newark, Delaware, US. Research Report No. CACR-16-02, 167pp.
  • Terry JP, Jankaew K, Dunne K. 2015. Coastal vulnerability to typhoon inundation in the Bay of Bangkok, Thailand? Evidence from carbonate boulder deposits on Ko Larn island. Estuarine, Coastal and Shelf Science. 165:261–269. doi: 10.1016/j.ecss.2015.05.028
  • Terry JP, Lau AY, Etienne S. 2013. Reef-platform coral boulders: evidence for high-energy marine inundation events on tropical coastlines. New York, NY: Springer.
  • Terry JP, Oliver GJH, Friess DA. 2016. Ancient high-energy storm boulder deposits on Ko Samui, Thailand, and their significance for identifying coastal hazard risk. Palaeogeography, Palaeoclimatology, Palaeoecology. 454:282–293. doi: 10.1016/j.palaeo.2016.04.046
  • Yamada M, Fujino S, Goto K. 2014. Deposition of sediments of diverse sizes by the 2011 Tohoku-oki tsunami at Miyako City, Japan. Marine Geology. 358:67–78. doi: 10.1016/j.margeo.2014.05.019
  • Yamashita K, Sugawara D, Takahashi T, Imamura F, Saito Y, Imato Y, Kai T, Uehara H, Kato T, Nkata K, et al. 2016. Numerical simulations of large-scale sediment transport caused by the 2011 Tohoku earthquake tsunami in Hirota Bay, southern Sanriku coast. Coastal Engineering Journal. 58:1640015-1–1640015-28. doi: 10.1142/S0578563416400155
  • Yoshii T, Tanaka S, Matsuyama M. 2018. Tsunami inundation, sediment transport, and deposition process of tsunami deposits on coastal lowland inferred from the Tsunami Sand Transport Laboratory Experiment (TSTLE). Marine Geology. 400:107–118. doi: 10.1016/j.margeo.2018.03.007
  • Yu K, Zhao J, Roff G, Lybolt M, Feng Y, Clark T, Li S. 2012. High-precision U-series ages of transported coral blocks on Heron Reef (southern Great Barrier Reef) and storm activity during the past century. Palaeogeography, Palaeoclimatology, Palaeoecology. 337–338:26–36.
  • Yu K, Zhao J, Shi Q, Meng Q. 2009. Reconstruction of storm/tsunami records over the last 4000 years using transported coral blocks and lagoon sediments in the southern South China Sea. Quaternary International. 195:128–137. doi: 10.1016/j.quaint.2008.05.004

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.